15.12.2012 Views

Journal of Cereal Science

Journal of Cereal Science

Journal of Cereal Science

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

1.<br />

Volume 49, Issue 1, Pages 1-162 (January 2009)<br />

IFC - Editorial Board<br />

Page IFC<br />

Rapid Communication<br />

2.<br />

Degradation <strong>of</strong> cereal beta-glucan by ascorbic acid induced oxygen radicals<br />

Pages 1-3<br />

Reetta Kivelä, Fred Gates, Tuula Sontag-Strohm<br />

Research Papers<br />

3.<br />

4.<br />

5.<br />

6.<br />

7.<br />

8.<br />

9.<br />

Variation in Granule Bound Starch Synthase I (GBSSI) loci amongst Australian wild cereal<br />

relatives (Poaceae)<br />

Pages 4-11<br />

F.M. Shapter, P. Eggler, L.S. Lee, R.J. Henry<br />

Effect <strong>of</strong> high temperature on albumin and globulin accumulation in the endosperm<br />

proteome <strong>of</strong> the developing wheat grain<br />

Pages 12-23<br />

William J. Hurkman, William H. Vensel, Charlene K. Tanaka, Linda Whitehand, Susan B. Altenbach<br />

Accumulation <strong>of</strong> mixed linkage (1 → 3) (1 → 4)-β-d-glucan during grain filling in barley: A<br />

vibrational spectroscopy study<br />

Pages 24-31<br />

Helene Fast Seefeldt, Andreas Blennow, Birthe Møller Jespersen, Bernd Wollenweber, Søren<br />

Balling Engelsen<br />

Mechanism <strong>of</strong> gas cell stabilization in bread making. I. The primary gluten–starch matrix<br />

Pages 32-40<br />

Baninder S. Sroan, Scott R. Bean, Finlay MacRitchie<br />

Mechanism <strong>of</strong> gas cell stabilization in breadmaking. II. The secondary liquid lamellae<br />

Pages 41-46<br />

Baninder S. Sroan, Finlay MacRitchie<br />

Expression <strong>of</strong> globulin-2, a member <strong>of</strong> the cupin superfamily <strong>of</strong> proteins with similarity to<br />

known food allergens, is increased under high temperature regimens during wheat grain<br />

development<br />

Pages 47-54<br />

Susan B. Altenbach, Charlene K. Tanaka, William J. Hurkman, William H. Vensel<br />

Biochemical markers: Efficient tools for the assessment <strong>of</strong> wheat grain tissue proportions in<br />

milling fractions<br />

Pages 55-64<br />

Youna Hemery, Valérie Lullien-Pellerin, Xavier Rouau, Joël Abecassis, Marie-Françoise Samson,<br />

Per Åman, Walter von Reding, Cäcilia Spoerndli, Cécile Barron


10.<br />

11.<br />

12.<br />

13.<br />

14.<br />

15.<br />

16.<br />

17.<br />

18.<br />

19.<br />

20.<br />

21.<br />

Lateral growth <strong>of</strong> a wheat dough disk under various growth conditions<br />

Pages 65-72<br />

Amy Penner, Leaelaf Hailemariam, Martin Okos, Osvaldo Campanella<br />

Digestibility <strong>of</strong> protein and starch from sorghum (Sorghum bicolor) is linked to biochemical<br />

and structural features <strong>of</strong> grain endosperm<br />

Pages 73-82<br />

Joshua H. Wong, Tsang Lau, Nick Cai, Jaswinder Singh, Jeffrey F. Pedersen, William H. Vensel,<br />

William J. Hurkman, Jeff D. Wilson, Peggy G. Lemaux, Bob B. Buchanan<br />

Fundamental study on protein changes taking place during malting <strong>of</strong> oats<br />

Pages 83-91<br />

Christina Klose, Beatus D. Schehl, Elke K. Arendt<br />

Effect <strong>of</strong> milling, pasta making and cooking on minerals in durum wheat<br />

Pages 92-97<br />

Francesco Cubadda, Federica Aureli, Andrea Raggi, Marina Carcea<br />

Starch granule size distribution <strong>of</strong> hard red winter and hard red spring wheat: Its effects on<br />

mixing and breadmaking quality<br />

Pages 98-105<br />

Seok-Ho Park, Jeff D. Wilson, Bradford W. Seabourn<br />

Total phenolics, flavonoids, antioxidant capacity in rice grain and their relations to grain<br />

color, size and weight<br />

Pages 106-111<br />

Yun Shen, Liang Jin, Peng Xiao, Yan Lu, Jinsong Bao<br />

Nucleotide polymorphisms in the waxy gene <strong>of</strong> NaN3-induced waxy rice mutants<br />

Pages 112-116<br />

Toong Long Jeng, Chang Sheng Wang, Tung Hai Tseng, Min Tze Wu, Jih Min Sung<br />

Amaranth (Amaranthus hypochondriacus) as an alternative crop for sustainable food<br />

production: Phenolic acids and flavonoids with potential impact on its nutraceutical quality<br />

Pages 117-121<br />

A.P. Barba de la Rosa, Inge S. Fomsgaard, Bente Laursen, Anne G. Mortensen, L. Olvera-<br />

Martínez, C. Silva-Sánchez, A. Mendoza-Herrera, J. González-Castañeda, A. De León-Rodríguez<br />

Relationship <strong>of</strong> milled grain percentages and flowering-related traits in rice<br />

Pages 122-127<br />

Rodante E. Tabien, Stanley Omar P.B. Samonte, Emmanuel R. Tiongco<br />

Texture, processing and organoleptic properties <strong>of</strong> chickpea-fortified spaghetti with insights<br />

to the underlying mechanisms <strong>of</strong> traditional durum pasta quality<br />

Pages 128-133<br />

Jennifer Ann Wood<br />

Impact <strong>of</strong> re-grinding on hydration properties and surface composition <strong>of</strong> wheat flour<br />

Pages 134-140<br />

M. Mohamad Saad, C. Gaiani, J. Scher, B. Cuq, J.J. Ehrhardt, S. Desobry<br />

Identification <strong>of</strong> novel haze-active beer proteins by proteome analysis<br />

Pages 141-147<br />

Takashi Iimure, Nami Nankaku, Megumi Watanabe-Sugimoto, Naohiko Hirota, Zhou Tiansu,<br />

Makoto Kihara, Katsuhiro Hayashi, Kazutoshi Ito, Kazuhiro Sato


22.<br />

23.<br />

Mixing properties and dough functionality <strong>of</strong> transgenic lines <strong>of</strong> a commercial wheat<br />

cultivar expressing the 1Ax1, 1Dx5 and 1Dy10 HMW glutenin subunit genes<br />

Pages 148-156<br />

Elena León, Santiago Marín, María J. Giménez, Fernando Piston, Marta Rodríguez-Quijano, Peter<br />

R. Shewry, Francisco Barro<br />

Wheat glutenin proteins assemble into a nanostructure with unusual structural features<br />

Pages 157-162<br />

Sarah H. Mackintosh, Susie J. Meade, Jackie P. Healy, Kevin H. Sutton, Nigel G. Larsen, Adam<br />

M. Squires, Juliet A. Gerrard


<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

Aims and Scope<br />

The <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> was established in 1983 to provide an international forum for the publication <strong>of</strong> original research papers <strong>of</strong> high standing covering all<br />

aspects <strong>of</strong> cereal science related to the functional and nutritional quality <strong>of</strong> cereal grains and their products.<br />

The journal also publishes concise and critical review articles appraising the status and future directions <strong>of</strong> specific areas <strong>of</strong> cereal science and short rapid communications<br />

that present news <strong>of</strong> important advances in research. The journal aims at topicality and at providing comprehensive coverage <strong>of</strong> progress in the field.<br />

Research areas include:<br />

Composition and analysis <strong>of</strong> cereal grains in relation to quality in end use<br />

Morphology, biochemistry, and biophysics <strong>of</strong> cereal grains relevant to functional and nutritional characteristics<br />

Structure and physicochemical properties <strong>of</strong> functionally and nutritionally important components <strong>of</strong> cereal grains such as polysaccharides, proteins, oils,<br />

enzymes, vitamins, and minerals<br />

Storage <strong>of</strong> cereal grains and derivatives and effects on nutritional and functional quality<br />

Genetics, agronomy, and pathology <strong>of</strong> cereal crops if there is a substantive relationship to end-use properties <strong>of</strong> cereal grains<br />

Functional and nutritional aspects <strong>of</strong> cereal-based foods and beverages, whether baked, fermented, or extruded<br />

Industrial products (e.g. starch derivatives, syrups, protein concentrates, and isolates) from cereal grains, and their technology<br />

J. Dexter<br />

Grain Research Laboratory<br />

Canadian Grain Commission<br />

Winnipeg, Canada<br />

Editorial Board<br />

N.-G. Asp<br />

University <strong>of</strong> Lund, Sweden<br />

D. E. Briggs<br />

University <strong>of</strong> Birmingham, UK<br />

G. B. Fincher<br />

Waite Agricultural Research<br />

Institute,<br />

University <strong>of</strong> Adelaide,<br />

Australia<br />

P. Colonna<br />

INRA, Laboratoire de Biochimie et Technologies<br />

des Glucides Nantes, France<br />

K. Denyer<br />

John Innes Centre, Norwich, UK<br />

P. J. Frazier<br />

Dalgety Food Technology Centre, Cambridge, UK<br />

R. A. Graybosch<br />

University <strong>of</strong> Nebraska, Lincoln, Nebraska, USA<br />

Zhonghu He<br />

International Maize and Wheat Improvement<br />

Center, Chinese Academy <strong>of</strong> Agricultural <strong>Science</strong>s,<br />

Beijing, China<br />

Founding Editors<br />

T. Galliard and J. D. Sch<strong>of</strong>ield<br />

Editor-in-Chief<br />

F. MacRitchie<br />

Grain <strong>Science</strong> and Industry Department<br />

Kansas State University<br />

Manhattan, Kansas, USA<br />

R. J. Hamer<br />

Wageningen Centre for<br />

Food <strong>Science</strong>s<br />

Wageningen,<br />

The Netherlands<br />

G. E. Inglett<br />

USDA, National Center for Agricultural<br />

Utilization Research, Peoria,<br />

Illinois, USA<br />

B. O. Juliano<br />

Philippines Rice Research Institute, Laguna,<br />

Philippines<br />

J. Kokini<br />

Cook College, Rutgers University,<br />

New Brunswick,<br />

New Jersey, USA<br />

P. Linko<br />

Helsinki University <strong>of</strong> Technology,<br />

Espoo, Finland<br />

A. W. MacGregor<br />

Livingston, Scotland, UK<br />

Editors<br />

Reviews Editor<br />

P. R. Shewry<br />

Rothamsted Research<br />

Harpenden, UK<br />

D. Lafiandra<br />

Dipartimento di Agrobiologia<br />

e Agrochimica<br />

University <strong>of</strong> Tuscia<br />

Viterbo, Italy<br />

J. R. N. Taylor<br />

University <strong>of</strong> Pretoria,<br />

South Africa<br />

Y. Popineau<br />

INRA, Unite de Recherche sur les<br />

Proteines Vegetales et leurs Interactions,<br />

Nantes, France<br />

K. Poutanen<br />

VTT Biotechnology,<br />

Espoo, Finland<br />

S. O. Serna Saldivar<br />

ITESM, Monterrey, Mexico<br />

B. A. Stone<br />

Department <strong>of</strong> Biochemistry<br />

La Trobe University<br />

Bundoora, Victoria, Australia<br />

B. Svensson<br />

BioCentrum-DTU,<br />

The Technical University <strong>of</strong><br />

Denmark, Kgs.<br />

Lyngby, Denmark<br />

Author enquiries<br />

For enquiries relating to the submission <strong>of</strong> articles (including electronic submission where available) please visit this journal’s homepage at http://www.elsevier.com/locate/jcs.<br />

You can track accepted articles at http://www.elsevier.com/trackarticle and set up e-mail alerts to inform you <strong>of</strong> when an article’s status has changed. Also accessible from<br />

here is information on copyright, frequently asked questions and more.<br />

Contact details for questions arising after acceptance <strong>of</strong> an article, especially those relating to pro<strong>of</strong>s, will be provided by the publisher.<br />

For a full and complete Guide for Authors, please go to: http://www.elsevier.com/locate/issn/0733-5210/authorinstructions


Rapid Communication<br />

Degradation <strong>of</strong> cereal beta-glucan by ascorbic acid induced oxygen radicals<br />

Reetta Kivelä *, Fred Gates 1 , Tuula Sontag-Strohm<br />

Department <strong>of</strong> Food Technology, University <strong>of</strong> Helsinki, Agnes Sjöberg Street 2, P.O. Box 66, FIN-00014 Helsinki, Finland<br />

article info<br />

Article history:<br />

Received 24 April 2008<br />

Received in revised form 15 August 2008<br />

Accepted 5 September 2008<br />

Keywords:<br />

Beta-glucan<br />

Degradation<br />

Oxygen radicals<br />

Ascorbic acid<br />

abstract<br />

We report degradation <strong>of</strong> cereal beta-glucan in aqueous solutions<br />

as a result <strong>of</strong> a treatment with ascorbic acid and its oxidation<br />

product dehydroascorbic acid. It was found that (1 / 3), (1 / 4)beta-D-glucan<br />

was depolymerised when treated with ascorbic acid<br />

or dehydroascorbic acid and FeSO4 in the absence <strong>of</strong> enzymatic<br />

activity or conditions that could cause acid hydrolysis. The degradation<br />

was inhibited by both glucose addition and by displacing<br />

oxygen with nitrogen. Non-enzymatic, ascorbate induced degradation<br />

<strong>of</strong> cell-wall polysaccharides is established in vitro and suggested<br />

to occur in plant cells during its elongation phase and<br />

ripening (Fry, 1998; Fry et al., 2002; Schopfer, 2001). However, this<br />

mechanism appears to be neglected in food technology in spite <strong>of</strong><br />

its potentially crucial role in reducing the physiological efficacy, and<br />

hence the validity <strong>of</strong> health claims, <strong>of</strong> high beta-glucan foods as<br />

well as influencing properties such as rheology and physical<br />

stability.<br />

Macromolecules are typically susceptible to degradation<br />

induced by reactive oxygen species (ROS). Particularly OH-radicals,<br />

one specific species <strong>of</strong> ROS, have been established to scission<br />

polysaccharides (Fry, 1998). One way to produce OH-radicals nonenzymatically<br />

is a Fenton type reaction (Reaction (1)). The key<br />

feature <strong>of</strong> Fenton chemistry is the catalytic nature <strong>of</strong> reduced<br />

Abbreviations: AH2, ascorbic acid; A, dehydroascorbic acid; BBG, beta-glucan<br />

isolate from barley; HPSEC, high performance size exclusion chromatography; Mw,<br />

weight average molecular mass; OBG, beta-glucan isolate from oat; ROS, reactive<br />

oxygen species, i.e. all kinds <strong>of</strong> chemically reactive molecules derived from ground<br />

state dioxygen ( 3 O2).<br />

* Corresponding author. Tel.: þ358 9 19158540; fax: þ358 9 19158460.<br />

E-mail address: reetta.kivela@helsinki.fi (R. Kivelä).<br />

1<br />

Present address: Department <strong>of</strong> Physics, University <strong>of</strong> Helsinki, P.O. Box 64,<br />

FIN-00014 Helsinki, Finland.<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.09.003<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 1–3<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

Degradation <strong>of</strong> cereal beta-glucan is usually attributed to enzymes or acid hydrolysis. However, there is<br />

evidence that polysaccharides are also susceptible to OH-radical induced depolymerisation, and that<br />

these radicals can be produced in cereal food systems. The role <strong>of</strong> Fenton type oxidation was demonstrated<br />

in pure beta-glucan solution (0.6%). An addition <strong>of</strong> ascorbic acid (10 mM) or its oxidation product,<br />

dehydroascorbic acid, in the presence <strong>of</strong> iron sulphate resulted in a significant decrease <strong>of</strong> the solution<br />

viscosity and molecular degradation <strong>of</strong> beta-glucan. The viscosity decrease was inhibited by introducing<br />

a OH-scavenger (glucose) in the solution or limiting oxygen level in the sample solution. This demonstrates<br />

the role <strong>of</strong> OH-radicals in beta-glucan scission and suggests oxidative cleavage to be a potential<br />

threat for the stability <strong>of</strong> beta-glucan in certain fibre enriched products.<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

metals (Fe 2þ ,Cu þ ,Zn þ ), which can produce aggressive OH-radicals<br />

from the modestly reactive H2O2 (Wardman and Candeias, 1996). A<br />

reducing agent is needed to reduce the catalysing metal to the<br />

active state and the dissolved oxygen to H2O2, thus it initiates the<br />

Fenton type reaction. All these reagents (dissolved oxygen, transition<br />

metal and compounds with reduction potential) are typically<br />

present in the aqueous phases <strong>of</strong> cereal foods, which also contains<br />

the soluble fibres (Slavin et al., 1999). In addition, beta-glucan is<br />

known to form complexes with metals (Platt and Clydesdale, 1984),<br />

which makes it a specific target for OH-radical induced degradation,<br />

as the short living radicals are produced in immediate vicinity<br />

<strong>of</strong> beta-glucan molecules (Chevion, 1988).<br />

Cu þ =Fe 2þ þ H 2O 2/ $ OH þ OH þ Cu 2þ =Fe 3þ<br />

Ascorbic acid (AH2) is an excellent reducing agent. It readily<br />

undergoes a two-step oxidation process forming dehydroascorbic<br />

acid (A) and ascorbate radical ( AH) as an intermediate. The<br />

ascorbate radical is a relatively stable free radical that has a capability<br />

<strong>of</strong> inhibiting oxidation chain reactions and thus can act as<br />

a secondary antioxidant. However, ascorbic acid is known to have<br />

a dual nature: in certain conditions it behaves as a pro-oxidant<br />

instead <strong>of</strong> antioxidant initiating the Fenton reaction due to the<br />

reducing capacity (Reactions (2a) and (2b)). The dual effect is<br />

considered to be concentration dependent, so that the rate <strong>of</strong><br />

antioxidant chain reactions is lowered and the pro-oxidant nature<br />

dominates at lower concentrations (Buettner and Jurkiewicz, 1996;<br />

Wardman and Candeias, 1996). Ascorbic acid is a common food<br />

constituent and commonly added as an antioxidant; it is also used<br />

as a flour improver which makes Fenton chemistry even more<br />

interesting in cereal foods.<br />

(1)


2<br />

AH 2 þ 2Cu 2þ =Fe 3þ /A þ 2H þ þ 2Cu þ =Fe 2þ<br />

AH 2 þ O 2/A þ H 2O 2<br />

(2a)<br />

(2b)<br />

Degradation <strong>of</strong> cereal beta-glucan by the chemical radical<br />

reactions may explain some <strong>of</strong> the controversial results obtained in<br />

various studies. It is not uncommon that a reduction in viscosity <strong>of</strong><br />

beta-glucan (i.e. molecular size reduction) or molecular degradation<br />

is observed in food systems such as bakery products and<br />

beverages (Åman et al., 2004; Beer et al., 1997). These changes have<br />

generally been attributed to enzymatic or acid hydrolysis or to high<br />

shear forces. However, in some cases enzymes seem to have been<br />

inactivated in heating processes or the conditions are not suitable<br />

for acid hydrolysis to occur, which requires a very low pH (1–2) and<br />

simultaneous heating (Johansson et al., 2006). In addition, there is<br />

no clear evidence to link the acid hydrolysis <strong>of</strong> beta-glucan to the<br />

organic acids present in foods.<br />

Technologically the degradation <strong>of</strong> beta-glucan is serious. In<br />

aqueous systems the breakdown <strong>of</strong> the polymer easily leads to<br />

product structural breakdown and instability. The degradation <strong>of</strong><br />

beta-glucan is also nutritionally crucial, since the health benefits <strong>of</strong><br />

beta-glucan are closely related to its rheology. In particular, it is<br />

well established that the ability <strong>of</strong> beta-glucan to attenuate postprandial<br />

blood glucose and insulin levels is related to the viscosity<br />

and molecular weight <strong>of</strong> beta-glucan (Wood et al., 2000). The<br />

structure <strong>of</strong> food matrices, in addition to the rheology, is proposed<br />

to affect cholesterol lowering <strong>of</strong> beta-glucan and liquid beta-glucan<br />

rich foods have been established to be especially effective (Kerckh<strong>of</strong>fs<br />

et al., 2003; Naumann et al., 2006). There is thus an<br />

increasing need to properly understand the threats for instability <strong>of</strong><br />

native beta-glucan in aqueous environments.<br />

In the present study effects <strong>of</strong> ascorbic acid (10 mM) and<br />

dehydroascorbic acid (10 mM) were examined in pure 0.6% betaglucan<br />

solutions when added ionic iron (0.01 mM FeSO4$H2O) was<br />

present. Also the roles <strong>of</strong> glucose (1 M, 18 w/w-%) and oxygen were<br />

studied. The beta-glucan solutions were prepared from high<br />

viscosity barley beta-glucan (BBG, Megazyme, purity >97%) or<br />

high viscosity oat beta-glucan (OBG, Megazyme, purity >97%) by<br />

wetting dry beta-glucan with ethanol and dispersing the blend in<br />

water. The dispersion was first boiled under stirring for 10 min and<br />

then continued for 3 h at 80 C before adjusting the volume in<br />

a volumetric flask. Viscosities <strong>of</strong> the stock solutions were<br />

controlled up to three weeks and were observed to maintain<br />

constant viscosity. Sodium azide (0.02 w/w-%) was added to<br />

prevent microbial enzyme activity. At this concentration, sodium<br />

azide was not shown to affect the studied reactions. All the<br />

reagents were added as solid compounds to the stock solution. The<br />

pH was adjusted to 4.8 0.1 to represent acidic conditions <strong>of</strong><br />

fermented and beverage-like foods and to enable ascorbic acid to<br />

exist in dissociated form (pKaAH2 ¼ 4.2) and the sample volumes<br />

were adjusted after pH adjustment. Viscosity measurements and<br />

oscillation tests were performed using a rheometer (Thermo-<br />

Haake RheoStress 600, Thermo Electron GmbH, Dreieich, Germany).<br />

Flow curves were obtained over a shear rate range <strong>of</strong> 0.3–<br />

300–0.3 s 1 using a double gap geometry (DG41). The mean<br />

apparent viscosities <strong>of</strong> three independent measurements<br />

(mean SD) at a shear rate <strong>of</strong> 10 s 1 as a function <strong>of</strong> time are<br />

shown. Dynamic rheological behaviour was characterised by<br />

a cone and plate geometry (35 mm/2 ) using a frequency sweep <strong>of</strong><br />

0.01–10 Hz and a stress <strong>of</strong> 0.02 Pa. All the rheological experiments<br />

were performed at þ10 0.1 C. Prior to measurement the<br />

samples were stored at þ6 0.1 C. The molecular weight <strong>of</strong> betaglucan<br />

was analysed by HPSEC with calc<strong>of</strong>luor-fluoresence and<br />

dual angle light scattering detection in 0.1 N sodium hydroxide<br />

solution (Suortti, 1993).<br />

R. Kivelä et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 1–3<br />

The rheological studies showed that 10 mM AH2 added together<br />

with 0.01 mM FeSO4$H2O caused an approximately 50% drop in<br />

viscosity <strong>of</strong> the barley beta-glucan (BBG) solution in 1 h, while the<br />

viscosity <strong>of</strong> BBG treated with FeSO4 alone remained unchanged<br />

(Fig. 1). The results <strong>of</strong> BBG solution alone are not shown, because<br />

the ionic iron treatment did not affect it. Consistent results were<br />

obtained with the oat beta-glucan solution (OBG, results not<br />

shown). The viscosity behaviour <strong>of</strong> these solutions changed from<br />

shear thinning to Newtonian in 2 h suggesting that the critical<br />

concentration, at which the entanglements between molecules or<br />

aggregates are lost, had been reached as a result <strong>of</strong> molecular size<br />

reduction (Fig. 2). HPSEC-studies showed that this reduction in<br />

molecule size was depolymerisation instead <strong>of</strong> other changes in<br />

physicochemical properties, since the weight average molecular<br />

mass (Mw) <strong>of</strong> beta-glucan was reduced from 520 ( 5) to 35<br />

( 2) 10 3 g/mol as a result <strong>of</strong> the ascorbic acid and iron treatment.<br />

The effect <strong>of</strong> the third reagent, oxygen, was demonstrated by<br />

limiting the amount <strong>of</strong> dissolved oxygen with nitrogen supplement<br />

during sample preparation and testing. After 4 h <strong>of</strong> N2-treatment,<br />

air was mixed into the nitrogen treated ascorbic acid-BBG sample in<br />

the measuring cap <strong>of</strong> the rheometer and this caused a drop in<br />

viscosity (Fig. 1). Thus all the reagents <strong>of</strong> Fenton chemistry, i.e.<br />

transition metal, dissolved oxygen and a reducing agent (Reactions<br />

(1)–(2b)), were shown to be needed for the depolymerisation <strong>of</strong><br />

beta-glucan in aqueous solution.<br />

Dehydroascorbic acid (10 mM) caused a 30% viscosity reduction<br />

when added with ionic iron to the beta-glucan solution (Fig. 1). The<br />

total decrease during the measuring time was appr. 50%, while<br />

ascorbic acid decreased the viscosity over 90%. The lower viscosity<br />

reduction, i.e. the rate <strong>of</strong> oxidative cleavage compared to the action<br />

<strong>of</strong> ascorbic acid agrees with the results <strong>of</strong> Fry (1998) with xyloglucan.<br />

Dehydroascorbic acid is likely to occur in food systems,<br />

since ascorbic acid undergoes a reversible oxidation reaction<br />

induced by light, heat and metal first forming semidehydroascorbic<br />

Apparent viscosity at10 1/s (mPas)<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15<br />

Time / h<br />

+1M glucose<br />

Control<br />

+10 mM AH2+<br />

1M glucose<br />

+10 mM A<br />

+10 mM AH2,<br />

limited oxygen<br />

+10 mM AH2<br />

Fig. 1. The effect <strong>of</strong> ascorbic acid (AH2) and dehydroascorbic acid (A) with ionic iron on<br />

the viscosity <strong>of</strong> beta-glucan (control, BBG, 0.6%) and the inhibition effects <strong>of</strong> oxygen<br />

limitation and glucose addition (1 M). The arrow shows the point <strong>of</strong> air introduction<br />

into the nitrogen treated sample. All the samples contained 0.01 mM FeSO4$7H2O.


Viscosity Pas<br />

1<br />

0,1<br />

Control<br />

0.5 h<br />

1 h<br />

2 h<br />

0,01<br />

0,001<br />

4 h<br />

6 h<br />

8 h<br />

10 h<br />

18 h<br />

0,1 1 10<br />

Shear rate 1/s<br />

100 1000<br />

Fig. 2. Flow curves <strong>of</strong> ascorbic acid (10 mM) treated beta-glucan solution (0.6%, BBG)<br />

from 0 h to 18 h and the control at 18 h. Both samples contained 0.01 mM FeSO4$7H2O.<br />

(AH ) acid and further dehydroascorbic acid (A). It has been<br />

reported that, like ascorbic acid, dehydroascorbic acid can also keep<br />

metals such as copper in their reduced form, but it cannot reduce<br />

oxygen to H2O2 (Fry, 1998). Yet as endogenic H2O2 usually occurs in<br />

water solutions and especially in natural solutions such as food<br />

systems, dehydroascorbic acid may also participate in the oxidative<br />

cleavage <strong>of</strong> beta-glucan.<br />

Glucose efficiently inhibited the ascorbic acid induced viscosity<br />

reduction <strong>of</strong> BBG solution (Fig. 1). Glucose was introduced to the<br />

study as a competitive OH-radical scavenger (Buxton et al., 1988). In<br />

the solutions with high concentration <strong>of</strong> glucose (1 M, appr. 20 w/<br />

w-%), the majority <strong>of</strong> indiscriminant OH-radicals will react with the<br />

glucose rather than with the high molecular weight polysaccharide<br />

(beta-glucan, 0.6 w/w-%). After the reaction with OH, glucose itself<br />

becomes a relatively stable free radical. In addition, the glucose<br />

radical may regenerate itself with a hydrogen atom from the<br />

solution, thus enhancing its stability and efficiency as a scavenger<br />

(Morelli et al., 2003). The viscosity <strong>of</strong> glucose–BBG solution was 17%<br />

higher than the viscosity <strong>of</strong> BBG solution alone, which is in accordance<br />

with that reported before (Autio et al., 1987), and it decreased<br />

20% due to the ascorbic acid treatment (Fig. 1). However, the<br />

viscosity <strong>of</strong> ascorbic acid–glucose–BBG solution remained<br />

unchanged as a function <strong>of</strong> time indicating that the proposed OHinduced<br />

degradation <strong>of</strong> beta-glucan was inhibited by glucose<br />

addition.<br />

The flow curves <strong>of</strong> ascorbic acid treated BBG solution showed<br />

hysteresis and noise after 5 h (Fig. 2). This may both indicate novel<br />

networks and particle formation. Gel formation was observed as<br />

increased turbidity in the relatively dilute ascorbic acid–BBG<br />

solution (0.6%) after four days storage, while the control BBG<br />

solution stayed clear. The gelation was shown by dynamic rheology,<br />

the elastic modulus (G 0 ) being nearly frequency-independent (ca.<br />

10 Pa at 1 Hz) and the loss modulus (G 00 ) being low (appr. 1 Pa at<br />

1 Hz) at frequencies 1–10 Hz. It is known that gelation rate as well<br />

as gel strength increase with decreasing molecular weight so that<br />

molecules with an Mw above 250 000 g/mol do not practically form<br />

gels (Lazaridou et al., 2003; Tosh et al., 2004). Such oxidative<br />

cleavage may bring new aspects for the phenomenon <strong>of</strong> gelation<br />

during food processing, since it can occur in enzyme and acid<br />

hydrolysis free conditions and the cleavage may result in different<br />

compounds as found in hydrolysis (Fry, 1998; Fry et al., 2002).<br />

In conclusion, it was found that ROS represents a threat to the<br />

stability <strong>of</strong> cereal beta-glucan in an aqueous environment, and that<br />

Fenton chemistry results in depolymerisation <strong>of</strong> beta-glucan. All<br />

R. Kivelä et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 1–3 3<br />

the reagents, i.e. a reducing agent (ascorbic acid), transition metal<br />

(iron) and oxygen, were shown to have a role in the depolymerisation<br />

<strong>of</strong> beta-glucan. The viscosity loss was inhibited by an oxygen<br />

radical scavenger (glucose), which supports the role <strong>of</strong> OH-radicals<br />

in the non-enzymatic cleavage <strong>of</strong> cereal beta-glucan. Common food<br />

additives as well as the minor compounds <strong>of</strong> the soluble fibre<br />

phases may be harmful for the stability <strong>of</strong> aqueous beta-glucan<br />

enriched foods. In addition, their role in the research <strong>of</strong> health<br />

benefits <strong>of</strong> beta-glucan in different processed foods needs proper<br />

understanding.<br />

Acknowledgements<br />

The authors thank Dr. Tapani Suortti (VTT, Finland) for performing<br />

the molecular weight measurement and data analysis and<br />

Pr<strong>of</strong>essor Hannu Salovaara and Dr. Laura Nyström for their important<br />

support in the process.<br />

References<br />

Autio, K., Myllymäki, O., Mälkki, Y., 1987. Flow properties <strong>of</strong> solutions <strong>of</strong> oat bglucans.<br />

<strong>Journal</strong> <strong>of</strong> Food <strong>Science</strong> 52, 1364–1366.<br />

Beer, M.U., Wood, P.J., Weisz, J., Fillion, N., 1997. Effect <strong>of</strong> cooking and storage on the<br />

amount and molecular weight <strong>of</strong> (1 / 3), (1 / 4)-b-D-glucan extracted from<br />

oat products by an in vitro digestion system. <strong>Cereal</strong> Chemistry 74, 705–709.<br />

Buettner, G.R., Jurkiewicz, B.A., 1996. Catalytic metals, ascorbate and free radicals:<br />

combinations to avoid. Radiation Research 145, 532–541.<br />

Buxton, G.V., Greenstock, C.L., Helman, W.P., Ross, A.B., 1988. Critical review <strong>of</strong> rate<br />

constants for reactions <strong>of</strong> hydrated electrons, hydrogen atoms and hydroxyl<br />

radicals ( OH/ O ) in aqueous solution. <strong>Journal</strong> <strong>of</strong> Physical and Chemical<br />

Reference Data 17, 513–769.<br />

Chevion, M., 1988. A site-specific mechanism for free radical induced biological<br />

damage: the essential role <strong>of</strong> redox-active transition metals. Free Radical<br />

Biology 5, 27–37.<br />

Fry, S.C., 1998. Oxidative scission <strong>of</strong> plant cell wall polysaccharides by ascorbateinduced<br />

hydroxyl radicals. Biochemical <strong>Journal</strong> 332, 507–515.<br />

Fry, S.C., Miller, J.G., Dumville, J.C., 2002. A proposed role for copper ions in cell wall<br />

loosening. Plant and Soil 247, 57–67.<br />

Johansson, L., Virkki, L., Anttila, H., Esselström, H., Tuomainen, P., Sontag-Strohm, T.,<br />

2006. Hydrolysis <strong>of</strong> b-glucan. Food Chemistry 97, 71–79.<br />

Kerckh<strong>of</strong>fs, D., Hornstra, G., Mensink, R.P., 2003. Cholesterol-lowering effect <strong>of</strong> betaglucan<br />

from oat bran in mildly hypercholesterolemic subjects may decrease<br />

when beta-glucan is incorporated into bread and cookies. The American <strong>Journal</strong><br />

<strong>of</strong> Clinical Nutrition 78, 221–227.<br />

Lazaridou, A., Biliaderis, C.G., Izydorczyk, M.S., 2003. Molecular size effects on<br />

rheological properties <strong>of</strong> oat beta-glucans in solution and gels. Food Hydrocolloids<br />

17, 693–712.<br />

Morelli, R., Russo-Volpe, S., Bruno, N., Lo Scalzo, R., 2003. Fenton-dependent<br />

damage to carbohydrates: free radical scavenging activity <strong>of</strong> some simple<br />

sugars. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 51, 7418–7425.<br />

Naumann, E., van Rees, A.B., Onning, G., Oste, R., Wydra, M., Mensink, R.P., 2006.<br />

Beta-glucan incorporated into a fruit drink effectively lowers serum LDLcholesterol<br />

concentrations. The American <strong>Journal</strong> <strong>of</strong> Clinical Nutrition 83, 601–<br />

605.<br />

Platt, S.R., Clydesdale, F.M., 1984. Binding <strong>of</strong> iron by cellulose, lignin, sodium phytate<br />

and beta-glucan, alone and in combination, under simulated gastrointestinal<br />

pH conditions. <strong>Journal</strong> <strong>of</strong> Food <strong>Science</strong> 49, 531–535.<br />

Schopfer, P., 2001. Hydroxyl radical-induced cell-wall loosening in vitro and in vivo:<br />

implications for the control <strong>of</strong> elongation growth. The Plant <strong>Journal</strong> 28, 679–<br />

688.<br />

Slavin, J.L., Martini, M.C., Jacobs, D.R., Marquart, L., 1999. Plausible mechanisms for<br />

the protectiveness <strong>of</strong> whole grains. The American <strong>Journal</strong> <strong>of</strong> Clinical Nutrition<br />

70, 459–463.<br />

Suortti, T., 1993. Size-exclusion chromatographic determination <strong>of</strong> beta-glucan with<br />

postcolumn reaction detection. <strong>Journal</strong> <strong>of</strong> Chromatography 19, 105–110.<br />

Tosh, S.M., Wood, P.J., Wang, Q., Weisz, J., 2004. Structural characteristics and<br />

rheological properties <strong>of</strong> partially hydrolyzed oat b-glucan: the effects <strong>of</strong><br />

molecular weight and hydrolysis method. Carbohydrate Polymers 55, 425–436.<br />

Wardman, P., Candeias, L.P., 1996. Fenton chemistry: an introduction. Radiation<br />

Research 145, 523–531.<br />

Wood, P.J., Beer, M.U., Butler, G., 2000. Evaluation <strong>of</strong> role <strong>of</strong> concentration and<br />

molecular weight <strong>of</strong> oat beta-glucan in determining effect <strong>of</strong> viscosity on<br />

plasma glucose and insulin following an oral glucose load. The British <strong>Journal</strong> <strong>of</strong><br />

Nutrition 84, 19–23.<br />

Åman, P., Rimsten, L., Andersson, R., 2004. Molecular weight distribution <strong>of</strong> bglucan<br />

in oat-based foods. <strong>Cereal</strong> Chemistry 81, 356–360.


Variation in Granule Bound Starch Synthase I (GBSSI) loci amongst<br />

Australian wild cereal relatives (Poaceae)<br />

F.M. Shapter *, P. Eggler, L.S. Lee 1 , R.J. Henry<br />

Grain Foods CRC, Centre for Plant Conservation Genetics, Southern Cross University, PO Box 157, Lismore, NSW 2480, Australia<br />

article info<br />

Article history:<br />

Received 30 July 2007<br />

Received in revised form 12 November 2007<br />

Accepted 20 June 2008<br />

Keywords:<br />

Australian native grasses<br />

Starch<br />

Crop wild relatives<br />

<strong>Cereal</strong><br />

Phylogenetic<br />

1. Introduction<br />

abstract<br />

Granule Bound Starch Synthase I (GBSSI), encoded by the Waxy<br />

gene, is responsible for the accumulation <strong>of</strong> amylose during the<br />

development <strong>of</strong> starch granules in cereal endosperm. GBSSI has<br />

been reported with changing enzyme nomenclature over the last<br />

two decades. The most recent, EC 2.4.1.242, clearly delineates this<br />

enzyme from previously recorded enzymes EC 2.4.1.11 (Taira et al.,<br />

1995) and EC 2.4.1.21 (Baldwin, 2001; Demeke et al., 1999), as being<br />

able to use both UDP-glucose and ADP-glucose as a substrate for<br />

starch synthesis. Although GBSSI and II share the same enzyme<br />

nomenclature and approximately 69% sequence identity, suggesting<br />

that they are homologous, the genes encoding them are<br />

situated at independent loci. The gene encoding GBSSI is predominantly<br />

expressed in endosperm and pollen whereas GBSSII is<br />

expressed in the leaf, culm and pericarp (Vrinten and Nakamura,<br />

2000). Equally well reported is that cereal species with a null allele<br />

at all Waxy loci will produce an amylose free endosperm. The<br />

* Corresponding author. Tel.: þ61 2 6620 3466; fax: þ61 2 6622 2080.<br />

E-mail address: frances.shapter@scu.edu.au (F.M. Shapter).<br />

1 Present address: Centre for Tropical Crops and Biocommodities, Queensland<br />

University <strong>of</strong> Technology, GPO Box 2434, Brisbane, Queensland 4001, Australia.<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.06.013<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 4–11<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

A complex cascade <strong>of</strong> enzymes is responsible for the development <strong>of</strong> starch granules in grain endosperm.<br />

Granule Bound Starch Synthase I (GBSSI), encoded by the Waxy gene, is a key enzyme <strong>of</strong> starch synthesis<br />

and determines the accumulation <strong>of</strong> amylose in the starch granules. The complete genomic GBSSI<br />

sequence was ascertained for eight Australian cereal wild relatives (CWR) to determine diversity within<br />

the gene. A phylogeny derived from the coding sequence <strong>of</strong> the entire Waxy gene was compared to<br />

established phylogenetic relationships. Starch granule morphology observed in conjunction with this<br />

phylogeny suggests that small polygonal starch granules arranged as compound granules are the<br />

ancestral state, evolving subsequently to bimodal starch granules and to larger simple granules. Genomic<br />

sequence length varied within the species from 2800 to 3572 bp. Most variation occurred within the<br />

intron sequences, the largest insertion showing strong homology to a retrotransposon. One wild species<br />

was determined to have a deletion in the 3 0 -end <strong>of</strong> exon 1 resulting in a putatively non-functional allele.<br />

Alignment <strong>of</strong> the amino acid sequence showed strong homology throughout the central fragments <strong>of</strong> the<br />

gene but broad variation in the transit peptides. All putative functional alleles maintained the reported<br />

active sites for glycogen synthesis, though with variations in other highly conserved areas <strong>of</strong> the gene.<br />

These variations within the wild relatives <strong>of</strong> cultivated cereals may provide novel sources <strong>of</strong> genetic<br />

diversity for future cereal improvement programs.<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

amylose status <strong>of</strong> cereal starches affects their functional properties<br />

and therefore their end uses (Domon et al., 2002; Gaines et al.,<br />

2000; Kim et al., 2003) and a simple, rapid iodine test for the<br />

presence/absence <strong>of</strong> amylose in endosperm has been recently<br />

optimised (Pedersen et al., 2004).<br />

Within cultivated cereals Waxy mutants are common from<br />

induced (Nakamura et al., 1995) and naturally occurring origins<br />

(Domon et al., 2002), with mutations remaining conserved in the<br />

genome over generations. The concentration <strong>of</strong> Waxy mutants<br />

occurring in cereal wild relatives (CWR) is as yet undetermined. Of<br />

the waxy and low-amylose cereal species sequenced to date, it<br />

appears that there is no specific polymorphism across all the<br />

species which results in the waxy phenotype. Rather, any disruption<br />

to the production or regulation <strong>of</strong> this enzyme is critical in<br />

cereal species (Mangalika et al., 2003; Rahman et al., 2000; Shure<br />

et al., 1983; Yan et al., 2000). Within any one cereal species,<br />

multiple alleles <strong>of</strong> the GBSSI coding gene are common and are<br />

denoted by a subscript or chromosomal reference which is species<br />

specific (Mangalika et al., 2003; Pedersen et al., 2007). In the case <strong>of</strong><br />

polyploid species such as wheat, the number <strong>of</strong> functional GBSSI<br />

coding alleles is correlated with the concentration <strong>of</strong> amylose in the<br />

grain (Mangalika et al., 2003).<br />

The phylogenetic utility <strong>of</strong> the GBSSI encoding gene has been<br />

repeatedly reported (Evans and Campbell, 2002; Mason-Gamer


et al., 1998; Small et al., 2004). Phylogenetic relationships between<br />

species provide an understanding <strong>of</strong> the breadth <strong>of</strong> their genetic<br />

diversity and have been utilised in crop-breeding programs to<br />

determine relationships between cultivated species and their wild<br />

cereal relatives (Williams et al., 2006). Initial phylogenetic analyses<br />

across this subset <strong>of</strong> Poaceae utilising a small central fragment <strong>of</strong><br />

the gene was insufficient to resolve the basal bifurcations and<br />

interspecies relationships (data not shown). As a result, more<br />

sequence was captured to enhance phylogenetic analysis. In order<br />

to ensure that phylogenetic predictions are resolving species relationships<br />

and not gene evolution, analysing more than one gene<br />

concurrently is prudent (Small et al., 2004). However, functional<br />

genes evolve under selective pressure and therefore their evolutionary<br />

histories can be correlated with their functional evolution<br />

(Ochieng et al., 2007). In this instance establishing the diversity and<br />

evolution <strong>of</strong> GBSSI is the primary objective.<br />

Following crop improvements made by plant breeders in the<br />

last century, cereal-breeding programs are increasingly targeting<br />

cereal wild relatives as a source <strong>of</strong> novel germplasm (Hajjar and<br />

Hodgkin, 2007). This is occurring in a time <strong>of</strong> increasing human<br />

populations and climatic instability. It has been proposed that<br />

meeting global food requirements by 2025 will require a 25%<br />

increase in grain production (Khush, 2001). The use <strong>of</strong> CWR in plant<br />

breeding programs will become increasingly important because <strong>of</strong><br />

their environmental adaptations (Ashikari and Matsuoka, 2006).<br />

Previously, hybridisation <strong>of</strong> CWR from the primary gene pool <strong>of</strong> the<br />

respective cultivated cereals has been used to improve agronomic<br />

traits such as growth habit, and pest and disease resistance. Recent<br />

advances in molecular and plant breeding techniques have shifted<br />

the focus from primary to secondary and/or tertiary gene pool<br />

species as sources <strong>of</strong> novel allelic variation for genes <strong>of</strong> importance<br />

(Hajjar and Hodgkin, 2007; Rao et al., 2003). Characterisation <strong>of</strong> the<br />

full coding region <strong>of</strong> GBSSI in selected native species may identify<br />

variants <strong>of</strong> the GBSSI coding gene. In this study GBSSI gene<br />

sequence is reported, and used as an indicator <strong>of</strong> the breadth <strong>of</strong><br />

genetic diversity within and between Australian CWR and cereals.<br />

Furthermore, alignment <strong>of</strong> the amino acid sequences derived from<br />

the genomic data generated identifies novel putative GBSSI amino<br />

acid sequences in a subset <strong>of</strong> the Australian CWR.<br />

2. Materials and methods<br />

2.1. Plant material and DNA extraction<br />

Representative species <strong>of</strong> the four major clades <strong>of</strong> Poaceae found<br />

in Australia were selected for their close relationship to a cereal<br />

species, large seed size, broad geographic distribution and/or, in the<br />

case <strong>of</strong> Microlaena stipoides and Astrebla lappacea, their proven<br />

potential to be cultivated. Seed was germinated and leaf material<br />

harvested for DNA extractions. Species identity was confirmed by<br />

a herbarium and voucher details are listed in Table 1. Field collection<br />

was undertaken in Northern NSW for A. lappacea and M.<br />

F.M. Shapter et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 4–11 5<br />

stipoides. Elymus scaber and Austrostipa aristiglumis were sourced<br />

from Native Seeds Pty Ltd Australia (http://www.nativeseeds.com.<br />

au). Native Oryza and Sorghum seeds were sourced from the<br />

Australian Tropical Crops and Forages Collection, Queensland<br />

Department <strong>of</strong> Primary Industries and Fisheries, Biloela, Australia<br />

(www.dpi.gov.au/auspgris/). DNA was extracted using a Qiagen<br />

DNeasy Plant Maxi Kit to ensure sufficient high-quality/highconcentration<br />

DNA from a single extraction. DNA quality was<br />

confirmed by gel electrophoresis and quantified using a NanoDrop Ò<br />

Technologies ND-1000 spectrophotometer.<br />

2.2. Presence <strong>of</strong> amylose<br />

The presence <strong>of</strong> amylose was determined by Pederson’s iodine<br />

staining technique (Pedersen et al., 2004). Controls for waxy<br />

(Sorghum bicolor cv. Batsuto Red) and non-waxy (S. bicolor cv.<br />

LR2409) endosperm (Pedersen et al., 2007) were used.<br />

2.3. Sequence alignment<br />

GBSSI coding DNA sequence <strong>of</strong> Oryza sativa (X65183), Triticum<br />

aestivum (AB019623), Hordeum vulgare (AB089162), Zea mays<br />

(X03935), Setaria italica (AB089141) and S. bicolor (AF488412.1) was<br />

retrieved from the NCBI Genbank Database (http://www.ncbi.nlm.<br />

nih.gov/). Sequence alignments were performed using ClustalW<br />

(EMBL – European Bioinformatics Institute; http://www.ebi.ac.uk/<br />

clustalw/).<br />

2.4. PCR amplification and DNA sequencing<br />

Universal primer pair C and E from McIntosh et al. (2005) was<br />

used in accordance with the described method, to amplify a small<br />

central fragment (CE) <strong>of</strong> the GBSSI gene in the wild grasses. Small<br />

fragment PCR products were purified using a Qiagen Gel Extraction<br />

Kit prior to DNA sequencing. Full gene amplification was obtained<br />

by PCR using AccuprimeÔ Hi-Fi (Invitrogen) according to the<br />

manufacturer’s guidelines. To increase specificity, DMSO and 50%<br />

glycerol were added to the PCR reaction at 3 and 6%, respectively,<br />

for some <strong>of</strong> the species (Table 2).<br />

Heteroduplexes can be induced by PCR <strong>of</strong> multi-copy loci, followed<br />

by mismatch repair during cloning (Jansen and Ledley, 1990;<br />

Jensen and Straus, 1993; Longeri et al., 2002). To ensure sequence<br />

integrity, reconditioning PCRs were performed prior to cloning <strong>of</strong><br />

the full gene fragments (Thompson et al., 2002). Reconditioning<br />

PCRs were spiked with 10% volume <strong>of</strong> the primary PCR product as<br />

template; otherwise PCR conditions were identical to that <strong>of</strong> the<br />

original PCR. Sequencing reactions utilised ABI PRISM Ò Big DyeÔ<br />

Terminator (BDT) V3.1 chemistry at one-eighth concentration<br />

reactions and the resulting amplicons analysed by capillary<br />

electrophoresis on an ABI 3730 Genetic Analyzer by Southern Cross<br />

Plant Genomics (SCPG, Lismore, Australia). Sequence data were<br />

Table 1<br />

Classification, collection record, herbarium voucher and Genbank accession number <strong>of</strong> eight native grass species included in this study<br />

Species Tribe Collection record a<br />

Herbarium voucher Genbank accession<br />

Astrebla lappacea Cynodonteae AC04-1003495 BRIAQ 751244 EF600041<br />

Austrostipa aristiglumis Stipeae AC04-1003487 BRIAQ 751222 EF600042<br />

Elymus scaber Triticeae AC04-1003502 BRIAQ 751227 EF600043<br />

Microlaena stipoides Erharteae AC04-1003504 BRIAQ 751226 EF600044<br />

Oryza australiensis Oryzeae AusTRCF310676 BRIAQ723945 EF600036<br />

Oryza rufipogon Oryzeae AusTRCF 309313 BRIAQ723944 EF600037<br />

Sorghum leiocladum Andropogoneae AusTRCF 300187 DNAD0155695 EF600038<br />

Sorghum nitidum Andropogoneae AusTRCF 302543 CANB479881 50 EF600039; 30 EF600040<br />

a<br />

Collection numbers prefixed with AC04 can be located through the Australian Plant DNA Bank (www.dnabank.com.au) and those prefixed with AusTRCF through the<br />

Australian Plant Genetic Resources Information Service (www.dpi.qld.gov.au/auspgris/).


6<br />

Table 2<br />

Full length gene primers and PCR conditions for A. aristiglumis (Aa), A. lappacea (Al), E. scaber (Es), S. leiocladum (Le), M. stipoides (Ms), O. australiensis (Oa) and O. rufipogon (Or)<br />

aligned and edited using SequencherÔ V4.5 (Gene Codes<br />

Corporation, Ann Arbor, USA).<br />

2.5. Genome walking<br />

Species-specific primers were designed using Primer<br />

PremierÔ V5.0 (Premier Bios<strong>of</strong>t International, Palo Alto, USA). To<br />

amplify the gene from the central CE fragment out to the 5 0 - and<br />

3 0 -UTR, respectively, the BD GenomeWalkerÔ Universal Kit (BD<br />

Biosciences, San Jose, USA) was used, as per the manufacturer’s<br />

instructions, except that all PCR amplification mixes were<br />

prepared at half volume. A mixture <strong>of</strong> 3% DMSO and 6% glycerol<br />

was added to the PCR mix and five cycles added to the standard<br />

PCR program, as per the manufacturer’s trouble-shooting guide.<br />

2.6. Cloning and DNA sequencing<br />

GenomeWalkerÔ and full gene PCR products were run on a 1%<br />

low-melt agarose gel and the individual PCR product bands <strong>of</strong><br />

interest excised. Immediately prior to ligation into the vector, the<br />

gel slice was melted in a thermocycler at 72 C for 1 min and then<br />

held at 37 C. Three microliter <strong>of</strong> molten gel was included in the<br />

standard ligation reaction mix <strong>of</strong> the pGEM Ò -T Easy Vector II<br />

System (Promega Corporation, Madison, USA) following the manufacturer’s<br />

recommended protocol. Clones were screened for size<br />

by PCR using Roche Taq (Roche Products Pty Limited, Australia) in<br />

the manufacturer’s recommended reaction mix. Target clones were<br />

isolated, then amplified using the TempliphiÔ System (GE<br />

Healthcare) as per the manufacturer’s instructions. Standard BDT<br />

V3.1 one-eighth sequencing reactions used 5 ml <strong>of</strong> diluted Templiphi<br />

as template.<br />

2.7. Exon/intron boundaries<br />

The putative location <strong>of</strong> each exon/intron boundary was determined<br />

by alignment with cDNA sequence from Genbank accessions<br />

<strong>of</strong> the CWR’s most closely related available cDNA. The start <strong>of</strong> the<br />

introns was then assigned as being the nearest GT to that position<br />

and ending on the most distal AG prior to the following exon,<br />

whilst maintaining the open reading frame (Murai et al., 1999;<br />

Ramakrishna et al., 2002).<br />

2.8. Species comparisons<br />

Neighbour-joining phylogenetic analyses were conducted in<br />

PAUP*Version 4.0b10 (Sw<strong>of</strong>ford, 2001), using Kimura-2 parameter,<br />

F.M. Shapter et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 4–11<br />

Primer name Sequence 50 –30 Tm ( C) Size<br />

GC (%) PCR Step 1<br />

(bp/mer)<br />

a<br />

PCR Step 2<br />

(cycle #/temp. C)<br />

a<br />

Polymerase<br />

(cycle #/temp. C)<br />

GBSSAaF1 AGTGCAGTCATCTTTCACCA 53.1 20 45.0 8/58–50 30/50 Accuprime DG<br />

GBSSAaR1 CAGAACCTCAAACTTATTAGCC 53.6 22 40.9<br />

GBSSAlF3 ACCGCAGCTACGTACTCCGCC 60.0 21 67.7 10/62–53 30/52 Accuprime DG<br />

GBSSAlR2 CACCGATCCCCCTTGTTCT 59.3 19 57.9<br />

GBSSEsF3 TGAGGATTCATGCTCTAGGGTTA 58.6 23 43.5 12/60–48 30/49 Accuprime<br />

GBSSEsR3 AAGAGGCAAGCGGCACA 57.8 17 58.8<br />

GBSSLeF5 GCACATGTCTTTTCTTGATGC 55.7 21 42.9 12/60–48 30/49 Accuprime<br />

GBSSLeR5 TGATCCATCATCCTTGTGCTA 56.1 21 42.9<br />

GBSSMsF1 AAAATTCTGTTATCCCAACCC 55.4 21 39.1 8/8–50 30/50 Accuprime DG<br />

GBSSMsR2 ATAATGGCTACGACAGACTCAC 53.9 22 45.5<br />

GBSSOaF1 ACAGCAAGAGCTAGACAACCG 57.4 21 52.4 8/58–50 30/50 Accuprime DG<br />

GBSSOaR2 AAACCACTGCTTCATTCGTCT 56.6 21 42.9<br />

GBSSOrF1 ATCTTCCACAGCAACAGCTA 53.1 20 45.0 10/60–50 30/50 Accuprime<br />

GBSSOrR2 ATCACTCCATATAGATCTCAGGC 54.8 23 43.5<br />

a<br />

Two step PCR was utilised: the first step a touchdown cycle, each cycle the annealing temperature decreasing 1 C within the given range, the second step at a constant<br />

annealing temperature for the given number <strong>of</strong> cycles.<br />

distance method and maximum parsimony settings. Support for<br />

branches was obtained from 1000 pseudoreplicates for 1920 <strong>of</strong><br />

1921 characters, using the TBR branch swapping algorithm.<br />

3. Results<br />

3.1. Gene sequence<br />

Initial alignment <strong>of</strong> the GBSSI coding sequence from Genbank<br />

determined that sites for universal primer design at the 5 0 - and 3 0 -<br />

ends <strong>of</strong> the gene were not available due to poor homology between<br />

the species in these regions. Direct sequencing analysis <strong>of</strong> purified<br />

PCR amplicons <strong>of</strong> a small central fragment (CE) <strong>of</strong> the GBSSI gene<br />

spanning exons 7 and 8 (McIntosh et al., 2005) produced poor<br />

sequence data for some species with ambiguous base calls indicating<br />

multiple alleles. This was expected as, apart from the two<br />

wild Oryza species, the CWR are polyploids and hence sequence<br />

data could only be resolved by cloning the PCR products prior to<br />

sequencing.<br />

Fragment CE sequence was used as a platform for the design <strong>of</strong><br />

bidirectional species-specific primers for use in the BD Genomewalking<br />

protocol. The first set <strong>of</strong> gene/species-specific primers<br />

produced gene fragments that did not reach the start and stop<br />

codons. Up to three sets <strong>of</strong> gene/species-specific primers were<br />

designed in the exon closest to the respective 5 0 - and 3 0 -ends <strong>of</strong> the<br />

sequences obtained in the current round <strong>of</strong> genome walking<br />

(Fig. 1). Addition <strong>of</strong> the recommended concentrations <strong>of</strong> DMSO and<br />

glycerol in the second and third rounds <strong>of</strong> genome walking resulted<br />

in improved amplification specificity.<br />

Of the eight species selected for full gene characterisation, seven<br />

were successfully amplified with full gene primers located in the 5 0 -<br />

and 3 0 -UTR. Due to the polyploid status <strong>of</strong> most <strong>of</strong> these species it<br />

was necessary to amplify the entire gene sequence in a single PCR<br />

fragment, then clone the amplicon. This ensured the sequence was<br />

obtained from a single genome.<br />

In the eighth species, Sorghum nitidum, genome walking<br />

produced two major gene fragments, overlapping at intron 4 and<br />

the start <strong>of</strong> exon 5. Polymorphisms between the overlapping fragments<br />

determined that these were two different alleles; indicating<br />

more than one genome may have been amplified. Although more<br />

than 20 primer pairs were utilised under various PCR conditions,<br />

full gene amplification was unsuccessful for this species. Consequently,<br />

S. nitidum has been entered into Genbank as two separate<br />

(5 0 and 3 0 ) sequences (Table 1). To determine a putative coding<br />

sequence for S. nitidum, the gene sequence was spliced, in silico, at


exon 5 and degenerate base calls inserted for the three polymorphic<br />

positions in the overlapping fragment.<br />

A. lappacea appeared to be missing the 3 0 portion <strong>of</strong> exon 1 and<br />

missing its expected exon 1/intron 1 boundary when compared to<br />

all other species examined. Exons 2 and 3 were as expected;<br />

however, the 3 0 boundary <strong>of</strong> exon 4 occurred 1 base earlier, causing<br />

a frame shift when translated. Other than exons 5 and 11, the exon<br />

sizes varied by between 1 and 29 bases, causing repeated frame<br />

shifts and encoding multiple stop codons commencing with the<br />

first in exon 5.<br />

Amongst all the species examined, the genomic base count<br />

ranged from 2760 bases (E. scaber) through to 3572 bases (Oryza<br />

rufipogon). This large variation in nucleotide number can be<br />

accounted for in the Triticeae by the absence <strong>of</strong> introns 4 and 7. The<br />

insertion <strong>of</strong> over 400 bases in intron 11 <strong>of</strong> O. rufipogon is partially<br />

responsible for its larger size (Fig. 1). The number <strong>of</strong> putative coding<br />

bases varies from 1812 base pairs (bp) in A. aristiglumis up to<br />

1830 bp for both Oryza australiensis and O. rufipogon, with the four<br />

cereal species investigated falling within this range. This translates<br />

to an addition/deletion <strong>of</strong> up to 14 amino acids.<br />

3.2. Transposable elements<br />

In Sorghum leiocladum the first set <strong>of</strong> genome walking gene/<br />

species-specific primers produced a 1300 bp fragment. Nine<br />

hundred bases <strong>of</strong> this fragment were homologous to exons 3 to 7 <strong>of</strong><br />

the O. sativa GBSSI coding sequence. At the 5 0 -end <strong>of</strong> the fragment<br />

the first 400 bases after the GenomeWalkerÔ AP2 primer sequence<br />

had no homology to the exons <strong>of</strong> any GBSSI coding gene in Genbank.<br />

These 400 bases were isolated and submitted to a BlastN<br />

search which returned a strong identity to a retrotransposon isolated<br />

from S. bicolor. Similarly, PCR amplification within the O.<br />

australiensis GenomeWalkerÔ library produced a fragment which<br />

had the GenomeWalkerÔ AP2 primer sequence at both ends <strong>of</strong> the<br />

fragment. The sequence showed no homology to the cultivated rice<br />

GBSSI exons, and was determined by BlastN search to be homologous<br />

to an O. australiensis retrotransposon. The 400 base insert in<br />

intron 12 <strong>of</strong> O. rufipogon did not return any matches when used as<br />

a BlastN query sequence.<br />

F.M. Shapter et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 4–11 7<br />

Fig. 1. Schematic diagram showing/illustrating the variations observed across eight Australian grass species and their cultivated cereal relatives. The location <strong>of</strong> central fragment<br />

(CE) and the genome-walking gene/species-specific primers are indicated for each species.<br />

3.3. Granule Bound Starch Synthase I<br />

Comparison <strong>of</strong> the putative amino acid sequences derived in this<br />

study with those for the cereals retrieved from Genbank showed<br />

little conservation within the 79 amino acid consensus positions <strong>of</strong><br />

the transit peptide (Fig. 2). The consensus <strong>of</strong> the mature protein was<br />

540 amino acids and averaged 75% identity. The conserved domain<br />

database (Marchler-Bauer et al., 2007) indicates two conserved<br />

domains exist in GBSSI; the glycosyl transferase groups five and<br />

one. Group five has 79.5% identity and group one maintains 71.4%<br />

identity. The C-terminus has 74.7% identity within the amino acid<br />

sequence. The ADPG active binding site near the N-terminus<br />

(Baldwin, 2001; Denyer et al., 2001) was identical in all species<br />

except A. lappacea and S. nitidum.<br />

A single amino acid change occurred in the spliced S. nitidum<br />

sequence. The deletion <strong>of</strong> part <strong>of</strong> exon 1 in the A. lappacea sequence<br />

translated to the loss <strong>of</strong> the entire binding site. Of the two<br />

conserved motifs near the C-terminus (Baldwin, 2001), Motif II had<br />

a proline substituted for the threonine in the two Oryza species.<br />

Motif III was conserved across all species.<br />

3.4. Phylogenetic outcomes<br />

Phylograms were developed containing all the possible<br />

boundary variations for the A. lappacea sequence (data not<br />

shown), including a sequence with no intron 1. All 16 permutations<br />

were monophyletic and remained a sister group to the<br />

remainder <strong>of</strong> the Poaceae examined. The representative coding<br />

sequence <strong>of</strong> A. lappacea used in the phylogenetic analyses was<br />

determined by having the least variation in exon sizes and the<br />

least amino acid sequence disruptions. Preliminary inclusion <strong>of</strong><br />

the spliced S. nitidum sequence resulted in its position within the<br />

topology changing when analyses between the second half <strong>of</strong> the<br />

gene sequence were compared to the full coding sequence. Using<br />

the entire sequence S. nitidum was placed between Z. mays and S.<br />

italica. When only the second half <strong>of</strong> the gene was analysed S.<br />

nitidum was clustered with S. leiocladum with stronger statistical<br />

support. As S. nitidum sequence data was not derived from


8<br />

a single genome it was excluded from the phylogenetic analysis<br />

presented.<br />

No suitable outgroup was available (from Genbank) for the<br />

phylogenetic analysis so all trees were unrooted. Neighbourjoining<br />

and maximum parsimony analyses <strong>of</strong> the full coding<br />

sequence produced trees with A. lappacea as an outgroup to the<br />

F.M. Shapter et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 4–11<br />

Fig. 2. Amino acid alignment <strong>of</strong> GBSSI and associated conserved domains.<br />

rest <strong>of</strong> the Poaceae examined (Fig. 3). Analyses produced a single<br />

congruent phylogeny with good bootstrap support except at the<br />

bifurcation <strong>of</strong> A. aristiglumis and M. stipoides which had lower<br />

bootstrap values. The basal bifurcation <strong>of</strong> the trees separates<br />

the BEP (Bambusoideae, Erhartoideae and Pooideae) from the<br />

PACCAD (Panicoideae, Arundinoideae, Chloridoideae,


63<br />

66<br />

100<br />

100<br />

61<br />

60<br />

50 changes<br />

51<br />

53<br />

Centothecoideae, Aristidoideae and Danthoniodeae) clades.<br />

Previous phylogenetic analyses utilising the Waxy gene did not<br />

incorporate the entire coding region and therefore a second tree<br />

was produced excluding characters 1–986 which resulted in no<br />

changes to the tree topology.<br />

3.5. Iodine test<br />

The iodine test for the presence or absence <strong>of</strong> amylose determined<br />

that all eight <strong>of</strong> the CWR had amylose present in their<br />

endosperm (Fig. 4).<br />

4. Discussion<br />

92<br />

91<br />

100<br />

100<br />

Alappacea<br />

100<br />

100<br />

100<br />

100<br />

Sitalica<br />

Oaustraliensis<br />

Orufipogon<br />

Osativa<br />

Mstipoides<br />

Aaristiglumis<br />

Zmays<br />

100<br />

100<br />

71<br />

75<br />

Sleiocladum<br />

Sbicolor<br />

Escaber<br />

T.aestivum<br />

Hvulgare<br />

Compound granules<br />

Bimodal granules<br />

Simple granules<br />

Compound granules<br />

Fig. 3. Neighbour-joining Tree <strong>of</strong> GBSSI complete coding sequence. Bootstrap values<br />

are calculated with 1000 replicates with neighbour-joining values presented above the<br />

branches and maximum parsimony values below. The type <strong>of</strong> starch granule<br />

morphology reported for each species is noted on the right <strong>of</strong> the phylogram.<br />

GBSSI was successfully amplified in seven <strong>of</strong> the eight species<br />

examined. For the eighth species, S. nitidum, screening multiple<br />

genome walker libraries was unable to isolate the same locus for<br />

the 5 0 - and 3 0 -gene fragments. As such the two overlapping sections<br />

<strong>of</strong> the gene that were sequenced did not provide matching UTR<br />

Fig. 4. Iodine test for the presence or absence <strong>of</strong> amylose. A magenta colour indicates<br />

a waxy phenotype (BR) and blue indicates the presence <strong>of</strong> amylose (LR). Aa – A.<br />

aristiglumis, Al–A. lappacea, Es–E. scaber, Ms–M. stipoides, Le–S. leiocladum, Ni–S.<br />

nitidum, Oa–O. australiensis, Or–O. rufipogon.<br />

F.M. Shapter et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 4–11 9<br />

sequence for full gene primers to be successfully applied.<br />

Comparison <strong>of</strong> the UTR <strong>of</strong> the other CWR and cereals sequenced<br />

showed no homology in this region even amongst closely related<br />

species. It is therefore not surprising that the sequence differentiation<br />

between genomes in this region <strong>of</strong> the gene is diverse enough<br />

to prevent allele specific primers from amplifying.<br />

The GBSSI homolog amplified in A. lappacea was missing<br />

a significant portion <strong>of</strong> exon 1 which translated to a putative active<br />

site in the protein. There was little conservation <strong>of</strong> the exon/intron<br />

boundaries throughout the gene and once translated, multiple stop<br />

codons were identified throughout the transcript. It is proposed<br />

that this sequence was derived from a non-functional ortholog.<br />

Null-alleles <strong>of</strong> this gene in diploids result in a waxy or amylose free<br />

phenotype. For example, an indigenous barley mutant with a large<br />

deletion including exon 1 was determined to have a waxy phenotype<br />

(Domon et al., 2002). Testing for amylose indicated that all<br />

eight CWR contained some amylose (Fig. 4) in their endosperm. The<br />

ploidy <strong>of</strong> A. lappacea is not clearly stated in the literature, though its<br />

base chromosome number <strong>of</strong> 2n ¼ 40 (Jozwik, 1969) would suggest<br />

some level <strong>of</strong> polyploidy. It can be assumed then, that A. lappacea<br />

has another GBSSI locus yet to be characterised, that is functional<br />

and being expressed in the endosperm. Wheat lines with multiple<br />

GBSSI loci still produced amylose in the endosperm in lower<br />

concentrations than normal when null-alleles did not occur<br />

concurrently at all GBSSI loci (Mangalika et al., 2003).<br />

Transposable elements have been previously identified in the<br />

intron sequence <strong>of</strong> the Waxy gene (Mason-Gamer et al., 1998; Wang<br />

et al., 1994). In the case <strong>of</strong> S. italica, insertion <strong>of</strong> transposable<br />

elements in the coding region is responsible for a waxy phenotype<br />

(Kawase et al., 2005). Multiple attempts to design genome-walking<br />

primers to amplify past the retrotransposon in the 5 0 -fragment <strong>of</strong> S.<br />

leiocladum were undertaken, though remained unsuccessful. It was<br />

hypothesised that the retrotransposon may be too large to amplify<br />

across in a single genome walking reaction. An alternative genomewalking<br />

library was used to amplify the 5 0 -end <strong>of</strong> the gene from<br />

another genome in which no retrotransposon was present in intron<br />

3. Retrotransposons have been reported as being specific within<br />

a genus or tribe (Bennetzen et al., 2004). Finding retrotransposable<br />

elements within the CWR genomes is to be expected as these<br />

transposable elements can comprise up to 50% <strong>of</strong> nuclear DNA<br />

content (Kumar and Bennetzen, 1999). The inclusion <strong>of</strong> transposable<br />

elements has also been associated with the larger genome size<br />

<strong>of</strong> O. australiensis (Vaughan et al., 2003).<br />

Sequence alignment <strong>of</strong> the complete gene sequence from<br />

multiple species identified hot spots <strong>of</strong> coding sequence divergence<br />

between the species. The first 200 and final 100 bp <strong>of</strong> the GBSSI<br />

coding gene were particularly poorly conserved within Poaceae.<br />

Amongst the putative functional alleles, variable intron sizes (Fig. 1)<br />

indicating insertions/deletions were observed, across the species<br />

with especially large variations within intron 12. Phylogenetic<br />

investigation used these indels to identify homoeologous copies <strong>of</strong><br />

the Waxy gene (Fortune et al., 2007). The gene variation observed in<br />

the current study concurs with comparative studies <strong>of</strong> other<br />

homologous grass genes, where exon/intron boundaries remain<br />

conserved, but the size <strong>of</strong> the introns varies (Feuillet and Keller,<br />

2002). The coding sequence also showed considerable diversity<br />

across the taxa sampled.<br />

Early protein studies determined that the molecular weight <strong>of</strong><br />

GBSSI varied from 58 to 60 kDa between species (Taira et al., 1995).<br />

Amongst the CWR, coding sequence diversity translates to the<br />

omission <strong>of</strong> 14 amino acids over the length <strong>of</strong> the protein and<br />

multiple polymorphisms in the amino acid sequence. Any changes<br />

to the amino acid sequence have the potential to cause novel<br />

phenotypic expression <strong>of</strong> this gene. Previous studies <strong>of</strong> diploid<br />

mutants containing a null Waxy allele revealed GBSSI to be<br />

responsible for amylose synthesis in the storage tissues <strong>of</strong> plants


10<br />

(Vrinten and Nakamura, 2000). Sequence variations, which translate<br />

to amino acid changes, may lead to the synthesis <strong>of</strong> novel<br />

starches in planta, thus creating the potential for plant breeders to<br />

manipulate starch synthesis through conventional breeding or<br />

transgenic modification <strong>of</strong> cereal crops. The production <strong>of</strong> cereals<br />

containing novel starches will reduce or eliminate the need for<br />

costly post-harvest modifications (Davis et al., 2003). Introgression<br />

<strong>of</strong> novel genes from CWR has already been responsible for major<br />

cereal production gains (Hajjar and Hodgkin, 2007; Langridge et al.,<br />

2001). The allelic variation observed in this small subset <strong>of</strong><br />

Australia’s native grasses shows that considerable functional<br />

diversity may exist in the native grass gene pool, which could be<br />

harvested for grain improvement programs.<br />

The basal bifurcation <strong>of</strong> the tree separates the BEP and PACCAD<br />

clades, as was reported from previous phylogenetic studies <strong>of</strong> the<br />

Poaceae (Kellogg, 2002; Preston and Kellogg, 2006; Tomlinson<br />

and Denyer, 2003). Mason-Gamer et al. (1998) concluded that<br />

sequencing through the introns <strong>of</strong> GBSSI was unnecessary to<br />

resolve relationships between genera, tribes and subfamilies within<br />

Poaceae. Sufficient resolution was obtained from cDNA data and the<br />

results from the current study support their findings. Similarly,<br />

Mason-Gamer et al. (1998) utilised just the 3 0 half <strong>of</strong> the gene,<br />

which can be universally amplified for analysis. Comparisons <strong>of</strong><br />

phylogenies developed from this smaller region <strong>of</strong> the gene<br />

compared to the full coding sequence did not result in any change<br />

in topology. This analysis indicates that time consuming and costly<br />

acquisition <strong>of</strong> sequence data from non-conserved areas <strong>of</strong> the Waxy<br />

gene does not provide any clearer resolution <strong>of</strong> phylogenetic<br />

inferences for the Poaceae.<br />

A previous scanning electron microscopy study (Shapter et al.,<br />

2008) identified the starch granule morphology <strong>of</strong> the eight species<br />

and their cereal relatives. Since these trees were developed using<br />

the entire coding region <strong>of</strong> the Waxy gene, and given the Waxy<br />

gene’s role in starch formation, comparative analysis <strong>of</strong> the two<br />

data sets was undertaken (Fig. 3). There are three main starch<br />

granule morphologies within the cereals: (1) compound granulesdsmall<br />

rigidly polygonal granules arranged in large spherical<br />

compound granules typical <strong>of</strong> rice; (2) simple granulesdindividual<br />

large spherical/lenticellular to polygonal granules found in maize<br />

and sorghum; (3) bimodal granule arrangement with a mix <strong>of</strong> small<br />

spherical (B-type) and large lenticellular (A-type) granules typical<br />

<strong>of</strong> wheat (Tomlinson and Denyer, 2003). S. italica is the earliest<br />

diverging lineage within the PACCAD clade (Fig. 3). It is also the only<br />

member <strong>of</strong> this clade from this study with compound granules.<br />

Early SEM investigations <strong>of</strong> starch granule morphology <strong>of</strong> other<br />

members <strong>of</strong> this clade outside the Andropogoneae, revealed they<br />

too had compound granules within their endosperm, and it was<br />

proposed that this phenotype is the ancestral state (Shapter et al.,<br />

2008; Tateoka, 1962). This hypothesis would appear to be correct as<br />

both simple and bimodal granules occur only in the more recently<br />

diverged lineages.<br />

While the association between starch granule morphology and<br />

Waxy phylogeny appears to indicate a possible link between GBSSI<br />

and starch granule formation, it seems unlikely as it has previously<br />

been reported that a loss <strong>of</strong> function in this gene causes no<br />

apparent change to starch granule appearance (Buleon et al., 1998;<br />

Kim et al., 2003). Conversely, lesions in the isoamylase gene in<br />

barley were correlated to a lack <strong>of</strong> A and B-type granules and the<br />

appearance <strong>of</strong> compound granules in the endosperm (Burton et al.,<br />

2002). Screening <strong>of</strong> the amino acid alignment <strong>of</strong> all the species<br />

examined in this study observed several amino acid changes in the<br />

conserved domains <strong>of</strong> the gene which also correlated with the<br />

starch granule morphology (Fig. 2). These separated either all three<br />

types <strong>of</strong> granule (V), just the species with simple granules (U) or<br />

just the bimodal granules (x). While it is likely that these amino acid<br />

associations are due to the evolutionary history <strong>of</strong> the taxa, rather<br />

F.M. Shapter et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 4–11<br />

than a direct genetic effect <strong>of</strong> the expression <strong>of</strong> a particular ortholog<br />

<strong>of</strong> GBSSI, it would be uncommon for an enzyme to have a single<br />

function (determination <strong>of</strong> amylose) with no other biochemical<br />

consequences. Perhaps as more is discovered about the role <strong>of</strong><br />

GBSSI in endosperm development, these associations will be found<br />

to have more than just an evolutionary connection.<br />

Acknowledgements<br />

Seed for this study was supplied by Australian Tropical Crops<br />

and Forages Collection, Queensland Department <strong>of</strong> Primary<br />

Industries and Fisheries, www.dpi.qld.gov.au/auspgris/and Native<br />

Seeds Pty Ltd Australia http://www.nativeseeds.com.au. Funding<br />

was provided by the Grain Foods Cooperative Research Centre.<br />

References<br />

Ashikari, M., Matsuoka, M., 2006. Identification, isolation and pyramiding <strong>of</strong><br />

quantitative trait loci for rice breeding. Trends in Plant <strong>Science</strong> 11, 344–350.<br />

Baldwin, P., 2001. Starch granule-associated proteins and polypeptides: a review.<br />

Starke. (Starch) 53, 475–503.<br />

Bennetzen, J.L., Coleman, C., Liu, R.Y., Ma, J.X., Ramakrishna, W., 2004. Consistent<br />

over-estimation <strong>of</strong> gene number in complex plant genomes. Current Opinion in<br />

Plant Biology 7, 732–736.<br />

Buleon, A., Colonna, P., Planchot, V., Ball, S., 1998. Starch granules: structure and<br />

biosynthesis. International <strong>Journal</strong> <strong>of</strong> Biological Macromolecules 23, 85–112.<br />

Burton, R.A., Jenner, H., Carrangis, L., Fahy, B., Fincher, G.B., Hylton, C., Laurie, D.A.,<br />

Parker, M., Waite, D., van Wegen, S., Verhoeven, T., Denyer, K., 2002. Starch<br />

granule initiation and growth are altered in barley mutants that lack isoamylase<br />

activity. The Plant <strong>Journal</strong> 31, 97–112.<br />

Davis, J.P., Supatcharee, N., Khandelwal, R.L., Chibbar, R.N., 2003. Synthesis <strong>of</strong> novel<br />

starches in planta: opportunities and challenges. Starke. (Starch) 55, 107–120.<br />

Demeke, T., Huel, P., Abdel-Aal, E.S.M., Baga, M., Chibbar, R.N., 1999. Biochemical<br />

characterization <strong>of</strong> the wheat waxy A protein and its effect on starch properties.<br />

<strong>Cereal</strong> Chemistry 76, 694–698.<br />

Denyer, K., Johnson, P., Zeeman, S., Smith, A.M., 2001. The control <strong>of</strong> amylose<br />

synthesis. <strong>Journal</strong> <strong>of</strong> Plant Physiology 158, 479–487.<br />

Domon, E., Saito, A., Takeda, K., 2002. Comparison <strong>of</strong> the waxy locus sequence from<br />

a non-waxy strain and two waxy mutants <strong>of</strong> spontaneous and artificial origins<br />

in barley. Genes & Genetic Systems 77, 351–359.<br />

Evans, R.C., Campbell, C.S., 2002. The origin <strong>of</strong> the apple subfamily (Maloideae;<br />

Rosaceae) is clarified by DNA sequence data from duplicated GBSSI genes.<br />

American <strong>Journal</strong> <strong>of</strong> Botany 89, 1478–1484.<br />

Feuillet, C., Keller, B., 2002. Comparative genomics in the grass family: molecular<br />

characterization <strong>of</strong> grass genome structure and evolution. Annals <strong>of</strong> Botany 89,<br />

3–10.<br />

Fortune, P.M., Schierenbeck, K.A., Ainouche, A.K., Jacquemin, J., Wendel, J.F.,<br />

Ainouche, M.L., 2007. Evolutionary dynamics <strong>of</strong> Waxy and the origin <strong>of</strong> hexaploid<br />

Spartina species (Poaceae). Molecular Phylogenetics and Evolution 43,<br />

1040–1055.<br />

Gaines, C.S., Raeker, M.O., Tilley, M., Finney, P.L., Wilson, J.D., Betchel, D.B., Martin, R.J.,<br />

Seib, P.A., Lookhart, G.L., Donelson, T., 2000. Associations <strong>of</strong> starch gel hardness,<br />

granule size, waxy allelic expression, thermal pasting, milling quality, and kernel<br />

texture <strong>of</strong> 12 s<strong>of</strong>t wheat cultivars. <strong>Cereal</strong> Chemistry 77, 163–186.<br />

Hajjar, R., Hodgkin, T., 2007. The use <strong>of</strong> wild relatives in crop improvement: a survey<br />

<strong>of</strong> developments over the last 20 years. Euphytica 156, 1–13.<br />

Jansen, R., Ledley, F.D., 1990. Disruption <strong>of</strong> phase during PCR amplification and<br />

cloning <strong>of</strong> heterozygous target sequences. Nucleic Acids Research 18, 5153–5156.<br />

Jensen, M., Straus, N., 1993. Effect <strong>of</strong> PCR conditions on the formation <strong>of</strong> heteroduplex<br />

and single-stranded DNA products in the amplification <strong>of</strong> bacterial<br />

ribosomal DNA spacer regions. PCR Methods Application 3, 186–194.<br />

Jozwik, F.X., 1969. Some systematic aspects <strong>of</strong> Mitchell grasses. Australian <strong>Journal</strong> <strong>of</strong><br />

Botany 17, 359–374.<br />

Kawase, M., Fukunaga, K., Kato, K., 2005. Diverse origins <strong>of</strong> waxy foxtail millet crops<br />

in east and southeast Asia mediated by multiple transposable element insertions.<br />

Molecular Genetics and Genomics 274, 131–140.<br />

Kellogg, E.A., 2002. Classification <strong>of</strong> the grass family. In: Mallett, K., Orchard, A.<br />

(Eds.), Flora <strong>of</strong> Australia. ABRS/CSIRO Australia, Melbourne, pp. 19–36.<br />

Khush, G.S., 2001. Challenges for meeting the global food and nutrient needs in the<br />

new millennium. Proceedings <strong>of</strong> the Nutrition Society 60, 15–26.<br />

Kim, W., Johnson, J.W., Graybosch, R.A., Gaines, C.S., 2003. Physicochemical properties<br />

and end-use quality <strong>of</strong> wheat starch as a function <strong>of</strong> waxy protein alleles.<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 37, 195–204.<br />

Kumar, A., Bennetzen, J., 1999. Plant retrotransposons. Annual Reviews in Genetics<br />

33, 479–532.<br />

Langridge, P., Lagudah, E.S., Holton, T.A., Appels, R., Sharp, P.J., Chalmers, K.J., 2001.<br />

Trends in genetic and genome analyses in wheat: a review. Australian <strong>Journal</strong> <strong>of</strong><br />

Agricultural Research 52, 1043–1077.<br />

Longeri, M., Zanotti, M., Damiani, G., 2002. Recombinant DRB sequences produced<br />

by mismatch repair <strong>of</strong> heteroduplexes during cloning in Escherichia coli. European<br />

<strong>Journal</strong> <strong>of</strong> Immunogenetics 29, 517–523.


Mangalika, W.H.A., Miura, H., Yamauchi, H., Noda, T., 2003. Properties <strong>of</strong> starches<br />

from near-isogenic wheat lines with different Wx protein deficiencies. <strong>Cereal</strong><br />

Chemistry 80, 662–666.<br />

Marchler-Bauer, A., Anderson, J.B., Derbyshire, M.K., DeWeese-Scott, C.,<br />

Gonzales, N.R., Gwadz, M., Hao, L., He, S., Hurwitz, D.I., Jackson, J.D., Ke, Z.,<br />

Krylov, D., Lanczycki, C.J., Liebert, C.A., Liu, C., Lu, F., Lu, S., Marchler, G.H.,<br />

Mullokandov, M., Song, J.S., Thanki, N., Yamashita, R.A., Yin, J.J., Zhang, D.,<br />

Bryant, S.H., 2007. CDD: a conserved domain database for interactive domain<br />

family analysis. Nucleic Acids Research 35, D237–D240.<br />

Mason-Gamer, R.J., Weil, C.F., Kellogg, E.A., 1998. Granule-bound starch synthase:<br />

structure, function, and phylogenetic utility. Molecular Biology and Evolution<br />

15, 1658–1673.<br />

McIntosh, S.R., Pacey-Millar, T., Henry, R.J., 2005. A universal protocol for identification<br />

<strong>of</strong> cereals. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 41, 37–46.<br />

Murai, J., Taira, T., Ohta, D., 1999. Isolation and characterization <strong>of</strong> the three Waxy<br />

genes encoding the granule-bound starch synthase in hexaploid wheat. Gene<br />

234, 71–79.<br />

Nakamura, T., Yamamori, M., Hirano, H., Hidaka, S., Nagamine, T., 1995. Production<br />

<strong>of</strong> waxy (amylose-free) wheats. Molecular and General Genetics 248,<br />

253–259.<br />

Ochieng, J.W., Muigai, A.W.T., Ude, G.N., 2007. Phylogenetics in plant biotechnology:<br />

principles, obstacles and opportunities for the resource poor. African <strong>Journal</strong> <strong>of</strong><br />

Biotechnology 6, 639–649.<br />

Pedersen, J.F., Bean, S.R., Funnell, D.L., Graybosch, R.A., 2004. Rapid iodine staining<br />

techniques for identifying the waxy phenotype in sorghum grain and waxy<br />

genotype in sorghum pollen. Crop <strong>Science</strong> 44, 764–767.<br />

Pedersen, J.F., Graybosch, R.A., Funnell, D.L., 2007. Occurrence <strong>of</strong> the waxy alleles<br />

wxa and wxb in waxy sorghum plant introductions and their effect on starch<br />

thermal properties. Crop <strong>Science</strong> 37.<br />

Preston, J.C., Kellogg, E.A., 2006. Reconstructing the evolutionary history <strong>of</strong><br />

paralogous APETALA1/FRUITFULL-like genes in grasses (Poaceae). Genetics 174,<br />

421–437.<br />

Rahman, S., Li, Z., Batey, I., Cochrane, M.P., Appels, R., Morell, M., 2000. Genetic<br />

alteration <strong>of</strong> starch functionality in wheat. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 31,<br />

91–110.<br />

Ramakrishna, W., Ma, J.X., SanMiguel, P., Emberton, J., Dubcovsky, J., Shil<strong>of</strong>f, B.A.,<br />

Jiang, Z.Y., Rostoks, N., Busso, C.S., Ogden, M., Linton, E., Kleinh<strong>of</strong>s, A.,<br />

Devos, K.M., Messing, J., Bennetzen, J.L., 2002. Frequent genic rearrangements in<br />

F.M. Shapter et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 4–11 11<br />

two regions <strong>of</strong> grass genomes identified by comparative sequence analysis.<br />

Comparative and Functional Genomics 3, 165–166.<br />

Rao, N.K., Reddy, L.J., Bramel, P.J., 2003. Potential <strong>of</strong> wild species for genetic<br />

enhancement <strong>of</strong> some semi-arid food crops. Genetic Resources and Crop<br />

Evolution 50, 707–721.<br />

Shapter, F.M., Lee, L.S., Henry, R.J., 2008. Endosperm and starch granule morphology<br />

in wild cereal relatives. Plant Genetic Resources 6, 85–97.<br />

Shure, M., Wessler, S., Fedor<strong>of</strong>f, N., 1983. Molecular identification and isolation <strong>of</strong><br />

the Waxy locus in maize. Cell 35, 225–233.<br />

Small, R.L., Cronn, R.C., Wendel, F., 2004. L.A.S. Johnson Review No. 2: use <strong>of</strong> nuclear<br />

genes for phylogeny reconstruction in plants. Australian Systematic Botany 17,<br />

145–170.<br />

Sw<strong>of</strong>ford, D.L., 2001. PAUP: Phylogenetic Analysis Using Parsimony (*and Other<br />

Methods), version 4.0b10. Sinauer Associates, USA.<br />

Taira, T., Fujita, N., Takaoka, K., Uematsu, M., Wadano, A., Kozaki, S., Okabe, S., 1995.<br />

Variation in the primary structure <strong>of</strong> waxy proteins (granule-bound starch<br />

synthase) in diploid cereals. Biochemical genetics 33, 269–281.<br />

Tateoka, T., 1962. Starch grains <strong>of</strong> endosperm in grass systematics. Botany Magazine<br />

Tokyo 75, 377–383.<br />

Thompson, J.R., Marcelino, L.A., Polz, M.F., 2002. Heteroduplexes in mixed-template<br />

amplifications: formation, consequence and elimination by ‘reconditioning<br />

PCR’. Nucleic Acids Research 30, 2083–2088.<br />

Tomlinson, K., Denyer, K., 2003. Starch synthesis in cereal grains. Advances in<br />

Botanical Research 40, 1–61.<br />

Vaughan, D.A., Morishma, H., Kadowaki, K., 2003. Diversity in the Oryza genus.<br />

Current Opinion in Plant Biology 6, 139–146.<br />

Vrinten, P.L., Nakamura, T., 2000. Wheat granule-bound starch synthase l and ll are<br />

encoded by separate genes that are expressed in different tissues. Plant<br />

Physiology 122, 255–263.<br />

Wang, Z.Y., Zheng, F.Q., Gao, J.P., Wang, X.Q., Wu, M., Zhang, J.L., Hong, M.M., 1994.<br />

Identification <strong>of</strong> two transposon-like elements in rice Wx gene. <strong>Science</strong> in<br />

China 37, 437–447.<br />

Williams, W., Verry, I.M., Ellison, N.W., 2006. A phylogenetic approach to germplasm<br />

use in clover breeding. In: Woodfield, D. (Ed.), 13th Australasian Plant<br />

Breeders Conference, Christchurch, NZ.<br />

Yan, L., Bhave, M., Fairclough, R., Konik, C., Rahman, S., Appels, R., 2000. The genes<br />

encoding granule-bound starch synthase at the waxy loci <strong>of</strong> the A, B and D<br />

progenitors <strong>of</strong> common wheat. Genome 43, 264–272.


Effect <strong>of</strong> high temperature on albumin and globulin accumulation in the<br />

endosperm proteome <strong>of</strong> the developing wheat grain q<br />

William J. Hurkman *, William H. Vensel, Charlene K. Tanaka, Linda Whitehand, Susan B. Altenbach<br />

U.S. Department <strong>of</strong> Agriculture, Agricultural Research Service, Western Regional Research Center, 800 Buchanan Street, Albany, CA 94710, USA<br />

article info<br />

Article history:<br />

Received 14 December 2007<br />

Received in revised form 6 June 2008<br />

Accepted 13 June 2008<br />

Keywords:<br />

Albumins<br />

Globulins<br />

Grain fill<br />

High temperature<br />

1. Introduction<br />

abstract<br />

High temperature during grain fill is one <strong>of</strong> the more significant<br />

environmental factors that affects wheat yield and flour quality<br />

(Skylas et al., 2002 and references therein). Although major<br />

processes in grain fill have been described (Laudencia-Chingcuanco<br />

et al., 2007; McIntosh et al., 2007; Vensel et al., 2005), little is<br />

known about the effects <strong>of</strong> high temperature on this developmental<br />

program. High temperature shortens the duration <strong>of</strong> grain fill and<br />

decreases the time to apoptosis and harvest maturity (Altenbach<br />

et al., 2003). Consistent with these events, transcripts for a-, g-, and<br />

u-gliadins, high molecular weight glutenin subunits (HMW-GS),<br />

and low molecular weight glutenin subunits (LMW-GS) accumulate<br />

and disappear earlier (Altenbach et al., 2002; Altenbach and<br />

Kothari, 2004). Transcripts for genes functioning in protein<br />

synthesis, starch synthesis, stress/defense, as well as storage<br />

accumulate earlier in response to high temperature (Altenbach and<br />

Abbreviations: dpa, days post-anthesis; HMW-GS, high molecular weight glutenin<br />

subunits; LMW-GS, low molecular weight glutenin subunits.<br />

q Disclaimer: The mention <strong>of</strong> a trademark or proprietary product does not<br />

constitute a guarantee or warranty <strong>of</strong> the product by the United States Department<br />

<strong>of</strong> Agriculture and does not imply its approval to the exclusion <strong>of</strong> other products<br />

that may be suitable.<br />

* Corresponding author. Tel.: þ1 510 559 5720; fax: þ1 510 559 5818.<br />

E-mail address: william.hurkman@ars.usda.gov (W.J. Hurkman).<br />

0733-5210/$ – see front matter Published by Elsevier Ltd.<br />

doi:10.1016/j.jcs.2008.06.014<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 12–23<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

The accumulation <strong>of</strong> KCl-soluble/methanol-insoluble albumins and globulins was investigated in the<br />

endosperm <strong>of</strong> developing wheat (Triticum aestivum, L. cv. Butte 86) grain produced under a moderate<br />

(24 C/17 C, day/night) or a high temperature regimen (37 C/28 C) imposed from 10 or 20 days postanthesis<br />

(dpa) until maturity. Proteins were separated by 2-DE and developmental pr<strong>of</strong>iles for nearly 200<br />

proteins were analyzed by hierarchical clustering. Comparison <strong>of</strong> protein pr<strong>of</strong>iles across physiologically<br />

equivalent stages <strong>of</strong> grain fill revealed that high temperature shortened, but did not substantially alter,<br />

the developmental program. Accumulation <strong>of</strong> proteins shifted from those active in biosynthesis and<br />

metabolism to those with roles in storage and protection against biotic and abiotic stresses. Few proteins<br />

responded transiently when plants were transferred to the high temperature regimens, but levels <strong>of</strong><br />

a number <strong>of</strong> proteins were altered during late stages <strong>of</strong> grain development. Specific protein responses<br />

depended on whether the high temperature regimens were initiated early or mid development. Some <strong>of</strong><br />

the heat responsive proteins have been implicated in gas bubble stabilization in bread dough and others<br />

are suspected food allergens.<br />

Published by Elsevier Ltd.<br />

Kothari, 2004; Hurkman et al., 2003). High temperatures during<br />

grain fill influence gluten protein accumulation levels. The relative<br />

amounts <strong>of</strong> certain a-gliadins and HMW-GS were higher and those<br />

for certain LMW-GS were lower in grain produced under a 37 C/<br />

28 C day/night regimen (DuPont et al., 2006b). In addition, accumulation<br />

rates increased more for the a-gliadins and HMW-GS than<br />

the LMW-GS in grain produced under this regimen (DuPont et al.,<br />

2006a). High temperature during grain fill also affects the accumulation<br />

levels <strong>of</strong> stress/defense proteins. Skylas et al. (2002)<br />

reported that heat shock, a shift from a 24 C/18 Ctoa40 C/25 C<br />

day/night regimen from 15–17 dpa, increased the number <strong>of</strong> small<br />

heat shock protein (HSP) is<strong>of</strong>orms in total protein extracts from<br />

endosperm. Majoul et al. (2003) reported that levels <strong>of</strong> HSPs as well<br />

as proteins that defend against reactive oxygen species (ROS) and<br />

desiccation increased in flour when developing grain was exposed<br />

to an extended high temperature day/night regimen <strong>of</strong> 34 C/10 C<br />

from anthesis to maturity. Majoul et al. (2004) isolated an albumin<br />

and globulin fraction from mature grain and found that the levels <strong>of</strong><br />

a number <strong>of</strong> proteins, including HSPs as well as enzymes involved<br />

in starch biosynthesis, carbohydrate metabolism, and ATP<br />

synthesis, responded to the high temperature regimen.<br />

In this study, we utilize proteome maps to develop a comprehensive<br />

picture <strong>of</strong> endosperm development and identify the effects<br />

<strong>of</strong> high temperature on protein composition during this crucial<br />

reproductive stage. Plants were grown under three temperature<br />

conditions – a moderate temperature regimen maintained


throughout grain development and a high temperature regimen<br />

initiated early or mid development and maintained until maturity.<br />

The high temperature regimens are more severe than might be<br />

encountered under field conditions, but were selected to accentuate<br />

protein responses. KCl-soluble/methanol-insoluble albumins<br />

and globulins were isolated from the endosperm at selected time<br />

points during grain development and separated by 2-DE. These<br />

proteins function in a wide range <strong>of</strong> cellular processes (Vensel et al.,<br />

2005), including the synthesis <strong>of</strong> gluten proteins and starch, major<br />

determinants <strong>of</strong> wheat yield and grain quality. Protein accumulation<br />

pr<strong>of</strong>iles and functions were compared across physiologically<br />

equivalent stages <strong>of</strong> grain fill to define major events <strong>of</strong> endosperm<br />

development and identify specific responses to high temperature.<br />

Protein levels in mature grain produced under the three temperature<br />

regimens were also compared to discover changes in protein<br />

composition related to grain quality.<br />

2. Materials and methods<br />

2.1. Plant material<br />

Triticum aestivum, L. cv. Butte 86, was grown in climatecontrolled<br />

greenhouses under moderate and high temperature<br />

regimens (Altenbach et al., 2007b). For the moderate temperature<br />

regimen, plants were grown at a maximum daytime temperature <strong>of</strong><br />

24 C and a minimum nighttime temperature <strong>of</strong> 17 C. For the high<br />

temperature regimen, plants were transferred at 10 or 20 days postanthesis<br />

(dpa) to a second greenhouse and grown at a maximum<br />

daytime temperature <strong>of</strong> 37 C and a minimum nighttime temperature<br />

<strong>of</strong> 28 C. Water and fertilizer (Plantex 20-20-20, 300 mg/day)<br />

were applied by drip irrigation. Natural light was supplemented<br />

with 100 W high-pressure sodium lamps to maintain a day length <strong>of</strong><br />

16 h. Heads were collected at selected time points during grain<br />

development. Under the moderate temperature regimen, heads<br />

were collected at 4-day intervals from 10 to 38 dpa and at 40 dpa.<br />

Because high temperature shortened the grain developmental<br />

program, heads were collected at 2-day intervals from 10 to 26 dpa<br />

for the high temperature regimen initiated at 10 dpa and from 20 to<br />

28 dpa for the high temperature regimen initiated at 20 dpa. The<br />

end point for each regimen was the oldest age that endosperm could<br />

be squeezed or scraped from the grain. Grains were removed from<br />

the heads and the region containing the embryo was excised with<br />

a razor blade. The endosperm was squeezed through the resultant<br />

opening in the pericarp/testa. Endosperm was transferred immediately<br />

into tubes cooled in liquid nitrogen and stored at 80 C.<br />

Endosperm was also collected from plants grown in three additional<br />

experiments in which a 37 C/28 Cora37 C/17 C high temperature<br />

regimen with 300 mg per day fertilizer was initiated at anthesis<br />

and a 37 C/28 C high temperature regimen with 150 mg per day<br />

fertilizer was initiated at 15 dpa.<br />

2.2. 2-DE analysis<br />

A KCl-soluble/methanol-insoluble albumin and globulin fraction<br />

was prepared from the endosperm and triplicate 2-D gels loaded with<br />

equal protein amounts from each time point were run as described<br />

previously (Hurkman and Tanaka, 2004; Vensel et al., 2005). Gels<br />

were digitized with a calibrated scanner (UMAX Powerlook III; Dallas,<br />

TX) at 300 dpi using the same settings for all gels. Computer s<strong>of</strong>tware<br />

(Progenesis PG240 Ver. 2006; Nonlinear Dynamics Limited, Newcastle<br />

uponTyne, UK) was used to match proteinpatterns and analyze<br />

the quantitative and qualitative differences across time points for the<br />

three temperature regimens; matching was validated manually for<br />

all spots. Protein accumulation pr<strong>of</strong>iles were constructed from<br />

the normalized spot volume (individual spot volume/total spot<br />

volume 100) data and validated manually.<br />

W.J. Hurkman et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 12–23 13<br />

2.3. Protein identification<br />

Proteins contained in 2-D gel spots were digested with trypsin<br />

using a DigestPro gel-spot-processing robot (Intavis, Lagenfeld,<br />

Germany). Peptides from the tryptic digests were identified using<br />

a QSTAR PULSAR i quadrupole time-<strong>of</strong>-flight (TOF) mass spectrometer<br />

(Applied Biosystems/MDS SCIEX, Toronto, Canada) as<br />

previously described (Vensel et al., 2005). The resulting AnalystQS<br />

wiff files were converted to MGF files using Mascot Daemon (http://<br />

www.matrixscience.com/) and data analysis <strong>of</strong> the MS/MS files was<br />

carried out using a locally installed copy <strong>of</strong> the Global Proteome<br />

Machine Organization (GPM) s<strong>of</strong>tware (http://www.thegpm.org).<br />

X!Tandem (version 2005.06.01.2), the spectrum modeler in the<br />

GPM s<strong>of</strong>tware, was used to match MS/MS fragmentation data to<br />

peptide sequences. MGF files were submitted in batch mode to<br />

X!Tandem using a script provided by Jayson Faulkner (University <strong>of</strong><br />

Michigan). X!Tandem was configured to search a file containing<br />

695,000 protein sequence entries from all proteins in the<br />

HarvEST:Wheat version 1.04 database (http://harvest.ucr.edu/<br />

HWheat104.exe), NCBI nonredundant green plant database, NCBI<br />

Triticum aestivum: UniGene Build #37, and wEST Database (http://<br />

wheat.pw.usda.gov/wEST). A local installation <strong>of</strong> the GPM and the<br />

GPMDB (Craig et al., 2004) was used to analyze, store, and display<br />

the data. The protein identification and mass spectrometry data are<br />

available as XML files (http://wheat.pw.usda.gov/pubs/2008). To<br />

access the data, click on Hurkman, scroll down to Compressed<br />

directory <strong>of</strong> individual XML files for gel spots, click to download,<br />

and then unzip the files. To view the files, access the GPM viewer<br />

(http://h777.thegpm.org/tandem/thegpm_upview.html), click<br />

Browse, select the file for the spot number <strong>of</strong> interest, and click<br />

View Models. Click Protein to see the protein sequence and identified<br />

peptides.<br />

2.4. Statistical analyses<br />

The protein accumulation data was analyzed by hierarchical<br />

cluster analysis to determine similarities and differences across the<br />

developmental time courses for the moderate and high temperature<br />

regimens. The data set consisted <strong>of</strong> accumulation pr<strong>of</strong>iles for<br />

193 proteins. Each pr<strong>of</strong>ile consisted <strong>of</strong> 23 time points across the<br />

three temperature regimens. In order to discover patterns independent<br />

<strong>of</strong> protein abundance, the data for each protein were<br />

converted to a percent scale (normalized volume <strong>of</strong> a given time<br />

point/total normalized volume for the 23 time points 100).<br />

Hierarchical cluster analysis was performed with SAS s<strong>of</strong>tware,<br />

PROC CLUSTER, using the average method. Statistical algorithms,<br />

including CCC, pseudo t, and pseudo F were used to choose the final<br />

number <strong>of</strong> clusters (SAS Institute Inc., 2004).<br />

To determine reproducibility <strong>of</strong> the data for effects <strong>of</strong> high<br />

temperature on endosperm protein levels late in grain development,<br />

correlations were computed for the data from the high<br />

temperature regimens initiated at anthesis, 10, 15, and 20 dpa. To<br />

ensure that the correlations were not unduly influenced by large<br />

observations, the Kendall correlation method, based on ranks<br />

across all proteins, was performed using the SAS s<strong>of</strong>tware PROC<br />

CORR (SAS Institute Inc., 2004). In addition Cronbach’s alpha was<br />

computed to estimate consistency among regimens.<br />

3. Results<br />

3.1. Endosperm proteome maps<br />

In a previous study (Vensel et al., 2005), we developed proteome<br />

maps for albumins and globulins isolated from the endosperm <strong>of</strong><br />

grain grown under moderate temperatures (24 C/17 C, day/<br />

night). The maps, which contain 254 proteins, were established


14<br />

using two developmental time points, 10 and 36 dpa, to maximize<br />

proteome coverage. In the present study, we mapped and identified<br />

35 additional endosperm proteins (Fig. S1). Spot numbers, accession<br />

numbers, percent coverage, expectation values, and peptide<br />

sequence data for these proteins are listed in Table S1. Of the 289<br />

proteins in the updated proteome maps, 193 were selected for<br />

pr<strong>of</strong>ile analysis. Spots that contained more than one protein, 90 in<br />

all, were eliminated from the experimental set, because normalized<br />

volumes for the individual proteins within these spots could not be<br />

determined. Since this study focuses on the albumins and globulins,<br />

the 6 spots containing gliadins and glutenins that co-purified<br />

with this fraction (Vensel et al., 2005) also were eliminated from<br />

the data set.<br />

3.2. Hierarchical cluster analysis<br />

Accumulation pr<strong>of</strong>iles based on normalized spot volumes were<br />

established for proteins in endosperm harvested from grain grown<br />

under the moderate and high temperature regimens (Fig. 1). The<br />

data set for the high temperature regimen initiated at 20 dpa<br />

includes spot volumes from the three initial time points (10, 14, and<br />

18 dpa) <strong>of</strong> the moderate temperature regimen under which the<br />

plants were grown before transfer to this high temperature regimen.<br />

When normalized protein volumes were summed for all time points<br />

across the three temperature regimens for each protein, the totals<br />

spanned four orders <strong>of</strong> magnitude, ranging from 0.13 for an is<strong>of</strong>orm<br />

<strong>of</strong> globulin-2 (#561) to 118 for protein disulfide isomerase (PDI,<br />

#118). Because <strong>of</strong> this wide dynamic range, the normalized volume<br />

data were converted to percentages so that hierarchical cluster<br />

analysis results would be independent <strong>of</strong> protein amount. In Fig. 1,<br />

the protein pr<strong>of</strong>ile data were converted to colors: white represents<br />

the lowest percentage range (0–0.75), shades <strong>of</strong> blue the intermediate<br />

percentage ranges (>0.75–40), and black the highest<br />

percentage range (>40). The identities <strong>of</strong> the proteins corresponding<br />

to the pr<strong>of</strong>iles depicted in Fig. 1 and their functions are listed in<br />

Table 1 by the order and cluster numbers determined by the hierarchical<br />

cluster analysis (Fig. S2). The order numbers in Table 1 may<br />

be used to identify proteins shown in Fig. 1. The spot numbers in<br />

Table 1 may be used to locate the spots on the proteome maps in<br />

Fig. S1 <strong>of</strong> this paper and Figs. 1 and 2 <strong>of</strong> Vensel et al. (2005).<br />

The hierarchical cluster analysis separated the 193 protein<br />

accumulation pr<strong>of</strong>iles into 39 clusters (Fig. 1 and Fig. S2), but nearly<br />

65% <strong>of</strong> the proteins in the analysis were members <strong>of</strong> four clusters (1,<br />

2, 3, and 5). Cluster 1 with 58 members was the largest cluster. The<br />

majority <strong>of</strong> these proteins functioned in carbohydrate metabolism,<br />

protein synthesis/assembly, and nitrogen metabolism. Cluster 5<br />

with 40 members was the second largest cluster. The majority <strong>of</strong><br />

the proteins in this cluster functioned in stress/defense, although<br />

some also functioned in carbohydrate and nitrogen metabolism.<br />

The next largest clusters were 2 and 3 with 16 and 13 members,<br />

respectively. In contrast to cluster 1, the majority <strong>of</strong> the proteins<br />

in these clusters functioned in transcription/translation. The<br />

remaining clusters, 4 and 6–39, were smaller, containing 1–10<br />

members, and the majority <strong>of</strong> these proteins functioned in carbohydrate<br />

metabolism, protein synthesis/assembly, storage and<br />

stress/defense. The most abundant proteins (total normalized<br />

volume >15, indicated by purple boxes in Fig. 1 and an H in Table 1)<br />

were members <strong>of</strong> the four largest clusters (1, 2, 3, and 5). The least<br />

abundant proteins (total normalized volume


W.J. Hurkman et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 12–23 15<br />

Fig. 1. Accumulation patterns <strong>of</strong> KCl-soluble/methanol-insoluble albumins and globulins in wheat endosperm. Proteins were isolated from grain produced under moderate (24 C/<br />

17 C) and high temperature conditions (37 C/28 C from 10 or 20 dpa). Pr<strong>of</strong>iles are shown as percent <strong>of</strong> total normalized volume. Hierarchical clusters (1–39) and accompanying<br />

protein order were determined by SAS analysis (see Fig. S1). Protein identifications are in Table 1. The most abundant and least abundant proteins, based on total normalized spot<br />

volume for the three treatments, are indicated in the ‘Abundance’ column. Differences in pr<strong>of</strong>iles between the 24 C/17 C and the 37 C/28 C from 10 or 20 dpa regimens are<br />

indicated in the ‘Changes with Heat’ columns.


16<br />

W.J. Hurkman et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 12–23<br />

Table 1<br />

Identities <strong>of</strong> endosperm proteins corresponding to the accumulation pr<strong>of</strong>iles shown in Fig. 1<br />

Order no. Cluster no. Spot no. Name Function a Peak timing b<br />

24/17 37/28 10 dpa 37/28 20 dpa<br />

1 1 264 Glyceraldehyde 3-P dehydrogenase (NAD) CM E E E H<br />

2 1 294 Malate dehydrogenase (NAD) CM E E E H<br />

3 1 262 Glyceraldehyde 3-P dehydrogenase (NAD) CM E E E H<br />

4 1 112 PPi-fructose 6-P 1-phosphotransferase b-subunit CM E E E<br />

5 1 252 Aldolase CM E E E H<br />

6 1 383 Cyclophilin A-2 PS E E E H<br />

7 1 97 Fructokinase CM E E E<br />

8 1 118 Protein disulfide isomerase PS E E E H<br />

9 1 176 Alanine amino transferase 2 NM E E E H<br />

10 1 138 Ketol-reductoisomerase NM E E E<br />

11 1 154 Aldehyde dehydrogenase CM E E E<br />

12 1 250 Reversibly glycosylated polypeptide CM E E E<br />

13 1 354 Triosephosphate isomerase CM E E E<br />

14 1 6 Carbamoyl phosphate synthetase NM E E E L<br />

15 1 69 Poly(A)-binding protein TT E E E<br />

16 1 72 10-Formyltetrahydr<strong>of</strong>olate synthetase NM E E E<br />

17 1 159 ADP-glucose PPase, SS CM E E E<br />

18 1 433 Gycine-rich RNA-binding protein TT E E E<br />

19 1 65 Heat shock protein 70 PS E E E<br />

20 1 145 Selenium binding protein S E E E<br />

21 1 103 Phosphoglycerate mutase, 2,3-bisphosphoglycerate-independent CM E E E<br />

22 1 104 Phosphoglycerate mutase, 2,3-bisphosphoglycerate-independent CM E E E<br />

23 1 96 Phosphoglucomutase CM E E E<br />

24 1 173 Alanine amino transferase 2 NM E E E<br />

25 1 135 Leucine amino peptidase PT E E E<br />

26 1 295 Malate dehydrogenase (NAD) CM E E E<br />

27 1 48 Methionine synthase NM E E E<br />

28 1 301 Guanine nucleotide-binding protein b subunit-like protein ST E E E<br />

29 1 415 Ubiquitin-protein ligase PT E E E<br />

30 1 444 40S Ribosomal protein S21 PS E E E<br />

31 1 157 UDP-glucose PPase CM E E E<br />

32 1 158 UDP-glucose PPPase CM E E E<br />

33 1 411 Gycine-rich RNA-binding protein TT E E E<br />

34 1 422 40S Ribosomal protein S12 PS E E E<br />

35 1 153 Aldehyde dehydrogenase S E E E<br />

36 1 655 S-Adenosylmethionine synthetase 2 NM E E E<br />

37 1 98 Acetohydroxyacid synthase NM E E E<br />

38 1 25 Aconitase CM E E E<br />

39 1 27 Aconitase CM E E E<br />

40 1 179 Enolase CM E E E<br />

41 1 185 Enolase CM E E E H<br />

42 1 161 Dihydrolipoamide dehydrogenase CM E E E<br />

43 1 82 Stress-induced protein, sti1-like S E E E<br />

44 1 117 Chaperonin 60 kDa b-subunit PS E E E<br />

45 1 1004 Fructose bisphosphate aldolase CM E E E<br />

46 1 373 20S Proteasome a-subunit B PT E E E<br />

47 1 278 TGF-b receptor-interacting protein 1 ST E E E<br />

48 1 306 Aldolase CM E E E<br />

49 1 163 Heat shock associated protein PS E E E<br />

50 1 74 DNAK-type molecular chaperone HSP70 PS E E E<br />

51 1 149 Heat shock associated protein PS E E E<br />

52 1 285 Legumin-like protein SP E E E<br />

53 1 61 Poly(A)-binding protein TT E E E L<br />

54 1 378 GSH-dependent dehydroascorbate reductase 1 S E E E<br />

55 1 447 Polyubiquitin 6 PT E E E<br />

56 1 49 Heat shock protein 80-2 PS E E E<br />

57 1 471 DNAK-type molecular chaperone HSP70 PT E E E<br />

58 2 342 Ascorbate peroxidase S E E E<br />

59 2 344 Ascorbate peroxidase S E E E<br />

60 2 119 Phosphoglycerate dehydrogenase-like protein CM E E E<br />

61 2 193 6-Phosphogluconate dehydrogenase CM E E E<br />

62 2 657 S-Adenosylmethionine synthetase 1 NM E E E<br />

63 2 70 Heat shock protein 70 PS E E E<br />

64 2 178 Enolase CM E E E<br />

65 2 53 Poly(A)-binding protein TT E E E L<br />

66 2 66 Poly(A)-binding protein TT E E E L<br />

67 2 57 Poly(A)-binding protein TT E E E L<br />

68 2 60 Poly(A)-binding protein TT E E E<br />

69 2 63 Poly(A)-binding protein TT E E E L<br />

70 2 31 Pyruvate Pi dikinase CM E E E<br />

71 2 55 Poly(A)-binding protein TT E E E<br />

72 2 493 UDP-glucose dehydrogenase PS E E E<br />

73 2 529 40S Ribosomal protein S19 PS E E E<br />

74 3 361 20S Proteasome a-subunit PT E E E<br />

Abundance c


Table 1 (continued )<br />

W.J. Hurkman et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 12–23 17<br />

Order no. Cluster no. Spot no. Name Function a Peak timing b<br />

24/17 37/28 10 dpa 37/28 20 dpa<br />

Abundance c<br />

75 3 399 Translationally controlled tumor protein TT E E E<br />

76 3 362 Elongation factor 1-b PS E E E<br />

77 3 94 RNA-binding protein, similarity TT E E E L<br />

78 3 182 Tubulin a-3 chain CD E E E<br />

79 3 412 Glycine-rich RNA-binding protein TT E E E<br />

80 3 201 SGT1 S E E E<br />

81 3 667 Ascorbate peroxidase S E E E<br />

82 3 348 Ascorbate peroxidase S E E E<br />

83 3 330 14-3-3 protein ST E E E<br />

84 3 640 Phosphoglycerate kinase CM E E E<br />

85 3 279 Adenosine kinase A E E E<br />

86 3 93 RNA-binding protein, similarity TT E E E<br />

87 4 42 Heat shock protein, 82 K PS E E E<br />

88 5 350 Small ras-related GTP-binding protein T M M M<br />

89 5 382 Dehydroascorbate reductase S M M M H<br />

90 5 356 Triosephosphate isomerase CM M M M<br />

91 5 110 2-Isopropylmalate synthase NM M M M<br />

92 5 205 26S Proteasome regulatory particle triple-A ATPase subunit 4 PT M M M<br />

93 5 232 Aspartate amino transferase NM M M M<br />

94 5 175 Globulin-2 SP M M M<br />

95 5 353 Avenin SP M M L H<br />

96 5 417 a-Amylase/trypsin inhibitor, CM3 S M M L H<br />

97 5 43 Late embryogenesis abundant protein-like S M L L L<br />

98 5 379 Expressed protein U M M M<br />

99 5 430 a-Amylase inhibitor 0.19 S M M M H<br />

100 5 111 PPi-fructose-6-P 1-phosphotransferase CM M M M<br />

101 5 436 a-Amylase inhibitor 0.53 S M M M<br />

102 5 384 Avenin N9 SP M L M H<br />

103 5 440 a-Amylase inhibitor Ima1, monomeric S M M M<br />

104 5 381 a-Amylase/subtilisin inhibitor S M L L H<br />

105 5 386 F10K1.21/F7A7_100 protein, similarity U M L L H<br />

106 5 245 Serpin WZS3 S M L M<br />

107 5 369 Avenin N9 SP E M E<br />

108 5 434 a-Amylase inhibitor 0.53 S M M L<br />

109 5 233 Serpin S E E M<br />

110 5 236 Aspartate amino transferase NM E E M<br />

111 5 231 Serpin WZS2 S E E M H<br />

112 5 246 Serpin S L M M H<br />

113 5 237 Serpin WZS3 S L M M<br />

114 5 33 Elongation factor 2 PS L M L<br />

115 5 305 Glyoxalase I S M L L<br />

116 5 128 ADP-glucose PPase, LS CM M M M<br />

117 5 132 ADP-glucose PPase, LS CM M M M<br />

118 5 130 ADP-glucose PPase, LS CM M M M<br />

119 5 148 Catalase isozyme 1 S M M M<br />

120 5 244 Aldolase CM M M M<br />

121 5 251 Formate dehydrogenase NM M M M<br />

122 5 249 Formate dehydrogenase NM M M M<br />

123 5 206 Eukaryotic initiation factor 4A TT M E M<br />

124 5 312 Purple acid phosphatase S L E M<br />

125 5 137 Leucine amino peptidase PT L L L<br />

126 5 418 Superoxide dismutase [Cu–Zn] S M L L<br />

127 5 102 Protein disulfide isomerase-like protein PS L M L<br />

128 6 141 Elongation factor 1-a PS L L L<br />

129 6 337 Seed globulin SP L M L<br />

130 6 336 Seed globulin SP L M L<br />

131 6 327 Seed globulin SP L M M<br />

132 6 335 Seed globulin SP L M M<br />

133 7 1009 Peroxidase 1 S L M M<br />

134 8 199 Triticin SP M L M<br />

135 9 255 Peroxidase 1 S L L L<br />

136 9 259 Peroxidase 1 S L L L<br />

137 9 263 Peroxidase BP1 S L L L<br />

138 9 371 Peroxiredoxin S L L L<br />

139 9 314 OSJNBb0118P14.5 U L L L<br />

140 9 324 OSJNBb0118P14.5 U L L L L<br />

141 9 168 Dihydrolipoamide dehydrogenase CM L L L<br />

142 9 212 DNAK-type molecular chaperone HSP70 PS L L L<br />

143 9 441 Avenin N9 SP L L L L<br />

144 9 639 Serpin S L L L<br />

145 10 407 Heat shock protein 16.9C PS M M M<br />

146 11 225 Heat shock protein 70 PS E E E<br />

147 11 338 DNA-binding protein HEXBP TT E E E<br />

148 12 666 Ascorbate peroxidase S E E E<br />

149 13 341 Chitinase-c S L L L<br />

(continued on next page)


18<br />

Table 1 (continued )<br />

Order no. Cluster no. Spot no. Name Function a Peak timing b<br />

24/17 37/28 10 dpa 37/28 20 dpa<br />

150 13 351 Chitinase-a S L L L<br />

151 13 282 Chitinase-a S L L L<br />

152 13 842 Late embryogenesis abundant protein-like S L L L L<br />

153 13 1001 Unknown protein [similar to late embryogenesis abundant protein-like] S L L L<br />

154 13 557 Embryo-specific protein U L L L L<br />

155 13 551 Late embryogenesis abundant protein-like S L L L<br />

156 13 427 Pathogenesis-related protein 4 S L L L<br />

157 14 318 Xylanase inhibitor I S L L L<br />

158 14 550 Sucrose synthase type 2 CM L L L<br />

159 14 633 Nonspecific lipid-transfer protein T L L L<br />

160 15 355 Seed globulin SP L L L<br />

161 15 591 Glucose and ribitol dehydrogenase CM L L L<br />

162 15 275 Glyceraldehyde 3-P dehydrogenase (NAD) CM L L L<br />

163 15 347 20S Proteasome a-subunit PT L L L<br />

164 15 420 Nucleoside diphosphate kinase I A L L L<br />

165 15 281 Globulin Beg 1 SP E L E<br />

166 16 147 Globulin-2 SP L L L<br />

167 17 376 Superoxide dismutase [Mn] S L L L<br />

168 18 155 Globulin-2 SP L L L<br />

169 18 620 Globulin Beg 1 SP L L L<br />

170 19 631 Globulin-like protein SP L L L L<br />

171 20 1012 Peroxidase 1 S L A M<br />

172 20 1014 Peroxidase 1 S L A M<br />

173 21 310 Glucose and ribitol dehydrogenase CM L M E<br />

174 22 283 TGF-b receptor-interacting protein 1 ST E A E<br />

175 23 359 Peroxidase 1 S E M E<br />

176 24 1010 Peroxidase 1 S M L M L<br />

177 25 419 Glycine-rich RNA-binding protein TT M E E<br />

178 26 672 Reversibly glycosylated polypeptide CM E E E<br />

179 27 357 Triosephosphate isomerase CM E E E<br />

180 28 869 60S Ribosomal protein L12 PS L L L<br />

181 29 125 b-Amylase CM L L L<br />

182 30 1007 Glyceraldehyde 3-P dehydrogenase (NAD) CM L A L<br />

183 31 559 Globulin-2 SP L L L<br />

184 31 851 Globulin-2 SP L L L L<br />

185 32 852 Globulin-2 SP L L A L<br />

186 33 871 Barwin S L L L<br />

187 34 561 Globulin-2 SP L L A L<br />

188 35 1022 Embryo globulin SP A L A L<br />

189 36 1002 Embryo globulin SP L L L L<br />

190 36 1003 Embryo globulin SP L L L<br />

191 37 311 Initiation factor 3g TT E E E L<br />

192 38 530 Cyclophilin A-1 PS E E E<br />

193 39 364 Ascorbate peroxidase S E E E L<br />

analysis <strong>of</strong> physiologically equivalent stages shows that the high<br />

temperature regimens did not substantially alter the overall pattern<br />

<strong>of</strong> protein accumulation during grain development.<br />

3.4. Transient effects <strong>of</strong> high temperature on protein pr<strong>of</strong>iles during<br />

grain development<br />

Developmental pr<strong>of</strong>iles were examined for transient responses<br />

following transfer <strong>of</strong> plants to the high temperature regimens<br />

initiated at 10 and 20 dpa. Proteins that increased (yellow boxes in<br />

Fig. 1) or decreased (purple boxes in Fig. 1) transiently in response<br />

to the high temperature regimens are listed in Table 2. Under the<br />

high temperature regimen initiated at 10 dpa, two HSP70 is<strong>of</strong>orms<br />

(225, 471), legumin-like protein (285), and triosephosphate isomerase<br />

(354) increased transiently early in development. HSP16.9<br />

(407), four globulin is<strong>of</strong>orms (327, 335–337), and a peroxidase<br />

(1009) increased while three is<strong>of</strong>orms <strong>of</strong> ADP-glucose PPase large<br />

subunit (128, 130, 132) decreased transiently during mid development<br />

under this regimen. Under the high temperature regimen<br />

initiated at 20 dpa, only two transient changes were detected.<br />

W.J. Hurkman et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 12–23<br />

Abundance c<br />

Protein clusters were determined using SAS as outlined in Section 2. Spot numbers correspond to 2-DE map locations in Fig. S1 and Figs. 1 and 2 in Vensel et al. (2005).<br />

a<br />

A, ATP interconversion; CD, cell division; CM, carbohydrate metabolism; NM, nitrogen metabolism; PS, protein synthesis/assembly; PT, protein turnover; S, stress/defense;<br />

SP, storage protein; ST, signal transduction; T, transport; TT, transcription/translation; U, unknown.<br />

b<br />

E, early; M, middle; L, late stages. See text for time points included for each <strong>of</strong> the temperature regimens.<br />

c<br />

H, high and L, low abundance proteins.<br />

HSP16.9 (407) increased and triticin (199) decreased transiently<br />

during mid development.<br />

Because the proteome maps were developed for endosperm<br />

proteins isolated from grain produced under the moderate<br />

temperature regimen, spots listed in Table 2 were excised from 2-D<br />

gels <strong>of</strong> endosperm proteins extracted from grain produced under<br />

the high temperature regimens and re-identified by MS/MS. The<br />

original protein identifications were confirmed for all spots. The<br />

peptide sequence data, accession numbers, percent coverage, and<br />

expectation values are listed in Table S2.<br />

3.5. Effects <strong>of</strong> high temperature regimens on endosperm protein<br />

levels late in grain development<br />

The high temperature regimens altered the accumulation levels<br />

<strong>of</strong> a number <strong>of</strong> proteins in the mature grain. Proteins that increased<br />

or decreased 1.8-fold or more relative to the moderate temperature<br />

regimen are listed in Table 3. Under the high temperature regimen<br />

initiated at 10 dpa, 31 proteins increased, the majority <strong>of</strong> which<br />

functioned in stress/defense and storage. The stress/defense


Average fresh weight/kernel (mg)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0<br />

10 20 30 40 50<br />

Days Post Anthesis (DPA)<br />

Fig. 2. Effect <strong>of</strong> high temperature on grain fresh weight. Wheat plants were grown<br />

under three day/night temperature regimens: -, 24 C/17 C; C, 37 C/28 C initiated<br />

at 10 dpa; and B, 37 C/28 C initiated at 20 dpa.<br />

proteins included a-amylase/subtilisin inhibitor (381), barwin/PR-4<br />

protein (427, 871), chitinase (282, 341, 351), late embryogenesis<br />

abundant (LEA) protein (551, 1001), and xylanase inhibitor protein<br />

(318). It should be noted that lipid-transfer protein (LTP, 633),<br />

which is classified as a transport protein, may also have a role in<br />

stress/defense (Jung et al., 2003; Wu et al., 2004). Nearly all <strong>of</strong> the<br />

storage proteins that increased in response to high temperature<br />

were is<strong>of</strong>orms <strong>of</strong> globulin-2 (147, 155, 559, 561, 851, 852) or globulin<br />

Beg 1 (281, 620, 1002, 1003). In addition, 32 proteins decreased<br />

Number <strong>of</strong> Proteins<br />

40<br />

30<br />

20<br />

10<br />

0<br />

30<br />

20<br />

10<br />

0<br />

30<br />

20<br />

10<br />

0<br />

Carb. Metab.<br />

N Metab.<br />

Pro. Syn./Assem.<br />

Pro. Turnover<br />

EARLY<br />

MID<br />

LATE<br />

Stress/Defense<br />

Function<br />

Storage Pro.<br />

Signal Trans.<br />

Transcrip./Transl.<br />

Fig. 3. Effect <strong>of</strong> high temperature during grain fill on the timing <strong>of</strong> biochemical<br />

processes in wheat endosperm. See text for specific time points included in the early<br />

(A), mid (B), and late (C) stages for each <strong>of</strong> the three temperature regimens. ,, 24 C/<br />

17 C; ,37 C/28 C initiated at 10 dpa; and -, 37 C/28 C initiated at 20 dpa.<br />

W.J. Hurkman et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 12–23 19<br />

A<br />

B<br />

C<br />

under this high temperature regimen and the majority <strong>of</strong> these<br />

function in carbohydrate metabolism, including starch biosynthesis,<br />

and protein synthesis.<br />

Under the high temperature regimen initiated at 20 dpa, only 10<br />

proteins increased and, like the regimen initiated at 10 dpa, the<br />

majority functioned in stress/defense and storage (Table 3). Eight <strong>of</strong><br />

these proteins also increased during the high temperature regimen<br />

initiated at 10 dpa, but the increases in sucrose synthase type 2<br />

(550), aspartate amino acid transferase (236), and Cu–Zn superoxide<br />

dismutase (418) were specific for the regimen initiated at<br />

20 dpa. In addition, 15 proteins decreased, including 7 that also<br />

decreased under the high temperature regimen initiated at 10 dpa.<br />

However, decreases in 40S ribosomal protein S12 (422) and LEA<br />

protein (842) were specific for the 20 dpa regimen. Six proteins,<br />

including five globulins (147, 561, 620, 631, 852) and a 40S ribosomal<br />

protein S21 (444), that increased under the high temperature<br />

regimen initiated at 10 dpa, decreased under the 20 dpa regimen.<br />

Like the proteins that responded transiently to high temperature,<br />

proteins whose levels were altered late in development were also reidentified<br />

by MS/MS (Table S2). The original protein identifications<br />

were confirmed for all but spot 69. This spot was initially identified as<br />

a poly(A)-binding protein, but was re-identified as formate tetrahydr<strong>of</strong>olate<br />

ligase. This is likely due to slight differences in the 2-DE<br />

pattern in this region <strong>of</strong> the gel, illustrating the importance <strong>of</strong> confirming<br />

identifications <strong>of</strong> proteins in the high temperature proteomes.<br />

Nine <strong>of</strong> the re-identified spots contained peptides from one<br />

or two additional proteins (superscript a in Table 3). Because the<br />

proteins in these spots comigrate in the 2-D gels, it is not possible to<br />

know which proteins are responding to high temperature. Nonetheless,<br />

these identifications are valuable because they provide<br />

additional candidates for heat responsive endosperm proteins.<br />

3.6. Protein responses to high temperature regimens initiated at<br />

anthesis and 15 dpa<br />

Accumulation pr<strong>of</strong>iles <strong>of</strong> albumins and globulins were analyzed<br />

in independent experiments in which grain was produced under<br />

a24 C/17 C regimen and one <strong>of</strong> three additional high temperature<br />

regimens: 37 C/17 C initiated at anthesis, 37 C/28 C initiated<br />

at anthesis, or 37 C/28 C initiated at 15 dpa. Cronbach’s<br />

alpha was computed for the control data from the last time point<br />

in these three experiments and the original experiment where<br />

heat was applied from 10 and 20 dpa. For this analysis, proteins<br />

that were absent at the last time point in one or more experiments<br />

were deleted, resulting in a data set <strong>of</strong> 163 proteins for<br />

which the calculated Cronbach’s alpha was 0.96 out <strong>of</strong> a maximum<br />

<strong>of</strong> 1.0. This high level <strong>of</strong> consistency among the controls justified<br />

a comparison <strong>of</strong> the data obtained for the four high temperature<br />

experiments. At the last time point for each protein, the ratio <strong>of</strong><br />

the normalized volume under the high temperature regimens to<br />

the moderate temperature regimen was calculated. Proteins that<br />

were absent at the last time point for one or more experiments<br />

were deleted, resulting in a data set containing 159 proteins, and<br />

the ratios were used to assign ranks to the proteins. Cronbach’s<br />

alpha was computed for these ranks at the last time point for the<br />

four high temperature experiments. The calculated Cronbach’s<br />

alpha <strong>of</strong> 0.60 showed a high degree <strong>of</strong> association among the four<br />

experiments. Individual correlations based on Kendall’s Tau rank<br />

correlation method showed that the highest correlation was<br />

between the 37 C/28 C regimens initiated at anthesis and 10 dpa<br />

(0.31 at p < 0.0001). The 37 C/28 C regimen initiated at 10 dpa<br />

was also highly correlated with the 37 C/17 C regimen initiated<br />

at anthesis (0.18 at p < 0.0007), but less correlated with the 37 C/<br />

28 C regimens initiated at 15 (0.12 at p < 0.022) and 20 dpa (0.19<br />

at p < 0.001). The 37 C/28 C regimens initiated at 15 and 20 dpa<br />

showed a high correlation with each other (0.22, significant at


20<br />

Table 2<br />

Proteins that undergo transient changes in accumulation levels in response to high temperatures during grain fill<br />

Spot no. Name Function a<br />

Order no. b<br />

Cluster no. Increase c<br />

Decrease<br />

10 dpa 20 dpa 10 dpa 20 dpa<br />

128 ADP-glucose PPase, LS CM 116 5 M<br />

130 ADP0-glucose PPase, LS CM 118 5 M<br />

132 ADP-glucose PPase, LS CM 117 5 M<br />

199 Triticin SP 134 8 M<br />

225 Heat shock protein 70 PS 146 11 E<br />

285 Legumin-like protein SP 52 1 E<br />

327 Seed globulin SP 131 9 M<br />

335 Seed globulin SP 132 6 M<br />

336 Seed globulin SP 130 6 M<br />

337 Seed globulin SP 129 6 M<br />

354 Triosephosphate isomerase CM 13 1 E<br />

407 Heat shock protein 16.9C PS 145 10 M M<br />

471 DNAK-type molecular chaperone HSP70 PT 57 1 E E<br />

1009 Peroxidase 1 S 133 7 M<br />

Identities <strong>of</strong> these proteins contained in these spots were confirmed by mass spectrometry. Accession numbers, percent coverage, expectation values, and peptide sequence<br />

data are listed in Table S2. XML files that contain the mass spectrometry data are available online (see Section 2.3).<br />

a<br />

Abbreviations for functional groups are: CM, carbohydrate metabolism; PS, protein synthesis/assembly; PT, protein turnover; S, stress/defense; SP, storage protein.<br />

b<br />

Order and cluster numbers were determined by hierarchal clustering. See Fig. 1 for accumulation pr<strong>of</strong>iles.<br />

c<br />

E, transient change early in development; M, transient change in mid development. See text for specific time points.<br />

p < 0.0001), lesser correlation with regimens initiated at 10 dpa as<br />

described above and even less with the 37 C/28 C regimen<br />

initiated at anthesis (0.10 and 0.16 at p < 0.056 and 0.004). Regimens<br />

initiated at 15 and 20 dpa were not related to the 37 C/<br />

17 C regimen initiated at anthesis.<br />

Like the 37 C/28 C regimen initiated at 10 dpa, a relatively<br />

large number <strong>of</strong> proteins responded to the high temperature regimens<br />

initiated at anthesis (Table 3). During the 37 C/28 C regimen<br />

28 proteins increased and 29 decreased and during the 37 C/17 C<br />

regimen 24 proteins increased and 22 decreased. A comparison <strong>of</strong><br />

the changes accompanying the high temperature regimens initiated<br />

at anthesis and 10 dpa revealed that 18 proteins increased and<br />

17 decreased under all three regimens. The majority <strong>of</strong> the proteins<br />

that increased functioned in storage and stress defense. Storage<br />

proteins included two globulin Beg 1 is<strong>of</strong>orms (281, 620), seven<br />

globulin-2 is<strong>of</strong>orms (147, 155, 559, 851, 852, 561, 631), and a seed<br />

globulin (355). The stress defense proteins were barwin/PR-4<br />

protein 4 (427), chitinase (351), LEA (551), and LTP (633). The<br />

majority <strong>of</strong> proteins that decreased functioned in carbohydrate<br />

metabolism and stress defense. Proteins with roles in carbohydrate<br />

metabolism were ADP-glucose PPase (130, 132), aldehyde dehydrogenase<br />

(154), and pyruvate Pi dikinase (31). Stress defense<br />

proteins were dehydroascorbate reductase (382), peroxidase<br />

(1009), purple acid phosphatase (312), and serpin (231). Like the<br />

37 C/28 C regimen initiated at 20 dpa, relatively few proteins<br />

responded when this regimen was initiated at 15 dpa (Table 3). Of<br />

the 11 proteins that increased and 17 that decreased under the<br />

regimen, only 4 proteins increased and 6 decreased under the<br />

37 C/28 C regimens initiated at 15 and 20 dpa. The proteins that<br />

increased have roles in stress/defense (341, chitinase; 427, barwin/<br />

PR-4 protein 4; 633, LTP) and storage (1002, globulin Beg 1). The<br />

proteins that decreased are involved in carbohydrate metabolism<br />

(193, 6-phosphogluconate dehydrogenase; 310, glucose and ribitol<br />

dehydrogenase), storage (631 and 852, globulin-2), protein<br />

synthesis (361, 20S proteosome a-subunit), and stress/defense<br />

(312, purple acid phosphatase).<br />

4. Discussion<br />

4.1. Endosperm development under the moderate<br />

temperature regimen<br />

The protein accumulation pr<strong>of</strong>iles developed in this study<br />

provide a dynamic picture <strong>of</strong> biological events that occur during<br />

wheat grain fill under a moderate temperature regimen. During<br />

W.J. Hurkman et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 12–23<br />

early grain fill, endosperm proteins have roles in seven functional<br />

processes: carbohydrate metabolism, nitrogen metabolism, protein<br />

synthesis/assembly, protein turnover, stress/defense, storage,<br />

signal transport, and transcription/translation. Based on the<br />

number <strong>of</strong> proteins, carbohydrate metabolism, protein synthesis/<br />

assembly, stress/defense, and transcription/translation are the<br />

principal functions at this stage <strong>of</strong> development. During mid grain<br />

fill, the number <strong>of</strong> stress/defense and storage proteins remains at<br />

earlier levels, but the number <strong>of</strong> proteins functioning in carbohydrate<br />

metabolism, nitrogen metabolism, protein synthesis/<br />

assembly, protein turnover, and transcription/translation<br />

decreases. Late in grain fill, the number <strong>of</strong> proteins that function in<br />

stress/defense and storage increase and proteins involved in<br />

nitrogen metabolism, signal transduction, and transcription/<br />

translation decrease to undetectable levels. This developmental<br />

program ensures that the endosperm synthesizes the requisite<br />

reserves for germination and seedling growth as well as proteins<br />

that protect these reserves from abiotic stresses and biotic invasion<br />

during grain desiccation and dormancy. Transcript analysis using<br />

micro arrays (Laudencia-Chingcuanco et al., 2007) and serial analysis<br />

<strong>of</strong> gene expression (SAGE) (McIntosh et al., 2007) revealed<br />

a similar sequence <strong>of</strong> events during wheat grain development.<br />

A more detailed picture <strong>of</strong> grain development emerges when<br />

the levels <strong>of</strong> specific proteins within the functional categories are<br />

compared. Early in grain fill, proteins involved in carbohydrate<br />

metabolism include those that function in glycolysis (glyceraldehyde-3-P<br />

dehydrogenase, phosphoglycerate kinase, phosphoglycerate<br />

mutase, enolase, pyruvate Pi dikinase), the citric acid cycle<br />

(malate dehydrogenase, aconitase), and starch biosynthesis (ADPand<br />

UDP-glucose PPase, phosphoglucomutase). During mid grain<br />

fill, enzymes functioning in glycolysis and the citric acid cycle were<br />

proportionately less abundant. ADP glucose PPase is present at<br />

lower levels, but UDP-glucose PPase and phosphoglucomutase<br />

decrease to undetectable levels. These decreases coincide with the<br />

decline in starch accumulation that accompanies this stage <strong>of</strong> grain<br />

development. Late in grain fill, proteins involved in carbohydrate<br />

metabolism function in starch (sucrose synthase) and glucose (bamylase,<br />

glucose and ribitol dehydrogenase) degradation rather<br />

than starch biosynthesis. Is<strong>of</strong>orms <strong>of</strong> glyceraldehyde-3-P dehydrogenase<br />

are again present. Since the wheat grain is undergoing<br />

desiccation at this developmental stage, this result is consistent<br />

with an earlier report that dehydration strongly increases the level<br />

<strong>of</strong> glyceraldehyde-3-P dehydrogenase in leaves <strong>of</strong> the resurrection<br />

plant, Craterostigma plantagineum, a species capable <strong>of</strong> withstanding<br />

severe desiccation (Velasco et al., 1994). However,


W.J. Hurkman et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 12–23 21<br />

Table 3<br />

Proteins that increased or decreased 1.8-fold or more in grain produced under high temperature regimens initiated at anthesis, 10, 15, or 20 dpa<br />

Spot Protein name Function Order Cluster 37 C/28 C 37 C/28 C 37 C/17 C 37 C/28 C<br />

no.<br />

no. no.<br />

Anthesis Anthesis 15 dpa<br />

10 dpa 20 dpa<br />

420 a<br />

Nucleoside diphosphate kinase A 164 15 2.09 B b<br />

275 a<br />

Glyceraldehyde 3-P dehydrogenase (NAD) CM 162 15 5.15 2.25 C C<br />

591 Glucose and ribitol dehydrogenase CM 161 15 2.79 C C<br />

110 2-Isopropylmalate synthase NM 91 5 2.57 B B<br />

444 40S Ribosomal protein S21 PS 30 1 2.58 2.33 , ,<br />

869 60S Ribosomal protein L12 PS 180 28 1.81 C<br />

163 Heat shock associated protein PS 49 1 1.83 1.83 C<br />

347 20S Proteasome a-subunit [proteasome subunit a type 6] PT 163 15 2.74 2.52 C<br />

381 a-Amylase/subtilisin inhibitor S 104 5 1.87 C<br />

871 Barwin [pathogenesis-related protein 4] S 186 33 8.29 - B C<br />

351 Chitinase-a [chitinase-c] S 150 13 2.39 C C C<br />

282 Chitinase-a [chitinase-c] S 151 13 2.21 C C<br />

341 Chitinase-c S 149 13 2.54 1.85 C C<br />

551 Late embryogenesis abundant protein-like S 155 13 3.00 C C<br />

427 Pathogenesis-related protein 4 [Barwin] S 156 13 4.22 1.82 C C C<br />

1001 Unknown protein [similar to late embryogenesis abundant<br />

protein]<br />

S 153 13 1.92 C C<br />

318 Xylanase inhibitor protein I S 157 14 1.86 C C<br />

1002 Globulin Beg 1 SP 189 36 8.33 1.83 C<br />

1003 Globulin Beg 1 SP 190 36 5.41 - B C<br />

281 Globulin Beg 1 [embryo globulin] SP 165 15 8.46 C C C<br />

620 Globulin Beg 1 [embryo globulin] SP 169 18 3.83 2.59 C C<br />

147 Globulin-2 SP 166 16 1.85 2.75 C C<br />

155 Globulin-2 SP 168 18 4.08 C C<br />

559 Globulin-2 SP 183 31 3.35 C C C<br />

851 Globulin-2 SP 184 31 2.92 2.38 - C<br />

852 Globulin-2 SP 185 32 3.86 - C ,<br />

561 Globulin-2 SP 187 34 3.78 - C<br />

631 Globulin-like protein [globulin-2] SP 170 19 2.07 6.43 - C B<br />

355 a<br />

Seed globulin SP 160 15 3.23 C C<br />

633 Nonspecific lipid-transfer protein T 159 14 2.53 2.54 C C C<br />

557 a<br />

Embryo-specific protein U 154 13 2.41 - C<br />

193 6-Phosphogluconate dehydrogenase CM 61 2 3.74 , ,<br />

132 ADP-glucose PPase, LS CM 117 5 2.20 B B B<br />

130 ADP-glucose PPase, LS CM 118 5 2.63 B B B<br />

154 a<br />

Aldehyde dehydrogenase CM 11 1 2.47 B ,<br />

1004 Aldolase CM 45 1 2.26<br />

310 Glucose and ribitol dehydrogenase CM 173 21 4.92 C C B<br />

31 Pyruvate Pi dikinase CM 70 2 3.34 B B<br />

158 UDP-glucose PPase CM 32 1 1.99 B B<br />

72 10-Formyltetrahydr<strong>of</strong>olate synthetase [formate tetrahydr<strong>of</strong>olate NM<br />

ligase]<br />

16 1 4.36 B B<br />

98 Acetohydroxyacid synthase NM 37 1 2.52 B ,<br />

138 a<br />

Ketol-reductoisomerase NM 10 1 3.25 B B B<br />

362 Elongation factor 1-b PS 76 3 7.30 , B B<br />

33 Elongation factor 2 PS 114 5 2.85 B B<br />

149 Heat shock associated protein PS 51 1 2.79 B B<br />

225 Heat shock protein 70 PS 146 11 2.01<br />

65 Heat shock protein 70 [DNAK-type molecular chaperone HSC70] PS 19 1 5.20 , B B<br />

118 Protein disulfide isomerase PS 8 1 2.08 B<br />

361 20S Proteasome a-subunit PT 74 3 4.83 1.81 , , B<br />

205 26S Proteasome PT 92 5 1.93 , C<br />

382 a<br />

Dehydroascorbate reductase S 89 5 2.09 B B<br />

1009 Peroxidase S 133 7 5.66 B ,<br />

312 a<br />

Purple acid phosphatase S 124 5 5.34 2.85 B B B<br />

246 Serpin S 112 5 2.27<br />

639 a<br />

Serpin S 144 9 2.98 2.14 B<br />

231 Serpin WZS2 S 111 5 1.92 B B<br />

441 Avenin N9 SP 143 9 1.85 1.95 , -<br />

285 Legumin-like protein SP 52 1 2.41<br />

330 14-3-3 Protein ST 83 3 2.50 , B<br />

301 Guanine nucleotide-binding protein ST 28 1 3.97 B B B<br />

60 Poly(A)-binding protein TT 68 2 2.80 1.87 , ,<br />

69 Poly(A)-binding protein [formate tetrahydr<strong>of</strong>olate ligase] TT 15 1 2.88 B<br />

550 Sucrose synthase type 2 CM 158 14 2.03<br />

236 Aspartate amino transferase NM 110 5 1.85<br />

422 40S Ribosomal protein S12 PS 34 1 4.35 , B<br />

418 Superoxide dismutase [Cu–Zn] S 126 5 2.32 C B<br />

842 Late embryogenesis abundant protein-like S 152 13 2.11 C C<br />

Identities <strong>of</strong> the proteins in these spots were confirmed by MS. Accession numbers, percent coverage, expectation values, and peptide sequence data are listed in Table S2.<br />

Brackets denote current database names for proteins identified in Vensel et al. (2005). Spot 69, previoulsy identified as poly(A)-binding protein, was identified as formate<br />

tetrahydr<strong>of</strong>olate ligase in this study. XML files that contain the mass spectrometry data are available online (see Section 2.3).<br />

a When reanalyzed by MS/MS, these spots contained peptides from one or two additional proteins, see Table S2.<br />

b Symbol key: B, decrease 1.8-fold or more; C, increase 1.8-fold or more; ,, absent under high temperature regimen; -, absent under moderate temperature regimen.


22<br />

a relationship between increased levels <strong>of</strong> this enzyme and drought<br />

tolerance has not been established.<br />

The major wheat storage proteins, the gliadins and glutenins,<br />

comprise more than 80% <strong>of</strong> endosperm protein and are found in the<br />

KCl-insoluble fraction. However, the KCl-soluble fraction contains<br />

a number <strong>of</strong> proteins with best matches to a variety <strong>of</strong> storage<br />

proteins, including avenin, triticin, legumin-like protein, seed<br />

globulin/19 kDa globulin, embryo globulin, globulin-2, or globulin<br />

Beg1. Avenin is a protein in oats that is similar to wheat gliadin.<br />

Triticin is a protein in wheat that is similar to pea legumin, an 11S<br />

globulin. Globulin-2 and globulin Beg1 are so named because they<br />

are similar to 7S globulins present in maize and barley, respectively.<br />

Like the gliadins and glutenins (DuPont et al., 2006a,b), the<br />

majority <strong>of</strong> these non-gluten storage proteins are most abundant<br />

late in grain fill. Eighteen globulins (one is<strong>of</strong>orm each <strong>of</strong> avenin and<br />

globulin-like protein; two is<strong>of</strong>orms <strong>of</strong> gobulin Beg1, three <strong>of</strong><br />

embryo globulins, six <strong>of</strong> globulin-2, five <strong>of</strong> seed globulin) were<br />

detected during late development while only two globulins (one<br />

is<strong>of</strong>orm each <strong>of</strong> avenin and legumin-like protein) were detected<br />

during early development and five (three is<strong>of</strong>orms <strong>of</strong> avenin and<br />

one each <strong>of</strong> globulin-2 and triticin) during mid development. The<br />

complement <strong>of</strong> globulin storage proteins is very complex and<br />

warrants further analysis, particularly since some <strong>of</strong> these proteins<br />

increase in response to high temperature.<br />

A range <strong>of</strong> stress/defense proteins is present throughout endosperm<br />

development, a finding in accord with transcript pr<strong>of</strong>ile<br />

analyses (McIntosh et al., 2007). A number <strong>of</strong> these proteins protect<br />

cells from reactive oxygen species (ROS) that are produced under<br />

normal growth conditions and can cause oxidative damage to<br />

proteins, DNA, and lipids. Thus, ascorbate peroxidase, peroxidase and<br />

GSH-dependent dehydroascorbate reductase accumulate early in<br />

grain fill. SGT1, a component <strong>of</strong> R-gene triggered disease resistance,<br />

and serpin, a serine protease inhibitor, are also present and may<br />

protect the developing grain against various pathogens. During<br />

mid grain fill, ROS-scavenging enzymes include dehydroascorbate<br />

reductase, peroxidase, catalase, and glyoxalase. Glyoxalase protects<br />

cells against reactive 2-oxoaldehydes generated by the triosephosphate<br />

isomerase reaction in glycolysis. Additional stress/<br />

defense proteins present at this stage <strong>of</strong> development include the<br />

a-amylase inhibitors and bifunctional a-amylase/trypsin inhibitors,<br />

which guard against digestive enzymes <strong>of</strong> insects and fungi. The<br />

levels <strong>of</strong> these inhibitors are actually much higher than those<br />

reported in this study, because many <strong>of</strong> these proteins partition into<br />

the KCl-soluble/methanol-soluble fraction (Wong et al., 2004). Late<br />

in grain fill, ROS-scavenging enzymes include peroxidase and<br />

superoxide dismutase. Pathogen resistance proteins present at this<br />

stage include serpin, chitinase, which hydrolyzes the structural<br />

carbohydrate <strong>of</strong> fungal cell walls, barwin/PR-4 protein, which is<br />

induced by fungal pathogens and binds chitin, and xylanase inhibitor<br />

protein, which inhibits a fungal enzyme that degrades plant cell<br />

walls. As the developing grain matures it undergoes desiccation, an<br />

event accompanied by the accumulation <strong>of</strong> LEA proteins, proteins<br />

known to increase in response to drought stress.<br />

4.2. Effect <strong>of</strong> high temperature on endosperm development<br />

The duration <strong>of</strong> grain fill is shortened by high temperature. Grain<br />

maturity is achieved earlier and is accompanied by corresponding<br />

decreases in kernel weight. These observations are in agreement<br />

with earlier studies inwhich maximum fresh weight, dry weight, and<br />

protein and starch content were lower in grain produced under high<br />

temperature regimens (37 C/17 Cor37 C/28 C) imposed from<br />

anthesis (Altenbach et al., 2002, 2003; DuPont et al., 2006b; Hurkman<br />

et al., 2003). Comparison <strong>of</strong> protein pr<strong>of</strong>iles revealed that<br />

although the high temperature regimens compressed the developmental<br />

program, they did not substantially alter it. Analysis <strong>of</strong> protein<br />

W.J. Hurkman et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 12–23<br />

pr<strong>of</strong>iles across physiologically equivalent stages under the three<br />

temperature regimens showed similar shifts in the timing <strong>of</strong> cellular<br />

functions during the transition from early to late development. The<br />

results in this study suggest that specific protein responses depend<br />

on when the high temperature regimens are imposed during<br />

development. Fewer proteins responded when the high temperature<br />

regimens were initiated during mid development (15 or 20 dpa) than<br />

during early development (anthesis or 10 dpa).<br />

Transient increases and decreases in protein levels were<br />

observed following transfer <strong>of</strong> plants to the high temperature<br />

regimens. These transient changes occurred over a period <strong>of</strong> days<br />

rather than hours as is typically observed in transcript studies.<br />

Under the regimen initiated at 10 dpa, HSP16.9, HSP70, peroxidase,<br />

triosephosphate isomerase, and globulins (legumin-like protein,<br />

seed globulin), increased transiently. The responses <strong>of</strong> HSP16.9 and<br />

HSP70 were also observed in previous studies on the effect <strong>of</strong> high<br />

temperature on wheat grain composition (Majoul et al., 2003,<br />

2004; Skylas et al., 2002). Since ROS production is elevated during<br />

abiotic and biotic stresses (reviewed, Suzuki and Mittler, 2006), an<br />

increase in peroxidase would be expected. The increase in triosephosphate<br />

isomerase agrees with findings that transcript<br />

(Minhas and Grover, 1999) and protein levels (Yan et al., 2005) for<br />

this enzyme increase in rice seedlings exposed to high temperature.<br />

The transient increase in globulins is interesting because it suggests<br />

that these proteins may have functional roles other than storage.<br />

One is<strong>of</strong>orm <strong>of</strong> ADP-glucose PPase decreased transiently under the<br />

high temperature regimen initiated at 10 dpa, but several is<strong>of</strong>orms<br />

<strong>of</strong> the large subunit <strong>of</strong> this enzyme decreased and remained at low<br />

levels following initiation <strong>of</strong> this high temperature regimen. These<br />

results are in keeping with the decrease in transcript levels for this<br />

enzyme and the reduction <strong>of</strong> starch accumulation that accompanies<br />

high temperatures during grain fill (Altenbach et al., 2003;<br />

Hurkman et al., 2003). Under the high temperature regimen initiated<br />

at 20 dpa, the only protein that increased transiently was an<br />

is<strong>of</strong>orm <strong>of</strong> HSP16.9. This finding agrees with the results <strong>of</strong> Skylas<br />

et al. (2002), where a number <strong>of</strong> HSP16.9 is<strong>of</strong>orms that increased<br />

during heat shock were absent in the mature grain.<br />

Among the proteins that responded to the high temperature<br />

regimens, seven were identified previously (Wong et al., 2003) as<br />

potential targets <strong>of</strong> thioredoxin, a widely distributed small disulfide<br />

protein that functions in redox regulation. These targets function in<br />

stress/defense (a-amylase/subtilisin inhibitor 381, serpin 231, 246,<br />

639, and dehydroascorbate reductase 382), storage (globulin Beg1<br />

620), and carbohydrate metabolism (pyruvate Pidikinase 31). A<br />

previous study showed that a number <strong>of</strong> potential thioredoxin<br />

targets responded to drought stress, including serpin and dehydroascorbate<br />

reductase (Hajheidari et al., 2007).<br />

4.3. High temperature and flour quality<br />

A striking feature <strong>of</strong> the high temperature regimens is the<br />

increase in the levels <strong>of</strong> stress/defense and globulin storage proteins<br />

in the endosperm <strong>of</strong> the mature grain. The stress/defense proteins<br />

include those known to respond to desiccation (glyceraldehyde-3-P<br />

dehydrogenase, LEA protein) and pathogen invasion (chitinase,<br />

xylanase inhibitor protein, a-amylase/subtilisin inhibitor, barwin/<br />

PR-4 protein, LTP, polyubiquitin, 20S proteosome a-subunit). Altenbach<br />

et al. (2007a,b) also reported that expression <strong>of</strong> transcripts for<br />

two <strong>of</strong> these genes, LTP and the PR-4 protein wheatwin, was<br />

enhanced under the 37 C/28 C regimen. The protein changes that<br />

occur late in grain development are <strong>of</strong> particular interest, because<br />

they may impact the functional properties <strong>of</strong> the resultant flour.<br />

Some <strong>of</strong> the proteins that increase in response to high temperature<br />

are present in dough liquor, a soluble fraction <strong>of</strong> wheat dough, or<br />

foams prepared from dough liquor (Salt et al., 2005) and may have<br />

roles in gas bubble stabilization in dough and crumb structure <strong>of</strong>


ead. These include LTP, a-amylase inhibitor, xylanase inhibitor,<br />

chitinase, serpin, HSP70, DNAK-type chaperone, peroxidase, glyceraldehyde-3-P<br />

dehydrogenase, nucleoside diphosphate kinase, and<br />

globulin-2. Some <strong>of</strong> these proteins, notably a-amylase inhibitor<br />

(Posch et al., 1995; Weiss et al., 1997), glyceraldehyde-3-P dehydrogenase,<br />

triosephosphate dehydrogenase, and serpin (Sander<br />

et al., 2001), also react with sera from patients with Baker’s asthma.<br />

LTP has been shown to react with sera from patients with immunoglobulin<br />

E-mediated food allergies, including wheat-dependent<br />

atopic dermatitis (Battais et al., 2005). In addition, many <strong>of</strong> the KClsoluble<br />

globulins and stress/defense proteins have structural similarities<br />

to food allergens (Jenkins et al., 2005). Further studies are<br />

warranted on the possible roles <strong>of</strong> these proteins in flour quality and<br />

their allergenic potential.<br />

Acknowledgements<br />

The authors thank Drs. F.M. DuPont and R. Thilmony for critical<br />

reading <strong>of</strong> the manuscript.<br />

Appendix. Supplementary material<br />

Supplementary data associated with this article can be found in<br />

the online version, at doi:10.1016/j.jcs.2008.06.014.<br />

References<br />

Altenbach, S.B., DuPont, F.M., Kothari, K.M., Chan, R., Johnson, E.L., Lieu, D., 2003.<br />

Temperature, water and fertilizer influence the timing <strong>of</strong> key events during<br />

grain development in a US spring wheat. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 37, 9–20.<br />

Altenbach, S.B., Kothari, K.M., 2004. Transcript pr<strong>of</strong>iles <strong>of</strong> genes expressed in<br />

endosperm tissue are altered by high temperature during wheat grain development.<br />

<strong>Cereal</strong> <strong>Science</strong> 40, 115–126.<br />

Altenbach, S.B., Kothari, K.M., Lieu, D., 2002. Environmental conditions during<br />

wheat grain development alter temporal regulation <strong>of</strong> major gluten protein<br />

genes. <strong>Cereal</strong> Chemistry 79, 279–285.<br />

Altenbach, S.B., Kothari, K.M., Tanaka, C.K., Hurkman, W.J., 2007a. Expression <strong>of</strong><br />

9 kDa lipid transfer protein genes in developing wheat grains is enhanced by<br />

high temperature but not by post-anthesis fertilizer. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong>,<br />

doi:10.1016/j.jcs.2007.03.010.<br />

Altenbach, S.B., Kothari, K.M., Tanaka, C.K., Hurkman, W.J., 2007b. Genes encoding<br />

the PR-4 protein wheatwin are developmentally regulated in wheat grains and<br />

respond to high temperature during grainfill. Plant <strong>Science</strong> 173, 135–143.<br />

Battais, F., Courcoux, P., Poplineau, Y., Kanny, G., Moneret-Vautrin, D.A., Denery-<br />

Papini, S., 2005. Food allergy to wheat: differences in immunoglobulin Ebinding<br />

proteins as a function <strong>of</strong> age and symptoms. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

42, 109–117.<br />

Craig, R., Cortens, J.P., Beavis, R.C., 2004. An open source system for analyzing,<br />

validating and storing protein identification data. <strong>Journal</strong> <strong>of</strong> Proteome Research<br />

3, 1234–1242.<br />

DuPont, F.M., Hurkman, W.J., Vensel, W.H., Chan, R., Lopez, R., Tanaka, C.K.,<br />

Altenbach, S.B., 2006a. Differential accumulation <strong>of</strong> sulfur-rich and sulfur-poor<br />

wheat flour proteins is affected by temperature and mineral nutrition during<br />

grain development. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 44, 101–112.<br />

DuPont, F.M., Hurkman, W.J., Vensel, W.H., Tanaka, C.K., Kothari, K.M., Chung, O.K.,<br />

Altenbach, S.B., 2006b. Protein accumulation and composition in wheat grains:<br />

effects <strong>of</strong> mineral nutrients and high temperature. European <strong>Journal</strong> <strong>of</strong><br />

Agronomy 25, 96–107.<br />

Hajheidari, M., Elivazi, A., Buchanan, B.B., Wong, J.H., Majidi, I., Salekdeh, G.H., 2007.<br />

Proteomics uncovers a role for redox in drought tolerance in wheat. <strong>Journal</strong> <strong>of</strong><br />

Proteome Research 6, 1451–1460.<br />

W.J. Hurkman et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 12–23 23<br />

Hurkman, W.J., McCue, K.F., Altenbach, S.B., Korn, A., Tanaka, C.K., Kothari, K.M.,<br />

Johnson, E.L., Bechtel, D.B., Wilson, J.D., Anderson, O.D., DuPont, F.M., 2003.<br />

Effect <strong>of</strong> temperature on expression <strong>of</strong> genes encoding enzymes for starch<br />

biosynthesis in developing wheat endosperm. Plant <strong>Science</strong> 164, 873–881.<br />

Hurkman, W.J., Tanaka, C.K., 2004. A simplified method for the isolation <strong>of</strong> wheat<br />

endosperm proteins for proteome analysis. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 40,<br />

294–299.<br />

Jenkins, J.A., Griffiths-Jones, S., Shewry, P.R., Breiteneder, H., et al., 2005. Structural<br />

relatedness <strong>of</strong> plant food allergens with specific reference to cross-reactive<br />

allergens: an in silico analysis. <strong>Journal</strong> <strong>of</strong> Allergy and Clinical Immunology 115,<br />

163–170.<br />

Jung, H.W., Kim, W., Hwang, B.K., 2003. Three pathogen-inducible genes<br />

encoding lipid transfer protein from pepper are differentially activated by<br />

pathogens, abiotic, and environmental stresses. Plant, Cell, and Environment<br />

26, 915–928.<br />

Laudencia-Chingcuanco, D.L., Stamova, B.S., You, F.M., Lazo, G.R., Beckles, D.M.,<br />

Anderson, O.D., 2007. Transcriptional pr<strong>of</strong>iling <strong>of</strong> wheat caryopsis development<br />

using cDNA microarrays. Plant Moleular Biology 63, 651–668.<br />

Majoul, T., Bancel, E., Triboi, E., Ben Hamida, J., Branlard, G., 2003. Proteomic analysis<br />

<strong>of</strong> the effect <strong>of</strong> heat stress on hexaploid wheat grain: characterization <strong>of</strong> heatresponsive<br />

proteins from total endosperm. Proteomics 3, 175–183.<br />

Majoul, T., Bancel, E., Triboi, E., Ben Hamida, J., Branlard, G., 2004. Proteomic analysis<br />

<strong>of</strong> heat stress on hexaploid wheat grain: characterization <strong>of</strong> heat-responsive<br />

proteins from non-prolamins fraction. Proteomics 3, 505–513.<br />

McIntosh, S., Watson, L., Bundock, P., Crawford, A., White, J., Cordeiro, G., Barbary, D.,<br />

Rooke, L., Henry, R., 2007. SAGE <strong>of</strong> the developing wheat caryopsis. Plant<br />

Biotechnology <strong>Journal</strong> 5, 69–83.<br />

Minhas, D., Grover, A., 1999. Transcript levels <strong>of</strong> genes encoding various glycolytic<br />

and fermentation enzymes change in response to abiotic stresses. Plant <strong>Science</strong><br />

146, 41–51.<br />

Posch, A., Weill, W., Wheeler, C., Dunn, M.J., Görg, A., 1995. Sequence analysis <strong>of</strong><br />

wheat grain allergens separated by two-dimensional electrophoresis with<br />

immobilized pH gradients. Electrophoresis 16, 1115–1119.<br />

Salt, L.J., Robertson, J.A., Jenkins, J.A., Mulholland, J., Clare Mills, E.N., 2005. The<br />

identification <strong>of</strong> foam-forming soluble proteins from wheat (Triticum aestivum)<br />

dough. Proteomics 5, 1612–1623.<br />

Sander, I., Flagge, A., Merget, R., Halder, T.M., Meyer, H.E., Baur, X., 2001. Identification<br />

<strong>of</strong> wheat flour allergens by means <strong>of</strong> 2-dimensional immunoblotting.<br />

<strong>Journal</strong> <strong>of</strong> Allergy and Clinical Immunology 107, 907–913.<br />

SAS Institute Inc., 2004. SAS OnlineDoc Ò 9.1.3. SAS Institute Inc., Cary, NC.<br />

Skylas, D.J., Cordwell, S.J., Hains, P.G., Larsen, M.R., Basseal, D.J., Walsh, B.J.,<br />

Blumenthal, C., Rathmell, W., Copeland, L., Wrigley, C.W., 2002. Heat shock <strong>of</strong><br />

wheat during grain filling: proteins associated with heat-intolerance. <strong>Journal</strong> <strong>of</strong><br />

<strong>Cereal</strong> <strong>Science</strong> 35, 175–188.<br />

Suzuki, N., Mittler, R., 2006. Reactive oxygen species and temperature stresses:<br />

a delicate balance between signaling and destruction. Physiologia Plantarum<br />

126, 45–51.<br />

Velasco, R., Salamini, F., Bartels, D., 1994. Dehydration and ABA increase mRNA<br />

levels and enzyme activity <strong>of</strong> cytosolic GAPDH in the resurrection plant Craterostigma<br />

plantagineum. Plant Molecular Biology 26, 541–546.<br />

Vensel, W.H., Tanaka, C.K., Cai, N., Wong, J.H., Buchanan, B.B., Hurkman, W.J., 2005.<br />

Developmental changes in the metabolic protein pr<strong>of</strong>iles <strong>of</strong> wheat endosperm.<br />

Proteomics 5, 1594–1611.<br />

Weiss, W., Huber, G., Engel, K.-H., Pethran, A., Dunn, M.J., Gooley, A.A., Görg, A., 1997.<br />

Identification and characterization <strong>of</strong> wheat grain albumin/globulin allergens.<br />

Electrophoresis 18, 826–833.<br />

Wong, J.H., Balmer, Y., Cai, N., Tanaka, C.K., Vensel, W.H., Hurkman, W.J.,<br />

Buchanan, B.B., 2003. Unraveling thioredoxin-linked metabolic processes <strong>of</strong><br />

cereal starchy endosperm using proteomics. FEBS Letters 547, 151–156.<br />

Wong, J.H., Cai, N., Tanaka, C.K., Vensel, W.H., Hurkman, W.J., Buchanan, B.B., 2004.<br />

Thioredoxin reduction alters the solubility <strong>of</strong> proteins <strong>of</strong> wheat starch<br />

endosperm: an early event in cereal germination. Plant Cell Physiology 45,<br />

407–415.<br />

Wu, G., Robertson, A.J., Liu, X., Zheng, P., Wilen, R.W., Nesbitt, N.T., Gusta, L.V., 2004.<br />

A lipid transfer protein gene BG-14 is differentially regulated by abiotic stress.<br />

<strong>Journal</strong> <strong>of</strong> Plant Physiology 161, 449–458.<br />

Yan, S., Tang, Z., Su, W., Sun, W., 2005. Proteomic analysis <strong>of</strong> salt stress-responsive<br />

proteins in rice root. Proteomics 5, 235–244.<br />

Zadoks, J.C., Chang, T.T., Konzak, C.F., 1974. A decimal code for the growth stages <strong>of</strong><br />

cereals. Weed Research 14, 415–421.


Accumulation <strong>of</strong> mixed linkage (1 / 3) (1 / 4)-b-D-glucan during grain<br />

filling in barley: A vibrational spectroscopy study<br />

Helene Fast Seefeldt a,b , Andreas Blennow c , Birthe Møller Jespersen b , Bernd Wollenweber a ,<br />

Søren Balling Engelsen b, *<br />

a University <strong>of</strong> Aarhus, Faculty <strong>of</strong> Agricultural <strong>Science</strong>s, Department <strong>of</strong> Genetics and Biotechnology, Forsøgsvej 1, DK-4200 Slagelse, Denmark<br />

b University <strong>of</strong> Copenhagen, Faculty <strong>of</strong> Life <strong>Science</strong>s, Department <strong>of</strong> Food <strong>Science</strong>, Quality & Technology, DK-1958 Frederiksberg C, Denmark<br />

c University <strong>of</strong> Copenhagen, Faculty <strong>of</strong> Life <strong>Science</strong>s, Department <strong>of</strong> Plant Biology and Biotechnology, VKR research centre Pro-Active Plants,<br />

DK-1871 Frederiksberg C, Denmark<br />

article info<br />

Article history:<br />

Received 20 November 2007<br />

Received in revised form 9 June 2008<br />

Accepted 30 June 2008<br />

Keywords:<br />

Infrared<br />

IR<br />

Near-infrared<br />

NIR<br />

(1 / 3) (1 / 4)-b-D-glucan<br />

Grain filling<br />

1. Introduction<br />

abstract<br />

In the quest for optimising cereals for soluble fibres and other<br />

health-promoting components, there is great interest in studying<br />

the metabolic changes during grain filling, such as cell wall fibre<br />

development in barley mutants differing in mixed linkage (1 / 3)<br />

(1 / 4)-b-D-glucan (BG) using high throughput methods. Inexpensive,<br />

spectroscopic methods based on vibrational spectroscopy<br />

<strong>of</strong>fer the possibility <strong>of</strong> fast and flexible analysis <strong>of</strong> a large number <strong>of</strong><br />

genotypes (Osborne, 2006). Near-infrared (NIR) and infrared (IR)<br />

spectroscopy measure the vibrations <strong>of</strong> molecular covalent bonds.<br />

The NIR region 14,300–4000 cm 1 (780–2500 nm) mainly gives<br />

Abbreviations: ADP-glucose, adenosine 5 0 diphosphate glucose; ATR, attenuated<br />

total reflection; BG, Mixed linkage (1 / 3) (1 / 4)-b-D-glucan; DAF, days after<br />

flowering; EISC, extended inverted signal correction; FT-IR, Fourier transform<br />

infrared; iPLS, interval partial least squares regression; MSC, multiplicative scatter<br />

correction; NIR, near-infrared reflectance; PC, principal component; PCA, principal<br />

component analysis; PLS, partial least squares regression; RMSECV, root mean<br />

square error <strong>of</strong> cross validation; SECV, standard error <strong>of</strong> cross validation; SEE,<br />

standard error <strong>of</strong> estimate; VIS, visual part <strong>of</strong> the NIR spectra.<br />

* Corresponding author. Tel.: þ45 3533 3205; fax: þ45 3533 3245.<br />

E-mail address: se@life.ku.dk (S.B. Engelsen).<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.06.012<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 24–31<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

The accumulation <strong>of</strong> mixed linkage barley (1 / 3) (1 / 4)-b-D-glucan (BG) during grain filling at eight<br />

stages was studied using standard reference methods and infrared spectroscopy. Two mutant barley<br />

genotypes having higher (starch mutant lys5f) and lower (high lysine mutant lys3a) BG content than the<br />

normal control Cork were studied. The Cork and lys3a genotypes showed a linear BG accumulation<br />

throughout the grain filling to reach a maximum <strong>of</strong> approximately 6 and 4% BG (w/w) dry matter,<br />

respectively. However, lys5f mutant exhibited an exponential increase in BG synthesis to a maximum <strong>of</strong><br />

approximately 18% BG (w/w) dry matter 30 days after flowering (DAF), seemingly compensating for<br />

a decreased synthesis <strong>of</strong> starch.<br />

The spectral information <strong>of</strong> the barley flour was compared to pure BG spectra and partial least squares<br />

regression (PLS) models were constructed for calibration to BG content. Informative regions in the nearinfrared<br />

(NIR) and the infrared (IR) spectra were identified for separation <strong>of</strong> temporal and genetic<br />

differences. Interval PLS yielded good calibration models to BG (R 2 ¼ 0.94 for NIR in the region<br />

1194–1240 nm, whereas the global PLS gave correlations with BG with R 2 ¼ 0.92 for IR).<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

information about overtones and combination tones (stretching<br />

and bending) involving anharmonic bonds primarily to hydrogen,<br />

whereas the IR region 4000–200 cm 1 (2500–50,000 nm) gives<br />

information about the fundamental vibrations. NIR spectroscopy is<br />

already well established for at/on-line quality control in the food<br />

and food ingredients industries (Zachariassen et al., 2005) and is<br />

able to provide information about chemical parameters such as<br />

water, protein and starch content as well as about physical<br />

parameters such as particle size and temperature. In cereals, NIR<br />

transmittance spectra <strong>of</strong> single seeds have proven very informative<br />

for the determination <strong>of</strong> different quality traits such as protein and<br />

fat content with high accuracy (Delwiche 1995; Pedersen et al.,<br />

2002). In fact, a new high capacity single-seed TriQ NIR sorting<br />

system (Bomill AB, Lund, Sweden) (Munck, 2007) has been developed<br />

utilising NIR spectra to diversify heterogeneous bulk lots <strong>of</strong><br />

wheat with regard to multivariate complex quality traits such as<br />

dough performance and baking value (Munck, 2007; Tønning,<br />

2008). Furthermore, NIR can be used to evaluate quality traits such<br />

as the fibre fraction <strong>of</strong> cereal cell walls (Blakeney and Flinn, 2005)<br />

and genetics (Jacobsen et al., 2005; Munck et al., 2004).<br />

While NIR spectroscopy has primarily been used in quantitative<br />

analysis <strong>of</strong> bulk components (in spite <strong>of</strong> its documented ability to<br />

predict complex qualitative traits like baking and malting quality


(Munck, 2007)), IR spectroscopy has mainly been used to study<br />

well-defined components such as plant cell wall polysaccharides<br />

(Chen et al., 1998; Kacurakova and Wilson, 2001; McCann et al.,<br />

1992; Robert et al., 2005; Séné et al., 1994). Recently, micro Fourier<br />

transformed infrared spectroscopy (FT-IR) was used to study the<br />

deposition <strong>of</strong> cell wall polysaccharides in wheat endosperm during<br />

grain development with emphasis on BG and arabinoxylans (Philippe<br />

et al., 2006). Also, NIR spectroscopy has proven valuable in<br />

monitoring plant physiological processes such as carbohydrate<br />

accumulation during grain filling (Gergely and Salgo, 2005). NIR<br />

screenings <strong>of</strong> normal barley and mutants have resulted in the<br />

discovery <strong>of</strong> genotypes with a strongly increased content <strong>of</strong> soluble<br />

BG fibres (Munck et al., 2004). The high BG barley mutant line lys5f<br />

is a structural (enzyme functional) low starch mutant (Munck et al.,<br />

2004; Munck and Møller, 2005) unable to transport ADP–glucose<br />

across the plastid envelope due to an inactive ADP–glucose transporter<br />

(Patron et al., 2004). Lys5f is thus disabled in efficient<br />

synthesis <strong>of</strong> starch, but compensates via an extremely high content<br />

(approximately 17% d.m.) <strong>of</strong> soluble and insoluble BG (Munck et al.,<br />

2004). The lys3a mutant is a regulatory mutant (Jacobsen et al.,<br />

2005) that primarily inhibits the synthesis <strong>of</strong> the hordein, but<br />

increases the synthesis <strong>of</strong> soluble proteins. It has approximately 2–<br />

3% BG compared to 3–4% in normal barley like Cork when grown in<br />

green houses (Munck et al., 2004).<br />

The aim <strong>of</strong> this study was thus to characterise and investigate<br />

two extreme recessive barley mutants with respect to BG and<br />

starch content during grain filling using classical reference methods<br />

as well as spectroscopic fingerprinting methods. The carbohydrate<br />

mutant lys5f with high BG content and the protein lys3a with low<br />

BG content were compared to the variety ‘Cork’ that has a normal<br />

content <strong>of</strong> BG.<br />

2. Experimental<br />

2.1. Plant material<br />

Three genotypes <strong>of</strong> barley were included in the study: a malt<br />

barley (Hordeum vulgare cv. Cork) and a barley mutant lys3a with<br />

alterations in the lys3 locus on chromosome five, tightly linked to<br />

adjacent BG synthesis suppressing genes (Munck et al., 2004). The<br />

mutant lys5f has a mutation in chromosome six. A ‘semifield’ pot<br />

experiment (72 pots, 16.5 cm diameter, 13 cm height) was carried<br />

out from April to August 2005 at the University <strong>of</strong> Aarhus, Research<br />

Centre Flakkebjerg, Denmark. Each pot was filled with 10 l <strong>of</strong><br />

sphagnum with 15% Perlite (Perlite, Denmark) added. Ten seeds <strong>of</strong><br />

each genotype were sown on April 20th and thinned to three<br />

seedlings per pot on May 31st. All pots were drip-watered<br />

throughout the experiment and standard pest control was performed<br />

against mildew and aphids when needed. Flowering was<br />

judged visually when 50% <strong>of</strong> the spikes showed clear pollen release,<br />

which occurred between the 26th and 27th <strong>of</strong> June. Spikes were<br />

harvested at eight time points during grain filling: 9, 13, 16, 20, 23,<br />

30, 39 and 47 days after flowering (DAF).<br />

2.2. Plant analysis<br />

The spike on the main tiller and the first spike <strong>of</strong> the side tillers<br />

were cut and immediately frozen in liquid nitrogen. After freezing,<br />

the kernels were detached from the spike, counted and weighed.<br />

Kernels were transferred to 80 C and freeze-dried within 3 weeks<br />

after harvest. One sample consisted <strong>of</strong> the seeds from two spikes<br />

from the same individual plant. The freeze-dried grains were milled<br />

(0.5 mm, Cyclotec 1093, Foss Tecator AB, Högenas, Sweden). A total<br />

<strong>of</strong> 91 samples were analysed (Cork, 31; lys5f, 31; and lys3a, 29). The<br />

total sample set consisted <strong>of</strong> three genotypes at eight temporal<br />

harvest points– two replicate spectra <strong>of</strong> each genotype from<br />

H.F. Seefeldt et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 24–31 25<br />

harvests 1–4 and six replicate spectra <strong>of</strong> each genotype from<br />

harvests 5–8 were analysed. This experimental design yielded<br />

a total <strong>of</strong> 96 samples (3 genotypes 4 (1–4 harvests) 2 replicates<br />

þ 3 genotypes 4 (5–8 harvests) 6 replicates). Due to<br />

a minor flaw in the experiment, three replicates <strong>of</strong> lys3a from<br />

harvests 1, 5 and 7, and two replicates <strong>of</strong> Cork harvest 5 and lys5f<br />

harvest 5 were lost. The ground material was stored in sealed<br />

plastic bottles at room temperature until analysis.<br />

2.3. Chemical analysis<br />

The content <strong>of</strong> soluble BG was analysed by fluorimetry (Calc<strong>of</strong>lour<br />

reagent type II, Scandinavian Brewery Laboratory, Frederiksberg,<br />

Denmark) (Munck et al., 1989). The high BG content flours<br />

showed major deviations in values when determined by Calc<strong>of</strong>lour,<br />

and hence the values from the last four time points in lys5f were<br />

double-checked with an enzymatic kit (Megazyme, Wicklow, Ireland)<br />

specific for mixed linkage BG.<br />

The total content <strong>of</strong> starch was analysed as follows: freeze-dried<br />

flour (10 mg) was washed three times with 1-ml aliquots <strong>of</strong> 80%<br />

ethanol in screw-cap Eppendorf tubes. Starch in the washed flour<br />

was gelatinised by the addition <strong>of</strong> KOH (400 ml, 0.2 M) and incubation<br />

at 95 C for 1 h. After cooling, 140 ml <strong>of</strong> 1 M acetic acid was<br />

added, the samples mixed and diluted 20-fold with water. The<br />

diluted sample (10 ml) was mixed with an equal volume <strong>of</strong> amyloglucosidase<br />

(10 U/ml, Fluka) solubilised in 50 mM Na acetate pH 5.0<br />

and starch was hydrolysed by incubation at 37 C for 2 h. A 210-ml<br />

aliquot <strong>of</strong> 50 mM Mops/KOH pH 7.3, 5 mM MgCl2, 1 mM EDTA,<br />

1 mM ATP, 1 mM NAD and 4 U/ml hexokinase was added and<br />

absorbance at 340 nm was registered. One microlitre <strong>of</strong> 500 U/ml<br />

glucose-6-phosphate dehydrogenase was added and the production<br />

<strong>of</strong> NADH was followed until steady from the absorbance at<br />

340 nm. Starch content in the original flour was calculated using<br />

glucose as standard. Amylose in the flour was determined by iodine<br />

complexation, as described by Bay-Smidt et al. (1999). The amylopectin<br />

chain length distribution <strong>of</strong> the starch was analysed by high<br />

performance anionic exchange chromatography with pulsed<br />

amperometric detection (HPAEC-PAD) as described by Blennow<br />

et al. (1998).<br />

FT-IR and NIR spectra were obtained from the pure substances:<br />

Cellulose (9004d34-6, Sigma-Aldrich Chemie, Steinheim, Germany),<br />

BG (Barley– Medium Viscosity, Megazyme, Wicklow, Ireland)<br />

and wheat starch with normal and high content <strong>of</strong> amylose<br />

(Ritmo, Sejet Plantbreeding, Horsens, Denmark). The starches were<br />

purified according to Blennow et al. (1998).<br />

2.4. FT-IR measurements<br />

All FT-IR spectra were acquired at room temperature using the<br />

Arid-Zone MB100 FT-IR spectrometer (ABB Bomen Inc., Quebec, PQ,<br />

Canada). The sampling was performed using an Attenuated Total<br />

Reflection (ATR) device with a diamond crystal (ZnSe, TR-plate, ARK<br />

0055-603, Spectra-Tech Inc., CT, USA) operating in the range 4000–<br />

750 cm 1 . Measurements were obtained using 64 scans at 4 cm 1<br />

resolution. Background scans were obtained using 128 scans. The<br />

scans were averaged and ratioed against a single-beam spectrum <strong>of</strong><br />

the clean ATR crystal and converted into absorbance units.<br />

2.5. NIR measurements<br />

A near-infrared instrument (Foss NIRSystems 6500, USA)<br />

was used in reflectance mode in the range <strong>of</strong> 400–2500 nm. All<br />

measurements were acquired at room temperature using a<br />

commercial ’’small sample cup’’ containing approximately 0.5 g <strong>of</strong><br />

flour. The spectra were recorded in 2-nm steps using a spinning<br />

sample module. Each sample was measured using 32 scans which


26<br />

were ratioed against 16 reference scans using a ceramic sample. The<br />

results were averaged as a log 1/R spectrum.<br />

2.6. Data analysis<br />

The spectra were analysed using the chemometric s<strong>of</strong>tware<br />

LatentiX 1.0 (http://www.latentix.com, Latent5, Copenhagen, Denmark)<br />

both for visual inspection <strong>of</strong> the spectra and calculation <strong>of</strong><br />

principal component analysis (PCA) (Wold et al., 1987) and partial<br />

least squares (PLS) regressions. The PCA was performed to visualise<br />

systematic spectral variation. Calibrations and predictions <strong>of</strong> BG<br />

and amylose based on the spectral information were made using<br />

PLS. Interval partial least squares regression (iPLS) (Nørgaard et al.,<br />

2000) was performed in order to reveal the most important spectral<br />

interval correlated with the calibration parameters. Prior to<br />

analysis, the spectral data was scatter-corrected using the extended<br />

inverted signal correction (EISC) (Martens et al., 2003; Pedersen<br />

et al., 2002) and all models are mean-centred. The iPLS was carried<br />

out using the iTOOLBOX (http://www.models.life.ku.dk) as<br />

a routine for MATLAB 6.0 (The Mathworks Inc., Natick, USA). All<br />

reported models are based on 30 intervals and are validated using<br />

full cross validation.<br />

3. Results and discussion<br />

The seeds showed normal fresh weight and water content<br />

during development (Fig. 1a,b) during grain filling (Gergely and<br />

Salgo, 2003; Jennings and Morton, 1962): Three distinct stages <strong>of</strong><br />

the grain filling process could be determined by analysis <strong>of</strong> the<br />

fresh grain weight (Fig. 1a). Until 20 DAF a rapid increase in grain<br />

weight occurred, followed by a lag phase with stable grain weight.<br />

After 30 DAF the grain weight decreased due to the drying and<br />

maturation <strong>of</strong> the seed. The first phase is characterised by a rapid<br />

influx <strong>of</strong> water and cell enlargement and continues until 10–15 DAF.<br />

During the second phase the seeds take up nutrients and start to<br />

synthesise starch and protein (Jenner et al., 1991). In the third phase<br />

the fresh weight decreases due to loss <strong>of</strong> water as indicated in<br />

Fig. 1b.<br />

3.1. Accumulation <strong>of</strong> BG during grain filling – chemical analysis<br />

Fig. 1c shows the development <strong>of</strong> BG synthesis between genotypes<br />

during grain filling. The analysis <strong>of</strong> the BG content <strong>of</strong> the three<br />

barley genotypes (Fig. 1c) revealed that lys5f from 16 to 30 DAF<br />

exhibited a very strong increase in BG accumulation to reach an<br />

extreme maximum BG content <strong>of</strong> approximately 18% at 30 DAF.<br />

During the same period lys3a and Cork exhibited slow linear<br />

accumulation <strong>of</strong> BG, which continued throughout the grain filling<br />

period until maturity at 47 DAF. The mutant lys3a had lower BG<br />

content as compared to both Cork and lys5f, reaching only<br />

approximately 4% at harvest. Cork displayed a normal content <strong>of</strong><br />

BG, reaching a maximum level <strong>of</strong> approximately 6% BG at maturity.<br />

During grain filling, Cork and lys3a did not start to accumulate BG<br />

until 16–20 DAF, whereas lys5f already started to accumulate BG<br />

between 9 and 13 DAF (Fig. 1c).<br />

In a previous study <strong>of</strong> three normal barley varieties (Coles,<br />

1979), a large increase in accumulation <strong>of</strong> BG starting at 19 DAF was<br />

found. In wheat, the cell walls mainly consist <strong>of</strong> arabinoxylans,<br />

whereas BG constitutes only 25%, but BG was accumulating from<br />

early grain filling until 10 DAF after which arabinoxylans were<br />

dominating (Philippe et al., 2006). These observations led to the<br />

hypothesis that BG could function as a structural element <strong>of</strong><br />

growing cell walls as well as storage material that could be<br />

hydrolysed during germination <strong>of</strong> the grain (Philippe et al., 2006).<br />

The starch synthesis in Cork and lys3a (Fig. 1d) followed the<br />

same pattern having a rapid increase from 9 DAF until 20 DAF and<br />

H.F. Seefeldt et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 24–31<br />

reaching a maximum <strong>of</strong> approximately 47% starch remaining at this<br />

level until 39 DAF. It should be noted that the levels <strong>of</strong> grain starch<br />

at maturation in our study were somewhat lower than generally<br />

found in barley grain as a result <strong>of</strong> a decrease in starch content <strong>of</strong><br />

approximately 10% the last 17 DAF. The mutant lys5f displayed<br />

a much slower increase throughout the entire grain filling period<br />

until 39 DAF to reach a maximum at 23% starch. A low final level <strong>of</strong><br />

starch in lys5f was previously reported (Munck et al., 2004).<br />

The high content <strong>of</strong> BG in lys5f can hence be interpreted as a new<br />

mechanism for compensating starch synthesis with BG synthesis<br />

(Munck et al., 2004). Interestingly, the total accumulation <strong>of</strong> starch<br />

and BG reveals three similar patterns for the genotypes with just an<br />

<strong>of</strong>fset in total accumulation. Thus, the rapid accumulation rate <strong>of</strong><br />

BG in lys5f corresponds largely to the rapid accumulation <strong>of</strong> starch<br />

in the non-starch genotypes lys3a and Cork. This confirms the<br />

observation <strong>of</strong> the starch and BG levels found in mature seed<br />

described by (Munck et al., 2004). The higher starch content found<br />

in Cork was followed by a high dry matter weight (Fig. 1e). The lys5f<br />

mutant had a markedly lower ratio <strong>of</strong> amylose/amylopectin (Fig. 1f)<br />

which is interesting from a starch functionality viewpoint (Tester<br />

et al., 2004). The chain length <strong>of</strong> amylopectin did, however, not<br />

differ between the three genotypes and it did not differ markedly<br />

over the developmental stages (data not shown).<br />

3.2. Near-infrared analysis <strong>of</strong> the barley grain filling<br />

The EISC pre-processed VIS/NIR spectra (400–2498 nm) <strong>of</strong> the<br />

91 barley flour samples are displayed in Fig. 2a. The first three<br />

harvests (9–16 DAF) represented by the light blue colours clearly<br />

differ in their NIR spectral pr<strong>of</strong>iles from the rest <strong>of</strong> the harvests (20–<br />

47 DAF, shown as dark red colours). The first three harvests are less<br />

intensive in the first overtone region from 1400–1850 nm and the<br />

O-H combination tone centred at 1940 nm. In general, the temporal<br />

differences influenced the <strong>of</strong>fset in the spectra, while the genetic<br />

differences influenced the shape <strong>of</strong> the spectra. In the visible part <strong>of</strong><br />

the spectrum, the grain filling is clearly evidenced by the gradual<br />

elimination <strong>of</strong> the chlorophyll peak at 672 nm (Fig. 2b), which<br />

reflects that the barley grains were visibly green until 30 DAF (the<br />

sixth harvest). Fig. 2c displays a score plot from a PCA <strong>of</strong> the 91 NIR<br />

spectra <strong>of</strong> the barley flour, explaining 92% <strong>of</strong> the variance. In the<br />

score plot, the temporal variation is separated into four welldefined<br />

groups. The first three harvests (9–16 DAF) are markedly<br />

different from the later harvests and located in the upper left corner<br />

<strong>of</strong> the PCA. The fourth and the fifth harvests (20 and 23 DAF), in<br />

which the BG is also primarily synthesised (Fig. 1c), are clustered<br />

with the lowest scores in the score plot, the sixth harvest (30 DAF)<br />

located in the lower right part <strong>of</strong> the spectra and the last two<br />

harvests (39–47 DAF) in the upper right corner <strong>of</strong> the PCA. By<br />

inspection <strong>of</strong> the loadings, the primary reason for the clear separation<br />

<strong>of</strong> the temporal grain fillings was found to be the chlorophyll<br />

peak at 672 nm (Table 1). As observed in Fig. 2c, the actual time <strong>of</strong><br />

flowering is difficult to determine– in the figure a few samples<br />

reside in their previous group (marked with stars in Fig. 2c),<br />

presumably due to inaccuracy in the DAF determination.<br />

Munck et al. (2004) found that the combination tone region<br />

2280–2360 nm contains unique information about the lys3<br />

mutants and Szczodrak et al. (1992) found that two wavelengths in<br />

the region between 2260 and 2380 nm correlate the best with BG.<br />

Fig. 2d shows that the NIR spectra <strong>of</strong> lys5f have a peak in the NIR<br />

spectra around 2345 nm (the second bold arrow). This peak is also<br />

present in the spectra <strong>of</strong> lys3a, but as a shoulder, whereas the<br />

spectra <strong>of</strong> Cork only have a very weak shoulder. On the other hand,<br />

the spectra <strong>of</strong> Cork have a characteristic shoulder at 2280 nm,<br />

indicating starch (first bold arrow in Fig. 3d, Table 1), according to<br />

Munck (2007). This shoulder is absent in the mutants lys5f and<br />

lys3a. As Cork and lys3a have almost identical starch contents, it


mg seed-1<br />

% dm<br />

mg seed-1<br />

110.0<br />

95.0<br />

80.0<br />

65.0<br />

50.0<br />

35.0<br />

22.0<br />

20.0<br />

18.0<br />

16.0<br />

14.0<br />

12.0<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

2.0<br />

0.0<br />

60.0<br />

50.0<br />

40.0<br />

30.0<br />

20.0<br />

10.0<br />

0.0<br />

A 80.0<br />

B<br />

C<br />

9 13 16 20 23 30 39 47<br />

Fresh weight<br />

9 13 16 20 23 30 39 47<br />

Beta-glucan<br />

9 13 16 20 23 30 39 47<br />

Dry matter<br />

could be questioned whether this peak corresponds to starch only.<br />

When comparing the average barley flour NIR spectra with the NIR<br />

spectra <strong>of</strong> pure cellulose, starch, BG and amylose (Fig. 3a), it is<br />

evident that a high degree <strong>of</strong> similarity exists with the starch<br />

spectrum except for the region 2200–2400 nm where the spectra<br />

show very distinct pr<strong>of</strong>iles including genotype specific patterns.<br />

From the lines indicated in Fig. 2d it can be observed that the<br />

three genotypes exhibit similar spectral properties during the first<br />

harvest (9 DAF) as well as during the second and third harvests. In<br />

order to avoid the dominance <strong>of</strong> the chlorophyll information, the<br />

visual part <strong>of</strong> the spectra was discarded and the NIR region 1100–<br />

2498 nm was examined separately (data not shown). The three<br />

genotypes are clustered together until 16 DAF. From the fourth<br />

harvest (20 DAF) and onwards, the three genotypes are separated<br />

into three distinct clusters. This is in excellent agreement with the<br />

H.F. Seefeldt et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 24–31 27<br />

%<br />

% dm<br />

60.0<br />

40.0<br />

20.0<br />

60.0<br />

45.0<br />

30.0<br />

15.0<br />

0.0<br />

40.0<br />

E F<br />

% starch<br />

35.0<br />

30.0<br />

25.0<br />

20.0<br />

15.0<br />

10.0<br />

D<br />

9 13 16 20 23 30 39 47<br />

Water content<br />

9 13 16 20 23 30 39 47<br />

Starch<br />

9 13 16 20 23 30 39 47<br />

Amylose<br />

Fig. 1. (A) Fresh seed weight (mg) (B) Water content (%). (C) BG content (% d.m.). (D) Starch content (% d.m.). (E) Seed dry weight (mg). (F) Amylose (% in starch). All values are<br />

denoted in relation to DAF (days after flowering). Solid, Cork; dotted, lys5f; broken, lys3a.<br />

rapid increase in the BG synthesis from 16 DAF to 23 DAF for lys5f<br />

(Fig. 1c) and the concomitant rapid increase in starch seen for lys3a<br />

and Cork. The loadings associated with the PCA reveal that the first<br />

PC is mainly spanned by information from moisture and starch (O-<br />

H stretching vibrations, 1440 and 1940 nm) (Table 1), which is in<br />

good accordance with the separation <strong>of</strong> grain filling time points<br />

along PC1. The third PC separates the genotypes into a Cork/lys3a<br />

cluster and a lys5f cluster. The corresponding loadings are starchrelated<br />

peaks (Munck et al., 2004; Szczodrak et al., 1992), which<br />

manifest the different syntheses in the three genotypes.<br />

3.3. Infrared investigation <strong>of</strong> barley grain filling<br />

The EISC-treated FT-IR spectra (1900–750 cm 1 ) <strong>of</strong> the 91 barley<br />

flour samples are shown in Fig. 4a. As was the case for the NIR


28<br />

Log 1/R<br />

Scores PC#2 (14.%)<br />

0.55<br />

0.5<br />

0.45<br />

0.4<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.10<br />

0.05<br />

0.00<br />

-0.05<br />

-0.10<br />

-0.15<br />

-0.20<br />

A<br />

400 800 1200 1600<br />

nm<br />

2000 2400<br />

1 2 3 4 5 6 7 8<br />

Levels <strong>of</strong> Harvest<br />

C<br />

1<br />

2<br />

2<br />

2 2<br />

1<br />

1<br />

232<br />

3<br />

3<br />

3<br />

78<br />

8<br />

87<br />

7 7<br />

8<br />

8878<br />

87 8<br />

7<br />

8<br />

7<br />

85 8<br />

87<br />

7<br />

7788<br />

87<br />

6<br />

6<br />

6<br />

6<br />

7 6 6<br />

6<br />

6 6 6<br />

6 6<br />

66<br />

6<br />

5<br />

4 5 5 5 5<br />

4 4 5<br />

5<br />

5 4 65<br />

4 5 5<br />

-0.5 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3<br />

Scores PC#1 (78.%)<br />

spectra, a clear temporal dependence is observed in the spectra<br />

(data not shown), but the difference between genotypes appears<br />

more pronounced than in the NIR spectra.<br />

The barley flour spectra represent typical broad absorption<br />

bands <strong>of</strong> polysaccharides in the region 1200–750 cm 1 (Kacurakova<br />

and Wilson, 2001) with a maximum absorption around<br />

1021 cm 1 . The glycosidic absorption band at 1160 cm 1 reported<br />

by Robert et al. (2005) and at 1150 cm 1 (Philippe et al., 2006) is<br />

found at 1152 cm 1 in our samples. Fig. 4c (raw spectra) shows the<br />

enlargement <strong>of</strong> the maximum absorption region in which a clear<br />

1<br />

H.F. Seefeldt et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 24–31<br />

Log 1/R<br />

0.36<br />

0.34<br />

0.32<br />

0.30<br />

0.28<br />

0.26<br />

0.24<br />

1log/R<br />

B<br />

D<br />

580 620 660<br />

nm<br />

700 740<br />

9 DAF<br />

2290 2310 2330 2350<br />

nm<br />

Fig. 2. The complete NIR spectra (400–2498 nm) <strong>of</strong> the three genotypes and eight harvest times. (A) The full, EISC-treated spectra. Squares indicate enlargements shown in B and D.<br />

(B) Magnification <strong>of</strong> the chlorophyll peak (672 nm). (C) PCA model <strong>of</strong> the full NIR region. Blue, first three harvests; Purple, fourth and fifth harvests; cerise, sixth harvest; pink,<br />

seventh and eighth harvests. (D) The region 2280–2360 nm. Open arrows indicate spectral features <strong>of</strong> interest. Black arrow indicates spectra <strong>of</strong> 9 DAF.<br />

Cork<br />

lys5f<br />

lys3a<br />

difference between the three genotypes is observed. At 1002 cm 1<br />

the high BG mutant lys5f has a different intensity compared to Cork<br />

and lys3a. Moreover, lys5f has a lower intensity <strong>of</strong> the shoulder<br />

around 1070 cm 1 , whereas lys3a and Cork have well-defined<br />

peaks at 1078 cm 1 . With regard to the temporal changes, a shift<br />

from 1036 cm 1 for 9 DAF towards 1021 cm 1 for the end <strong>of</strong> the<br />

grain filling is observed in the spectra (Fig. 4c). The anomer specific<br />

peaks in the range from 800 to 950 cm 1 contain information about<br />

BG as well as other polysaccharide cell wall components (Philippe<br />

et al., 2006). Fig. 4d displays an enlargement <strong>of</strong> this region as<br />

Table 1<br />

The most informative peaks and their putative alignments as determined from loading plots corresponding to PCAs made on the defined regions<br />

Region<br />

NIR method<br />

PC Peak 1 Peak 2 Peak 3 Peak 4 Alignment<br />

500–750 1 672 þ400 þ1440 þ1930 672: chlorophyll a, 1440: C–H, O–H in starch,1940 H2O<br />

500–750 2 672 400 þ500 þ560<br />

1100–2498 1 þ1444 1856 þ1930 1670 1440 O–H starch, H2O, 1940 H2O<br />

1100–2498 3 þ2250 2477 þ2355 þ2384 2252, 2461 O–H starch, 2353 C–H cellulose, 2380: R–OH<br />

2280–2360 1 2308 2348 2310 CH2, 2347 HC¼CHCH2<br />

2280–2360 2 þ2288 þ2314 2360 2280: CH3, 2310 C–H CH2, 2352: C–H cellulose<br />

IR method<br />

1900–750 1 þ1010 960 900–990: aromatic, 1000–1260: C–O alcohol,<br />

1900–750 2 1049 þ1147 þ1002 þ1094 1002 ring stretching (arabinoxylans a ), 1160–1210:C–C(O)–C ester<br />

The peaks with the highest intensity are listed first. þ indicates a positive loading, while indicates a negative loading. Units for region: NIR, nm; IR, cm 1 .<br />

a From (Philippe et al., 2006). All other alignments from (Osborne et al., 1993).


logR/1<br />

0.45<br />

0.4<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

A B<br />

Average NIR flour spectra<br />

Betaglucan<br />

Cellulose<br />

Wheat starch<br />

Wheat amylose<br />

1400 1600 1800 2000 2200 2400<br />

nm<br />

Absorbance<br />

0.16<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

BG<br />

peak<br />

Average IR flour spectra<br />

Betaglucan<br />

Cellulose<br />

Wheat starch<br />

Wheat amylose<br />

800 900 1000 1100 1200 1300 1400 1500 1600 1700<br />

cm-1 Fig. 3. IR and NIR spectra <strong>of</strong> pure seed carbohydrates compared to an averaged IR and NIR flour spectrum. (A) IR spectra. The arrow indicates the BG and cellulose specific peak. (B)<br />

NIR spectra.<br />

Absorbance Absorbance<br />

0.11<br />

0.1<br />

0.09<br />

0.08<br />

0.07<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.11<br />

0.10<br />

0.09<br />

0.08<br />

0.07<br />

0.06<br />

0.05<br />

A Cork<br />

0.10 B<br />

lys5f<br />

0.08<br />

lys3a<br />

0.06<br />

800 1000 1200 1400 1600 1800<br />

C<br />

9-13 DAF<br />

cm -1<br />

Cork<br />

lys5f<br />

lys3a<br />

cm-1 980 1000 1020 1040 1060 1080 1100 1120<br />

H.F. Seefeldt et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 24–31 29<br />

Scores PC#2 (1.4 %)<br />

0.04<br />

0.02<br />

0<br />

-0.02<br />

-0.04<br />

-0.06<br />

2.derivative<br />

7<br />

4<br />

8<br />

8 628 2<br />

67<br />

7 78<br />

6 5<br />

53<br />

8<br />

6<br />

3 3<br />

78<br />

6 4 7<br />

356 74<br />

7 786<br />

5 5<br />

8<br />

878<br />

4<br />

86 7<br />

54<br />

6 5<br />

6<br />

5 5<br />

7<br />

78<br />

7<br />

8<br />

78 2<br />

66 8<br />

665<br />

56<br />

5 3<br />

63<br />

2<br />

24<br />

-0.36 -0.34 -0.32 -0.3 -0.28 0.26 -0.24 -0.22 -0.2<br />

D<br />

Cork<br />

lys5f<br />

lys3a<br />

Scores PC#1 (97 %)<br />

57<br />

1<br />

1<br />

Cork<br />

+ lys5f<br />

lys3a<br />

cm-1 820 840 860 880 900 920 940<br />

Fig. 4. FT-IR spectra <strong>of</strong> the region 750–1900 cm 1 . (A) EISC-treated spectra. Squares are indicated magnified in C and D. (B) PCA <strong>of</strong> the second derivative <strong>of</strong> the spectra (not meancentred).<br />

(C) EISC-pre-processed spectra <strong>of</strong> the region <strong>of</strong> maximum absorption. Arrows indicate 9 and 13 DAF. (D) The second-derivative spectra <strong>of</strong> the region from 800 to 940 cm 1 .


30<br />

RMSECV<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

314568775617645654625615486531<br />

1098 1290 1482 1672 1856 2040 2224 2408<br />

Wavelength(nm)<br />

second-derivative spectra which reveal that lys5f has less intensive<br />

peaks compared to Cork and lys3a. Cork and lys3a display a relatively<br />

strong a-anomer band at 855 cm 1 , whereas the high BG lys5f<br />

and the early stages <strong>of</strong> grain filling show a much weaker a-glucan<br />

peak. This is in good accordance with the starch levels in Cork and<br />

lys3a compared to lys5f. In the case <strong>of</strong> the b-anomer sensitive peak<br />

normally present at approximately 890 cm 1 (Fig. 4d), we observed<br />

higher background intensity in lys5f than in the other two genotypes.<br />

From approximately 23 DAF, the intensity is the highest<br />

corresponding to the highest level <strong>of</strong> BG. Pure barley BG shows<br />

a peak at 895 cm 1 , not present in the average grain spectra (see<br />

Fig. 3b).<br />

A PCA was applied to the full second-derivative IR spectra<br />

(Fig. 4b) and the first PC explains 97% <strong>of</strong> the variance. Both the first<br />

and the second PC separate the genotypes. The corresponding<br />

loading plot (Table 1) revealed that PC1 is mainly spanned by the<br />

peak at 1010 cm 1 and PC2 at 1049 cm 1 , which are the dominating<br />

peaks <strong>of</strong> starch in the second-derivative spectra <strong>of</strong> the pure<br />

substances (Fig. 3b). Thus BG seems to play a less pivotal role in<br />

discriminating the genotypes compared to starch.<br />

3.4. Relationships between the chemical and spectroscopic methods<br />

In order to study the relationships between the spectral data<br />

and grain filling, a range <strong>of</strong> calibration models were investigated.<br />

Global PLS models based on the entire data set were used, as well as<br />

iPLS models, where sub-regions are tested against the global<br />

model. Only a few studies have attempted to construct an NIR<br />

calibration to BG content. The quality <strong>of</strong> the predictions has varied<br />

greatly. A correlation coefficient <strong>of</strong> 0.69 and an RMSE <strong>of</strong> 0.557% was<br />

achieved in the BG range 5.8–8.4% (Czuchajowska et al., 1992). In<br />

another study, an 1-VR value that is a determination for cross<br />

validation and measure <strong>of</strong> goodness <strong>of</strong> fit was found to be 0.92 and<br />

an SECV <strong>of</strong> 0.45 in the BG range 0.09–5.12% (Blakeney and Flinn,<br />

2005). One major problem has been the limited BG range for the<br />

calibrations. However, Szczodrak et al. (1992) succeeded in developing<br />

calibration models to BG with a correlation coefficient <strong>of</strong><br />

0.871 and an SEE (standard error <strong>of</strong> estimate) <strong>of</strong> 0.677 using the<br />

calibration range 2.7–9.5%.<br />

In this study, the BG variation (calibration range) was further<br />

expanded in order to establish a quantitative model between the<br />

NIR spectra and the BG content. The global NIR model predicts BG<br />

with a squared correlation coefficient <strong>of</strong> 0.84 and an RMSECV <strong>of</strong><br />

2.60 using two PLS components with three outliers removed (lys5f<br />

from 9 DAF and two from 30 DAF). However, application <strong>of</strong> iPLS<br />

H.F. Seefeldt et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 24–31<br />

A B<br />

Predicted (BG%)<br />

20<br />

15<br />

10<br />

5<br />

0<br />

r = 0.97<br />

RMSECV = 1.57<br />

revealed that the region 1194–1240 nm yielded an R 2 <strong>of</strong> 0.94 and<br />

RMSECV <strong>of</strong> 1.6% using four PLS components (Fig. 5a,b). This subregion<br />

covers second overtones <strong>of</strong> the C-H stretching and shows<br />

that BG can be predicted well in this region.<br />

PLS calibrations to the FT-IR region and BG were also studied.<br />

The global FT-IR models predicted BG slightly better than the global<br />

NIR model with a squared correlation coefficient <strong>of</strong> 0.92 and<br />

a RMSECV value <strong>of</strong> 1.83 using six PLS components and removing<br />

two outliers (lys5f from 9 DAF and from 30 DAF) from the model.<br />

Subsequently an iPLS was applied to the IR spectra but no intervals<br />

were able to improve the calibration performance when compared<br />

to the global model. It thus appear that the information regarding<br />

BG is hidden across the IR spectra from 1900 to 750 cm 1 , and not<br />

confined to the beta specific peak at approx 890 cm 1 , whereas the<br />

BG information in the NIR spectra can with advantage be correlated<br />

with the region <strong>of</strong> C-H stretching at 1194–1240 nm.<br />

Correlations with the amylose content were predicted with<br />

a squared correlation coefficient <strong>of</strong> 0.81 and an RMSECV <strong>of</strong> 2.46 for<br />

the FT-IR and 0.77 and an RMSECV <strong>of</strong> 2.71 for NIR. The poorer<br />

correlations with amylose probably reflect the uncertainty <strong>of</strong> the<br />

amylose determinations more than the ability <strong>of</strong> the spectra to<br />

predict amylose or starch content.<br />

4. Conclusions<br />

0 5 10 15 20<br />

Measured (BG%)<br />

Fig. 5. RMSECV correlation coefficients <strong>of</strong> iPLS models <strong>of</strong> NIR and FT-IR spectra. (A) The iPLS performed on 30 intervals superimposed on the NIR spectra. Many intervals are better<br />

calibrators (smaller RMSECV values) for BG than the global model using two PLS components. Numbers on bars indicate PLS components used. The red marked interval 1194–<br />

1240 nm is the best calibrator. (B) Measured versus predicted plot <strong>of</strong> the model in the interval 1194–1240 nm and with three outliers expelled: two replicates <strong>of</strong> lys5f, 30 DAF and<br />

one lys5f at 9 DAF.<br />

A rapid increase in grain fresh weight during grain filling is<br />

concomitant with the increase <strong>of</strong> starch and BG. The high BG<br />

mutant lys5f shows an exponential increase <strong>of</strong> BG until 23 DAF and<br />

reaches a final content <strong>of</strong> 18%. This is in contrast to the low BG<br />

mutant lys3a and the conventional malting barley Cork having<br />

linear phases <strong>of</strong> BG deposition and reaching maxima <strong>of</strong> only 4 and<br />

6%, respectively. Starch deposition in these latter genotypes was<br />

exponential, reaching maxima <strong>of</strong> 47% compared to a linear increase<br />

in lys5f leading to a maximum <strong>of</strong> 23% starch.<br />

The NIR and FT-IR spectra are complete chemical fingerprints <strong>of</strong><br />

the grain and thus more complex to resolve than single chemical<br />

analysis <strong>of</strong> grains. Data extraction relies on the use <strong>of</strong> chemometrics<br />

to reveal hidden information. In the case <strong>of</strong> NIR, the genetic variation<br />

was mainly resolved from the spectral region between 2260<br />

and 2380 nm. Moreover, the NIR spectra showed mostly temporal<br />

differences primarily due to complex changes in moisture. In the<br />

NIR region 2260–2380 nm, all the barley flour spectra differed and<br />

the average grain spectrum was influenced by both the starch/<br />

amylose and the BG contents, which underlines the importance <strong>of</strong><br />

this NIR region for breeding purposes.


Both NIR and IR spectra gave good correlations with BG. The NIR<br />

sub-region 1194–1240 nm gave the best correlation to BG with an<br />

R 2 <strong>of</strong> 0.94 which is promising for using NIR to predict BG in<br />

heterogeneous samples such as seeds and flour with a wide span in<br />

BG content.<br />

Fast and non-destructive spectroscopic fingerprinting is obviously<br />

advantageous for studying physiological processes such as<br />

grain filling, as the techniques are non-destructive, fast and sensitive<br />

and it is possible to do real-time analysis.<br />

Acknowledgements<br />

The DIAS Competence Fund and the Ministry <strong>of</strong> Food Agriculture<br />

and Fisheries are greatly acknowledged for financial support to the<br />

project (FFS05-9: Build Your Food). Betina Sørensen (University <strong>of</strong><br />

Aarhus) and Lisbeth T. Hansen and Louise Nancke (Faculty <strong>of</strong> Life<br />

<strong>Science</strong>s, University <strong>of</strong> Copenhagen) are acknowledged for technical<br />

support in semi-field experiments and chemical analysis.<br />

Gilda Kischinovsky is acknowledged for pro<strong>of</strong>reading. The authors<br />

wish to thank Lars Munck for valuable comments and for sharing<br />

his unique barley mutant collection.<br />

References<br />

Bay-Smidt, A.M., Blennow, A., Bojko, M., Møller, B.L., 1999. The effect <strong>of</strong> phosphate<br />

and amylopectin molecular size on peak viscosity <strong>of</strong> starch pastes. Annual<br />

Transactions <strong>of</strong> the Nordic Rheology Society 7, 31–38.<br />

Blakeney, A.B., Flinn, P.C., 2005. Determination <strong>of</strong> non-starch polysaccharides in<br />

cereal grains with near-infrared reflectance spectroscopy. Molecular Nutrition &<br />

Food Research 49, 546–550.<br />

Blennow, A., Bay-Smidt, A.M., Wischmann, B., Olsen, C.E., Møller, B.L., 1998. The<br />

degree <strong>of</strong> starch phosphorylation is related to the chain length distribution <strong>of</strong><br />

the neutral and the phosphorylated chains <strong>of</strong> amylopectin. Carbohydrate<br />

Research 307, 45–54.<br />

Chen, L., Carpita, N.C., Reiter, W.-D., Wilson, R.H., Jeffries, C., McCann, M.C., 1998. A<br />

rapid method to screen for cell-wall mutants using discriminant analysis <strong>of</strong><br />

Fourier transform infrared spectra. The Plant <strong>Journal</strong> 16, 385–392.<br />

Coles, G., 1979. Relationship <strong>of</strong> mixed-link beta-glucan accumulation to accumulation<br />

<strong>of</strong> free sugars and other glucans in the developing barley endosperm.<br />

Carlsberg Research Communications 44, 439–453.<br />

Czuchajowska, Z., Szczodrak, J., Pomeranz, Y., 1992. Characterization and estimation<br />

<strong>of</strong> barley polysaccharides by near-infrared spectroscopy.1. Barleys, starches, and<br />

betaDglucans. <strong>Cereal</strong> Chemistry 69, 413–418.<br />

Delwiche, S.R., 1995. Single wheat kernel analysis by near-infrared transmittance:<br />

protein content. <strong>Cereal</strong> Chemistry 72, 11–16.<br />

Gergely, S., Salgo, A., 2003. Changes in moisture content during wheat maturation –<br />

what is measured by near infrared spectroscopy? <strong>Journal</strong> <strong>of</strong> Near Infrared<br />

Spectroscopy 11, 17–26.<br />

Gergely, S., Salgo, A., 2005. Changes in carbohydrate content during wheat maturation<br />

– what is measured by near infrared spectroscopy? <strong>Journal</strong> <strong>of</strong> Near<br />

Infrared Spectroscopy 13, 9–17.<br />

Jacobsen, S., Søndergaard, I., Møller, B., Desler, T., Munck, L., 2005. A chemometric<br />

evaluation <strong>of</strong> the underlying physical and chemical patterns that support near<br />

infrared spectroscopy <strong>of</strong> barley seeds as a tool for explorative classification <strong>of</strong><br />

endoperm genes and gene combinations. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 42, 281–299.<br />

Jenner, C.F., Ugalde, T.D., Aspinall, D., 1991. The physiology <strong>of</strong> starch and protein<br />

deposition in the endosperm <strong>of</strong> wheat. Australian <strong>Journal</strong> <strong>of</strong> Plant Physiology<br />

18, 211–226.<br />

H.F. Seefeldt et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 24–31 31<br />

Jennings, A.C., Morton, R.K., 1962. Changes in carbohydrate, protein and non-protein<br />

nitrogenous compounds <strong>of</strong> developing wheat grain. Australian <strong>Journal</strong> <strong>of</strong> Biological<br />

<strong>Science</strong>s 16, 318–331.<br />

Kacurakova, M., Wilson, R.H., 2001. Developments in mid-infrared FT-IR spectroscopy<br />

<strong>of</strong> selected carbohydrates. Carbohydrate Polymers 44, 291–303.<br />

Martens, H., Nielsen, J.P., Engelsen, S.B., 2003. Light scattering and light absorbance<br />

separated by extended multiplicative signal correction. Application to nearinfrared<br />

transmission analysis <strong>of</strong> powder mixtures. Analytical Chemistry 75,<br />

394–404.<br />

McCann, M.C., Hammouri, M., Wilson, R., Belton, P., Roberts, K., 1992. Fourier<br />

transform infrared microspectroscopy is a new way to look at plant cell walls.<br />

Plant Physiology 100, 1940–1947.<br />

Munck, L., 2007. Breeding for Quality Traits in <strong>Cereal</strong>s – A Revised Outlook on Old<br />

and New Tools for Integrated Breeding. <strong>Cereal</strong>s. Springer Verlag, New York.<br />

Munck, L., Møller, B., 2005. Principal component analysis <strong>of</strong> near infrared spectra as<br />

a tool <strong>of</strong> endosperm mutant characterisation and in barley breeding for quality.<br />

Czech <strong>Journal</strong> <strong>of</strong> Genetics and Plant Breeding 41, 89–95.<br />

Munck, L., Jørgensen, K.G., Ruudhansen, J., Hansen, K.T., 1989. The EBC methods for<br />

determination <strong>of</strong> high molecular-weight beta-glucan in barley, malt, wort and<br />

beer. <strong>Journal</strong> <strong>of</strong> the Institute <strong>of</strong> Brewing 95, 79–82.<br />

Munck, L., Møller, B., Jacobsen, S., Søndergaard, I., 2004. Near infrared spectra<br />

indicate specific mutant endosperm genes and reveal a new mechanism for<br />

substituting starch with (1-3,1-4) betaglucan in barley. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

40, 213–222.<br />

Nørgaard, L., Saudland, A., Wagner, J., Nielsen, J.P., Munck, L., Engelsen, S.B., 2000.<br />

Interval partial least-squares regression (iPLS). A comparative chemometric<br />

study with an example from near-infrared spectroscopy. Applied Spectroscopy<br />

54, 413–419.<br />

Osborne, B.G., 2006. Applications <strong>of</strong> near infrared spectroscopy in quality screening<br />

<strong>of</strong> early-generation material in cereal breeding programmes. <strong>Journal</strong> <strong>of</strong> Infrared<br />

Spectroscopy 14, 93–101.<br />

Osborne, B.G., Fearn, T., Hindle, P.H., 1993. Theory <strong>of</strong> near infrared spectrophotometry.<br />

Practical NIR Spectroscopy with Applications in Food and Beverage<br />

Analysis, second ed. Longman Scientific and Technical, Harlow, UK, pp. 13-35.<br />

Patron, N.J., Greber, B., Fahy, B.F., Laurie, D.A., Parker, M.L., Denyer, K., 2004. The lys5<br />

mutations <strong>of</strong> barley reveal the nature and importance <strong>of</strong> plastidial ADP-Glc<br />

transporters for starch synthesis in cereal endosperm. Plant Physiology 135,<br />

2088–2097.<br />

Pedersen, D.K., Martens, H., Nielsen, J.P., Engelsen, S.B., 2002. Near-infrared<br />

absorption and scattering separated by extended inverted signal correction<br />

(EISC): analysis <strong>of</strong> near-infrared transmittance spectra <strong>of</strong> single wheat seeds.<br />

Applied Spectroscopy 56, 1206–1214.<br />

Philippe, S., Robert, P., Barron, C., Saulnier, L., Guillon, F., 2006. Deposition <strong>of</strong> cell<br />

wall polysaccharides in wheat endosperm during grain development: Fourier<br />

transform-infrared microspectroscopy study. <strong>Journal</strong> <strong>of</strong> Agricultural and Food<br />

Chemistry 54, 2303–2308.<br />

Robert, P., Marquis, M., Barron, C., Guillon, F., Saulnier, L., 2005. FT-IR investigation<br />

<strong>of</strong> cell wall polysaccharides from cereal grains. Arabinoxylan infrared assignment.<br />

<strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 53, 7014–7018.<br />

Séné, C.F.B., McCann, M.C., Wilson, R.H., Grinter, R., 1994. Fourier-transform raman<br />

and Fourier-transform infrared spectroscopy – an investigation <strong>of</strong> five higher<br />

plant cell walls and their components. Plant Physiology 106, 1623–1631.<br />

Szczodrak, J., Czuchajowska, Z., Pomeranz, Y., 1992. Characterization and estimation<br />

<strong>of</strong> barley polysaccharides by near-infrared spectroscopy. 2. Estimation <strong>of</strong> total<br />

betaDglucans. <strong>Cereal</strong> Chemistry 69, 419–423.<br />

Tester, R.F., Karkalas, J., Qi, X., 2004. Starch – composition, fine structure and<br />

architecture. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 39, 151–165.<br />

Tønning, E., 2008. Wheat baking quality in a process analytical perspective.<br />

Sampling, diversification, prediction and chemometric method development.<br />

PhD thesis, University <strong>of</strong> Copenhagen. ISBN: 87-91949-23-8.<br />

Wold, S., Esbensen, K.H., Geladi, P., 1987. Principal component analysis. Chemometrics<br />

and Intelligent Laboratory Systems 2, 37–52.<br />

Zachariassen, C.B., Larsen, J., van den Berg, F., Engelsen, S.B., 2005. Use <strong>of</strong> NIR<br />

spectroscopy and chemometrics for on-line process monitoring <strong>of</strong> ammonia in<br />

low methoxylated amidated pectin production. Chemometrics and Intelligent<br />

Laboratory Systems 76, 149–161.


Mechanism <strong>of</strong> gas cell stabilization in bread making. I. The primary<br />

gluten–starch matrix q<br />

Baninder S. Sroan a, *, Scott R. Bean b , Finlay MacRitchie a<br />

a Department <strong>of</strong> Grain <strong>Science</strong> and Industry, Kansas State University, Manhattan, KS 66506, USA<br />

b USDA-ARS, Grain Marketing and Production Research Center, Manhattan, KS 66502, USA<br />

article info<br />

Article history:<br />

Received 19 December 2007<br />

Received in revised form 28 June 2008<br />

Accepted 8 July 2008<br />

Keywords:<br />

Bread making<br />

Gas cell stability<br />

Liquid lamella<br />

Gluten–starch matrix<br />

Dough rheology<br />

1. Introduction<br />

abstract<br />

During mixing, gas is occluded and concentrated in the liquid<br />

phase <strong>of</strong> dough in the form <strong>of</strong> small nuclei. No further occlusion <strong>of</strong><br />

Abbreviations: MALLS, multiangle laser light scattering; MDDT, mixograph<br />

dough development time; MT, threshold molecular weight; MW, molecular weight;<br />

MW, weight average molecular weight; MWD, molecular weight distribution; Rmax,<br />

extensograph maximum resistance to extension; SDS, sodium dodecyl sulfate; SE-<br />

HPLC, size-exclusion high-performance liquid chromatography; UPP, unextractable<br />

polymeric protein.<br />

q Mention <strong>of</strong> firm names or trade products does not constitute endorsement by<br />

the U.S. Department <strong>of</strong> Agriculture over others not mentioned.<br />

* Corresponding author. Present address: FritoLay, 7701 Legacy Dr., 3T-124, Plano,<br />

TX 75024, USA. Tel.: þ1 972 334 4526; fax: þ1 972 334 2329.<br />

E-mail address: baninder@k-state.edu (B.S. Sroan).<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.07.003<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 32–40<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

Expansion <strong>of</strong> dough and hence bread making performance is postulated to depend on a dual mechanism<br />

for stabilization <strong>of</strong> inflating gas bubbles. Two flours were used in this study, one from the wheat variety<br />

Jagger (Jagger) and the other from a composite <strong>of</strong> s<strong>of</strong>t wheat varieties (S<strong>of</strong>t). Thin liquid lamellae (films),<br />

stabilized by adsorbed surface active compounds, act as an auxiliary to the primary gluten–starch matrix<br />

in stabilizing expanding gas cells and this mechanism operates when discontinuities begin to appear in<br />

the gluten–starch matrix during later proving and early baking stages. Contributions <strong>of</strong> the liquid<br />

lamellae stability to dough expansion were assessed using flours varying in their lipid content. Incremental<br />

addition <strong>of</strong> natural lipids back into defatted flour caused bread volume to decrease, and, after<br />

reaching a minimum, to increase. Strain hardening is a key rheological property responsible for stabilizing<br />

the primary gluten–starch matrix. Jagger gave higher test-bake loaf volume than S<strong>of</strong>t and higher<br />

strain hardening index for dough. The different lipid treatments were found to have negligible effects on<br />

strain hardening index. Image analysis <strong>of</strong> crumb grain revealed that differences in number <strong>of</strong> gas cells<br />

and average cell elongation with different lipid treatments were insignificant. The evidence agrees with<br />

a dual mechanism to stabilize the gas cells in bread dough. To understand dough rheology at a molecular<br />

level, rheological properties <strong>of</strong> doughs were varied by addition <strong>of</strong> flour protein fractions prepared by pH<br />

fractionation. Fractions were characterized by SE-HPLC and MALLS. The molecular weight distribution<br />

(MWD) <strong>of</strong> fractions progressively shifted to higher values as the pH <strong>of</strong> fractionation decreased. Mixograph<br />

dough development time paralleled the MWD. However, the strain hardening index and the testbake<br />

loaf volume increased with increasing MWD up to a point (optimum), after which they declined. At<br />

a given strain rate, the behavior at the optimum is thought to result from slippage <strong>of</strong> the maximum<br />

number <strong>of</strong> statistical segments between entanglements, without disrupting the entangled network <strong>of</strong><br />

polymeric proteins. Shift <strong>of</strong> MWD to molecular weight higher than the optimum results in a stronger<br />

network with reduced slippage through entanglement nodes, whereas a shift to lower molecular weights<br />

will decrease the strength <strong>of</strong> the network due to a lesser number <strong>of</strong> entanglements per chain.<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

gas occurs in succeeding stages (sheeting, molding, etc.) <strong>of</strong> the<br />

bread making process (Baker and Mize, 1946). However, these<br />

subsequent stages (punching, sheeting and molding) cause subdivision<br />

<strong>of</strong> already existing gas cells, thus improving their number<br />

and size distribution. Gas nuclei expand during fermentation due to<br />

release <strong>of</strong> fermentation gases, and during baking due to expansion<br />

<strong>of</strong> these gases as temperature increases.<br />

Wheat flour dough, due to its unique visco-elastic properties, is<br />

capable <strong>of</strong> stabilizing the expanding gas cells. The classical view has<br />

emphasized only the gluten–starch matrix as the cell membrane<br />

that stabilizes the expanding gas cells (Bloksma, 1990; Hoseney,<br />

1992). However, scanning electron micrographs <strong>of</strong> dough at the end<br />

<strong>of</strong> the proving stage appear to show the existence <strong>of</strong> intact gas cells<br />

even when discontinuities in the gluten–starch matrix begin to<br />

appear (Gan et al., 1990). Based on these electron micrographs <strong>of</strong>


the dough, Gan and co-workers (Gan et al., 1990, 1995) proposed<br />

that the expanding gas cells appear to be stabilized against coalescence<br />

and disproportionation by a primary gluten–starch matrix<br />

with a secondary liquid lamella on its inner side, enveloping the gas<br />

cell. The dual film hypothesis seems plausible in view <strong>of</strong> evidence<br />

from previous studies, which show presence <strong>of</strong> a continuous liquid<br />

phase in dough (MacRitchie, 1976a) and the effect <strong>of</strong> surface active<br />

components on bread loaf volume (DeStefanis and Ponte, 1976;<br />

MacRitchie, 1976a,b; MacRitchie and Gras, 1973). In such cases,<br />

where there is an independent mechanism <strong>of</strong> gas cell stabilization<br />

apart from the gluten–starch matrix, the dough rheology should<br />

not be affected by variations in natural lipid levels. MacRitchie and<br />

Gras (1973) reported that Alveograms are not affected by variations<br />

in the natural lipid fraction <strong>of</strong> the flour. A Stable Micro Systems<br />

dough inflation system mounted on a texture analyzer (TAXT<br />

2plus), used in this study to understand biaxial extensional<br />

rheology <strong>of</strong> wheat flour dough, works on a principle similar to the<br />

Alveograph, with the advantage <strong>of</strong> inflating bubbles at constant<br />

strain rates. The amounts <strong>of</strong> lipid in the study by MacRitchie and<br />

Gras (1973) were as low as 1–1.5% and had significant effects on<br />

baking performance. It is probably due to their surface activity that<br />

these compounds (lipids and proteins) are adsorbed at the gas–<br />

liquid interface <strong>of</strong> the liquid lamellae and affect stability <strong>of</strong> the gas<br />

cells (Mills et al., 2003; Paternotte et al., 1993; Ross and MacRitchie,<br />

1995). Analogous variations in foaming properties <strong>of</strong> the dough<br />

liquor, and bread loaf volume and crumb structure (MacRitchie,<br />

1976a) provide reasonable evidence for their action at the interface.<br />

Loaf volume can be defined as the extent <strong>of</strong> dough expansion<br />

(Gandikota and MacRitchie, 2005), which depends upon how thin<br />

the gluten–starch matrix can be stretched before reaching its<br />

expansion limit or point <strong>of</strong> rupture. Rheology <strong>of</strong> the gluten–starch<br />

matrix is important in the bread making process as this determines<br />

extensibility and strength. The proving and baking stages <strong>of</strong> bread<br />

making are characterized by fast biaxial expansion <strong>of</strong> gas cells,<br />

expanding at strain rates <strong>of</strong> 0.001–0.0001/s and 0.01–0.001/s,<br />

respectively (Dobraszczyk, 1997). During expansion, the gluten–<br />

starch matrix around gas cells expands biaxially to large strains<br />

(>100%) due to excess pressure produced in the gas cells by<br />

diffusion <strong>of</strong> carbon dioxide during proving, and by thermal<br />

expansion <strong>of</strong> gases during baking. This causes thinning <strong>of</strong> gas cell<br />

walls and, if a gas cell continues to expand along this thin region, it<br />

may rupture. However, if the stress in the thin region increases<br />

more than proportionally to strain, the thin region <strong>of</strong> a cell wall or<br />

(gluten–starch matrix) will resist further deformation and the gas<br />

cell will continue to expand along thicker parts <strong>of</strong> the cell wall. This<br />

localized increase <strong>of</strong> stress in response to strain, preventing failure<br />

<strong>of</strong> the gas cell walls, is called strain hardening (Dobraszczyk and<br />

Roberts, 1994; van Vliet et al., 1992), and is a necessary rheological<br />

property for obtaining good bread volume.<br />

A power law relation between stress and Hencky strain, as<br />

described by the equation below, was used by Dobraszczyk and<br />

Roberts (1994) to determine strain hardening.<br />

s [ K3 n<br />

where, s is the true stress, K is a power law constant, 3 is strain and<br />

n is a strain hardening index. The index n must be greater than 1 in<br />

order to have a curved relation between stress and Hencky strain<br />

(Fig. 1). Doughs with a strain hardening index <strong>of</strong> 1 or higher have<br />

the potential to give good loaf volume by allowing gas cells to<br />

expand without undergoing disproportionation and coalescence<br />

(Dobraszczyk and Roberts, 1994). In later studies, Dobraszczyk and<br />

co-workers (Dobraszczyk and Salmanowicz, 2008; Dobraszczyk<br />

et al., 2003) found that an exponential relationship shows better fit<br />

to data, particularly in the case <strong>of</strong> doughs for which bubbles inflate<br />

B.S. Sroan et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 32–40 33<br />

Fig. 1. Stress vs. strain (Hencky) curves <strong>of</strong> S<strong>of</strong>t and Jagger wheat flours showing<br />

differences in their tendencies to strain harden. The S<strong>of</strong>t wheat flour with poor bread<br />

making potential fails at relatively lower strains.<br />

to strains greater than 2. Chin and Campbell (2005) also used an<br />

exponential relationship to study dough rheology by bubble inflation.<br />

Strain hardening index is a reliable criterion for differentiating<br />

flours based on bread making potential (Dobraszczyk and Roberts,<br />

1994; Dobraszczyk and Salmanowicz, 2008; Dobraszczyk et al.,<br />

2003; Tronsmo et al., 2003;). This is illustrated in Fig. 1 for the two<br />

flours used in this study.<br />

At a molecular level, strain hardening can be explained by a well<br />

established polymer entanglement network theory (MacRitchie<br />

and Lafiandra, 1997; Singh and MacRitchie, 2001). Strain hardening<br />

is believed to originate from entanglement coupling <strong>of</strong> large glutenin<br />

molecules (Singh and MacRitchie, 2001), with molecular<br />

weight (MW) greater than a threshold MW (MT), MT being the<br />

minimum molecular weight at which stable entanglements are<br />

formed (Bersted and Anderson, 1990). Various studies substantiate<br />

this theory. Dough strength in terms <strong>of</strong> Rmax (extensograph<br />

maximum resistance to extension) shows positive correlation <strong>of</strong><br />

the glutenin fraction with MW greater than MT, which is estimated<br />

to be approximately 250,000 Da (Bangur et al., 1997). The shift in<br />

MW distribution (MWD) <strong>of</strong> this fraction towards higher MW or<br />

increase in relative proportion <strong>of</strong> this fraction will increase the<br />

number <strong>of</strong> entanglements per chain and reduce the MW between<br />

entanglements (Termonia and Smith, 1988). This will lead to an<br />

increase in strength (Gupta et al., 1990). A similar phenomenon is<br />

observed in wheat flour doughs, where the strain hardening index<br />

(n) seems to have a positive curvilinear relation with failure strain.<br />

As n approaches higher limits, the failure strain seems to become<br />

constant (Dobraszczyk and Roberts, 1994; Dobraszczyk et al., 2003).<br />

Successive addition <strong>of</strong> glutenin fractions with increasing molecular<br />

weight to a base flour at constant protein level causes loaf volume<br />

to increase, and, after reaching a maximum, to decrease (Lundh and<br />

MacRitchie, 1989; MacRitchie, 1987; MacRitchie et al., 1991). The<br />

optimum here possibly indicates a balance between strength and<br />

extensibility, beyond which high entanglement network density,<br />

resulting in decrease in failure strain, might cause lower loaf<br />

volumes (Singh and MacRitchie, 2001; Termonia and Smith, 1988).<br />

The objectives <strong>of</strong> this study were: firstly, to investigate and seek<br />

evidence for the presence <strong>of</strong> liquid lamellae and their ability to<br />

stabilize gas cells; i.e. whether or not the dual film hypothesis is<br />

plausible; secondly, to understand the rheology <strong>of</strong> the gluten–<br />

starch matrix required to stabilize the expanding gas cells and to<br />

understand the underlying molecular structure–function relationship<br />

as explained by polymer entanglement network theory<br />

(MacRitchie and Lafiandra, 1997; Singh and MacRitchie, 2001).


34<br />

2. Experimental<br />

2.1. Materials<br />

Jagger flour and s<strong>of</strong>t wheat flour (a blend <strong>of</strong> s<strong>of</strong>t wheat varieties)<br />

were two untreated and unbleached flours used in this study. The<br />

flours were milled in a Buhler mill (73% milling extraction rate) and<br />

were evaluated for certain quality characteristics which are shown<br />

in Table 1. These flours were stored at 20 C until use. For the sake<br />

<strong>of</strong> brevity, Jagger wheat flour and s<strong>of</strong>t wheat flour have been<br />

referred to as Jagger and S<strong>of</strong>t, respectively. Reagents were<br />

purchased from Sigma–Aldrich, USA. Chlor<strong>of</strong>orm was <strong>of</strong> HPLC<br />

grade, whereas all other chemicals used were <strong>of</strong> ACS grade. Distilled<br />

deionized water, sterilized in an autoclave, was used in all stages <strong>of</strong><br />

the experiments.<br />

2.2. Analytical procedures<br />

Moisture content was determined as per AACC method 44-15 A<br />

(AACC, 2001). Protein content was determined by the nitrogen<br />

combustion method using a LECO FP-2000 nitrogen/protein<br />

analyzer with a factor <strong>of</strong> 5.7 to convert N to protein. Lipid content<br />

was determined by cold extraction with chlor<strong>of</strong>orm (HPLC grade)<br />

using the procedure <strong>of</strong> MacRitchie and Gras (1973).<br />

2.3. Dough mixing properties<br />

Mixing properties were evaluated using a 10 g Mixograph<br />

(National Manufacturing Co., Lincoln, NE). Mixing parameters (peak<br />

development time and weakening angle) were used for comparison.<br />

The procedure used was similar to AACC method 55-40 (AACC,<br />

2001) except that sodium chloride (1.5% w/w on a flour weight<br />

basis) was added.<br />

2.4. Preparation <strong>of</strong> gluten proteins<br />

Gluten protein was prepared from defatted Jagger flour<br />

following the procedure used by MacRitchie (1985). The wet gluten<br />

mass was freeze-dried and ground to a particle size less than<br />

150 microns. Powdered gluten was stored at a temperature <strong>of</strong><br />

20 C until its use.<br />

2.5. pH fractionation <strong>of</strong> gluten protein<br />

Powdered gluten prepared from Jagger was subjected to pH<br />

fractionation, using the procedure <strong>of</strong> Gupta et al. (1990). Gluten was<br />

stirred in water at pH 5.3 (35 g/1000 ml) for 10 min and centrifuged<br />

at 2300 g for 10 min. The residue, after collection <strong>of</strong> supernatant,<br />

was subjected to two more extractions at the same pH. The total<br />

Table 1<br />

Physico-chemical analysis and SE-HPLC relative composition <strong>of</strong> polymeric proteins<br />

<strong>of</strong> Jagger and S<strong>of</strong>t wheat flours<br />

Jagger wheat<br />

flour<br />

S<strong>of</strong>t wheat<br />

flour<br />

Protein (%) (14% moisture basis) 10.4 0.02 9.2 0.03<br />

Lipid Content (%) (14% moisture basis) 0.89 0.03 0.93 0.01<br />

Mixograph midline peak development time<br />

(min)<br />

7.21 0.06 9.25 0.00<br />

Mixograph absorption (%) 63 60<br />

Area (%) under chromatogram curve<br />

TPP 36.4 0.01 37.0 0.05<br />

EPP 43.6 0.02 39.0 0.11<br />

UPP 56.4 2.77 61.1 3.40<br />

TPPdtotal polymeric protein; EPPdextractable polymeric protein; UPPdunextractable<br />

polymeric protein.<br />

B.S. Sroan et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 32–40<br />

combined supernatant was named as pH 5.3 gluten fraction. The<br />

residue was subjected to further fractionation in a similar manner<br />

by lowering <strong>of</strong> pH to 4.9 by addition <strong>of</strong> HCl. The gluten protein was<br />

thus fractionated into 6 fractions by sequential lowering <strong>of</strong> pH.<br />

Fractions were obtained at pH 5.3, 4.9, 4.1, 3.5 and 3.1, obtaining two<br />

fractions at pH 3.1 i.e. supernatant <strong>of</strong> pH 3.1 and residue <strong>of</strong> pH 3.1.<br />

All fractions were then freeze-dried. Freeze-dried fractions were<br />

ground to particle sizes less than 150 microns. The fractions were<br />

stored in polyethylene bags in containers with desiccant at<br />

a temperature <strong>of</strong> 20 C until their use.<br />

Fractions were analyzed for percentage protein and moisture.<br />

Protein size characterization <strong>of</strong> the fractions was done using<br />

SE-HPLC and MALLS. Fractions retained their functional properties<br />

(dough mixing properties) when reconstituted in proportions<br />

relative to their yield, with starch (Midsol-50, MGP Ingredients,<br />

Inc., Hutchison, KS) to form a flour with protein content equal to<br />

that <strong>of</strong> the original flour (Jagger).<br />

2.6. Addition <strong>of</strong> protein fractions to flour<br />

Addition <strong>of</strong> protein fractions to defatted base flour (Jagger and<br />

S<strong>of</strong>t) was done at a level <strong>of</strong> 1% (dry protein) on a total flour weight<br />

basis.<br />

2.7. Size-exclusion high-performance liquid chromatography<br />

(SE-HPLC) and multiangle laser light scattering (MALLS)<br />

Size characterization <strong>of</strong> gluten proteins was done using SE-HPLC<br />

(Hewlett-Packard 1100 system) in conjunction with MALLS (Multiangle<br />

light scattering detector DAWN EOS <strong>of</strong> Wyatt Technology<br />

Corp., Santa Barbara, CA). The basic procedure for SE-HPLC is<br />

described elsewhere (Batey et al., 1991). Proteins were fractionated<br />

with a Biosep SEC-4000 column (Phenomenex, Torrance, CA) using;<br />

50 mM disodium orthophosphate buffer (NaPhos), pH 7.0, containing<br />

1% sodium dodecyl sulfate (SDS), as mobile phase at a flow<br />

rate <strong>of</strong> 0.5 ml/min. Proteins were detected at 214 nm. Injection<br />

volume <strong>of</strong> samples for total and extractable protein analysis was<br />

80 ml and that for unextractable protein analysis was 40 ml. Therefore,<br />

results obtained from chromatograms <strong>of</strong> unextractable<br />

proteins were multiplied by a factor <strong>of</strong> 2. Reduced injection volume<br />

for unextractable protein analysis was used to avoid issues in<br />

MALLS analysis due to the high concentration <strong>of</strong> very large glutenin<br />

protein.<br />

SEC-HPLC data was analyzed using s<strong>of</strong>tware program Chem-<br />

Station (Agilent Technologies, USA). MALLS data was analyzed with<br />

s<strong>of</strong>tware program ASTRA 4.50 (Wyatt Technology Corp.) using 0.31<br />

as the dn/dc value, as per procedure by Bean and Lookhart (2001).<br />

2.8. Sample preparation for SE-HPLC<br />

The basic procedure for SE-HPLC is based on the method <strong>of</strong><br />

(Batey et al., 1991. Samples were prepared to analyze total,<br />

extractable and unextractable polymeric protein (Gupta et al.,<br />

1993). Samples for total and extractable protein analysis were<br />

weighed in micr<strong>of</strong>uge tubes. Sample size was determined based on<br />

protein content <strong>of</strong> each sample such that protein content in all<br />

samples was kept constant, using Jagger flour as standard and the<br />

quantity for Jagger being 10.0 mg 0.1 mg. Weighed samples were<br />

suspended in 50 mM NaPhos, pH 6.9, þ0.5% SDS and solubilized by<br />

vortexing for 10 min. For total protein analysis, to achieve solubilization<br />

<strong>of</strong> the largest molecular size fraction, samples were sonicated<br />

(Singh et al., 1990) at room temperature at an output <strong>of</strong> 6 W<br />

for 15 s. No sonication step was included for extractable protein<br />

analysis. The sonicator probe was placed at 1/3 distance from the<br />

bottom <strong>of</strong> the micr<strong>of</strong>uge tube. Micr<strong>of</strong>uge tubes with protein<br />

suspensions were then centrifuged at 12,000 g for 20 min. The


supernatant was decanted in HPLC vials and sealed. To ensure<br />

stability <strong>of</strong> prepared samples, the vials with supernatant were heat<br />

treated in a water bath at 85 C for 5 min to inhibit any intrinsic<br />

proteolytic activity. After heat treatment, vials were cooled with<br />

crushed ice and analyzed by SE-HPLC. Residue from extractable<br />

protein was used for unextractable protein analysis. The same<br />

procedure was followed, except that suspensions were sonicated<br />

for 25 s at an output <strong>of</strong> 6 W.<br />

2.9. Lipid extraction from flour<br />

Flour lipids were extracted using three batch extractions with<br />

chlor<strong>of</strong>orm in a glass beaker, followed by Buchner filtration through<br />

Whatman No. 1 filter paper (MacRitchie and Gras, 1973). 200 g <strong>of</strong><br />

flour and 400 ml <strong>of</strong> chlor<strong>of</strong>orm were used for each extraction. The<br />

defatted flour was spread out on a flat glass tray in a fume hood for<br />

12 h to allow evaporation <strong>of</strong> solvent.<br />

2.10. Addition <strong>of</strong> lipids to defatted flour<br />

For incorporation <strong>of</strong> intact natural flour lipids into defatted flour,<br />

the parent flour was mixed with defatted flour in different<br />

proportions, for both the flours, to give different flour lipid levels.<br />

2.11. Test baking<br />

Test loaves (35 g flour) were baked using a modified rapid bake<br />

test (MacRitchie and Gras, 1973). A lean formulation was used with<br />

no added shortening; flour (100%), sugar (6%), sodium chloride<br />

(1.5%), instant yeast (2.7%), potassium bromate (30 ppm), water and<br />

mix time (as optimized from Mixograph analysis). Loaf volumes<br />

were measured by rapeseed displacement after cooling for 20 min.<br />

2.12. Image analysis<br />

Image analysis <strong>of</strong> crumb grain <strong>of</strong> baked loaves was carried out<br />

12 h after baking, with a C-Cell, an image analyzing s<strong>of</strong>tware and<br />

equipment (Calibre Control International Ltd., UK). Loaves were<br />

sliced using a rotary disc blade (unserrated Graef Ò blade) cutter.<br />

Image analysis was performed on central slices <strong>of</strong> 15 mm thickness,<br />

as soon as possible to avoid any shrinkage <strong>of</strong> crumb grain. Image<br />

analysis parameters (number <strong>of</strong> cells and average cell elongation)<br />

were used for comparison between different treatments.<br />

2.13. Biaxial extensional rheology<br />

Biaxial extensional rheological properties <strong>of</strong> the doughs were<br />

measured with a Stable Micro Systems dough inflation system<br />

mounted on a texture analyzer (TAXT 2plus) by means <strong>of</strong> the<br />

procedure established by Dobraszczyk (1997). Doughs for rheological<br />

testing were mixed in the same mixer as used for bake tests,<br />

using the same water absorption, mixing times and sodium chloride<br />

addition. After mixing, dough pieces were squashed by hand<br />

on a sheeting board without putting too much stress on the dough,<br />

and then allowed to relax for 5 min. They were then sheeted, rolled<br />

out slowly with several passes and rotated by 90 degrees after each<br />

pass. Sheeting was done for 5 min with relaxation <strong>of</strong> 10 s between<br />

each pass. Sheeting in all directions prevents anisotropic effects<br />

during dough inflation, allowing dough pieces to expand uniformly<br />

into spherical shapes. After sheeting, dough pieces were relaxed for<br />

20 min. They were then cut into circular discs using a 55 mm cookie<br />

cutter, squashed to a height <strong>of</strong> 2.67 mm for 20 s. Sample dough<br />

pieces (in pots) were then proved at 35 C for 25 min. During<br />

sample preparation, dough pieces were protected against loss <strong>of</strong><br />

moisture using a fine coating <strong>of</strong> mineral oil (Saybold viscosity 335/<br />

358) and covering with shrink wrap film. Mineral oil <strong>of</strong> lower<br />

B.S. Sroan et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 32–40 35<br />

viscosity seems to penetrate dough pieces and may affect rheological<br />

measurements.<br />

Dough pieces were inflated at a flow rate <strong>of</strong> 500 cm 3 /min at<br />

a strain rate <strong>of</strong> 0.1/s. Rheological parameters (peak stress, failure<br />

strain and strain hardening index) were used to compare between<br />

different treatments. Strain hardening index was calculated by<br />

fitting an exponential curve to the stress–strain (Hencky) curve,<br />

after transferring data to Micros<strong>of</strong>t Excel.<br />

2.14. Statistical analysis<br />

Results were analyzed using analysis <strong>of</strong> variance (ANOVA).<br />

ANOVA was performed using a general linear model procedure to<br />

determine significant differences and interactions for the various<br />

treatments. Means were compared by using Fishers LSD procedure<br />

(a ¼ 0.05). Statistical analysis was performed using proc GLM in SAS<br />

(version 9.1; SAS Institute Inc., Cary, NC) s<strong>of</strong>tware. Duplicates were<br />

prepared for each treatment and the order <strong>of</strong> treatment was not<br />

significant.<br />

3. Results and discussion<br />

This section is further divided in two sub-sections in accordance<br />

with objectives <strong>of</strong> the study.<br />

3.1. General mechanism <strong>of</strong> gas cell stability<br />

(investigating dual film hypothesis)<br />

3.1.1. Physico-chemical analysis <strong>of</strong> flours<br />

Two wheat flours, Jagger and S<strong>of</strong>t, used for this study were<br />

analyzed for their chemical composition and physical dough mixing<br />

properties. Moisture (wet weight basis), protein (14% moisture<br />

basis) and lipid (14% moisture basis) contents <strong>of</strong> the flours are given<br />

in Table 1. Jagger (10.4%) was nearly 1% higher in protein content<br />

than S<strong>of</strong>t (9.2%). However, S<strong>of</strong>t gave higher mixing time (9.25 min)<br />

compared to Jagger (7.21 min). Higher percentage <strong>of</strong> unextractable<br />

polymeric protein (UPP) (Table 1) in S<strong>of</strong>t explains the higher mixing<br />

requirements for this flour. This is not typical <strong>of</strong> s<strong>of</strong>t wheat flours,<br />

which normally are considered weak. However, some s<strong>of</strong>t varieties<br />

like Caldwell are relatively stronger than others. S<strong>of</strong>t wheat flour<br />

was milled from a composite <strong>of</strong> s<strong>of</strong>t wheat varieties grown in<br />

Missouri during 2004–2005 and obtained from grain elevators at<br />

Kansas City, MO. Increase in mixing requirements in terms <strong>of</strong><br />

Mixograph dough development time (MDDT) with increase in<br />

percent UPP has been reported previously by Gupta et al. (1993).<br />

Data on polymeric proteins was obtained by SE-HPLC analysis <strong>of</strong><br />

the flours. Lipid content <strong>of</strong> Jagger and S<strong>of</strong>t was 0.89% and 0.93%,<br />

respectively.<br />

3.1.2. Effect <strong>of</strong> natural flour lipid level variation on baking<br />

performance<br />

3.1.2.1. Bread making. Significant differences (P < 0.0001) in loaf<br />

volumes <strong>of</strong> the two flours were observed in response to varying<br />

levels <strong>of</strong> flour lipids (Fig. 2). Incremental addition <strong>of</strong> flour lipids<br />

back into defatted parent flour caused bread volume to decrease,<br />

and after reaching a minimum, to increase. At all levels <strong>of</strong> lipid<br />

addition, relatively lower volumes were observed for S<strong>of</strong>t<br />

compared to Jagger. For S<strong>of</strong>t, minimum volume is reached at<br />

a relatively lower percentage <strong>of</strong> natural flour lipids (w30%) in<br />

comparison to Jagger (w50%) (Fig. 2).<br />

The results <strong>of</strong> loaf volume variations in response to natural flour<br />

lipid level variations agreed qualitatively with those <strong>of</strong> MacRitchie<br />

and co-workers (MacRitchie, 1976a; MacRitchie and Gras, 1973;<br />

McCormack et al., 1991). Loaf volume seems to be governed by<br />

expansion capacity <strong>of</strong> the gas cells. The expansion capacity <strong>of</strong> the<br />

gas cells can be defined as the extent to which gas cells can expand


36<br />

Fig. 2. Loaf volume vs. intact natural flour lipids, added to defatted Jagger (- - -) and<br />

S<strong>of</strong>t (d) wheat flours as percentage <strong>of</strong> natural flour lipids.<br />

without failure; i.e. the value reached after which the loaf does not<br />

increase in volume. This maximum for doughs giving lower loaf<br />

volumes is achieved during the late proving stage, and for those<br />

giving higher loaf volumes is achieved during the early baking<br />

stage. This was determined by measuring oven spring (data not<br />

shown). Oven spring was observed only in doughs giving higher<br />

loaf volume.<br />

Scanning electron micrographs by Gan et al. (1990) appeared to<br />

show the presence <strong>of</strong> intact gas cells with discontinuities in gluten–<br />

starch matrix at advanced stages <strong>of</strong> proving. This means that the<br />

expansion capacity <strong>of</strong> the gas cells is not just controlled by the<br />

gluten–starch matrix and there is possibly a secondary factor<br />

contributing to it. This secondary factor as hypothesized by Gan<br />

et al. (1990) is a liquid lamella present on the inner side <strong>of</strong> the gas<br />

cell. When liquid lamellae also fail, the presence <strong>of</strong> discontinuities<br />

in the gluten–starch matrix leads to coalescence <strong>of</strong> gas cells, thus<br />

decreasing volume. It is quite possible that lipids, due to their<br />

surface action, will be affecting stability <strong>of</strong> liquid lamellae, thus<br />

causing variation in loaf volume. However, this needs to be investigated<br />

further in order to make sure that the action <strong>of</strong> flour lipids is<br />

independent <strong>of</strong> the rheological properties <strong>of</strong> the gluten–starch<br />

matrix.<br />

Relatively lower loaf volumes at all levels <strong>of</strong> lipid addition in<br />

S<strong>of</strong>t, could be attributed to inherent differences in gluten quality <strong>of</strong><br />

the two flours, as previously suggested by MacRitchie (1978). The<br />

differences in the two flours are clear from SEC-MALLS data (Table<br />

5). A new desirable relative proportion <strong>of</strong> polymeric proteins<br />

greater than 250,000 Da in Jagger is probably responsible for better<br />

loaf volumes as explained in Section 3.2.2.3.<br />

3.1.2.2. Crumb structure. Image analysis <strong>of</strong> bread crumb (Table 2)<br />

showed that addition <strong>of</strong> different levels <strong>of</strong> flour lipids resulted in<br />

insignificant variations in number <strong>of</strong> gas cells (P ¼ 0.02) and<br />

average cell elongation (P ¼ 0.94 for Jagger and P ¼ 0.12 for S<strong>of</strong>t).<br />

These non-significant differences in number <strong>of</strong> gas cells further<br />

make it evident that differences in loaf volumes are due to differences<br />

in expansion capacity <strong>of</strong> the gas cells (Table 2), and factors<br />

like gas cell concentration are not involved. Negligible variations in<br />

average cell elongation indicated that lipids might not be causing<br />

any variations in rheology <strong>of</strong> the gluten–starch matrix, since cell<br />

elongation is thought to be associated with dough rheology (Gandikota<br />

and MacRitchie, 2005).<br />

3.1.3. Effect <strong>of</strong> natural flour lipid level variation on biaxial<br />

extensional rheology <strong>of</strong> gluten–starch matrix<br />

In order to investigate possible independent action <strong>of</strong> lipids in<br />

causing loaf volume variations, to indicate presence <strong>of</strong> liquid<br />

B.S. Sroan et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 32–40<br />

Table 2<br />

Crumb structure responses <strong>of</strong> Jagger and S<strong>of</strong>t wheat flour breads to natural flour<br />

lipid levels<br />

%<strong>of</strong><br />

natural<br />

flour<br />

lipids<br />

Jagger wheat flour S<strong>of</strong>t wheat flour<br />

Number <strong>of</strong> cells Average cell<br />

elongation<br />

(C-cell score)<br />

Number <strong>of</strong> cells Average cell<br />

elongation<br />

(C-cell score)<br />

0 2000.0 11.3 a 1.58 0.01 b 2544.5 187.4 a,b 1.69 0.01 a,b<br />

20 1935.0 222.03 a,b 1.59 0.05 b 2521.0 110.3 a,b,c 1.68 0.01 a,b,c<br />

40 1880.5 78.5 a,b 1.64 0.06 a,b 2210.5 115.3 d 1.65 0.03 b,c<br />

50 1898.0 59.4 a,b 1.68 0.04 a 2209.0 79.2 d 1.65 0.02 c<br />

60 1788.5 81.3 a,b 1.67 0.05 a,b 2347.5 2.1 b,c,d 1.70 0.02 a<br />

80 1907.5 2.1 a,b 1.62 0.01 a,b 2196.0 86.3 d 1.66 0.00 a,b,c<br />

100 1749.0 41.0 b 1.63 0.01 a,b 2229.0 114.6 d 1.69 0.03 a,b<br />

Values represent mean standard deviation for duplicate determinations.<br />

Means with the same letter within columns are not significantly different (p > 0.05).<br />

lamellae, biaxial extensional rheology tests were conducted on<br />

defatted Jagger and S<strong>of</strong>t doughs, and at flour lipid levels where<br />

maximum and minimum loaf volumes were observed.<br />

Biaxial extensional rheological parameters (maximum stress,<br />

failure strain and strain hardening index) were higher for Jagger<br />

doughs (Table 3) in comparison to S<strong>of</strong>t doughs. Different lipid<br />

treatments did not influence these parameters within doughs<br />

prepared from each flour type (Table 3). Though minor differences<br />

were observed for strain hardening indices and failure strain <strong>of</strong> the<br />

doughs at some lipid levels, no specific trend was observed, thus<br />

attributing the variation to slight experimental scatter.<br />

Higher values <strong>of</strong> biaxial extensional rheological parameters for<br />

Jagger verify that, rheologically, Jagger is a better bread making<br />

flour. Higher strain hardening index and failure strains ensure that<br />

gas cells in Jagger doughs expand to a greater extent than S<strong>of</strong>t<br />

doughs. Similar results that differentiate between bread making<br />

potential <strong>of</strong> flours have been reported by Dobraszczyk and coworkers<br />

(Dobraszczyk and Roberts, 1994; Dobraszczyk and<br />

Salmanowicz, 2008; Dobraszczyk et al., 2003;). On the other hand,<br />

variations in natural lipid levels in a particular flour that produce<br />

significant variations in loaf volume (Fig. 2) made no difference to<br />

biaxial extensional rheology <strong>of</strong> the dough from that flour. This<br />

provides clear evidence <strong>of</strong> the role <strong>of</strong> lipids as surface active<br />

compounds stabilizing liquid lamellae, a phenomenon independent<br />

<strong>of</strong> dough rheology.<br />

This study, for the first time provides clear evidence on the<br />

presence <strong>of</strong> liquid lamellae as an independent stabilizing mechanism<br />

working auxiliary to the gluten–starch matrix in stabilizing<br />

expanding gas cells during bread making. The study also demonstrates<br />

that lipids at their natural levels do not affect biaxial<br />

extensional rheology as determined by dough inflation.<br />

Table 3<br />

Mean bubble inflation rheological response <strong>of</strong> Jagger and S<strong>of</strong>t wheat flour doughs to<br />

natural flour lipid levels<br />

% <strong>of</strong> natural flour<br />

lipid<br />

Max. stress (Kpa) Failure strain<br />

(Hencky)<br />

Strain hardening<br />

index<br />

Jagger wheat flour<br />

0 571.62 305.45 a 2.58 0.18 a 2.21 0.12 b,c<br />

60 491.68 237.21 a 2.60 0.12 a 2.38 0.064 a,b<br />

100 598.62 112.05 a 2.66 0.04 a 2.44 0.06 a<br />

S<strong>of</strong>t wheat flour<br />

0 120.80 12.76 a 1.79 0.00 a 1.76 0.01 a<br />

40 125.42 12.14 a 1.85 0.02 a 1.74 0.01 a<br />

100 102.40 5.45 a 1.66 0.05 b 1.64 0.04 c<br />

Values represent mean standard deviation for duplicate determinations.<br />

Means with the same letter within columns are not significantly different (p > 0.05).


3.2. Mechanism <strong>of</strong> stability <strong>of</strong> the gluten–starch matrix<br />

3.2.1. Gluten fractionation<br />

Yield (%) and protein content (%) <strong>of</strong> each fraction from the pH<br />

fractionation are given in Table 4. Cumulative percentage <strong>of</strong> total<br />

polymeric protein extracted is plotted as a function <strong>of</strong> pH <strong>of</strong> solution<br />

in Fig. 3. This was calculated using yield data and percentage<br />

total polymeric protein <strong>of</strong> each fraction as given in Table 4.<br />

Approximately 35% <strong>of</strong> total polymeric proteins and most <strong>of</strong> the<br />

gliadins were solubilized and extracted at pH 5.3. The relative<br />

proportion <strong>of</strong> total polymeric proteins and unextractable polymeric<br />

proteins increased in subsequent fractions (Table 4), with the<br />

highest values in the pH 3.1 residue. SE-HPLC analyses <strong>of</strong> individual<br />

fractions (Fig. 4) also show the subsequent shift in MWD <strong>of</strong> the<br />

fractions to higher MW. This is clear in Table 5, where addition <strong>of</strong><br />

these fractions to the base flours causes similar variations in<br />

percentage <strong>of</strong> polymeric protein greater than 250,000 Da. The<br />

solubility <strong>of</strong> gluten proteins depends on their molecular weight<br />

(Singh and MacRitchie, 2001). As the pH is lowered, relatively larger<br />

sized gluten proteins become soluble, and can be separated into<br />

different fractions by stepwise reduction <strong>of</strong> pH (Gupta et al., 1990;<br />

MacRitchie, 1985).<br />

3.2.2. Reconstitution studies<br />

Extracted fractions were added to defatted base flours (Jagger<br />

and S<strong>of</strong>t) at a level <strong>of</strong> 1% (dry protein) on a total flour weight basis.<br />

MacRitchie (1987) found significant rheological differences at these<br />

levels <strong>of</strong> addition <strong>of</strong> protein fractions. The reconstituted flours were<br />

analyzed for their MWD, dough mixing properties, baking performance<br />

and biaxial extensional rheology.<br />

3.2.2.1. SE-HPLC and MALLS (SEC-MALLS) analysis. Cut-<strong>of</strong>f analysis,<br />

as described elsewhere (Bangur et al., 1997), <strong>of</strong> SE-HPLC chromatograms<br />

overlaid with MALLS signal, <strong>of</strong> all Jagger and S<strong>of</strong>t wheat<br />

flours (base and reconstituted) at 4 min intervals, helped to determine<br />

the elution time at which the polymers with MW <strong>of</strong><br />

250,000 Da were being eluted. The chromatograms were integrated<br />

to determine the monomeric (mainly gliadins) to polymeric<br />

(mainly glutenins) ratio, and relative proportion <strong>of</strong> polymeric<br />

proteins greater than 250,000 Da (Table 5). Addition <strong>of</strong> the gliadin<br />

rich (pH 5.3) fraction caused a decrease in the relative proportion <strong>of</strong><br />

polymeric proteins greater than 250,000 Da. However, with addition<br />

<strong>of</strong> later fractions (pH4.9 to pH 3.1 residue), progressive increase<br />

in the proportion <strong>of</strong> polymeric proteins greater than 250,000 Da<br />

was observed. Use <strong>of</strong> the relative proportion <strong>of</strong> polymeric proteins<br />

greater than MW <strong>of</strong> 250,000 is based on a previous estimate by<br />

Bangur et al. (1997), which also seems to hold true in this study and<br />

is explained in Section 3.2.2.3.<br />

Table 4<br />

Yield, moisture and protein content <strong>of</strong> gluten protein fractions extracted by pH<br />

fractionation, and their SE-HPLC relative composition <strong>of</strong> polymeric proteins (all<br />

numbers in percentages)<br />

Fractions Yield<br />

(%)<br />

Moisture<br />

content<br />

(% wet basis)<br />

Protein<br />

content<br />

(%) (14% m.b.)<br />

Area (%) under<br />

chromatogram curve<br />

TPP UPP<br />

pH 5.3 50.4 4.5 0.08 83.8 0.05 31.6 1.59 f 23.9 0.13 d<br />

pH 4.9 20.5 4.6 0.16 78.0 0.08 52.6 0.08 e 72.6 1.48 c<br />

pH 4.1 10.6 4.1 0.15 79.6 0.12 64.1 0.10 d 78.3 2.85 b<br />

pH 3.5 4.2 4.2 0.37 78.5 0.07 65.5 0.01 c 77.6 2.31 b<br />

pH 3.1<br />

supernatant<br />

3.5 4.3 0.09 73.2 0.14 68.0 0.21 b 83.2 0.93 a<br />

pH 3.1 residue 10.8 4.9 0.26 45.8 0.18 69.5 0.49 a 86.6 0.08 a<br />

TPPdtotal polymeric protein; UPPdunextractable polymeric protein<br />

Values represent mean standard deviation for duplicate determinations.<br />

Means with the same letter within columns are not significantly different<br />

(p > 0.05).<br />

B.S. Sroan et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 32–40 37<br />

Fig. 3. Cumulative percentage <strong>of</strong> total polymeric protein extracted as a function <strong>of</strong> final<br />

pH <strong>of</strong> supernatant <strong>of</strong> Jagger gluten.<br />

Elution <strong>of</strong> proteins through a SE-HPLC column is based on their<br />

hydrodynamic radii. Proteins with similar MW may vary in the<br />

conformation and thus the hydrodynamic radii in solution. Such<br />

proteins will not elute at the same time through the SE-HPLC<br />

column. An overlaid MALLS signal provides a true measure <strong>of</strong> MW at<br />

different elution times (Bean and Lookhart, 2001), providing nearly<br />

precise estimation <strong>of</strong> the relative proportion <strong>of</strong> polymeric proteins<br />

greater than 250,000 Da.<br />

3.2.2.2. Dough mixing properties. To observe variations in mixing<br />

properties <strong>of</strong> the flours on addition <strong>of</strong> fractions, Mixograph analysis<br />

was performed with constant water absorption; i.e. 63% for Jagger<br />

wheat flour and 60% for S<strong>of</strong>t wheat flour. Mixograph dough<br />

development time (MDDT) was recorded from Mixograph traces <strong>of</strong><br />

Jagger and S<strong>of</strong>t wheat flours. MDDT is plotted for each fraction as<br />

a function <strong>of</strong> the total protein that has been extracted (Fig. 5).<br />

Addition <strong>of</strong> the first 35% (approximately) <strong>of</strong> total polymeric protein<br />

extracted from Jagger gluten (corresponding to the pH 5.3 fraction)<br />

caused a decrease in mixing requirements. In contrast, addition <strong>of</strong><br />

subsequent fractions caused an increase in mixing requirements.<br />

These trends for MDDTs can be explained on the basis <strong>of</strong> SEC-<br />

MALLS data (Table 5). As expected, the decrease in the monomeric<br />

to polymeric ratio and increase in relative proportion <strong>of</strong> polymeric<br />

proteins greater than 250,000 Da lead to an increase in MDDT and<br />

corresponding decrease in breakdown. At a molecular level, mixing<br />

Fig. 4. SE-HPLC chromatograms <strong>of</strong> total protein <strong>of</strong> Jagger gluten protein fraction. As pH<br />

is reduced, the percentage <strong>of</strong> polymeric protein increases and MWD shifts towards<br />

higher molecular weight.


38<br />

Table 5<br />

Parameters calculated from SEC-MALLS chromatograms <strong>of</strong> Jagger wheat flour with<br />

addition <strong>of</strong> gluten protein fractions, added at 1% (dry protein) level (on flour weight<br />

basis)<br />

Fraction added Jagger wheat flour S<strong>of</strong>t wheat flour<br />

MON/POL a<br />

%<br />

POL > 250,000 b<br />

MON/POL a<br />

%<br />

POL > 250,000 b<br />

None (Control) 1.47 0.00 a 65.9 0.01 d 1.41 0.01 a 76.8 0.07 d<br />

pH 5.3 1.48 0.01 a 35.0 0.05 f 1.43 0.03 a 46.7 0.10 f<br />

pH 4.9 1.36 0.01 b,c 66.9 0.18 c 1.27 0.07 b 68.1 0.22 e<br />

pH 4.1 1.34 0.00 c,d 58.0 0.04 e 1.28 0.01 b 99.7 0.14 b,c<br />

pH 3.5 1.37 0.00 b 67.1 0.02 c 1.30 0.01 b 100.0 0.00 a<br />

pH 3.1<br />

supernatant<br />

1.33 0.01 d 85.1 0.12 b 1.16 0.00 c 96.1 0.04 c<br />

pH 3.1 residue 1.32 0.00 d 100 0.00 a 1.16 0.02 c 100.0 0.00 a<br />

Values represent mean standard deviation for duplicate determinations.<br />

Means with the same letter within columns are not significantly different (p > 0.05).<br />

a<br />

Monomeric to polymeric protein ratio.<br />

b<br />

Percentage <strong>of</strong> polymeric proteins 250,000 Da.<br />

is characterized by extension (stretching) <strong>of</strong> glutenin polymers and<br />

entangling <strong>of</strong> these stretched polymers (MacRitchie, 1986). This is<br />

achieved by work input and mixing intensity that must be above<br />

certain minimum critical levels (Kilborn and Tipples, 1972; Tipples<br />

and Kilborn, 1975). At a given work input and intensity <strong>of</strong> mixing,<br />

MDDT probably relates to the extension and entangling <strong>of</strong> the<br />

largest glutenin molecules (Singh and MacRitchie, 2001). The work<br />

input and mixing intensity <strong>of</strong> the flour is therefore a function <strong>of</strong> its<br />

MWD. It has been observed that when the percentage <strong>of</strong> UPP<br />

increases and/or the MWD <strong>of</strong> the polymeric fraction is shifted to<br />

higher MWs, the MDDT increases, as requirements for work input<br />

increase (Gupta et al., 1993).<br />

3.2.2.3. Baking performance. Loaf volumes (Fig. 6) decreased on<br />

addition <strong>of</strong> the pH 5.3 fraction. However, on addition <strong>of</strong> subsequent<br />

fractions that lead to increase in MDDT, increase in loaf volume was<br />

observed, which after reaching a maximum, again decreased. This<br />

maximum was achieved in the case <strong>of</strong> Jagger on addition <strong>of</strong> total<br />

polymeric proteins that were extracted after the first 60% and<br />

before the first 80% corresponding to the pH 4.1 fraction. In the case<br />

<strong>of</strong> S<strong>of</strong>t, the maximum occurred on addition <strong>of</strong> total polymeric<br />

proteins that were extracted after the first 35% and before the first<br />

70%, corresponding to the pH 4.9 fraction (Figs. 3 and 6). It seems<br />

that the optimum balance <strong>of</strong> strength and extensibility was achieved<br />

on addition <strong>of</strong> these fractions to base flours, thus giving best<br />

baking performance.<br />

Bangur et al. (1997) reported that the ratio <strong>of</strong> polymeric proteins<br />

with MWs above the threshold molecular weight (MT) to those (i.e.<br />

Fig. 5. Effect on Mixograph peak times <strong>of</strong> Jagger (- - -) and S<strong>of</strong>t (d) wheat flours on<br />

addition <strong>of</strong> gluten protein fractions at a 1% (dry protein) level (on flour weight basis).<br />

B.S. Sroan et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 32–40<br />

Fig. 6. Effect on loaf volumes (d) and strain hardening index (- - -) <strong>of</strong> Jagger (:) and<br />

S<strong>of</strong>t (-) wheat flours on addition <strong>of</strong> gluten protein fractions at a 1% (dry protein) level<br />

(on flour weight basis).<br />

polymeric proteins) with MWs below MT was approximately 60:40.<br />

The MT for polymeric proteins was estimated to be 250,000 Da.<br />

Polymeric proteins greater in molecular weight than 250,000 Da<br />

confer strength to the entangled gluten protein network and those<br />

with molecular weight less than this may counter the strength by<br />

acting as diluents (or plasticizers). This 60:40 proportion is nearly<br />

achieved on addition <strong>of</strong> the pH 4.1 fraction in Jagger wheat flour<br />

and pH 4.9 fraction in S<strong>of</strong>t wheat flour (Table 5) giving maximum<br />

loaf volumes and strain hardening index. Any shift in the proportion<br />

shows deleterious effects on baking performance (Fig. 6).<br />

3.2.2.4. Biaxial extensional rheology. Significant differences<br />

(P < 0.0001) in biaxial extensional rheological parameters (strain<br />

hardening index and failure strain) were recorded for addition <strong>of</strong><br />

fractions to the base flours. Variations in strain hardening index<br />

paralleled those <strong>of</strong> loaf volume, for both flours (Fig. 6). Analogous<br />

variations in strain hardening index and loaf volume show that this<br />

parameter is a good determinant <strong>of</strong> rheological requirements for<br />

the gluten–starch matrix in terms <strong>of</strong> bread making. As reported in<br />

some previous studies (Dobraszczyk and Roberts, 1994; Dobraszczyk<br />

et al., 2003; Tronsmo et al., 2003), in this case also, strain<br />

hardening shows a high degree <strong>of</strong> correlation with loaf volume and<br />

failure strain (Fig. 7). The Mixograph parameter (MDDT), on the<br />

other hand, correlates well with MWD but does not predict baking<br />

performance equally well.<br />

The concept <strong>of</strong> strain hardening explains the requirements for<br />

maximum inflatability <strong>of</strong> gas cells in bread making without failure,<br />

to give good loaf volume and crumb grain. In a gluten–starch<br />

matrix, gluten proteins form a continuous network and the<br />

rheology <strong>of</strong> the dough is that <strong>of</strong> a continuous gluten protein<br />

network. It is only the fraction <strong>of</strong> glutenins with MW greater than<br />

MT that confers strength, as suggested by Bangur et al. (1997),<br />

whereas smaller ones act as diluents, preventing additional physical<br />

constraints or entanglements and enhancing extensibility<br />

(Bersted and Anderson, 1990; Singh and MacRitchie, 2001). To get<br />

maximum inflatability <strong>of</strong> the gas cells, the gluten–starch matrix<br />

around them must stretch to its maximum extensibility without<br />

breaking, i.e. the glutenins/polymeric proteins (mainly with<br />

MW > MT; i.e. 250,000 Da) must be stretched to their maximum<br />

length through entanglements, as described by entanglement<br />

network theory (MacRitchie and Lafiandra, 1997; Singh and<br />

MacRitchie, 2001). Addition <strong>of</strong> the gliadin rich fraction, i.e. the pH<br />

5.3 fraction (Fig. 4), significantly increases diluent concentration<br />

that will lead to disentanglement <strong>of</strong> the gluten protein network<br />

without sufficient elongation when subjected to extensional forces,<br />

thus leading to earlier failure <strong>of</strong> the gas cells. As MWD shifts


towards larger glutenins, the strength increases (Gupta et al., 1993).<br />

On doing so, it passes through maximum extensibility or strain<br />

(Hencky), beyond which the balance shifts towards strength and<br />

the gluten protein network is no longer as extensible. This leads to<br />

decrease in strain hardening index as well as loaf volume upon<br />

addition <strong>of</strong> the latter extracted fractions (Fig. 6).<br />

4. Conclusion<br />

Results <strong>of</strong> this study provide clear evidence for the presence <strong>of</strong><br />

liquid lamellae, thus supporting the dual film hypothesis. The liquid<br />

lamellae act as a secondary stabilizing mechanism and is on the<br />

inner side <strong>of</strong> the gluten–starch matrix enveloping the gas cells (Gan<br />

et al., 1995).<br />

Stability <strong>of</strong> the gluten–starch matrix, which is the primary<br />

stabilizing factor for expanding gas cells against disproportionation<br />

and coalescence, depends on its tendency to strain harden. The<br />

phenomenon <strong>of</strong> strain hardening appears to depend on the balance<br />

between strength and extensibility <strong>of</strong> the entangled network <strong>of</strong><br />

polymeric proteins <strong>of</strong> wheat flour. Extensibility ensures slippage <strong>of</strong><br />

the maximum number <strong>of</strong> statistical segments between entanglements<br />

(Singh and MacRitchie, 2001), whereas strength prevents<br />

disruption <strong>of</strong> the entangled network <strong>of</strong> polymeric proteins. Thus, to<br />

ensure stability <strong>of</strong> gas cells, the dough needs to be sufficiently<br />

extensible to respond to gas pressure but also strong enough to<br />

resist collapse. Differences in gluten quality, as demonstrated in<br />

this study can significantly affect the bread making potential.<br />

Strength is conferred by the fraction <strong>of</strong> polymeric proteins having<br />

molecular weight greater or equivalent to MT (250,000), and the<br />

fraction <strong>of</strong> gluten protein smaller than MT may counter the strength<br />

by acting as diluents. The optimum balance seems to exist when the<br />

relative proportions <strong>of</strong> polymeric proteins greater and smaller than<br />

MT are roughly 60:40. Shift in the balance to either side will<br />

decrease loaf volume. Increase in smaller proteins (less than MT)<br />

may decrease stability <strong>of</strong> the gluten–starch matrix due to a lesser<br />

number <strong>of</strong> entanglements per chain. On the other hand, increase in<br />

strength conferring proteins may prevent sufficient expansion <strong>of</strong><br />

the gluten–starch matrix required to increase loaf volume due to<br />

reduced slippage <strong>of</strong> gluten polymers through entanglement nodes<br />

as a result <strong>of</strong> increase in number <strong>of</strong> entanglements per chain.<br />

B.S. Sroan et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 32–40 39<br />

Fig. 7. Scatter plots <strong>of</strong> strain hardening index vs. failure strain and loaf volume (cc) for different Jagger and S<strong>of</strong>t wheat flour doughs exhibiting high degree <strong>of</strong> correlation.<br />

The secondary stabilizing mechanism involves thin liquid<br />

lamellae stabilized by adsorbed surface active compounds (lipids<br />

and proteins) at the gas–liquid interface. Liquid lamellae prevent<br />

coalescence and disproportionation <strong>of</strong> gas cells when they come in<br />

close contact with each other during the late proving and early<br />

baking stages <strong>of</strong> bread making i.e. when discontinuities begin to<br />

appear in the gluten–starch matrix. The study demonstrates that<br />

the flour lipids at their natural levels do not influence rheological<br />

properties <strong>of</strong> the gluten–starch matrix surrounding the gas cells, as<br />

measured by the dough inflation system. Nevertheless, the small<br />

amounts in which these lipids are naturally present are sufficient to<br />

influence surface properties. The exact mechanism by which these<br />

lipids and other surface active components (proteins) stabilize<br />

liquid lamellae is discussed in the succeeding paper (Part II).<br />

References<br />

AACC, 2001. Approved Methods <strong>of</strong> the AACC, tenth ed. American Association <strong>of</strong><br />

<strong>Cereal</strong> Chemists, St. Paul, Minnesota, USA.<br />

Baker, J.C., Mize, M.D., 1946. Gas occlusion during dough mixing. <strong>Cereal</strong> Chemistry<br />

23, 39–51.<br />

Bangur, R., Batey, I.L., McKenzie, E., MacRitchie, F., 1997. Dependence <strong>of</strong> extensograph<br />

parameters on wheat protein composition measured by SE-HPLC. <strong>Journal</strong><br />

<strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 25, 237–241.<br />

Batey, I.L., Gupta, R.B., MacRitchie, F., 1991. Use <strong>of</strong> size-exclusion high-performance<br />

liquid chromatography in the study <strong>of</strong> wheat flour proteins: an improved<br />

chromatographic procedure. <strong>Cereal</strong> Chemistry 68, 207–209.<br />

Bean, S.R., Lookhart, G.L., 2001. Factors influencing the characterization <strong>of</strong> gluten<br />

proteins by size-exclusion chromatography and multiangle laser light scattering<br />

(SEC-MALLS). <strong>Cereal</strong> Chemistry 78, 608–618.<br />

Bersted, B.H., Anderson, T.G., 1990. Influence <strong>of</strong> molecular weight and molecular<br />

weight distribution on the tensile properties <strong>of</strong> amorphous polymers. <strong>Journal</strong> <strong>of</strong><br />

Applied Polymer <strong>Science</strong> 39, 499–514.<br />

Bloksma, A.H., 1990. Dough structure, rheology, and baking quality. <strong>Cereal</strong> Foods<br />

World 35, 237–244.<br />

Chin, N.L., Campbell, G.M., 2005. Dough aeration and rheology. II. Effects <strong>of</strong> flour<br />

type, mixing speed and total work input on aeration and rheology <strong>of</strong> bread<br />

dough. <strong>Journal</strong> <strong>of</strong> the <strong>Science</strong> <strong>of</strong> Food and Agriculture 85, 2194–2202.<br />

DeStefanis, V.A., Ponte Jr., J.G., 1976. Studies on breadmaking properties <strong>of</strong> wheatflour<br />

nonpolar lipids. <strong>Cereal</strong> Chemistry 53, 636–642.<br />

Dobraszczyk, B.J., 1997. Development <strong>of</strong> a new dough inflation system to evaluate<br />

doughs. <strong>Cereal</strong> Foods World 42, 516–519.<br />

Dobraszczyk, B.J., Roberts, C.A., 1994. Strain hardening and dough gas cell-wall<br />

failure in biaxial extension. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 20, 265–274.<br />

Dobraszczyk, B.J., Salmanowicz, B.P., 2008. Comparison and predictions <strong>of</strong> baking<br />

volume using large deformation rheological properties. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong><br />

<strong>Science</strong> 47, 292–301.


40<br />

Dobraszczyk, B.J., Smewing, J., Albertini, M., Maesmans, G., Sch<strong>of</strong>ield, J.D., 2003.<br />

Extensional rheology and stability <strong>of</strong> gas cell walls in bread doughs at elevated<br />

temperatures in relation to breadmaking performance. <strong>Cereal</strong> Chemistry 80,<br />

218–224.<br />

Gan, Z., Angold, R.E., Williams, M.R., Ellis, P.R., Vaughan, J.G., Galliard, T., 1990. The<br />

microstructure <strong>of</strong> gas retention <strong>of</strong> bread dough. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 12,<br />

15–24.<br />

Gan, Z., Ellis, P.R., Sch<strong>of</strong>ield, J.D., 1995. Gas cell stabilization and gas retention in<br />

wheat bread dough. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 21, 215–230.<br />

Gandikota, S., MacRitchie, F., 2005. Expansion capacity <strong>of</strong> doughs: methodology and<br />

applications. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 42, 157–163.<br />

Gupta, R.B., Bekes, F., Gras, P.W., MacRitchie, F., 1990. Functionality <strong>of</strong> glutenin,<br />

gliadin and secalin fractions as measured by extensograph and mixograph. In:<br />

Gluten Proteins. Association <strong>of</strong> <strong>Cereal</strong> Research, Detmold, pp 550–559.<br />

Gupta, R.B., Khan, K., MacRitchie, F., 1993. Biochemical basis <strong>of</strong> flour properties in<br />

bread wheats. I. Effects <strong>of</strong> variation in the quality and size distribution <strong>of</strong><br />

polymeric protein. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 18, 23–41.<br />

Hoseney, R.C., 1992. Physical chemistry <strong>of</strong> bread dough. In: Schwartzberg, H.G.,<br />

Hartel, R.W. (Eds.), Physical Chemistry <strong>of</strong> Foods. Marcel Dekker, Inc., New York,<br />

USA, pp. 443–457.<br />

Kilborn, R.H., Tipples, K.H., 1972. Factors affecting mechanical dough development.<br />

I. Effect <strong>of</strong> mixing intensity and work input. <strong>Cereal</strong> Chemistry 49, 34–47.<br />

Lundh, G., MacRitchie, F., 1989. Size exclusion HPLC characterization <strong>of</strong> gluten<br />

protein fractions varying in breadmaking potential. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 10,<br />

247–253.<br />

MacRitchie, F., 1976a. The liquid phase <strong>of</strong> dough and its role in baking. <strong>Cereal</strong><br />

Chemistry 53, 318–326.<br />

MacRitchie, F., 1976b. Monolayer compression barrier in emulsion and foam<br />

stability. <strong>Journal</strong> <strong>of</strong> Colloid and Interface <strong>Science</strong> 56, 53–56.<br />

MacRitchie, F., 1978. Differences in baking quality between wheat flours. <strong>Journal</strong> <strong>of</strong><br />

Food Technology 13, 187–194.<br />

MacRitchie, F., 1985. Studies <strong>of</strong> the methodology for fractionation and reconstitution<br />

<strong>of</strong> wheat flours. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 3, 221–230.<br />

MacRitchie, F., 1986. Physicochemical processes in mixing. In: Blanshard, J.M.V.,<br />

Frazier, P.J., Galliard, T. (Eds.), Chemistry and Physic <strong>of</strong> Baking. Royal Society <strong>of</strong><br />

Chemistry, London, UK, pp. 132–146.<br />

B.S. Sroan et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 32–40<br />

MacRitchie, F., 1987. Evaluation <strong>of</strong> contributions from wheat protein fractions to<br />

dough mixing and breadmaking. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 6, 259–268.<br />

MacRitchie, F., Gras, P.W., 1973. The role <strong>of</strong> flour lipids in baking. <strong>Cereal</strong> Chemistry<br />

50, 292–302.<br />

MacRitchie, F., Kasarda, D.D., Kuzmicky, D.D., 1991. Characterization <strong>of</strong> wheat<br />

protein fractions differing in contributions to breadmaking quality. <strong>Cereal</strong><br />

Chemistry 68, 122–130.<br />

MacRitchie, F., Lafiandra, D., 1997. Structure–functionality relationships <strong>of</strong> wheat<br />

proteins. In: Damodaran, S., Paraf, A. (Eds.), Food Proteins and Their Applications.<br />

Marcel Dekker, Inc., New York, USA, pp. 293–324.<br />

McCormack, G., Panozzo, J., MacRitchie, F., 1991. Contributions to breadmaking <strong>of</strong><br />

inherent variations in lipid content and composition <strong>of</strong> wheat cultivars. II.<br />

Fractionation and reconstitution studies. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 13, 263–274.<br />

Mills, E.N.C., Wilde, P.J., Salt, L.J., Skeggs, P., 2003. Bubble formation and stabilization<br />

in bread dough. Food and Bioproducts Processing 81, 189–193.<br />

Paternotte, T.A., Orsel, R., Hamer, R.J., 1993. Interactions between flour proteins and<br />

flour lipids at the liquid/air interface. In: Gluten Proteins. Association <strong>of</strong> <strong>Cereal</strong><br />

Research, Detmold, Germany, pp. 207–217.<br />

Ross, A., MacRitchie, F., 1995. Interactions <strong>of</strong> wheat proteins, carbohydrates and<br />

lipids. In: Gaonkar, A.G. (Ed.), Ingredient Interactions Effects on Food Quality.<br />

Marcel Deker, New York, USA, pp. 321–356.<br />

Singh, N.K., Donovan, G.R., Batey, I.L., MacRitchie, F., 1990. Use <strong>of</strong> sonication and<br />

size-exclusion high-performance liquid chromatography in the study <strong>of</strong> wheat<br />

flour proteins. I. Dissolution <strong>of</strong> total proteins in the absence <strong>of</strong> reducing agents.<br />

<strong>Cereal</strong> Chemistry 67, 150–161.<br />

Singh, H., MacRitchie, F., 2001. Application <strong>of</strong> polymer science to properties <strong>of</strong><br />

gluten. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 33, 231–243.<br />

Termonia, Y., Smith, P., 1988. Kinetic model for tensile deformation <strong>of</strong> polymers. 2.<br />

Effect <strong>of</strong> entanglement spacing. Macromolecules 21, 2184–2189.<br />

Tronsmo, K.M., Færgestad, E.M., Sch<strong>of</strong>ield, J.D., Magnus, E.M., 2003. Wheat protein<br />

quality in relation to baking performance evaluated by the Chorleywood bread<br />

process and a hearth bread baking test. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 38, 205–215.<br />

Tipples, K.H., Kilborn, R.H., 1975. Unmixingdthe disorientation <strong>of</strong> developed bread<br />

doughs by slow speed mixing. <strong>Cereal</strong> Chemistry 52, 248–262.<br />

van Vliet, T., Janssen, A.M., Bloksma, A.H., Walstra, P.,1992. Strain hardening <strong>of</strong> dough<br />

as a requirement for gas retention. <strong>Journal</strong> <strong>of</strong> Texture Studies 23, 439–460.


Mechanism <strong>of</strong> gas cell stabilization in breadmaking. II. The secondary liquid<br />

lamellae<br />

Baninder S. Sroan *, Finlay MacRitchie<br />

Department <strong>of</strong> Grain <strong>Science</strong> and Industry, Kansas State University, Manhattan, KS 66506, USA<br />

article info<br />

Article history:<br />

Received 19 December 2007<br />

Received in revised form 28 June 2008<br />

Accepted 9 July 2008<br />

Keywords:<br />

Breadmaking<br />

Gas cell stability<br />

Liquid lamella<br />

Gluten–starch matrix<br />

Coalescence<br />

Disproportionation<br />

1. Introduction<br />

abstract<br />

Concentration <strong>of</strong> uniformly sized gas bubbles in the liquid phase<br />

<strong>of</strong> dough during mixing is essential for their sufficient expansion<br />

without failure during the breadmaking process for good loaf<br />

volume and crumb grain. There is evidence that the liquid lamellae<br />

surrounding expanding gas cells act as a secondary protection<br />

together with the primary gluten–starch matrix, thus preventing<br />

their coalescence and disproportionation as outlined in Part-I. The<br />

secondary stabilizing mechanism <strong>of</strong> thin liquid lamellae (films)<br />

operates when gas bubbles come in close contact during later<br />

proving and early baking when discontinuities begin to appear in<br />

the gluten–starch matrix. Stability <strong>of</strong> the liquid lamellae depends<br />

on surface active compounds (proteins and lipids) (Baker et al.,<br />

1946; MacRitchie, 1976a; Mills et al., 2003), that are adsorbed at the<br />

gas–liquid interface. Analogous variations in foaming properties <strong>of</strong><br />

the dough liquor, and bread loaf volume and crumb structure<br />

(MacRitchie, 1976a) provide plausible evidence <strong>of</strong> surface action <strong>of</strong><br />

these compounds at the gas–liquid interface. Contributions <strong>of</strong><br />

liquid film stability to dough expansion can be assessed from<br />

different studies. It has been shown that flour lipids, though<br />

present in very small amounts (w1–1.5%) have significant effects on<br />

* Corresponding author. Present address: FritoLay, 7701 Legacy Dr., 3T-124, Plano,<br />

TX 75024, USA. Tel.: þ1 972 334 4526; fax: þ1 972 334 2329.<br />

E-mail address: baninder@k-state.edu (B.S. Sroan).<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.07.004<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 41–46<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

The study for the first time demonstrates that flour lipids at their natural levels do not affect dough<br />

rheology as measured by bubble inflation, thus indicating the presence <strong>of</strong> liquid lamellae as an independent<br />

secondary gas cell stabilizing mechanism in bread dough. The liquid lamellae, stabilized by<br />

adsorbed surface active compounds, plays its role during the later proving and early baking stage, when<br />

discontinuities occur in the gluten–starch matrix surrounding gas bubbles. To study this secondary<br />

stabilizing mechanism, different lipid fractions were added incrementally to the defatted flours. No<br />

effects were observed on the rheological properties <strong>of</strong> the dough. However, large effects on the loaf<br />

volume were measured. The additives used were the total flour lipid and its polar and non-polar fractions<br />

and the fatty acids palmitic, linoleic and myristic. Polar lipids and palmitic acid had positive or little effect<br />

on loaf volume, respectively. Non-polar lipid, linoleic and myristic acids had negative effects on loaf<br />

volume. The different effects <strong>of</strong> the lipid fractions are thought to be related to the type <strong>of</strong> monolayer that<br />

is formed. Polar lipid and palmitic acid form condensed monolayers at the air/water interface whereas<br />

non-polar lipid, linoleic and myristic acids form expanded monolayers.<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

baking performance in terms <strong>of</strong> volume and crumb structure<br />

(MacRitchie and Gras, 1973; Mills et al., 2003). Polar lipids improve<br />

bread loaf volume and crumb grain, whereas non-polar lipids have<br />

the opposite effects (MacRitchie and Gras, 1973; Ponte and DeStefanis,<br />

1969). DeStefanis and Ponte (1976) demonstrated that an<br />

unsaturated free fatty acid (linoleic acid), a component <strong>of</strong> the nonpolar<br />

lipid fraction is responsible for detrimental effects on loaf<br />

volume. Mills et al. (2003) reviewed some <strong>of</strong> the more recent<br />

studies on effects <strong>of</strong> surface active components <strong>of</strong> wheat flour on<br />

breadmaking performance.<br />

The most probable mechanism by which these surface active<br />

components control stability <strong>of</strong> gas cells is through formation <strong>of</strong><br />

monolayers at the gas–liquid interface (Mills et al., 2003; Ross and<br />

MacRitchie, 1995). Saturated free fatty acids, like palmitic and<br />

stearic, form condensed monolayers, which are highly incompressible<br />

with high surface elasticity and these molecules do not<br />

desorb easily from the monolayers. These condensed monolayers<br />

generate substantial elastic restoring forces, which resist destabilization<br />

<strong>of</strong> liquid lamellae when changes in interfacial area occur<br />

(MacRitchie, 1976b). Molecules like digalactosyl diglyceride<br />

(DGDG) with strong polar head groups are expected to orient<br />

themselves strongly at the interface as condensed monolayers.<br />

Unsaturated free fatty acids like linoleic acid give expanded<br />

monolayers, which are relatively compressible and slightly soluble<br />

in the subphase, leading to instability <strong>of</strong> the liquid lamellae<br />

(MacRitchie, 1976b). Rich in surface active compounds (lipids,


42<br />

proteins, arabinoxylans, added surfactants, etc.), the liquid phase <strong>of</strong><br />

dough is a very competitive environment in terms <strong>of</strong> adsorption <strong>of</strong><br />

these compounds at the gas–liquid interface (Mills et al., 2003; Ross<br />

and MacRitchie, 1995). Proteins (soluble or insoluble) in wheat flour<br />

are capable <strong>of</strong> forming condensed monolayers. Pure protein<br />

monolayers possess high surface elasticity. When lipids are introduced<br />

in the system, surface elasticity drops as lipids compete with<br />

proteins for the gas–liquid interface (Paternotte et al., 1993; Salt<br />

et al., 2006). It seems that the type <strong>of</strong> monolayer formed will either<br />

stabilize or destabilize the freshly occluded gas cells, and any<br />

changes in monolayer state will influence the gas cell stability at all<br />

stages <strong>of</strong> breadmaking.<br />

Therefore, the stability <strong>of</strong> liquid lamellae determines the ease<br />

with which gas cells are concentrated during mixing and ensure<br />

their permanence during the entire breadmaking process.<br />

The objective <strong>of</strong> this part <strong>of</strong> the study was to understand the<br />

possible mechanism by which liquid lamellae are stabilized. This<br />

part <strong>of</strong> the study provides further evidence in support <strong>of</strong> independent<br />

existence <strong>of</strong> liquid lamellae as a secondary stabilizing<br />

mechanism supporting expanding gas cells. It shows for the first<br />

time that lipids, at their natural levels, have no effect on the<br />

rheological properties <strong>of</strong> the starch/protein (gluten–starch) matrix<br />

as measured by bubble inflation and their effects are therefore<br />

independent <strong>of</strong> the stabilizing mechanism discussed in Part-I.<br />

2. Experimental<br />

Materials and various procedures (analytical, dough mixing<br />

properties, test baking, image analysis, biaxial extensional rheology<br />

and statistical analysis) used in the study are explained in detail in<br />

Part-I.<br />

2.1. Free fatty acids<br />

Palmitic acid (99%), linoleic acid (60%), and myristic acid (99%)<br />

were purchased from Sigma-Aldrich, USA. Values in brackets are <strong>of</strong><br />

minimum purity by gas chromatography as per certificate <strong>of</strong> analysis<br />

by manufacturer. Linolenic acid was the main impurity in<br />

linoleic acid.<br />

2.2. Lipid extraction from flour<br />

Natural flour lipids were extracted using three batch extractions<br />

with chlor<strong>of</strong>orm in a glass beaker, followed by Buchner filtration<br />

through Whatman No. 1 filter paper (MacRitchie and Gras, 1973).<br />

200 g <strong>of</strong> flour and 400 ml <strong>of</strong> chlor<strong>of</strong>orm were used for each<br />

extraction. The defatted flour was spread out on a flat glass tray in<br />

a fume hood for 12 h to allow evaporation <strong>of</strong> solvent.<br />

2.3. Lipid fractionation<br />

Natural lipids <strong>of</strong> both flours in chlor<strong>of</strong>orm were collected<br />

together and concentrated under vacuum using a rotary evaporator<br />

at a temperature


Fig. 2. Loaf volume vs. different lipid types, added to defatted Jagger wheat flour as<br />

percentage <strong>of</strong> natural flour lipids.<br />

depression for small additions followed by increase in loaf volume.<br />

This increase in volume was appreciably larger than what was<br />

observed for the same levels <strong>of</strong> natural flour lipids. Addition <strong>of</strong><br />

increasing levels <strong>of</strong> the non-polar lipids and linoleic acid caused<br />

a continuous decrease in loaf volume for both the flours.<br />

Incremental addition <strong>of</strong> palmitic acid produced no change in loaf<br />

volume <strong>of</strong> Jagger flour and slight depression in s<strong>of</strong>t wheat flour.<br />

Incremental addition <strong>of</strong> myristic acid caused depression in loaf<br />

volume at initial levels, however, at considerably higher levels <strong>of</strong><br />

addition, some reversal was observed. This increase, in the case<br />

<strong>of</strong> myristic acid, leveled <strong>of</strong>f at the volumes equivalent to those <strong>of</strong><br />

Fig. 3. Loaf volume vs. different lipid types, added to defatted s<strong>of</strong>t wheat flour as<br />

percentage <strong>of</strong> natural flour lipids.<br />

B.S. Sroan, F. MacRitchie / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 41–46 43<br />

intact flours (100% <strong>of</strong> natural flour lipids). Maximum addition <strong>of</strong><br />

myristic acid in Jagger was 400% <strong>of</strong> natural flour lipid content. This<br />

was a very high level <strong>of</strong> addition in comparison to other lipid types<br />

and free fatty acids used in the study; this was done to have a clear<br />

picture <strong>of</strong> the loaf volume trend in this case.<br />

The results for natural flour lipids and their polar and non-polar<br />

fractions agreed qualitatively with those <strong>of</strong> MacRitchie and coworkers<br />

(McCormack et al., 1991; MacRitchie, 1977; MacRitchie,<br />

1978; MacRitchie and Gras, 1973), and those <strong>of</strong> unsaturated (linoleic)<br />

and saturated (palmitic) free fatty acids concurred with results<br />

<strong>of</strong> DeStefanis and Ponte (1976). Effects <strong>of</strong> different lipid types on<br />

loaf volume and crumb structure correlated with results from film<br />

balance studies reported by various workers (Gaines, 1966;<br />

MacRitchie, 1990). This substantiated our earlier results (Part-I),<br />

providing evidence for the presence <strong>of</strong> liquid lamellae around<br />

expanding gas cells. The results also indicated that the most<br />

probable mechanism by which different lipid types influenced the<br />

gas cell stability was through their adsorption as monolayers at the<br />

gas–liquid interface.<br />

The monolayers formed are either condensed or expanded<br />

(Gaines, 1966; MacRitchie, 1990). Condensed monolayers are<br />

characterized by close packing <strong>of</strong> surface active molecules, with<br />

polar head groups oriented towards water and non-polar hydrocarbon<br />

chains towards air. Expanded monolayers are formed due to<br />

loose packing <strong>of</strong> surface active molecules, occupying a much higher<br />

area per molecule. Condensed monolayers are relatively incompressible<br />

and elastic compared to expanded ones and are less easily<br />

desorbed than expanded monolayers (MacRitchie, 1976b). On<br />

expansion <strong>of</strong> gas cells, condensed monolayers provide elastic<br />

restoring forces, contributing to resistance to collapse <strong>of</strong> liquid<br />

lamellae with change <strong>of</strong> interfacial area (MacRitchie, 1976b).<br />

In defatted flours, proteins are the only surface active components<br />

forming monolayers. This results in a higher loaf volume and<br />

a finer and more uniform crumb structure as proteins provide quite<br />

stable monolayers and do not desorb easily from the gas–liquid<br />

interface. This has been attributed to suitable configurations<br />

adopted by protein molecules at the interface and to their large<br />

sizes (Larsson et al., 2006; MacRitchie, 1990). Addition <strong>of</strong> different<br />

lipid types may lead to the formation <strong>of</strong> mixed monolayers (Mills<br />

et al., 2003; Ross and MacRitchie, 1995; Salt et al., 2006). The<br />

occurrence <strong>of</strong> minima in loaf volume on addition <strong>of</strong> natural flour<br />

lipid and its polar fraction is analogous to the decrease in stability <strong>of</strong><br />

mixed films reported by Paternotte et al. (1993) on passage from<br />

a pure protein film to a lipid (monogalactosyl monoglyceride)–<br />

protein stabilized film. Variations in percentage <strong>of</strong> natural flour<br />

lipids at which the minimum volume is reached could be attributed<br />

to differences in protein composition <strong>of</strong> the two flours. A study by<br />

Salt et al. (2006) also indicated that mixed protein–lipid interfaces<br />

are relatively less elastic and therefore more unstable compared to<br />

pure protein ones, which possess high surface elasticity. Palmitic<br />

acid, a saturated free fatty acid, and polar lipids having digalactosyl<br />

diglyceride (DGDG), due to their tendency to form condensed<br />

monolayers (Gaines, 1966; MacRitchie, 1990), had either no effect<br />

or positive effect on loaf volumes, respectively. On the other hand,<br />

the probable formation <strong>of</strong> an expanded monolayer by linoleic acid<br />

resulted in detrimental effects on loaf volume. Myristic acid is<br />

a saturated free fatty acid but, due to its shorter chain length, forms<br />

expanded monolayers (MacRitchie, 1990), thus decreasing loaf<br />

volume. At higher levels <strong>of</strong> myristic acid addition, a reversal <strong>of</strong> the<br />

effect is observed (Figs. 2 and 3). This could be attributed to the<br />

presence <strong>of</strong> impurities in myristic acid, which is w99% pure by gas<br />

chromatographic analysis. However, such purity analysis does not<br />

indicate surface chemical purity <strong>of</strong> surface active compounds like<br />

myristic acid (MacRitchie, 1990). The presence <strong>of</strong> impurities, even<br />

in minor amounts, can influence surface properties. This might<br />

have been the case at higher levels <strong>of</strong> addition <strong>of</strong> myristic acid in


44<br />

B.S. Sroan, F. MacRitchie / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 41–46<br />

Table 1<br />

Crumb structure responses <strong>of</strong> Jagger and s<strong>of</strong>t wheat flour breads to different lipid types and levels, added to defatted flours as percentage <strong>of</strong> natural flour lipids<br />

% Of natural flour lipids Jagger wheat flour S<strong>of</strong>t wheat flour<br />

Number <strong>of</strong> cells Average cell elongation (C-cell score) Number <strong>of</strong> cells Average cell elongation (C-cell score)<br />

Natural flour lipids<br />

0% 2000.0 11.3 a 1.58 0.01 b 2544.5 187.4 a,b 1.69 0.01 a,b<br />

20% 1935.0 222.03 a,b 1.59 0.05 b 2521.0 110.3 a,b,c 1.68 0.01 a,b,c<br />

40% 1880.5 78.5 a,b 1.64 0.06 a,b 2210.5 115.3 d 1.65 0.03 b,c<br />

50% 1898.0 59.4 a,b 1.68 0.04 a 2209.0 79.2 d 1.65 0.02 c<br />

60% 1788.5 81.3 a,b 1.67 0.05 a,b 2347.5 2.1 b,c,d 1.70 0.02 a<br />

80% 1907.5 2.1 a,b 1.62 0.01 a,b 2196.0 86.3 d 1.66 0.00 a,b,c<br />

100%<br />

Flour polar lipids<br />

1749.0 41.0 b 1.63 0.01 a,b 2229.0 114.6 d 1.69 0.03 a,b<br />

0% 2361.5 7.8 a 1.56 0.03 c,d 2107.0 26.9 b 1.60 0.02 c<br />

20% 2307.0 157.0 a 1.57 0.01 b,c,d 1876.5 37.5 c 1.66 0.02 a,b<br />

40% 2001.0 24.0 b 1.60 0.04 a,b,c 2389.0 0.0 a 1.64 0.00 b<br />

66% 1985.0 0.0 b 1.62 0.00 a 2183.0 0.0 b 1.67 0.00 a,b<br />

132% 2285.0 0.0 a 1.54 0.00 d 2180.5 23.3 b 1.69 0.03 a<br />

200%<br />

Flour non-polar lipids<br />

2385.0 0.0 a 1.61 0.00 a,b 2092.0 216.4 b 1.67 0.01 a,b<br />

0% 2436.0 113.1 a 1.57 0.01 b 2110.0 22.6 a 1.60 0.02 a<br />

15% 2306.5 126.6 a 1.57 0.00 b 2036.5 95.5 a,b 1.59 0.01 a<br />

30% 1989.0 69.3 b,c 1.60 0.04 b 1832.0 144.2 b 1.68 0.05 a<br />

60% 2202.5 187.4 a,b 1.61 0.01 a,b 1950.0 104.7 a,b 1.66 0.05 a<br />

132% 1937.5 47.4 b,c 1.65 0.01 a 2053.0 14.1 a,b 1.64 0.01 a<br />

200%<br />

Linoleic acid<br />

1761.5 0.0 c 1.61 0.00 a,b – –<br />

0% 2452.5 26.2 a 1.63 0.06 a 2060.5 29.0 a,b 1.60 0.00 c<br />

66% 1803.0 220.6 b 1.68 0.02 a 2290.0 158.4 a 1.68 0.02 b<br />

132% 1796.5 67.2 b 1.72 0.04 a 1926.0 116.0 b 1.69 0.02 a,b<br />

200% 1473.5 169.0 b 1.70 0.04 a 2019.0 144.2 a,b 1.73 0.01 a<br />

250%<br />

Palmitic acid<br />

1505.5 13.4 b 1.72 0.00 a 1994.0 120.2 a,b 1.71 0.02 a<br />

0% 2477.5 139.3 a 1.61 0.04 a 2060.5 29.0 a 1.60 0.00 b<br />

66% 2438.5 103.9 a 1.64 0.02 a 2043.0 69.3 a 1.64 0.01 a,b<br />

132% 2522.5 171.8 a 1.63 0.06 a 1940.5 68.6 a 1.66 0.06 a,b<br />

200% 2331.0 14.1 a 1.58 0.00 a 1880.5 17.7 a 1.72 0.04 a<br />

250%<br />

Myristic acid<br />

2285.5 126.5 a 1.59 0.01 a 1829.5 204.4 a 1.65 0.03 b<br />

0% 2328.0 63.6 a 1.62 0.01 a 2100.0 135.8 a 1.68 0.01 b<br />

66% 2368.0 190.9 a 1.62 0.02 a 1954.0 24.0 a 1.68 0.01 b<br />

132% 2404.5 54.5 a 1.61 0.01 a 2263.5 54.5 a 1.68 0.01 b<br />

200% 2447.5 150.6 a 1.64 0.00 a 2270.5 147.8 a 1.68 0.01 b<br />

250% 2244.5 17.7 a 1.65 0.00 a 2156.5 188.8 a 1.68 0.01 b<br />

330% 2324.0 84.9 a 1.66 0.00 a – –<br />

400% 2632.0 35.4 a 1.65 0.00 a – –<br />

Values represent mean standard deviation for duplicate determinations.<br />

Means with the same letter within columns are not significantly different (p > 0.05).<br />

Fig. 4. C-cell images <strong>of</strong> Jagger wheat flour bread slices showing differences in gas cell expansion at different polar lipid levels (represented as percentages) added to defatted flour as<br />

percentage <strong>of</strong> natural flour lipids.


Fig. 5. C-cell images <strong>of</strong> Jagger wheat flour bread slices showing differences in gas cell<br />

expansion at different non-polar lipid levels (represented as percentages) added to<br />

defatted flour as percentage <strong>of</strong> natural flour lipids.<br />

baking studies. Therefore analysis <strong>of</strong> surface chemical purity<br />

(Lunkenheimer and Miller, 1987) <strong>of</strong> added surfactants, though not<br />

performed in this study, will be <strong>of</strong> great advantage in future studies.<br />

3.1.2. Crumb structure<br />

Image analysis <strong>of</strong> bread crumb showed that addition <strong>of</strong> different<br />

lipid types and variations in their levels resulted in negligible<br />

differences in number <strong>of</strong> gas cells and average cell elongation (Table<br />

1). These insignificant differences in the number <strong>of</strong> gas cells for<br />

a particular flour when treated with different levels <strong>of</strong> the same<br />

lipid type and free fatty acid, suggest that higher loaf volumes were<br />

due to an increase in expansion capacity <strong>of</strong> gas cells and not due to<br />

their number. That is, the film stabilizers (polar lipids and palmitic<br />

acid) allowed gas cells to expand, increasing the volume (Fig. 4).<br />

With improvement in expansion capacity, the size <strong>of</strong> gas cells<br />

increased and crumb structure changed from fine to more<br />

uniformly distributed larger gas cells. The inability <strong>of</strong> gas cells to<br />

expand (Fig. 5) also explains the earlier observations <strong>of</strong> relatively<br />

fine crumb appearance with addition <strong>of</strong> non-polar lipids reported<br />

by Ponte and DeStefanis (1969).<br />

However, an exception was observed in the case <strong>of</strong> linoleic acid<br />

addition to Jagger flour. Incremental additions <strong>of</strong> linoleic acid to<br />

Jagger flour caused small but continuous decrease in number <strong>of</strong> gas<br />

cells (Table 1). Linoleic acid might also be detrimental to initial<br />

concentration and stability <strong>of</strong> gas cells during mixing, thus<br />

hindering the ease with which gas cells are occluded into the liquid<br />

phase <strong>of</strong> dough. This initial adsorption depends upon the diffusion<br />

coefficient <strong>of</strong> surface active compounds (lipids and proteins) in the<br />

liquid phase <strong>of</strong> the dough (MacRitchie, 1990), and will vary from<br />

flour to flour. The nature <strong>of</strong> surface active compounds which has<br />

been adsorbed initially at the interface to form a monolayer will<br />

stabilize or destabilize freshly occluded gas cells, thus affecting the<br />

ease with which they are concentrated (Larsson et al., 2006).<br />

Negligible variations in average cell elongation indicated that<br />

different lipid types and free fatty acids at the levels <strong>of</strong> addition<br />

used might not be influencing rheology <strong>of</strong> the gluten–starch<br />

matrix, since cell elongation is thought to be associated with dough<br />

rheological properties (Gandikota and MacRitchie, 2005).<br />

3.2. Effect <strong>of</strong> variations in lipid types and free fatty acids, and their<br />

levels on biaxial extensional rheology <strong>of</strong> gluten–starch matrix<br />

Biaxial extensional rheological tests were performed on Jagger<br />

and s<strong>of</strong>t doughs with different lipid types and their levels, in order<br />

B.S. Sroan, F. MacRitchie / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 41–46 45<br />

Table 2<br />

Mean bubble inflation rheological responses <strong>of</strong> Jagger wheat flour doughs to<br />

different lipid types and levels, added to defatted flours as percentage <strong>of</strong> natural<br />

flour lipids<br />

% Of natural flour<br />

lipid<br />

to investigate possible independent action <strong>of</strong> lipids on baking<br />

performance (loaf volume and crumb structure). Different lipid<br />

types and free fatty acids at the levels used did not influence biaxial<br />

extensional rheological parameters (maximum stress, failure strain<br />

and strain hardening index) <strong>of</strong> the doughs (Tables 2 and 3).<br />

However, higher values were observed for Jagger than s<strong>of</strong>t wheat<br />

flour. Though minor differences were observed for strain hardening<br />

indices within a particular flour at some treatments, no specific<br />

trend was observed, thus attributing the variation to slight experimental<br />

scatter.<br />

As for the case in Part-I <strong>of</strong> this study, higher values <strong>of</strong> biaxial<br />

extensional rheological parameters were observed in the case <strong>of</strong><br />

Jagger flour, indicating its superior breadmaking potential<br />

(Dobraszczyk et al., 2003). Inability <strong>of</strong> different lipid types and free<br />

fatty acids to cause any effect on biaxial extensional rheology,<br />

demonstrates that their action is independent <strong>of</strong> rheology <strong>of</strong> the<br />

Table 3<br />

Mean bubble inflation rheological responses <strong>of</strong> s<strong>of</strong>t wheat flour doughs to different<br />

lipid types and levels, added to defatted flours as percentage <strong>of</strong> natural flour lipids<br />

% Of natural flour<br />

lipid<br />

Max. stress<br />

(kPa)<br />

Max. stress<br />

(kPa)<br />

Failure strain<br />

(Hencky)<br />

Failure strain<br />

(Hencky)<br />

Strain hardening<br />

index<br />

Natural flour lipids<br />

0% 571.62 305.45 a 2.58 0.18 a 2.21 0.12 b,c<br />

60% 491.68 237.21 a 2.60 0.12 a 2.38 0.064 a,b<br />

100% 598.62 112.05 a 2.66 0.04 a 2.44 0.06 a<br />

Flour polar lipids<br />

40% 480.85 67.26 a 2.56 0.04 a 2.18 0.03 c<br />

200% 660.22 130.74 a 2.64 0.05 a 2.22 0.04 b,c<br />

Flour non-polar lipids<br />

60% 936.93 560.68 a 2.74 0.16 a 2.44 0.09 a<br />

Linoleic acid<br />

132% 650.21 18.07 a 2.66 0.04 a 2.36 0.10 a,b,c<br />

Palmitic acid<br />

132% 598.55 239.23 a 2.62 0.12 a 2.23 0.09 b,c<br />

Myristic acid<br />

250% 613.71 90.70 a 2.65 0.03 a 2.32 0.10 b,c<br />

Over all average 623.58 236.59 2.63 0.10 2.31 0.12<br />

Values represent mean standard deviation for duplicate determinations.<br />

Means with the same letter within columns are not significantly different (p > 0.05).<br />

Strain hardening<br />

index<br />

Natural flour lipids<br />

0% 120.80 12.76 a 1.79 0.00 a 1.76 0.01 a<br />

40% 125.42 12.14 a 1.85 0.02 a 1.74 0.01 a<br />

100% 102.40 5.45 a 1.66 0.05 b 1.64 0.04 c<br />

Flour polar lipids<br />

20% 105.49 3.52 a 1.68 0.02 b 1.66 0.00 b,c<br />

200% 130.17 1.61 a 1.84 0.02 a 1.72 0.02 a,b<br />

Flour non-polar lipids<br />

60% 108.71 27.51 a 1.83 0.02 a 1.73 0.01 a,b<br />

Linoleic acid<br />

132% 104.35 12.49 a 1.81 0.06 a 1.75 0.07 a<br />

Palmitic acid<br />

132% 116.21 0.52 a 1.82 0.01 a 1.79 0.01 a<br />

Myristic acid<br />

250% 115.44 23.14 a 1.78 0.02 a 1.72 0.03 a<br />

Over all average 114.19 13.69 1.79 0.07 1.72 0.05<br />

Values represent mean standard deviation for duplicate determinations.<br />

Means with the same letter within columns are not significantly different (p > 0.05).


46<br />

gluten–starch matrix and that they act purely as surface active<br />

compounds at these levels <strong>of</strong> addition.<br />

4. Conclusion<br />

This study for the first time demonstrates that the lipids at their<br />

natural levels do not affect dough rheology and gas cell stabilization<br />

by the gluten–starch matrix. Nevertheless, their ability to modify<br />

loaf volume and crumb structure supports the dual film hypothesis<br />

<strong>of</strong> Gan et al. (Gan et al., 1990, 1995). It suggests the presence <strong>of</strong><br />

liquid lamellae, providing an independent mechanism <strong>of</strong> gas cell<br />

stabilization. The effects <strong>of</strong> different surface active components<br />

may be explained by the type <strong>of</strong> monolayer that they form.<br />

References<br />

Baker, J.C., Parker, H.K., Mize, M.D., 1946. Supercentrifugates from dough. <strong>Cereal</strong><br />

Chemistry 23, 539–544.<br />

DeStefanis, V.A., Ponte Jr., J.G., 1976. Studies on breadmaking properties <strong>of</strong> wheatflour<br />

nonpolar lipids. <strong>Cereal</strong> Chemistry 53, 636–642.<br />

Dobraszczyk, B.J., Smewing, J., Albertini, M., Maesmans, G., Sch<strong>of</strong>ield, J.D., 2003.<br />

Extensional rheology and stability <strong>of</strong> gas cell walls in bread doughs at elevated<br />

temperatures in relation to breadmaking performance. <strong>Cereal</strong> Chemistry 80,<br />

218–224.<br />

Gaines, G.L., 1966. Insoluble monolayers at liquid–gas interfaces. In: Prigogine, I.<br />

(Ed.), Interscience Monographs on Physical Chemistry. John Wiley and Sons,<br />

Inc., New York, USA.<br />

Gan, Z., Angold, R.E., Williams, M.R., Ellis, P.R., Vaughan, J.G., Galliard, T., 1990. The<br />

microstructure <strong>of</strong> gas retention <strong>of</strong> bread dough. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 12,15–24.<br />

Gan, Z., Ellis, P.R., Sch<strong>of</strong>ield, J.D., 1995. Gas cell stabilization and gas retention in<br />

wheat bread dough. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 21, 215–230.<br />

B.S. Sroan, F. MacRitchie / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 41–46<br />

Gandikota, S., MacRitchie, F., 2005. Expansion capacity <strong>of</strong> doughs: methodology and<br />

applications. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 42, 157–163.<br />

Larsson, K., Quinn, P., Sato, K., Tiberg, F., 2006. Lipids: Structure, Physical Properties<br />

and Functionality. The Oily Press, Bridgwater, England.<br />

Lunkenheimer, K., Miller, R., 1987. A criterion for judging the purity <strong>of</strong> adsorbed<br />

surfactant layers. <strong>Journal</strong> <strong>of</strong> Colloid Interface <strong>Science</strong> 120, 176–183.<br />

McCormack, G., Panozzo, J., MacRitchie, F., 1991. Contributions to breadmaking <strong>of</strong><br />

inherent variations in lipid content and composition <strong>of</strong> wheat cultivars. II.<br />

Fractionation and reconstitution studies. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 13,<br />

263–274.<br />

MacRitchie, F., 1976a. The liquid phase <strong>of</strong> dough and its role in baking. <strong>Cereal</strong><br />

Chemistry 53, 318–326.<br />

MacRitchie, F., 1976b. Monolayer compression barrier in emulsion and foam<br />

stability. <strong>Journal</strong> <strong>of</strong> Colloid and Interface <strong>Science</strong> 56, 53–56.<br />

MacRitchie, F., 1977. Flour lipids and their effects in baking. <strong>Journal</strong> <strong>of</strong> the <strong>Science</strong> <strong>of</strong><br />

Food and Agriculture 28, 53–58.<br />

MacRitchie, F., 1978. Differences in baking quality between wheat flours. <strong>Journal</strong> <strong>of</strong><br />

Food Technology 13, 187–194.<br />

MacRitchie, F., 1990. Chemistry at Interfaces. Academic Press, Inc., San Diego, CA,<br />

USA.<br />

MacRitchie, F., Gras, P.W., 1973. The role <strong>of</strong> flour lipids in baking. <strong>Cereal</strong> Chemistry<br />

50, 292–302.<br />

Mills, E.N.C., Wilde, P.J., Salt, L.J., Skeggs, P., 2003. Bubble formation and stabilization<br />

in bread dough. Food and Bioproducts Processing 81, 189–193.<br />

Paternotte, T.A., Orsel, R., Hamer, R.J., 1993. Interactions between flour proteins and<br />

flour lipids at the liquid/air interface. In: Gluten Proteins. Association <strong>of</strong> <strong>Cereal</strong><br />

Research, Detmold, Germany, pp. 207–217.<br />

Ponte Jr., J.G., DeStefanis, V.A., 1969. Note on separation and baking properties <strong>of</strong><br />

polar and nonpolar wheat flour lipids. <strong>Cereal</strong> Chemistry 46, 325–329.<br />

Ross, A., MacRitchie, F., 1995. Interactions <strong>of</strong> wheat proteins, carbohydrates and<br />

lipids. In: Gaonkar, A.G. (Ed.), Ingredient Interactions Effects on Food Quality.<br />

Marcel Deker, New York, USA, pp. 321–356.<br />

Salt, L.J., Wilde, P.J., Georget, D., Wellner, N., Skeggs, P.K., Mills, E.N.C., 2006.<br />

Composition and surface properties <strong>of</strong> dough liquor. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

43, 284–292.


Expression <strong>of</strong> globulin-2, a member <strong>of</strong> the cupin superfamily <strong>of</strong> proteins with<br />

similarity to known food allergens, is increased under high temperature<br />

regimens during wheat grain development<br />

Susan B. Altenbach *, Charlene K. Tanaka, William J. Hurkman, William H. Vensel<br />

USDA-ARS Western Regional Research Center, 800 Buchanan Street, Albany, CA 94710, USA<br />

article info<br />

Article history:<br />

Received 27 February 2008<br />

Received in revised form 28 June 2008<br />

Accepted 8 July 2008<br />

Keywords:<br />

Flour quality<br />

Gene expression<br />

Proteomics<br />

Quantitative RT-PCR<br />

1. Introduction<br />

abstract<br />

Wheat is one <strong>of</strong> the major crops grown throughout the world<br />

with a primary use in human nutrition. In 2005, wheat was grown<br />

on more than 200 million hectares and more than 600 million tons<br />

<strong>of</strong> grain were produced worldwide (http://faostat.fao.org/). A large<br />

variety <strong>of</strong> breads, noodles and baked goods are made from wheat<br />

flour because <strong>of</strong> the unique viscoelastic properties that result when<br />

the flour is mixed with water. These viscoelastic properties are<br />

determined largely by the gluten proteins, the major storage<br />

proteins that comprise 80–85% <strong>of</strong> wheat flour protein. The gluten<br />

proteins, consisting <strong>of</strong> gliadins and glutenin subunits, are characterized<br />

by unusually high levels <strong>of</strong> proline and glutamine and have<br />

very low solubilities in water or salt solutions. These proteins have<br />

been studied extensively because <strong>of</strong> their importance in flour<br />

functionality. The remaining 15–20% <strong>of</strong> the endosperm protein is<br />

Abbreviations: 2-DE, two-dimensional gel electrophoresis; 2-DE/MS, twodimensional<br />

gel electrophoresis/mass spectrometry; DPA, days’ post-anthesis;<br />

ESTs, expressed sequence tags; MS/MS, tandem mass spectrometry; NPK,<br />

nitrogen–phosphorous–potassium; qRT-PCR, quantitative reverse-transcriptase<br />

polymerase chain reaction; SDS-PAGE, sodium dodecyl sulfate-polyacrylamide gel<br />

electrophoresis.<br />

* Corresponding author. Tel.: þ1 510 559 5614; fax: þ1 510 559 5818.<br />

E-mail address: altnbach@pw.usda.gov (S.B. Altenbach).<br />

0733-5210/$ – see front matter Published by Elsevier Ltd.<br />

doi:10.1016/j.jcs.2008.07.005<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 47–54<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

Twenty-three expressed sequence tags (ESTs) from the US spring wheat Butte 86 were identified that<br />

encode proteins similar to a globulin-2 protein from maize embryos. The ESTs assembled into three<br />

contigs, two <strong>of</strong> which include the entire coding region for the mature protein. The encoded proteins<br />

contain two cupin domains and show significant identities with 7S seed proteins from other species that<br />

are known or putative food allergens. Quantitative reverse-transcriptase polymerase chain reaction (qRT-<br />

PCR) with primers specific for two <strong>of</strong> the sequences demonstrated that the globulin-2 genes are<br />

expressed late in grain development and that transcript levels increase when grain is produced under<br />

high temperature conditions. Transcripts were detected in both whole grain and endosperm, but levels<br />

were significantly higher in whole grain and highest in embryo. In wheat flour, at least 17 protein spots<br />

that differ in both size and pI were identified as globulin-2 by 2-DE/MS. Seven <strong>of</strong> the spots increased<br />

more than 2-fold in relative proportion when grain was produced under high temperature regimens. The<br />

data suggest that both transcriptional and post-translational mechanisms are involved in the response <strong>of</strong><br />

globulin-2 to high temperatures.<br />

Published by Elsevier Ltd.<br />

a heterogeneous mixture <strong>of</strong> proteins, most <strong>of</strong> which are soluble in<br />

water or dilute salt solutions. Within the salt-soluble protein fraction<br />

are a number <strong>of</strong> proteins related to the predominant storage<br />

proteins characterized in seeds from dicotyledonous plants. Some<br />

<strong>of</strong> these proteins have been shown to have sedimentation values<br />

similar to the 7S vicilins from pea and to share antigenic determinants<br />

with the pea vicilins (Robert et al., 1985). A 7S þ 3S globulin<br />

fraction characterized by Robert et al. (1985) contained major<br />

proteins <strong>of</strong> w75, 55, 36–40, 24,17–20 and 10 kDa when analyzed by<br />

SDS-PAGE.<br />

Recently, a 2-DE/MS approach was used to identify more than<br />

200 proteins contained in a salt-soluble fraction <strong>of</strong> wheat endosperm<br />

from the US spring wheat Butte 86 (Vensel et al., 2005).<br />

Seven spots in the 55 kDa region <strong>of</strong> the 2-D gels were determined to<br />

be most similar to a 7S vicilin-like protein characterized from maize<br />

embryos, referred to as globulin-2 (Wallace and Kriz, 1991). These<br />

proteins were found in wheat endosperm at 36 DPA but not at 10<br />

DPA. In a subsequent proteomic analysis <strong>of</strong> endosperm from<br />

developing grain, the globulin-2 proteins were shown to accumulate<br />

during late stages <strong>of</strong> wheat grain development. Interestingly,<br />

these proteins also were among a group that increased in relative<br />

proportion when grain development occurred under a high<br />

temperature regimen <strong>of</strong> 37/28 C (day/night) rather than under a<br />

moderate temperature regimen <strong>of</strong> 24/17 C(Hurkman et al., 2009).<br />

Since these data suggest biological roles for the globulin-2 proteins


48<br />

in the response <strong>of</strong> the grain to abiotic stress, we surveyed databases<br />

<strong>of</strong> wheat expressed sequence tags (ESTs) to identify the complement<br />

<strong>of</strong> globulin-2 genes from the US spring wheat Butte 86.<br />

Additionally, we examined the expression <strong>of</strong> two <strong>of</strong> the globulin-2<br />

genes in developing grain and endosperm under different<br />

controlled environmental regimens using quantitative RT-PCR and<br />

evaluated the relative proportions <strong>of</strong> globulin-2 proteins in flour<br />

produced from grain subjected to moderate and high temperature<br />

regimens.<br />

2. Experimental<br />

2.1. Growth <strong>of</strong> plants and tissue collection<br />

The US hard red spring wheat Triticum aestivum cv. Butte 86 was<br />

grown in a climate-controlled greenhouse as described by Altenbach<br />

et al. (2007). During grain development, plants were placed<br />

under either a moderate (24/17 C day/night) or a high (37/28 C<br />

day/night) temperature regimen. Levels <strong>of</strong> fertilizer (Plantex<br />

20–20–20 NPK) were adjusted at anthesis so that plants received<br />

300 mg per day (1 ), 150 mg per day (0.5 )orno(0 ) NPK during<br />

grain development. Developing grain and endosperm were<br />

collected at various intervals after anthesis under both temperature<br />

regimens. Embryo tissue was harvested from 30 DPA grain from<br />

plants grown under the 24/17 C regimen. Awns, glumes, stems and<br />

leaves were obtained during middle stages <strong>of</strong> grain development<br />

under the moderate regimen. Roots were harvested from seedlings<br />

grown on water-saturated sand at room temperature for 20 days.<br />

All tissues were frozen in liquid nitrogen and stored at 80 C<br />

until use.<br />

2.2. RNA preparation and qRT-PCR<br />

Total RNA was isolated from all tissues as described previously<br />

for endosperm (Altenbach, 1998), treated with RQ1 RNAse-free<br />

DNAse (Promega, Madison, WI), and further purified by extraction<br />

with phenol:chlor<strong>of</strong>orm:isoamylalcohol (24:24:1) and ethanol<br />

precipitation. RNA was reverse-transcribed using the QuantiTect<br />

Reverse Transcription Kit (QIAGEN, Valencia, CA) according to the<br />

manufacturer’s directions. Amplification reactions were carried out<br />

in a volume <strong>of</strong> 25 ml containing cDNA, 0.3 mM <strong>of</strong> forward and<br />

reverse primers and SYBR Green Supermix (Biorad Laboratories,<br />

Hercules, CA) using a BioRad iCycler with an initial denaturation<br />

step <strong>of</strong> 95 C for 3 min, followed by 40 cycles <strong>of</strong> 95 C for 10 s and<br />

55 C for 45 s. At the end <strong>of</strong> the PCR cycles, a melting curve was<br />

generated and analyzed by the iCycler s<strong>of</strong>tware.<br />

Oligonucleotide primers for globulin-2 genes were based on<br />

expressed sequence tags (ESTs) from developing grain <strong>of</strong> Butte 86<br />

available from NCBI. Primers were selected using Beacon Designer<br />

4.0 (Premier Bios<strong>of</strong>t International, Palo Alto, CA) and synthesized by<br />

Operon (Huntsville, AL). Forward primer QF123 has the sequence 5 0<br />

GGTCCAAATGGCTCCGC 3 0 and reverse primer QR123 has the<br />

sequence 5 0 CAAGAAGGACTGCGGGC 3 0 . Forward primer QF126 has<br />

the sequence 5 0 GGTCGTCATGCTCCTCAAC 3 0 and reverse primer<br />

QR126 has the sequence 5 0 GACGCTGAAGAAGGACTGTG 3 0 . Primers<br />

for an 18S rRNA used as a reference were reported in Altenbach<br />

et al. (2007). Amplification efficiencies for each primer pair were<br />

calculated from standard curves generated in three independent<br />

experiments using the iCycler s<strong>of</strong>tware. Each standard curve had<br />

a minimum <strong>of</strong> five points and R values greater than 0.99. Amplification<br />

products were evaluated by gel electrophoresis on 4%<br />

Metaphor agarose gels (VWR International, Brisbane, CA).<br />

For quantification <strong>of</strong> transcripts, PCR reactions were carried out<br />

in triplicate using cDNA from the equivalent <strong>of</strong> 10 ng RNA from each<br />

time point. The 18S rRNA served as a reference RNA and was<br />

amplified in parallel with target genes in all experiments. The Ct<br />

S.B. Altenbach et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 47–54<br />

value was determined for each RNA sample and primer pair by the<br />

iCycler s<strong>of</strong>tware. Mean normalized expression and standard errors<br />

were determined for each target gene relative to the 18S reference<br />

using the Q-Gene Core Module (available at http://www.genequantification.de/download.html#qgene).<br />

Amplification efficiencies<br />

<strong>of</strong> each primer pair were taken into account for all calculations<br />

using equation #3 <strong>of</strong> Muller et al. (2002). Mean normalized<br />

expression was plotted as a function <strong>of</strong> chronological age for each<br />

experiment.<br />

2.3. Protein extraction, 2-DE and MS/MS analyses<br />

Endosperm was collected from developing grain and a KClsoluble/methanol-insoluble<br />

protein fraction was isolated and<br />

quantified as described previously (Hurkman and Tanaka, 2004;<br />

Vensel et al., 2005). Equal amounts <strong>of</strong> protein (18 mg) from each<br />

time point were separated by 2-DE in triplicate. Gels were digitized<br />

with a calibrated scanner (UMAX PowerLook III, Dallas, TX) at<br />

300 dpi using the same setting for all gels and analyzed using<br />

a computerized image analysis system (Progenesis PG240 version<br />

2006, Non-Linear Dynamics Limited, Newcastle upon Tyne, UK).<br />

Normalized volumes (individual spot volume/total spot volume<br />

100) were determined for each spot, averaged over the three<br />

gels, and plotted against the chronological age <strong>of</strong> the grain in DPA.<br />

Grain produced under different temperature regimens was<br />

tempered to 15% moisture and milled to flour by the USDA-ARS<br />

Wheat Quality Laboratory (Manhattan, KS) using a Brabender<br />

Quadrumat Jr. mill (Hackensack, NJ). Methods for extraction and<br />

analysis <strong>of</strong> proteins from flour were the same as those used for<br />

endosperm.<br />

For identification <strong>of</strong> proteins by tandem mass spectrometry<br />

(MS/MS), spots were excised from a representative gel and reduction,<br />

alkylation, reagent removal, and tryptic digestion were carried<br />

out automatically by a DigestPro xyz robot (INTAVIS Bioanalytical<br />

Instruments AG, Bergisch Gladbach, Germany). The DigestPro<br />

collection tray containing the tryptic peptides was placed in an<br />

autosampler that was interfaced with a QSTAR pulsar I hybrid<br />

quadrupole-TOF instrument (Applied Biosystems/MDX Sciex,<br />

Toronto, Canada) configured with an ESI source. Acquisition <strong>of</strong><br />

tandem mass spectrometry data was as follows. From an initial<br />

survey scan <strong>of</strong> mass range m/z 400–2000, the most abundant<br />

doubly or triply charged ion above a threshold <strong>of</strong> 20 counts was<br />

selected for fragmentation. Collision induced dissociation <strong>of</strong> the<br />

mass-selected ion was carried out using UHP nitrogen. Following<br />

the 3S MS/MS fragmentation period, the MS survey scan was<br />

repeated until another MS/MS period was triggered. Wiff data files<br />

were created for each sample by the Analyst QS version 1.1<br />

s<strong>of</strong>tware, converted to MGF text files using Mascot Daemon (http://<br />

www/matrixscience.com/) and submitted in batch mode to<br />

a locally installed copy <strong>of</strong> X!Tandem (Craig and Beavis, 2004) using<br />

a script provided by Jayson Faulkner (University <strong>of</strong> Michigan).<br />

X!Tandem was configured to search files containing sequences<br />

from the NCBI non-redundant collection: nr-Arabidopsis-thaliana.fasta,<br />

nr-other-Viridiplantae.fasta, nr-Oryza-sativa.fasta, in addition<br />

to entries from all proteins in the HarvEST:Wheat version 1.04<br />

database (http://harvest.ucr.edu/HWheat104.exe), NCBI Triticum<br />

aestivum: UniGene Build #37, and wEST Database (http://wheat.pw.<br />

usda.gov/wEST). The results were visualized using a locally<br />

installed copy <strong>of</strong> the Global Proteome Machine (GPM) (http://<br />

thegpm.org/). Trypsin was selected as the cleavage enzyme. The<br />

results were searched with a fragment ion mass tolerance <strong>of</strong><br />

1.00 Da and a parent ion tolerance <strong>of</strong> 0.40 Da. Oxidation <strong>of</strong> methionine<br />

was specified as a variable modification (see also: Vensel<br />

et al., 2002, 2005). Identifications accepted as valid had log e values<br />

less than 4.0.


3. Results<br />

3.1. Accumulation <strong>of</strong> globulin-2 in developing endosperm under<br />

different temperature regimens<br />

A number <strong>of</strong> protein spots were identified as globulin-2 in<br />

proteomic maps <strong>of</strong> KCl-soluble/MeOH-insoluble proteins from<br />

developing endosperm from Butte 86 (Vensel et al., 2005). Six spots<br />

(referred to as 147, 155, 559, 561, 851, and 852 in Vensel et al., 2005)<br />

ranged from about 54.8 to 56.1 kDa by 2-DE and had pIs between<br />

6.2 and 7.0. Another spot (631 in Vensel et al., 2005) was about<br />

15 kDa and had a pI <strong>of</strong> about 5.7. Accumulation pr<strong>of</strong>iles for these<br />

seven globulin-2 proteins during grain development under<br />

moderate and high temperature regimens were summarized as<br />

part <strong>of</strong> a large-scale proteomic analysis by Hurkman et al. (2009)<br />

and are shown in more detail in Fig. 1. Under moderate temperatures<br />

(24/17 C), these proteins accumulated very late in grain<br />

development, between 32 and 34 DPA, and each protein spot<br />

S.B. Altenbach et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 47–54 49<br />

encompassed less than 0.2% <strong>of</strong> the normalized volumes <strong>of</strong> all spots<br />

in the KCl-soluble/MeOH-insoluble fraction. When a high temperature<br />

regimen <strong>of</strong> 37/28 C was applied from 10 DPA until maturity,<br />

the proteins accumulated earlier, between 22 and 24 DPA. In<br />

addition, maximum normalized spot volumes were 1.9- to 4.1-fold<br />

higher than in grain produced under the 24/17 C regimen (Fig. 1<br />

and Hurkman et al., 2009).<br />

3.2. Identification <strong>of</strong> ESTs encoding globulin-2 from Butte 86<br />

A collection <strong>of</strong> contigs assembled from more than 116,000 wheat<br />

ESTs by Chao et al. (2006) was surveyed to identify sequences that<br />

encode the wheat globulin-2 proteins. The EST collection includes<br />

3639 ESTs from Butte 86 developing grain. Information about both<br />

contigs and ESTs is available at http://wheat.pw.usda.gov/westsql/.<br />

Three contigs, NSFT03P2_Contig18428, NSFT03P2_Contig17295 and<br />

NSFT03P2_Contig17366, encode proteins similar to a globulin-2<br />

characterized previously in maize (Wallace and Kriz, 1991). Amino<br />

Fig. 1. Accumulation <strong>of</strong> globulin-2 proteins in endosperm from grain produced under moderate and high temperatures. In each panel the solid lines denote endosperm produced<br />

under the 24/17 C regimen and the dashed lines denote endosperm produced under the 37/28 C regimen. The high temperature regimen was imposed from 10 DPA to maturity<br />

and plants were supplied with 0.5 NPK. Bars indicate the standard error for each time point among triplicate gels. Accumulation data for spots 155, 559, 851, 147, 631, 852 and 561<br />

from Vensel et al. (2005) are presented in panels A, B, C, D, E, F and G, respectively.


50<br />

acid sequences <strong>of</strong> the proteins encoded by the contigs are shown in<br />

Fig. 2. Contig 18428 includes nine ESTs from Butte 86 (BQ806323,<br />

BQ807065, BQ806087, BQ805228, BQ806420, BQ805507,<br />

BQ806904, BQ804552, BQ804667) that represent the entire coding<br />

region. Contig 17295 includes four ESTs from Butte 86 (BQ805639,<br />

BQ839104, BQ805390, BQ807159) that cover the 5 0 portion <strong>of</strong> the<br />

coding region and one EST (BQ805520) that represents the 3 0 end.<br />

Nine ESTs from Butte 86 were included in contig 17366 (BQ839060,<br />

BQ805920, BQ807178, BQ838786, BQ839055, BQ838678, BQ807089,<br />

BQ806653, BQ805350) and all nine represent the 5 0 half <strong>of</strong> the<br />

coding region. It is likely that this contig is missing the 3 0 end <strong>of</strong> the<br />

coding region.<br />

When evaluated with the Signal P algorithm (http://www.cbs.<br />

dtu.dk/services/SignalP/), proteins encoded by the three contigs<br />

are predicted to contain signal peptide cleavage sites between the<br />

alanine and the serine residues at positions 26 and 27 in contigs<br />

17295 and 17366 and positions 20 and 21 in contig 18428 (Bendtsen<br />

et al., 2004), suggesting that these proteins enter the secretory<br />

pathway (Fig. 2). Contigs 18428 and 17295 encode proteins with<br />

predicted molecular weights <strong>of</strong> 53,495 and 53,773, respectively,<br />

excluding the signal peptide. Both proteins contain abundant<br />

arginine, glycine and glutamine residues. An InterProScan (http://<br />

www.ebi.ac.uk/InterProScan/) showed that each protein contains<br />

two cupin domains (Fig. 2) that are able to form barrel-like<br />

structures typical <strong>of</strong> the cupin superfamily <strong>of</strong> proteins.<br />

Since many plant proteins within the cupin superfamily are food<br />

allergens (Mills et al., 2002), the protein encoded by contig 17295<br />

was used to search the Food Allergy Research and Resource Program<br />

(FARRP) Protein AllergenOnline Database version 8.0 (http://www.<br />

allergenonline.com/databasefasta.asp). A sliding 80-mer search<br />

revealed matches <strong>of</strong> >35% identity between this globulin-2 and 18<br />

7S proteins that are known or putative allergens (Table 1). These<br />

include proteins from sesame, various tree nuts (English walnut,<br />

black walnut, hazelnut, cashew) and legumes (lentil, pea, soybean,<br />

peanut). Current Codex Alimentarius guidelines indicate that<br />

proteins with identities to known allergens that are >35% over 80<br />

amino acid regions may be allergenic (Goodman, 2006).<br />

S.B. Altenbach et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 47–54<br />

3.3. Selection <strong>of</strong> primers for qRT-PCR and analysis <strong>of</strong> gene<br />

expression<br />

ESTs BQ806323, BQ805639 and BQ839060 from cv Butte 86<br />

represent the first half <strong>of</strong> the coding sequences <strong>of</strong> contigs 18428,<br />

17295 and 17366, respectively, and were used for the design <strong>of</strong><br />

gene-specific primers. Primer QF123 was a perfect match with<br />

BQ806323, but had only 13/17 nucleotides identical to BQ839060<br />

and BQ805639 (Fig. 3). Primer QR123 was a reverse complement <strong>of</strong><br />

BQ806323, but had only 13/17 nucleotides that complemented<br />

either BQ839060 or BQ805639. QF123/QR123 amplified a 110 bp<br />

fragment that was verified to correspond to BQ839060 by digestion<br />

with AlwI, HaeII, and HpaII. The average amplification efficiency <strong>of</strong><br />

QF123/QR123 was 96.7%. Primer QF126 was a perfect match with<br />

both BQ805639 and BQ839060 but had only 16/19 nucleotides in<br />

common with BQ806323 (Fig. 3). Primer QR126 was a reverse<br />

complement <strong>of</strong> BQ805639, but had only 18 and 17 <strong>of</strong> 20 residues<br />

that complemented BQ806323 and BQ839060, respectively. QF126/<br />

QR126 amplified a 100 bp fragment that was verified to correspond<br />

to BQ805639 by digestion with HpaII, MboII, AciI, AlwI and HaeII.<br />

The average amplification efficiency <strong>of</strong> QF126/QR126 was 93.7%.<br />

Transcript accumulation for the two globulin-2 genes was<br />

assessed in developing grain produced under either moderate or<br />

high temperature regimens from anthesis to maturity (Fig. 4). In the<br />

absence <strong>of</strong> post-anthesis NPK (Fig. 4A, B), transcripts corresponding<br />

to both genes were first detected at 11 DPA under moderate<br />

temperatures and reached maximum levels at 26 DPA. Under the<br />

high temperature regimen, transcripts began to accumulate earlier<br />

in grain development, by 6 DPA, and reached maximum levels at 22<br />

DPA. Changes in the timing <strong>of</strong> globulin-2 transcript accumulation<br />

were consistent with changes in the timing <strong>of</strong> developmental<br />

events due to high temperature (Altenbach et al., 2003) and with<br />

changes in the timing <strong>of</strong> expression <strong>of</strong> other genes (Altenbach and<br />

Kothari, 2004; Altenbach et al., 2007, 2008). Maximum levels <strong>of</strong><br />

transcripts corresponding to BQ806323 and BQ805639 were 4.1and<br />

4.9-fold higher, respectively, under high temperatures than<br />

under moderate temperatures. In a second growth experiment,<br />

Fig. 2. Comparison <strong>of</strong> protein sequences encoded by NSFT03P2_Contig18428, NSFT03P2_Contig17295 and NSFT03P2_Contig17366. Identical amino acid residues in each protein are<br />

enclosed in boxes. Brackets denote portions <strong>of</strong> the coding regions in each contig that correspond to ESTs from Butte 86. The arrow indicates putative signal peptide cleavage sites.<br />

Cupin domains in each protein are shown in bold type and putative N-linked glycosylation sites are indicated by asterisks.


Table 1<br />

Comparison <strong>of</strong> globulin-2 encoded by NSFT03P2_Contig17295 with known and<br />

putative allergens in the FARRP Protein AllergenOnline Database Version 8.0 using<br />

a sliding 80-mer search<br />

GenBank<br />

accession #<br />

Source Best %<br />

identity<br />

13183177 Sesame indicum (sesame) 51.90 285<br />

6580762 Juglans regia (English walnut) 48.10 269<br />

31321944 Juglans nigra (black walnut) 47.50 251<br />

19338630 Corylus avellana (hazelnut) 47.50 204<br />

21666498 Anacardium occidental (cashew) 43.80 173<br />

21914823 Anacardium occidental (cashew) 43.80 179<br />

29539109 Lens culinaris (lentil) 42.51 112<br />

46560474 Arachis hypogaea (peanut) 41.29 62<br />

42414627 Pisum sativum (pea) 41.24 97<br />

42414629 Pisum sativum (pea) 41.24 105<br />

1168390 Arachis hypogaea (peanut) 40.70 75<br />

46560472 Arachis hypogaea (peanut) 40.70 58<br />

1168391 Arachis hypogaea (peanut) 40.70 79<br />

46560476 Arachis hypogaea (peanut) 40.70 79<br />

29539111 Lens culinaris (lentil) 40.00 123<br />

256427 Glycine max (soybean) 38.79 45<br />

18536 Glycine max (soybean) 36.29 11<br />

169929 Glycine max (soybean) 35.03 2<br />

# Hits > 35% a<br />

Proteins with >35% identity over 80 amino acids to wheat globulin-2 are shown.<br />

a <strong>of</strong> 426 80-mers.<br />

plants were supplied with post-anthesis NPK (Fig. 4C, D). In this<br />

experiment, transcripts for BQ806323 were first detected by 12 DPA<br />

and reached maximum levels at 34 DPA under moderate temperatures<br />

(Fig. 4C) while transcripts for BQ805639 were detectable at<br />

low levels at 7 DPA and reached maximum levels at 34 DPA<br />

(Fig. 4D). Under the high temperature regimen, transcripts for both<br />

genes were first detected at 7 DPA and reached maximum levels at<br />

18 DPA, the last time point sampled. Maximum levels <strong>of</strong> BQ806323<br />

and BQ805639 transcripts were 4.1- and 2.1-fold higher, respectively,<br />

under the high temperature regimen than under moderate<br />

temperatures.<br />

The accumulation <strong>of</strong> globulin-2 transcripts also was assessed in<br />

developing endosperm (Fig. 4E, F). Transcripts for both BQ806323<br />

and BQ805639 were detected at low levels at both 7 and 14 DPA<br />

under moderate temperatures and reached maximum levels at 34<br />

DPA. When high temperatures were applied from anthesis to<br />

maturity, transcripts for both genes were detected at the first time<br />

point sampled at 5 DPA and increased to maximum levels by 20<br />

DPA. Transcript levels at the 20 DPA time point under the high<br />

temperature regimen were 3.9- and 5.3-fold higher for BQ806323<br />

Fig. 3. Comparison <strong>of</strong> the nucleotide sequences <strong>of</strong> BQ806323, BQ839060 and<br />

BQ805639 in regions used for gene-specific primers. QF123 is identical to nucleotides<br />

481–497 <strong>of</strong> BQ806323 and QR123 is the reverse complement <strong>of</strong> nucleotides 573–590<br />

<strong>of</strong> BQ806323. QF126 is identical to nucleotides 532–550 <strong>of</strong> BQ805639 and QR126 is the<br />

reverse complement <strong>of</strong> nucleotides 612–631 <strong>of</strong> BQ805639. Nucleotides that differ in<br />

other ESTs are underlined.<br />

S.B. Altenbach et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 47–54 51<br />

and BQ805639, respectively, than at the 34 DPA time point under<br />

moderate temperatures.<br />

In the experiments shown in Fig. 4, globulin-2 transcript levels<br />

relative to the 18S reference RNA were higher in whole grain than in<br />

endosperm. Fig. 5 shows a direct comparison <strong>of</strong> globulin-2 transcript<br />

levels in whole grain at 30 DPA and in embryo and endosperm<br />

dissected from 30 DPA grain produced under the moderate<br />

temperature regimen. In this analysis, transcripts corresponding to<br />

BQ806323 and BQ805639 were 10.1- and 7.8-fold higher, respectively,<br />

in whole grain than in endosperm. BQ806323 and BQ805639<br />

transcripts were most abundant in embryo, about 2.5- and 2.6-fold<br />

higher, respectively, in embryo than in whole grain. Consistent with<br />

these findings is the fact that a significant proportion <strong>of</strong> the ESTs<br />

that comprise the globulin-2 contigs were derived from an embryo<br />

cDNA library, including 9/22 for contig 18428, 4/11 for contig 17295<br />

and 2/11 for contig 17366. The collection <strong>of</strong> ESTs assembled into<br />

contigs by Chao et al. (2006) included 4147 sequences from endosperm<br />

and 1916 sequences from embryo in addition to the 3639<br />

from Butte 86 whole grain. RNA prepared from awns, glumes,<br />

stems, leaves and roots also was analyzed in the same experiment.<br />

Transcripts corresponding to BQ805639 were detected at very low<br />

levels in roots (Fig. 5B), but were not detectable in awns, glumes,<br />

stems or leaves. BQ806323 was not expressed in any <strong>of</strong> the nongrain<br />

samples (data not shown).<br />

3.4. Globulin-2 proteins in flour produced under moderate<br />

and high temperatures<br />

A proteomic analysis was undertaken to determine the levels <strong>of</strong><br />

globulin-2 proteins in white flour milled from grain produced<br />

under moderate and high temperatures with and without postanthesis<br />

NPK. This analysis revealed a complex collection <strong>of</strong> globulin-2<br />

proteins in the KCl-soluble/MeOH-insoluble fraction<br />

(Fig. 6A–C). Seventeen spots in the 52–55 kDa region <strong>of</strong> the 2-D gels<br />

were identified as globulin-2. Spots 219, 209, 199, 200, 193, 190 and<br />

258 in Fig. 6C correspond to spots 147, 155, 559, 561, 851, 852 and<br />

175 identified in Vensel et al. (2005). MS/MS data used to identify<br />

the remaining spots (192, 202, 205, 221, 222, 228, 236, 239, 320,<br />

and 210) as globulin-2 proteins are summarized in Table S1. One <strong>of</strong><br />

the spots (210) contained more than one protein and was excluded<br />

from further analysis. Collectively, the 16 remaining spots comprise<br />

about 2.9 and 3.1% <strong>of</strong> the total protein in the fraction when grain<br />

was produced under a moderate temperature regimen with or<br />

without post-anthesis NPK, respectively. When grain was produced<br />

under a high temperature regimen from anthesis to maturity, the<br />

16 spots comprised a larger percentage <strong>of</strong> the total protein in the<br />

fraction, 5.2% when plants were supplied with NPK and 5.0% when<br />

plants did not receive post-anthesis NPK. Of the 16 spots, nine<br />

showed statistically significant increases with high temperatures<br />

(P < 0.05) under at least one <strong>of</strong> the fertilizer regimens. Seven spots<br />

increased more than 2-fold in response to high temperature under<br />

both fertilizer regimens (192, 193, 199, 202, 205, 209 and 320)<br />

(Fig. 6C, D). The largest change with high temperature was a 8.3fold<br />

increase detected for spot 199 when grain was produced<br />

without NPK. When grain was produced with NPK, the largest<br />

response was a 4.8-fold increase for spot 320. Four spots (200, 221,<br />

228 and 239) showed small decreases with high temperatures. The<br />

other three spots identified as globulin-2 proteins encompassed<br />

similar proportions <strong>of</strong> the KCl-soluble/MeOH-insoluble flour fraction<br />

under all four environmental regimens (222, 236 and 258).<br />

Five <strong>of</strong> the seven spots that increased more than 2-fold (193, 199,<br />

202, 209 and 320) also increased in flour produced under high<br />

temperatures in a second growth experiment in which plants were<br />

supplied with 0.5 NPK (data not shown).


52<br />

Fig. 4. Accumulation <strong>of</strong> transcripts corresponding to BQ806323 (A, C, E) and BQ805639 (B, D, F) in developing grain (A–D) or endosperm (E, F) produced under a 24/17 C regimen<br />

(solid lines) or a 37/28 C regimen (dashed lines). Plants in A and B did not receive post-anthesis NPK, while plants in C and D received 1 NPK and plants in E and F received 0.5<br />

NPK. Bars indicated the standard error among triplicate reactions.<br />

4. Discussion<br />

In dicots, the 7S globulins are embryo proteins that serve<br />

a storage function and are the primary source <strong>of</strong> amino acids for the<br />

developing seedling. In wheat, the 7S globulins make relatively<br />

small contributions to the storage reserves <strong>of</strong> the grain and the<br />

primary storage proteins are the gliadins and glutenin subunits.<br />

While Halford and Shewry (2007) suggested that globulin proteins<br />

S.B. Altenbach et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 47–54<br />

in cereals have no biological role other than storage, data presented<br />

here suggests that some <strong>of</strong> the wheat globulin-2 proteins may be<br />

involved in the response <strong>of</strong> the developing grain to high temperature<br />

stress. Transcript pr<strong>of</strong>iles demonstrate that two closely-related<br />

globulin-2 genes are expressed in whole grain and endosperm late<br />

in grain development and that levels <strong>of</strong> transcripts increase in<br />

response to high temperatures. Globulin-2 proteins also encompass<br />

a greater proportion <strong>of</strong> the total KCl-soluble/MeOH-insoluble<br />

Fig. 5. Accumulation <strong>of</strong> transcripts corresponding to BQ806323 (A) and BQ805639 (B) in embryo, endosperm and whole grain, all at 30 DPA, and in root. Bars indicate the standard<br />

error among triplicate reactions.


protein in endosperm and flour from grain produced under high<br />

temperature conditions.<br />

These findings are consistent with previous work on the<br />

homologous globulin-2 protein from maize. In an early study, Kriz<br />

and Wallace (1991) reported maize variants lacking globulin-2 and<br />

suggested that this protein was not essential for seed development,<br />

maturation or germination. More recent studies using proteomic<br />

approaches have shown that globulin-2 proteins respond to stress<br />

in maize. Kollipara et al. (2002) found that globulin-2 was among<br />

the proteins up-regulated in desiccation-tolerant maize hybrids<br />

and Labra et al. (2006) noted that a number <strong>of</strong> proteins identified as<br />

globulin-2 were up-regulated in maize plantlets germinated in the<br />

presence <strong>of</strong> potassium dichromate.<br />

Analysis <strong>of</strong> wheat ESTs indicates that the wheat globulin-2<br />

proteins are encoded by a simple gene family consisting <strong>of</strong> three<br />

members. Yet at least 17 protein spots have been identified as<br />

globulin-2 in wheat flour using 2-DE/MS, suggesting that the<br />

proteins undergo post-translational modifications that result in<br />

changes in size and pI. It is likely that the wheat globulin-2 proteins,<br />

like 7S seed storage proteins from legumes, are glycosylated. Robert<br />

et al. (1985) noted charge heterogeneity among proteins isolated<br />

from a 7S þ 3S fraction <strong>of</strong> globulins from wheat grain in the 55 kDa<br />

region <strong>of</strong> 2-D gels and observed that 55 kDa proteins also were<br />

bound to a concanavalin A sepharose column. Additionally, proteins<br />

encoded by contigs 18428, 17295 and 17366 contain putative sites<br />

for N-linked glycosylation, defined as N–X–T, where X can be any<br />

amino acid except P (Fig. 2). Using affinity blotting, Kriz (1989) also<br />

noted that the homologous globulin-2 protein from maize was able<br />

S.B. Altenbach et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 47–54 53<br />

Fig. 6. 2-DE analysis <strong>of</strong> proteins from flour produced under a 24/17 C (A) and a 37/28 C (B) regimen with 1 NPK. Boxes indicate regions <strong>of</strong> each gel that contain spots identified<br />

by MS/MS as globulin-2. The boxed region <strong>of</strong> the gel from panel B is enlarged in panel C and spots identified as globulin-2 are numbered. Numbers that are underlined refer to spots<br />

that increase in relative proportion under high temperature conditions. Panel D compares the average normalized volumes for each globulin-2 spot from plants subjected to the 24/<br />

17 C and 37/28 C regimens, with and without post-anthesis NPK. The solid and open bars denote the 24/17 C regimens with or without NPK, respectively. The hatched and<br />

stippled bars denote the 37/28 C regimen, with or without NPK, respectively. Asterisks denote changes with high temperature under each NPK regimen that are statistically<br />

significant (P < 0.05).<br />

to bind Con A. Some 7S globulins also undergo proteolytic processing<br />

(reviewed by Müntz, 1996). It is interesting that not all <strong>of</strong><br />

the globulin-2 protein spots identified by 2-DE increased in<br />

response to high temperatures. In fact, decreases in the levels <strong>of</strong><br />

several globulin-2 proteins were detected in flour. Thus, both<br />

transcriptional and post-translational mechanisms may be<br />

involved in the response <strong>of</strong> globulin-2 genes to high temperatures<br />

in the developing grain.<br />

Transcripts for globulin-2 were present at lower levels in the<br />

endosperm than in whole grain. Halford and Shewry (2007) noted<br />

that 7S globulins from cereals are expressed predominantly in the<br />

embryo and in the aleurone layer <strong>of</strong> the endosperm. Our analyses<br />

did not distinguish the aleurone tissue, which is a single cell layer in<br />

wheat, from the starchy endosperm tissue. Nonetheless, transcript<br />

levels clearly increased in both the whole grain and the endosperm<br />

samples. Globulin-2 proteins were present in endosperm tissue<br />

extruded from grain from which the embryo had been removed. In<br />

addition, globulin-2 proteins were present in milled white flour<br />

that generally contains minimal amounts <strong>of</strong> germ and bran.<br />

Increases in the relative proportion <strong>of</strong> globulin-2 proteins in<br />

flour produced under high temperatures may have important<br />

implications given the potential allergenicity <strong>of</strong> these proteins. The<br />

KCl-soluble/MeOH-insoluble proteins account for about 11% <strong>of</strong> the<br />

total flour protein (Hurkman and Tanaka, 2004) and globulin-2<br />

proteins comprise as much as 6% <strong>of</strong> this fraction under high<br />

temperature conditions. Wheat is among the top eight foods<br />

responsible for human food allergies. Allergies to wheat are characterized<br />

by a broad spectrum <strong>of</strong> symptoms including urticaria,


54<br />

gastrointestinal distress, asthma and anaphylaxis and are due to<br />

sensitivities to a variety <strong>of</strong> proteins in both the gluten protein and<br />

albumin/globulin fractions (Battais et al., 2005). Since the wheat<br />

globulin-2 shows significant identity to other proteins that are<br />

known or putative allergens, further studies on the allergenicity <strong>of</strong><br />

these proteins are warranted. IgE binding studies using sera from<br />

patients with characterized wheat allergies would be key to<br />

determining whether the globulin-2 proteins should be classified<br />

as food allergens.<br />

An additional question is whether changes in the relative<br />

proportions <strong>of</strong> the globulin-2 proteins might alter the functional<br />

properties <strong>of</strong> the flour. It is generally accepted that the gluten<br />

proteins contribute elasticity and extensibility properties essential<br />

for the formation <strong>of</strong> wheat flour doughs. Interactions between the<br />

gluten protein matrix and other flour components also may affect<br />

rheological properties but have been little studied. Recently, the<br />

globulin-2 proteins were identified in a proteomic study <strong>of</strong> a dough<br />

liquor fraction prepared from wheat flour (Salt et al., 2005). It is<br />

interesting that the levels <strong>of</strong> the globulin-2 proteins in the dough<br />

liquor increased when salt or salt plus ascorbate was included in<br />

the dough mixture. Salt is frequently included in commercial<br />

processes to improve breadmaking properties and has been shown<br />

to influence water absorption, dough development time, dough<br />

stability and loaf volume. It is not known whether the globulin-2<br />

proteins are associated with the gluten matrix in the absence <strong>of</strong><br />

salt. Further study is required to determine whether these proteins<br />

play roles in flour functionality.<br />

Acknowledgements<br />

The authors thank Drs. Frances DuPont and Ann Blechl for<br />

critical review <strong>of</strong> the manuscript. Mention <strong>of</strong> a specific product<br />

does not constitute an endorsement and does not imply a recommendation<br />

over other suitable products.<br />

Appendix. Supplementary data<br />

Supplementary data associated with this article can be found, in<br />

the online version, at doi:10.1016/j.jcs.2008.07.005.<br />

References<br />

Altenbach, S.B., 1998. Quantification <strong>of</strong> individual low-molecular-weight glutenin<br />

subunit transcripts in developing wheat grains by competitive RT-PCR.<br />

Theoretical and Applied Genetics 97, 413–421.<br />

Altenbach, S.B., DuPont, F.M., Kothari, K.M., Chan, R., Johnson, E.L., Lieu, D., 2003.<br />

Temperature, water and fertilizer influence the timing <strong>of</strong> key events during<br />

grain development in a US spring wheat. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 37, 9–20.<br />

Altenbach, S.B., Kothari, K.M., 2004. Transcript pr<strong>of</strong>iles <strong>of</strong> genes expressed in<br />

endosperm tissue are altered by high temperature during wheat grain<br />

development. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 40, 115–126.<br />

S.B. Altenbach et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 47–54<br />

Altenbach, S.B., Kothari, K.M., Tanaka, C.K., Hurkman, W.J., 2007. Genes encoding the<br />

PR-4 protein wheatwin are developmentally regulated in wheat grains and<br />

respond to high temperatures during grainfill. Plant <strong>Science</strong> 173, 135–143.<br />

Altenbach, S.B., Kothari, K.M., Tanaka, C.K., Hurkman, W.J., 2008. Expression <strong>of</strong> 9kDa<br />

non-specific lipid transfer protein genes in developing wheat grains is<br />

enhanced by high temperatures but not by post-anthesis fertilizer. <strong>Journal</strong> <strong>of</strong><br />

<strong>Cereal</strong> <strong>Science</strong> 47, 201–213.<br />

Battais, F., Courcoux, P., Popineau, Y., Kanny, G., Moneret-Vautrin, D.A., Denery-<br />

Papini, S., 2005. Food allergy to wheat: differences in immunoglobin Ebinding<br />

proteins as a function <strong>of</strong> age or symptoms. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

42, 109–117.<br />

Bendtsen, J.D., Nielsen, H., von Heijne, G., Brunak, S., 2004. Improved prediction <strong>of</strong><br />

signal peptides: SignalP 3.0. <strong>Journal</strong> <strong>of</strong> Molecular Biology 340, 783–795.<br />

Chao, S., Lazo, G.R., You, F., Crossman, C.C., Hummel, D.D., Lui, N., Laudencia-<br />

Chingcuanco, D., Anderson, J.A., Close, T.J., Dubcovsky, J., Gill, B.S., Gill, K.S.,<br />

Gustafson, J.P., Kianian, S.F., Lapitan, N.L.V., Nguyen, H.T., Sorrells, M.E.,<br />

McGuire, P.E., Qualset, C.O., Anderson, O.D., 2006. Use <strong>of</strong> a large-scale Triticeae<br />

expressed sequence tag resource to reveal gene expression pr<strong>of</strong>iles in hexaploid<br />

wheat (Triticum aestivum L.). Genome 49, 531–544.<br />

Craig, R., Beavis, R.C., 2004. TANDEM: matching proteins with mass spectra.<br />

Bioinformatics 20, 1466–1467.<br />

Goodman, R.E., 2006. Practical and predictive bioinformatics methods for the<br />

identification <strong>of</strong> potentially cross-reactive protein matches. Molecular Nutrition<br />

and Food Research 50, 655–660.<br />

Halford, N.G., Shewry, P.R., 2007. The structure and expression <strong>of</strong> cereal storage<br />

protein genes. Plant Cell Monographs 8, 195–218.<br />

Hurkman, W.J., Tanaka, C.K., 2004. Improved methods for separation <strong>of</strong> wheat<br />

endosperm proteins and analysis by two-dimensional gel electrophoresis.<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 40, 295–299.<br />

Hurkman, W.J., Vensel, W.H., Tanaka, C.K., Whitehand, L., Altenbach, S.B., 2009.<br />

Effect <strong>of</strong> high temperature on albumin and globulin accumulation in the<br />

endosperm proteome <strong>of</strong> the developing wheat grain. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

49, 12–23.<br />

Kollipara, K.P., Saab, I.N., Wych, R.D., Lauer, M.J., Singletary, G.W., 2002. Expression<br />

pr<strong>of</strong>iling <strong>of</strong> reciprocal maize hybrids divergent for cold germination and<br />

desiccation tolerance. Plant Physiology 129, 974–992.<br />

Kriz, A.L., 1989. Characterization <strong>of</strong> embryo globulins encoded by the maize Glb<br />

genes. Biochemical Genetics 27, 239–251.<br />

Kriz, A.L., Wallace, N.H., 1991. Characterization <strong>of</strong> the maize globulin-2 gene and<br />

analysis <strong>of</strong> two null alleles. Biochemical Genetics 29, 241–254.<br />

Labra, M., Gianazza, E., Waitt, R., Eberini, I., Sozzi, A., Regondi, S., Grassi, F., Agradi, E.,<br />

2006. Zea mays L. protein changes in response to potassium dichromate<br />

treatments. Chemosphere 62, 1234–1244.<br />

Mills, E.N.C., Jenkins, J., Marigheto, N., Belton, P.S., Gunning, A.P., Morris, V.J., 2002.<br />

Allergens <strong>of</strong> the cupin superfamily. Biochemical Society Transactions 30,<br />

925–929.<br />

Müntz, K., 1996. Proteases and proteolytic cleavage <strong>of</strong> storage proteins in developing<br />

and germinating dicotyledonous seed. <strong>Journal</strong> <strong>of</strong> Experimental Botany 47,<br />

605–622.<br />

Muller, P.Y., Janovjak, H., Miserez, A.R., Dobbie, Z., 2002. Processing <strong>of</strong> gene<br />

expression data generated by quantitative real-time RT-PCR. BioTechnology 32,<br />

1372–1379.<br />

Robert, L.S., Adeli, K., Altosaar, I., 1985. Homology among 3S and 7S globulins from<br />

cereals and pea. Plant Physiology 78, 812–816.<br />

Salt, L.J., Robertson, J.A., Jenkins, J.A., Mulholland, F., Mills, E.N.C., 2005. The identification<br />

<strong>of</strong> foam-forming soluble proteins from wheat (Triticum aestivum)<br />

dough. Proteomics 5, 1612–1623.<br />

Vensel, W.H., Harden, L., Tanaka, C.K., Hurkman, W.J., Haddon, W.F., 2002. Identification<br />

<strong>of</strong> wheat endosperm proteins by MALDI mass spectrometry and LC-MS/<br />

MS. <strong>Journal</strong> <strong>of</strong> Biomolecular Techniques 13, 95–100.<br />

Vensel, W.H., Tanaka, C.K., Cai, N., Wong, J.H., Buchanan, B.B., Hurkman, W.J., 2005.<br />

Developmental changes in the metabolic protein pr<strong>of</strong>iles <strong>of</strong> wheat endosperm.<br />

Proteomics 5, 1594–1611.<br />

Wallace, N.H., Kriz, A.L., 1991. Nucleotide sequence <strong>of</strong> a cDNA clone corresponding<br />

to the maize globulin-2 gene. Plant Physiology 95, 973–975.


Biochemical markers: Efficient tools for the assessment <strong>of</strong> wheat grain tissue<br />

proportions in milling fractions<br />

Youna Hemery a , Valérie Lullien-Pellerin a , Xavier Rouau a , Joël Abecassis a , Marie-Françoise Samson a ,<br />

Per Åman b , Walter von Reding c ,Cäcilia Spoerndli c ,Cécile Barron a, *<br />

a<br />

INRA, UMR 1208 ‘‘Agropolymers Engineering and Emerging Technologies’’, INRA-CIRAD-UMII-Supagro, F-34000 Montpellier, France<br />

b<br />

Department <strong>of</strong> Food <strong>Science</strong>, Swedish University <strong>of</strong> Agriculture <strong>Science</strong> (SLU), Box 7051, S-750 07 Uppsala, Sweden<br />

c<br />

Bühler AG, CH-9240 Uzwil, Switzerland<br />

article info<br />

Article history:<br />

Received 10 April 2008<br />

Received in revised form 3 July 2008<br />

Accepted 8 July 2008<br />

Keywords:<br />

Wheat<br />

Grain<br />

Bran<br />

Aleurone<br />

Outer layers<br />

Fractionation<br />

Milling<br />

Processes<br />

1. Introduction<br />

abstract<br />

Numerous epidemiological studies have demonstrated the<br />

health benefits <strong>of</strong> consuming more whole-grain foods (Jacobs and<br />

Steffen, 2003; Liu, 2007). However, all the wheat grain parts are not<br />

health-promoting, for example the outermost parts have been<br />

Abbreviations: ARs, alkylresorcinols; Cr, crousty cultivar; FAt, 4-O-8 0 , 5 0 -5 00<br />

dehydrotriferulic acid (ferulic acid trimer); P2O5, phosphorus pentaoxyde; p-CA,<br />

para-coumaric acid; Tg, Tiger cultivar; WGA, wheat germ agglutinin.<br />

* Corresponding author. Tel.: þ33 4 99 61 31 04; fax: þ33 4 99 61 30 76.<br />

E-mail address: barron@supagro.inra.fr (C. Barron).<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.07.006<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 55–64<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

To produce safe and healthy whole wheat food products, various grain or bran dry fractionation<br />

processes have been developed recently. In order to control the quality <strong>of</strong> the products and to adapt these<br />

processes, it is important to be able to monitor the grain tissue proportions in the different milling<br />

fractions produced. Accordingly, a quantitative method based on biochemical markers has been developed<br />

for the assessment <strong>of</strong> grain tissue proportions in grain fractions. Grain tissues that were quantified<br />

were the outer pericarp, an intermediate layer (including the outer pericarp, the testa and the hyaline<br />

layer), the aleurone cell walls, the aleurone cell contents, the endosperm and the germ, for two grain<br />

cultivars (Tiger and Crousty). Grain tissues were dissected by hand and analysed. Biochemical markers<br />

chosen were ferulic acid trimer, alkylresorcinols, para-coumaric acid, phytic acid, starch and wheat germ<br />

agglutinin, for outer pericarp, intermediate layer, aleurone cell walls, aleurone cell contents, endosperm<br />

and germ respectively. The results <strong>of</strong> tissue quantification by hand dissection and by calculation were<br />

compared and the sensitivity <strong>of</strong> the method was regarded as good (mean relative errors <strong>of</strong> 4% and 8% for<br />

Crousty and Tiger outer layers respectively). The impact <strong>of</strong> the analytical variability (maximum 13%<br />

relative error on coarse bran) was also regarded as acceptable. Wheat germ agglutinin seems to be<br />

a promising marker <strong>of</strong> wheat germ: even if the quantification method was not able to quantify the germ<br />

proportions in milling fractions, it was able to classify these fractions according to their germ content.<br />

The efficiency <strong>of</strong> this method was tested, by assessing the grain tissue proportions <strong>of</strong> fractions exhibiting<br />

very different compositions such as flour, bran and aleurone-rich fractions obtained from three different<br />

grain or bran dry fractionation processes (conventional milling, debranning process, production <strong>of</strong><br />

aleurone-rich fractions from coarse bran). By calculation <strong>of</strong> the composition <strong>of</strong> the different products<br />

generated, it was possible to study the distribution <strong>of</strong> the different tissues among fractions resulting from<br />

the different fractionation processes. This quantitative method is thus a useful tool for the monitoring<br />

and improvement <strong>of</strong> processes, and allows the effects <strong>of</strong> these processes to be understood and their<br />

adaption to reach the objectives.<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

shown to concentrate the majority <strong>of</strong> the grain contaminants, like<br />

microorganisms, mycotoxins, pesticide residues and heavy metals<br />

(Aureli and D’Egidio, 2007; Fleurat-Lessard et al., 2007; Laca et al.,<br />

2006). On the other hand, the wheat aleurone layer has been shown<br />

to have great nutritional interest, and to concentrate most <strong>of</strong> the<br />

minerals and vitamins <strong>of</strong> the wheat grain (Pomeranz, 1988). Buri<br />

et al. (2004) reported that wheat aleurone layer contains interesting<br />

proportions <strong>of</strong> proteins, b-glucans, phenolic compounds and<br />

other phytochemicals (lignans, sterols). Zhou et al. (2004) pointed<br />

out that antioxidant, including phenolic acids, are concentrated in<br />

the aleurone layer <strong>of</strong> wheat bran, and Mateo Anson et al. (2008)<br />

showed that the higher the proportion <strong>of</strong> aleurone material in<br />

wheat fractions, the higher the antioxidant capacity observed for


56<br />

these fractions. Amrein et al. (2003) found that aleurone-rich<br />

fractions exhibited better in vitro digestibility and colonic fermentability<br />

than wheat bran, and Bach Knudsen et al. (1995)<br />

observed that the digestibility <strong>of</strong> minerals, protein and non-starch<br />

polysaccharides is much higher in bran fractions rich in aleurone<br />

than in fractions rich in pericarp and testa. These results suggest<br />

that it could be interesting to produce aleurone-rich fractions for<br />

use as food ingredients. As a consequence, new processes are<br />

developed in order to exploit all the nutritional benefits <strong>of</strong> whole<br />

grain and to produce new wheat foods and wheat-based ingredients<br />

with enhanced nutritional quality (Hemery et al., 2007). For<br />

example, depending on the desired product, a process can aim at<br />

discarding the pericarp to obtain whole grains containing less<br />

contaminants, while other processes may be developed in order to<br />

produce bran fractions highly concentrated in aleurone material<br />

(Buri et al., 2004). It is <strong>of</strong>ten difficult to exactly monitor the<br />

distribution <strong>of</strong> the different grain tissues among fractions during<br />

processing, as no simple method exists to quantify the respective<br />

proportions <strong>of</strong> these tissues in fractions. However, the monitoring<br />

<strong>of</strong> tissue proportions in the different fractions is essential, as it<br />

allows to control the quality <strong>of</strong> the products and to consequently<br />

adapt the processes. Therefore quantitative tools are needed.<br />

Different compounds can be measured to evaluate bran<br />

contamination in wheat flours and fractions. Ash content and<br />

measurement <strong>of</strong> flour colour are widely used in the milling industry<br />

as indicators <strong>of</strong> flour purity. The amino acid composition <strong>of</strong> the<br />

various tissues was also studied but did not allow the quantification<br />

<strong>of</strong> the grain tissues in milling fractions (Jensen and Martens, 1983).<br />

Some authors (Lempereur et al., 1998; Pussayanawin et al., 1988)<br />

suggested using the concentration <strong>of</strong> ferulic acid to quantify bran in<br />

flours and semolinas, and alkylresorcinols have more recently been<br />

shown to be good markers <strong>of</strong> wheat bran content in foods (Chen<br />

et al., 2004). Such analyses may be useful to evaluate the total outer<br />

layer content but they do not allow to distinguish between the<br />

different outer tissues. Another way to evaluate the different grain<br />

tissues in wheat fractions consists <strong>of</strong> the use <strong>of</strong> specific fluorescence<br />

properties <strong>of</strong> the outer layers. Indeed, the aleurone cell walls<br />

display blue fluorescence under UV-light, due to the presence <strong>of</strong><br />

ferulic acid, whereas the pericarp shows green fluorescence under<br />

blue light (Fulcher et al., 1972; Jensen et al., 1982; Symons and<br />

Dexter, 1996). Based on these fluorescence properties, commercial<br />

equipment has been developed to determine the amount <strong>of</strong> aleurone<br />

in flours. Multispectral fluorescence image analysis <strong>of</strong> grain<br />

sections coupled with classification techniques have also been<br />

developed to more precisely quantify the proportions <strong>of</strong> the<br />

different parts <strong>of</strong> the grain (Baldwin et al., 1997; Courcoux et al.,<br />

2002), but have not yet been applied to powdery samples. These<br />

imaging methods can be good tools for on-line use, but their main<br />

disadvantage would be their lack <strong>of</strong> specificity, as they do not allow<br />

to quantify other tissues than aleurone and pericarp. Moreover, all<br />

these methods allow determination <strong>of</strong> bran proportions in flours<br />

during milling, but they may not be adaptable to other fractionation<br />

systems (such as progressive abrasion or bran fractionation).<br />

Peyron et al. (2002) and Antoine et al. (2004) carried out<br />

biochemical analyses <strong>of</strong> isolated wheat grain tissues and used the<br />

differences in chemical composition between these tissues to<br />

assess the histological composition <strong>of</strong> technological fractions. They<br />

used compounds such as phenolic acids, phytic acid, and starch as<br />

biochemical markers. Indeed, these compounds were shown to be<br />

either present exclusively in one part <strong>of</strong> the grain (starch in starchy<br />

endosperm), or present in greater amounts in one particular tissue<br />

(phytic acid in aleurone cell contents and some phenolic acids in<br />

the cell walls). These biochemical markers were used to determine<br />

the amounts <strong>of</strong> aleurone layer and pericarp in flours and other<br />

milling fractions (Peyron et al., 2002), and in the samples obtained<br />

from a wheat bran fractionation process (Antoine et al., 2004).<br />

Y. Hemery et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 55–64<br />

Antoine et al. (2004) used different markers for aleurone cell walls<br />

and aleurone cell contents to assess the histological composition <strong>of</strong><br />

bran fractions and to evaluate the dissociation and the accessibility<br />

<strong>of</strong> aleurone cellular components. This method was reported to<br />

provide accurate quantification <strong>of</strong> the histological composition <strong>of</strong><br />

samples and to be versatile, as it can be refined and adapted<br />

depending on the type <strong>of</strong> sample analysed, from either bran or<br />

whole-grain fractionation. However, it did not allow the quantification<br />

<strong>of</strong> testa (this tissue was deduced by subtraction and thus was<br />

perhaps overestimated), and it neither allowed the detection <strong>of</strong> the<br />

germ. Having a marker for wheat germ would nevertheless be very<br />

useful as it is either a part <strong>of</strong> the grain that needs to be excluded to<br />

avoid lipid oxidation and rancidity, or a nutritionally interesting byproduct<br />

that could be followed during fractionation processes in<br />

order to get germ-rich fractions.<br />

The aim <strong>of</strong> this study was first to improve the accuracy <strong>of</strong> the<br />

biochemical markers method by introducing new markers <strong>of</strong> testa<br />

and germ and by developing a better quantification <strong>of</strong> aleurone cell<br />

walls and cell contents, to evaluate the sensitivity <strong>of</strong> the method,<br />

and then to validate its use as a tool to assess the grain tissue<br />

proportions in technological fractions exhibiting very distinct<br />

compositions, obtained by different grain or bran fractionation<br />

processes.<br />

2. Materials and methods<br />

2.1. Preparation <strong>of</strong> grain tissues and fractions<br />

2.1.1. Wheat samples<br />

Two common wheat (Triticum aestivum L.) cultivars differing in<br />

kernel hardness (hard wheat: cv. Tiger, and s<strong>of</strong>t wheat: cv. Crousty),<br />

harvested in 2005 in Germany (Tiger) and France (Crousty) were<br />

used in this study.<br />

2.1.2. Isolation <strong>of</strong> wheat grain tissues for biochemical analyses<br />

The structure and composition <strong>of</strong> the different wheat grain<br />

tissues have already been described (Barron et al., 2007; Hemery<br />

et al., 2007; Surget and Barron, 2005). To obtain isolated grain<br />

tissues, the grains ends (germ and brush) were removed with<br />

a razor blade and the remaining parts were immersed in distilled<br />

water for 12–16 h at 20 C. A crease incision was made and the<br />

endosperm was removed using a scalpel. The crease area was<br />

removed and three different strips comprising different tissues<br />

were separated from these ‘‘whole outer layers’’ using a scalpel.<br />

Antoine et al. (2003) showed that the outer strip corresponds to<br />

the outer pericarp (epidermis and hypodermis), the inner strip<br />

corresponds to the aleurone layer, and the intermediate layer is<br />

a composite <strong>of</strong> several tissues (inner pericarp, testa and nucellar<br />

tissue). Scutellum and embryonic axis were also dissected. Nonadhering<br />

outer tissues surrounding the embryonic axis were<br />

removed and then the embryonic axis was separated from dry<br />

grains with a needle. The scutellum was removed with a scalpel<br />

after soaking the grain in distilled water for 12–16 h at 20 C.<br />

These hand-isolated tissues were then used for the biochemical<br />

analyses.<br />

2.1.3. Determination <strong>of</strong> the relative amount <strong>of</strong> tissues within the<br />

grain<br />

Similar hand dissections <strong>of</strong> 10 grains were performed in triplicate<br />

for each cultivar, to quantify the relative amounts <strong>of</strong> the<br />

different tissues, as described by Barron et al. (2007). All recovered<br />

tissues were weighed after drying at 25 C over phosphorus pentaoxide<br />

(P2O5). Outer layer tissues proportions were deduced from<br />

the combined weights <strong>of</strong> the aleurone, intermediate layer and<br />

outer pericarp, dissected from whole outer layers without the<br />

crease.


2.1.4. Production <strong>of</strong> the conventional milling fractions<br />

White flour (0.55% ash content), whole meal flour, fine bran and<br />

coarse bran fractions were produced using a conventional milling<br />

process, carried out on a Test-mill in the Department <strong>of</strong> Safety &<br />

Quality <strong>of</strong> <strong>Cereal</strong>s, Federal Research Centre for Nutrition and Food<br />

(BFEL), Germany.<br />

2.1.5. Production <strong>of</strong> bran and flour fractions by peeling, pearling<br />

and milling<br />

Eight different bran and flour fractions were provided by Bühler<br />

A.G. Uzwil, Switzerland. These fractions were obtained by two<br />

different debranning processes before milling (Fig. 1), by friction<br />

(peeling) and by abrasion (pearling), according to the published<br />

Bühler A.G. patent applications (Eugster and Gerschwiler, 2006).<br />

Peeling fractions were produced using a Bühler DC Peeler taking <strong>of</strong>f<br />

3–3.5% <strong>of</strong> the wheat kernel. Peeled wheat kernels were used as raw<br />

material for the pearling processing, using a Bühler stone polishing<br />

machine taking <strong>of</strong>f approximately 3% more <strong>of</strong> the grain, resulting in<br />

pearled kernels and the pearling fractions. Peeled and pearled<br />

wheat kernels were tempered for 12–15 h to a moisture content <strong>of</strong><br />

15.5%, and then milled on a Bühler laboratory mill to 76% extraction<br />

rate flour (based on whole kernels before peeling and pearling) to<br />

give the 76% flours after peeling and after pearling, and the bran<br />

fractions after peeling and after pearling. For the production <strong>of</strong><br />

reconstituted ‘whole-grain’ flour (100% extraction <strong>of</strong> the remaining<br />

kernel) from peeled and pearled kernels, coarse bran as well as<br />

bottom flour were reduced to particle sizes 180 mm and Aleurone 1 180 mm<br />

and Aleurone 2 < 180 mm) using electrostatic separation. The Byproduct<br />

1 and By-product 2 fractions correspond to the by-products<br />

<strong>of</strong> the Aleurone 1 and Aleurone 2 fractions, respectively.<br />

Peeling fraction<br />

Milling<br />

100% flour<br />

after peeling<br />

Wheat grains<br />

Peeling (friction)<br />

Y. Hemery et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 55–64 57<br />

Grains after peeling<br />

76% flour<br />

after peeling<br />

Milling<br />

Bran after peeling<br />

2.2. Biochemical analyses<br />

Dissected tissues were dried at 25 C over P2O5, and then ground<br />

under cryogenic conditions for 4 min (with a Spex CertiPrep 6750<br />

lab impact grinder), and dried again before chemical analyses.<br />

Milling fractions (except flours) were ground in a ball mill (Dangoumeau,<br />

France) for 4 min.<br />

2.2.1. Isolation <strong>of</strong> aleurone cell wall material<br />

The amount <strong>of</strong> cell walls in Tiger and Crousty hand-isolated<br />

aleurone layers was assessed by gravimetric determination <strong>of</strong><br />

insoluble cell wall material after proteolysis, using an adaptation <strong>of</strong><br />

the method described by Brillouet et al. (1988). The dried and<br />

ground sample was suspended in hexane and centrifuged, the<br />

supernatant was discarded, these steps were repeated once, and<br />

the final pellet was air dried. The dried pellet was suspended in<br />

water with 1.5% SDS and 5 mM Na-bisulfite, a pronase solution was<br />

added (0.5 ml, 1 mg pronase/ml water), the mixture was agitated<br />

for 1 h at room temperature and then centrifuged. The pellet was<br />

washed extensively with water with intermittent centrifugation,<br />

and then successively suspended in 80% ethanol, absolute ethanol,<br />

acetone and ether, with centrifugation between two successive<br />

suspensions. The final pellet corresponding to the aleurone cell<br />

wall material was dried at 25 C over P2O5 for 72 h, and weighed<br />

with 0.1 mg accuracy. This protocol was only used to evaluate the<br />

cell wall content in aleurone layers, and not for biochemical<br />

analyses.<br />

2.2.2. Starch<br />

Total starch content <strong>of</strong> fractions was measured using Megazyme<br />

kits (Megazyme International Ireland Ltd., Ireland) according to<br />

approved AACC method 76-13 (AACC, 2000). Samples were analysed<br />

in duplicate with c.v.


58<br />

2.2.4. Phenolic acids<br />

Ester-linked phenolic acids were saponified under Argon<br />

(oxygen-free) at 35 C in 2 N sodium hydroxide. An internal standard<br />

(2,3,5-trimethoxy-(E)-cinnamic acid (TMCA), T-4002, Sigma<br />

Chemical Co., St Louis, USA) was added before adjusting pH to 2.<br />

Phenolic acids were then extracted with diethylether and quantified<br />

by RP-HPLC as described by Antoine et al. (2003). The response<br />

factors <strong>of</strong> the para-coumaric acid (p-CA) and the 4-O-8 0 , 5 0 -5 00<br />

dehydrotriferulic acid (ferulic acid trimer, FAt) relative to the<br />

internal standard were determined at 320 nm with pure<br />

compounds. All analyses were performed at least in duplicate, with<br />

c.v. 12 times more concentrated in the aleurone layer than in the<br />

outer pericarp, intermediate layer and germ (including scutellum).<br />

Thus, as p-CA is known to be bound to cell wall polysaccharides<br />

(Fincher and Stone, 1986; Grabber et al., 2004; Rhodes et al., 2002),<br />

this compound can be used as a marker <strong>of</strong> aleurone cell walls.<br />

However, as the amounts <strong>of</strong> p-CA found in the outer pericarp and<br />

the intermediate layer are not negligible, these amounts have to be<br />

taken into account and subtracted from the total p-CA content <strong>of</strong><br />

a fraction in order not to bias the calculation <strong>of</strong> the proportion <strong>of</strong><br />

aleurone cell walls in this fraction.<br />

Antoine et al. (2004) assumed that the sum <strong>of</strong> the outer pericarp,<br />

intermediate layer, aleurone and endosperm proportions in<br />

the studied fractions was equal to 100% (i.e. they supposed that the<br />

bran fractions did not contain germ), and then deduced the intermediate<br />

layer proportion by subtraction. If fractions containing germ<br />

were analysed using this method, the results may be biased and the<br />

intermediate layer proportion would be overestimated. A specific<br />

marker is thus needed to allow an accurate quantification <strong>of</strong> the<br />

intermediate layer proportion in fractions. Landberg et al. (2008)<br />

Table 1<br />

Composition <strong>of</strong> hand-isolated outer layers and peripheral tissues (reference values)<br />

Tissues Cultivar FAt<br />

(mg/g)<br />

Total ARs<br />

(mg/g)<br />

p-CA<br />

(mg/g)<br />

Phytic acid<br />

(mg/g)<br />

Maximum c.v. 10% 10% 6% 9% 6%<br />

Outer pericarp Tiger 1.35 0.08 0.01 – –<br />

Crousty 1.15 0.10 0.02 – –<br />

Intermediate<br />

layer a<br />

Tiger 0.06 16.4 0.11 – –<br />

Crousty 0.04 16.2 0.09 – –<br />

Aleurone layer Tiger 0.12 0.03 0.29 152.3 –<br />

Crousty 0.10 nd 0.22 156.3 –<br />

Starch<br />

(% w/w)<br />

Endosperm Tiger nd 0.02 nd – 78.3<br />

Crousty nd 0.01 nd – 78.1<br />

Total outer<br />

layers<br />

Tiger 0.34 4.07 0.16 70.7 –<br />

Crousty 0.28 3.93 0.13 78.4 –<br />

Whole grains Tiger 0.04 0.47 0.03 13.7 644.4<br />

Crousty 0.03 0.42 0.02 11.8 656.5<br />

nd: non-detected; FAt: ferulic acid dehydrotrimer; ARs: alkylresorcinols; p-CA: paracoumaric<br />

acid.<br />

a Composed <strong>of</strong> the hyaline layer þ testa þ inner pericarp.


studied the location <strong>of</strong> alkylresorcinols in cereal grains and found<br />

that more than 99% <strong>of</strong> wheat grain total ARs are concentrated in<br />

the hand isolated so-called intermediate layer, and more precisely<br />

in the cuticle <strong>of</strong> the testa, showing that these phenolic lipids could<br />

be used as specific biochemical markers. The total amount <strong>of</strong><br />

ARs was therefore chosen as a new marker <strong>of</strong> intermediate layer.<br />

Wheat germ agglutinin is a lectin found exclusively in wheat germ<br />

(Miller and Bowles, 1982). It was therefore selected as a potential<br />

marker <strong>of</strong> wheat germ in milling fractions and was used as a qualitative<br />

indicator to classify milling fractions according to their germ<br />

content.<br />

3.1.2. Quantification <strong>of</strong> tissue proportions and validation by<br />

comparison with hand-dissection<br />

After selection <strong>of</strong> the biochemical markers and assessment<br />

<strong>of</strong> the biochemical composition <strong>of</strong> isolated tissues and fractions,<br />

the grain tissue proportions can be estimated. By combining the<br />

biochemical marker contents <strong>of</strong> hand-isolated grain tissues and<br />

the marker contents <strong>of</strong> the different technological fractions, the<br />

following relations allow calculation <strong>of</strong> the grain tissue proportions<br />

within fractions.<br />

%p ¼½FAtŠF =½FAtŠp 100<br />

%i ¼½ARsŠF =½ARsŠi 100<br />

%aw ¼ ½p-CAŠF %p ½p-CAŠp %i ½p-CAŠi %ac ¼ ½Phytic acidŠF =½Phytic acidŠac 100<br />

%e ¼½StarchŠF =½StarchŠe 100<br />

with:<br />

½p-CAŠ aw ¼½p-CAŠ a =½awŠ a<br />

½Phytic acidŠ ac ¼½Phytic acidŠ a =½acŠ a<br />

.<br />

½p-CAŠ aw 100<br />

where:<br />

(%p), (%i), (%ac), (%aw), and (%e): proportion <strong>of</strong> outer pericarp,<br />

intermediate layer, aleurone cell content, aleurone cell walls, and<br />

starchy endosperm in studied fractions.<br />

[FAt]: ferulic acid dehydrotrimer content; [ARs]: total alkylresorcinols<br />

content; [p-CA]: para-coumaric acid content; [phytic<br />

acid]: phytic acid content; [starch]: starch content, in studied<br />

fractions (F), outer pericarp (p), intermediate layer (i), total aleurone<br />

(a), aleurone cell walls (aw), aleurone cell content (ac), and<br />

starchy endosperm (e).<br />

[aw]a and [ac]a: concentration <strong>of</strong> cell walls and cell contents in<br />

the aleurone layer.<br />

The approximate proportions <strong>of</strong> germ in fractions (%g) was<br />

calculated as follows:<br />

%g ¼½WGAŠ F =½WGAŠ grain<br />

½germŠ grain<br />

where [WGA]F and [WGA]grain are the WGA content in fractions and<br />

ground whole grain respectively, and [germ]grain is the proportion<br />

<strong>of</strong> germ (scutellum plus embryonic axis) obtained by hand<br />

dissection <strong>of</strong> grains.<br />

In order to separately quantify the proportions <strong>of</strong> aleurone cell<br />

walls and cell contents in the different fractions, the results have to<br />

be calculated in relation to the amount <strong>of</strong> markers found in aleurone<br />

cell walls and aleurone cell contents (expressed in mgp-CA/<br />

galeurone cell walls and in mgphytates/galeurone cell contents), and not in<br />

relation to the amount <strong>of</strong> markers found in the whole aleurone<br />

layer (expressed in mgmarker/galeurone), otherwise these proportions<br />

would be over-estimated. Thus, the amount <strong>of</strong> aleurone cell walls<br />

was assessed by gravimetric determination <strong>of</strong> insoluble cell wall<br />

material. Crousty and Tiger aleurone layers were found to contain<br />

42.4% 2.5% and 48.0% 2.6% cell walls, respectively. This high<br />

proportion <strong>of</strong> cell walls in the aleurone layer corresponds to<br />

Y. Hemery et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 55–64 59<br />

microscopy observations carried out in previous studies (Antoine<br />

et al., 2003; Bacic and Stone, 1981; Evers and Reed, 1988; Rhodes<br />

et al., 2002), that showed thick aleurone cell walls accounting for an<br />

important proportion <strong>of</strong> aleurone cell total volume.<br />

To test the efficiency <strong>of</strong> the described tissue proportion calculation<br />

method, the composition <strong>of</strong> whole grains and whole outer<br />

layers was calculated and compared to the composition assessed by<br />

hand dissection and weighing. The comparison <strong>of</strong> the results<br />

obtained for Crousty and Tiger grains by both methods (markers<br />

and dissection) is shown in Table 2. Almost no difference was<br />

found between the endosperm amounts, and the total outer layer<br />

proportions obtained by both methods were very close (0.6% and<br />

0.1% absolute error for Crousty and Tiger grains respectively). The<br />

proportions <strong>of</strong> peripheral tissues were found to be a little more<br />

variable (from 0.4% to 1.4% absolute error) depending on the<br />

method used, but these variations in peripheral tissue proportions<br />

are due to the fact that these tissues are present in small amounts in<br />

the ground grains analysed, and that biochemical analyses are less<br />

precise on small amounts. These results show that the biochemical<br />

markers method allows the distinct quantification <strong>of</strong> tissues<br />

proportions in wheat fractions that contain very few peripheral<br />

layers, with an absolute error <strong>of</strong> 1.5%. About 2.5% <strong>of</strong> the wheat<br />

grain was not quantified with this method. This proportion is<br />

comparable to the proportion <strong>of</strong> germ found in wheat grain by<br />

hand-dissection. Fig. 2 presents the tissue proportions obtained for<br />

Crousty and Tiger whole outer layers. The tissues proportions<br />

calculated using the biochemical markers are very close to those<br />

assessed by hand dissection. The mean relative errors are 4% and 8%<br />

for Crousty and Tiger respectively, and the total <strong>of</strong> the tissue<br />

proportions are 98% and 96%. These results show that this method<br />

allows effective quantification, and suggest that it may be very well<br />

adapted for the assessment <strong>of</strong> grain tissue proportions in fractions<br />

rich in bran tissues.<br />

3.1.3. Impact <strong>of</strong> analytical variability and possible simplification <strong>of</strong><br />

the method<br />

This method is based on the measurement <strong>of</strong> 5 different<br />

compounds (FAt, ARs, p-CA, phytic acid and starch) on two sample<br />

sets (i.e. the hand-isolated reference tissues and the technological<br />

fractions), it implies therefore 10 quantitative determinations. The<br />

random analytical error on the tissue proportions was calculated<br />

for Tiger and Crousty coarse bran, by taking into account the<br />

reproducibility (i.e. the mean c.v. <strong>of</strong> the dosage) <strong>of</strong> each measurement.<br />

Coarse bran was chosen for the calculation <strong>of</strong> variability<br />

because this fraction contains all <strong>of</strong> the grain tissues. If these<br />

calculations had been carried out for other samples, the results<br />

could have been different as the random analytical error depends<br />

on the composition <strong>of</strong> the studied fraction, due to the fact that the<br />

reproducibility varies from one dosage to another. The effects on<br />

Table 2<br />

Comparison <strong>of</strong> Crousty and Tiger grains composition (%), using either biochemical<br />

markers or hand-dissection<br />

Grains tissues Crousty Tiger<br />

Markers Dissection Markers Dissection<br />

Total outer layers 13.7 13.1 14.7 14.6<br />

Outer pericarp 2.7 3.5 3.1 4.3<br />

Intermediate layer 2.6 3.2 2.9 3.3<br />

Aleurone cell walls 4.1 2.7 4.1 3.4<br />

Aleurone cell contents 4.4 3.7 4.7 3.6<br />

Starchy endosperm 84.1 84.1 82.4 82.7<br />

Embryo þ scutellum – 2.8 – 2.7<br />

Total <strong>of</strong> grain tissues 97.8 100 97.1 100<br />

Unknown fraction 2.2 – 2.9 –<br />

The ‘‘unknown fractions’’ is the difference between 100% and the sum <strong>of</strong> all the grain<br />

tissues proportions.


60<br />

100%<br />

0%<br />

98%<br />

29% 28% 24%<br />

21%<br />

24%<br />

100% 96%<br />

100%<br />

25%<br />

21% 22% 23%<br />

24% 25%<br />

24% 27% 25% 30%<br />

Markers Dissection Markers Dissection<br />

CROUSTY TIGER<br />

outer pericarp intermediate layer<br />

aleurone cell walls aleurone cell contents<br />

22%<br />

Fig. 2. Comparison <strong>of</strong> Crousty and Tiger outer layers composition (%), using either<br />

biochemical markers or hand-dissection.<br />

under-estimates or over-estimates <strong>of</strong> the biochemical markers<br />

content in reference hand-isolated tissues and/or in coarse bran<br />

were calculated. The mean, low (mean value 1 standard deviation)<br />

and high (mean value þ 1 standard deviation) values <strong>of</strong> each<br />

biochemical marker were calculated for each grain tissue and for<br />

coarse bran. Then, the grain tissue proportions resulting from the<br />

combinations <strong>of</strong> all these values were calculated. The lowest variability<br />

was observed for the aleurone cell walls (4%) and the highest<br />

for the intermediate layer (22%). The maximum relative error (i.e.<br />

random analytical error) observed for coarse bran was 13% (similar<br />

results were obtained for both cultivars). The impact <strong>of</strong> analytical<br />

variability on tissue proportion calculation can thus be considered<br />

as acceptable.<br />

To determine the tissue distribution in fractions obtained from<br />

wheat grain processing, four different biochemical assays need to<br />

be carried out, i.e. a rapid and simple colorimetric method to estimate<br />

the phytic acid content, an enzymatic assay to measure starch<br />

content which is standardized as a commercial kit, and two more<br />

complex and time consuming assays to analyse either the phenolic<br />

acids or the alkylresorcinols composition and content. However, if<br />

analysis <strong>of</strong> the composition in different phenolic acids appeared<br />

necessary in order to distinguish between distinct tissues as the<br />

pericarp and the aleurone layer, only the total content <strong>of</strong> alkylresorcinols<br />

is required as all <strong>of</strong> these compounds were found to be<br />

concentrated in the intermediate layer <strong>of</strong> wheat grains (Landberg<br />

et al., 2008). Thus, in order to simplify and reduce the time for<br />

biochemical analyses, we tested the potential <strong>of</strong> a colorimetric<br />

method for the determination <strong>of</strong> alkylresorcinols using Fast Blue B<br />

reagent as described by Tluscik et al. (1981). Choice <strong>of</strong> the most<br />

effective solvent and <strong>of</strong> the extraction conditions were adapted to<br />

fractionation samples after preliminary studies (cf. Section 2). A<br />

very good correlation was observed between determination <strong>of</strong> total<br />

AR content using either the colorimetric method or the sum <strong>of</strong> each<br />

<strong>of</strong> the distinct AR molecules separated by gas chromatography<br />

(colorimetric method ¼ 0.91 GC method, R 2 ¼ 0.98). This result<br />

agrees with previous studies (Andersson et al., in press) and shows<br />

that it is possible to develop a simpler and quicker assay for<br />

determination <strong>of</strong> wheat grain tissue proportions in fractions<br />

obtained from processing. Therefore, only the analysis <strong>of</strong> phenolic<br />

acids appears as the limiting factor when trying to simplify the<br />

assays. However it seems possible, in future, to develop specific<br />

Y. Hemery et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 55–64<br />

immunological tests in order to further simplify the determination<br />

<strong>of</strong> the biochemical markers.<br />

3.2. Biochemical markers as a tool to monitor fractionation<br />

processes<br />

The method was then used to assess the histological proportion<br />

<strong>of</strong> fractions exhibiting very different compositions, obtained by<br />

three different grain or bran dry fractionation processes, in order to<br />

verify if this method effectively allows study <strong>of</strong> the distribution <strong>of</strong><br />

the different tissues among fractions, and to evaluate if this method<br />

can be a useful tool.<br />

3.2.1. Conventional milling process: whole-grain fractionation<br />

The calculated grain tissue composition <strong>of</strong> whole grains, whole<br />

meal flours, white flours, fine brans and coarse brans is shown in<br />

Table 3. The composition <strong>of</strong> the fractions made <strong>of</strong> both cultivars was<br />

quite similar, except for fine brans. Indeed, Tiger fine bran contains<br />

twice as much endosperm as Crousty fine bran, may be due to<br />

differences in bran brushing steps. Whole grain and whole meal<br />

flour exhibit very similar composition, and the white flour (0.55%<br />

ash content) contains only 1–2% <strong>of</strong> non-endosperm tissues. For<br />

most <strong>of</strong> the fractions, the sum <strong>of</strong> the tissue proportion is superior to<br />

85%, but for fine brans, totals <strong>of</strong> 73% and 67% are observed for Tiger<br />

and Crousty respectively, suggesting that a non-negligible part <strong>of</strong><br />

these fractions was not quantified. In conventional milling<br />

processes, wheat germs are <strong>of</strong>ten found in high amounts in the fine<br />

bran fractions (Posner and Li, 1991). Fig. 3 shows that there is<br />

a relation between the approximate germ proportion and the ‘nonquantified’<br />

proportion in fractions, suggesting that the ‘non-quantified’<br />

30% in fine brans may effectively be composed <strong>of</strong> germ. Fig. 4<br />

presents the proportions <strong>of</strong> each <strong>of</strong> the peripheral layers relative to<br />

the total amount <strong>of</strong> outer layers in bran fractions. Crousty and Tiger<br />

coarse brans exhibit tissue proportions similar to those found for<br />

whole hand-isolated outer layers, except for the outer pericarp<br />

proportion (about 19% outer pericarp in coarse bran compared to<br />

25% in hand-isolated outer layers). This difference may be due to<br />

the high friability <strong>of</strong> outer pericarp, which is a brittle material that<br />

can be easily removed from the other layers during milling and<br />

fragmented into finer particles not recovered in coarse brans<br />

(Antoine et al., 2003, 2004). Fig. 4 also shows that the proportions<br />

<strong>of</strong> cell walls in the total aleurone layer were almost the same for<br />

coarse and fine bran as for whole hand-isolated outer layers,<br />

showing that the aleurone cell walls and cell contents were not<br />

dissociated in these fractions (i.e. the aleurone cells were not<br />

damaged).<br />

Table 3<br />

Grain tissues proportions (%) in fractions produced by a conventional milling process<br />

(Tiger and Crousty cv.)<br />

Whole<br />

grain<br />

Whole<br />

meal<br />

flour<br />

White<br />

flour<br />

Outer Intermediate Aleurone layer Starchy Total<br />

pericarp layer<br />

Total Cell Cell<br />

endosperm<br />

walls content<br />

Tiger 3.1 2.9 8.8 4.1 4.7 82.4 97<br />

Crousty 2.7 2.6 8.5 4.1 4.4 84.1 98<br />

Tiger 3.0 3.2 8.1 3.9 4.2 69.9 84<br />

Crousty 2.4 2.7 8.5 3.7 4.8 75.0 89<br />

Tiger 0.1 0.1 1.2 0.5 0.7 99.9 101<br />

Crousty 0.2 0.1 1.5 0.5 1.0 97.0 99<br />

Fine bran Tiger 6.1 6.6 17.4 7.9 9.5 42.9 73<br />

Crousty 7.1 8.0 26.6 12.6 14.0 25.3 67<br />

Coarse<br />

bran<br />

Tiger 15.9 23.0 42.5 20.9 21.6 8.0 89<br />

Crousty 15.1 21.9 46.3 21.3 24.9 8.0 91


Germ proportion (%)<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

y = 0,81x<br />

R 2 = 0,74<br />

0 5 10 15 20 25 30 35 40 45<br />

Unknown proportion (%)<br />

Fig. 3. Comparison <strong>of</strong> the approximate germ proportions with the ‘unknown<br />

proportions’ in milling fractions. The ‘unknown proportion’ in a fraction is the difference<br />

between 100% and the sum <strong>of</strong> all the calculated grain tissues proportions.<br />

3.2.2. Whole-grain fractionation by peeling, pearling, and milling<br />

This process, by sequential removal <strong>of</strong> grain outer layers by<br />

debranning prior to milling, aims at producing less contaminated<br />

wheat grains and whole-grain flours, by removing the outermost<br />

layers <strong>of</strong> the grain. The first debranning step carried out by friction<br />

removes 3–4% <strong>of</strong> the grain by weight. As shown in Table 4, this<br />

removed fraction (i.e. the peeling fraction) contains 53% and 60% <strong>of</strong><br />

outer pericarp for Tiger and Crousty respectively. It also contains<br />

various proportions <strong>of</strong> all the other grain tissues, including endosperm,<br />

probably due to the presence <strong>of</strong> broken grains in the peeling<br />

fractions and to the fact that the grain ends are more worn away<br />

during debranning. This first debranning step by peeling removes<br />

the major part <strong>of</strong> the outer pericarp <strong>of</strong> the grain (up to 60%). Due to<br />

the shape <strong>of</strong> wheat grains, removing all the outer pericarp by<br />

peeling does not seem possible as about 25% <strong>of</strong> this tissue is located<br />

in the crease (Evers and Millar, 2002). The pearling fractions,<br />

that correspond to the fraction removed <strong>of</strong>f the grain during the<br />

second debranning step (about 3% additional debranning), contains<br />

30–35% aleurone material and 25–30% starchy endosperm,<br />

showing that all the peripheral layers have been removed on some<br />

parts <strong>of</strong> the grains: the grain ends are more eroded and the grains<br />

are rounded by pearling. These pearling fractions also contain 7–13%<br />

<strong>of</strong> germ. As a large part <strong>of</strong> the outer pericarp was removed during<br />

the friction step, the other bran fractions and the flour fractions<br />

obtained afterwards do not contain much <strong>of</strong> this tissue (Table 4). As<br />

a consequence, the 100% flours from peeling contain less outer<br />

layer particles than the whole meal flours, and the 100% flours from<br />

pearling, which are made from grains that have undergone an<br />

additional debranning step (pearling), contain even less outer<br />

layers. All these 100% flours contain 3–4% germ, that is close to the<br />

amount <strong>of</strong> germ quantified in whole grains (Table 2). The 76% flours<br />

from peeling and pearling are both almost only composed <strong>of</strong> starchy<br />

endosperm (only 2% <strong>of</strong> non-endosperm particles).<br />

Fig. 4 presents the proportions <strong>of</strong> each <strong>of</strong> the outer layers relative<br />

to the total amount <strong>of</strong> peripheral layers in bran fractions. The<br />

pearling fractions contain the most important proportion <strong>of</strong> intermediate<br />

layer (8–10% more than in hand-isolated whole outer<br />

layers), due to the removal <strong>of</strong> outer pericarp by peeling and to the<br />

fact that a part <strong>of</strong> the aleurone layer probably remains attached to<br />

the grains and is found in the bran fraction after pearling.<br />

Surprisingly, bran fractions after peeling and after pearling exhibit<br />

quite similar tissues proportions. The differences between these<br />

two fractions are probably reduced because they both contain the<br />

bran present in the crease area. This crease bran represents a nonnegligible<br />

proportion <strong>of</strong> the grain total outer layers and is not easily<br />

removed neither by peeling nor by pearling (Dexter and Wood,<br />

1996). Fig. 4. shows that for almost all the bran fractions, the ratio<br />

aleurone cell walls/aleurone cell contents is very close to that<br />

observed for coarse bran and hand-isolated whole outer layers,<br />

suggesting that aleurone material is not dissociated and that the<br />

aleurone cells are not broken. Only the peeling fractions exhibit<br />

a surprisingly high proportion <strong>of</strong> aleurone cell walls compared to<br />

their proportion <strong>of</strong> aleurone cell contents. This may be explained by<br />

the fact that the cell walls, being linked to the other outer layers, are<br />

torn <strong>of</strong>f the grain by peeling, leaving the cell contents on the grain.<br />

But this high proportion <strong>of</strong> aleurone cell walls in peeling fractions<br />

seems quite excessive when compared to the proportions <strong>of</strong><br />

intermediate layers in these fractions, suggesting that it may have<br />

been over-estimated due to the high content <strong>of</strong> para-coumaric acid<br />

in these fractions. If the proportions <strong>of</strong> total aleurone layer is<br />

calculated by taking into account only the phytic acid content and<br />

not the p-CA content <strong>of</strong> the peeling fractions, the total aleurone<br />

content in these fractions is found to be 3–4%. This could be more<br />

25 29 31 34 27 30 26 27 32 38<br />

29 27 33 35<br />

23 22<br />

26<br />

30<br />

26 26<br />

8 4<br />

26<br />

26<br />

24 27<br />

27 29<br />

26<br />

26<br />

25<br />

24<br />

22<br />

20<br />

19<br />

17<br />

28<br />

20<br />

26<br />

18<br />

63 68<br />

32<br />

9<br />

30<br />

6<br />

35<br />

11<br />

34<br />

11<br />

30<br />

10<br />

28<br />

8<br />

Tg Cr Tg Cr Tg Cr Tg Cr Tg Cr Tg Cr Tg Cr<br />

Hand-isolated<br />

whole bran<br />

Y. Hemery et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 55–64 61<br />

Fine bran Coarse bran peeling<br />

fraction<br />

bran fraction<br />

after peeling<br />

pearling<br />

fraction<br />

outer pericarp intermediate layer aleurone cell walls aleurone cell contents<br />

bran fraction<br />

after pearling<br />

Fig. 4. Study <strong>of</strong> the dissociation <strong>of</strong> the outer layers in the different Tiger and Crousty bran fractions: histological composition <strong>of</strong> the particles <strong>of</strong> outer tissues present in these<br />

fractions (i.e. endosperm excluded). The proportions <strong>of</strong> outer pericarp, intermediate layer, and aleurone layer are re-calculated relatively to the total amount <strong>of</strong> these tissues in bran<br />

fractions.


62<br />

Table 4<br />

Grain tissues proportions (%) in fractions produced by a whole-grain fractionation process including peeling, pearling, and milling steps (Tiger and Crousty cv.)<br />

plausible, as these fractions are also supposed to contain 9–17% <strong>of</strong><br />

germ. The high amount <strong>of</strong> p-CA quantified in the peeling fractions<br />

might be due to an external contamination. Indeed, contaminating<br />

materials such as straw and chaff can be found in the debranning<br />

fraction, and several authors showed that wheat straw is very rich<br />

in para-coumaric acid (Benoit et al., 2006; Sun et al., 2001; Tapin<br />

et al., 2006).<br />

3.2.3. Bran fractionation process: production <strong>of</strong> aleurone-rich<br />

fractions<br />

The wheat aleurone layer has been shown to have great nutritional<br />

interest, and to concentrate most <strong>of</strong> the micronutrients<br />

(minerals and vitamins), antioxidants (mainly phenolic acids) and<br />

other phytochemicals <strong>of</strong> the wheat grain (Hemery et al., 2007).<br />

Thus, it can be interesting to produce bran fractions highly<br />

concentrated in aleurone material, to transform a by-product <strong>of</strong><br />

flour into a high nutritional value food ingredient. In the present<br />

study, various aleurone-rich bran fractions were studied (Table 5).<br />

The Aleurone 2 180 mm Tiger 22.9 13.6 55.3 28.3 27.0 2.0 94<br />

Crousty 19.1 13.1 64.7 31.0 33.7 2.4 99<br />

Aleurone 1 180 mm Tiger 9.1 12.9 74.2 32.7 41.5 2.6 99<br />

Crousty 11.2 12.7 83.2 38.9 44.3 2.7 110<br />

Aleurone 2 < 180 mm Tiger 7.8 11.1 79.3 37.3 42.0 2.9 101<br />

Crousty 8.5 12.0 89.1 42.4 46.7 3.0 113<br />

By-product 1 Tiger 30.9 18.9 29.7 20.5 9.2 3.6 83<br />

Crousty 36.6 18.6 29.4 19.9 9.6 4.4 89<br />

By-product 2 Tiger 45.6 11.0 27.6 15.9 11.7 1.6 86<br />

Crousty 33.2 12.7 32.9 20.2 12.7 2.0 81


fractions (compared to Aleurone 1 fractions) after the second<br />

fractionation step.<br />

For almost all the aleurone fractions, the ratio aleurone cell<br />

walls/aleurone cell contents is really close to the ratio observed for<br />

hand-isolated aleurone layers, suggesting that aleurone material is<br />

not dissociated (i.e. that the aleurone cells are not broken). This<br />

observation are in agreement with microscopy observations<br />

(Amrein et al., 2003; Buri et al., 2004) showing that these aleuronerich<br />

fractions are composed <strong>of</strong> clusters <strong>of</strong> intact cells. All these<br />

observations are in agreement with the observations made by<br />

several authors. Indeed, the higher protein and beta-glucan<br />

contents observed by Amrein et al. (2003), the higher phenolic<br />

acids content observed by Zhou et al. (2004), and the higher<br />

minerals and phytic acid content observed by Buri et al. (2004) for<br />

Aleurone 2 than for Aleurone 1, can be explained by the enrichment<br />

in aleurone material <strong>of</strong> Aleurone 2 compared to Aleurone 1.<br />

4. Conclusion<br />

The evaluation <strong>of</strong> the method described here, in comparison<br />

with tissue quantification by hand dissection, has shown its efficacy<br />

(mean relative errors <strong>of</strong> 4% and 8% for Crousty and Tiger outer layers<br />

respectively), and the impact <strong>of</strong> the analytical variability<br />

(maximum 13% relative error on coarse bran) was regarded as<br />

acceptable. Some improvements can be made to the current<br />

method. For example, a better quantification <strong>of</strong> the germ content is<br />

needed, as the germ proportion is <strong>of</strong>ten not negligible in bran<br />

factions and fractions coming from debranning processes. Wheat<br />

germ agglutinin seems to be a promising marker <strong>of</strong> germ. The<br />

quantification method does not currently enable the accurate<br />

quantification <strong>of</strong> the germ proportions in milling fractions, but it<br />

gives qualitative information and allows classification <strong>of</strong> these<br />

fractions according to their germ content (Fig. 3). This method<br />

should be improved to add WGA as another marker to complete the<br />

current method.<br />

In the milling industry, batches <strong>of</strong> different grain cultivars are<br />

frequently used as starting material for dry fractionation processes.<br />

In this study, the tissue proportions have been deduced from<br />

reference values assessed on exactly the same wheat cultivars from<br />

which the fractions were produced: the genetic and agronomic<br />

variability was thus not taken into account. Some <strong>of</strong> the constituents<br />

chosen as biochemical markers have specific functions in<br />

wheat grain. Therefore, the natural variability <strong>of</strong> compounds such<br />

as phytic acid, phenolic acids, and alkylresorcinols among a wheat<br />

population could be very high. For example, Barrier-Guillot et al.<br />

(1996) showed that the amount <strong>of</strong> Phytic P can vary from 0.92 to<br />

2.8 mg/g dm for common wheat with various agronomic and<br />

genetic characteristics. Within the European HEALTHGRAIN project<br />

(Poutanen et al., 2008), other authors analysed the numerous<br />

wheat varieties in the diversity screen <strong>of</strong> the project and showed<br />

that the total phenolic acid content varies from 326 to 1171 mg/g dm<br />

in the 130 winter wheat lines and from 456 to 892 mg/g dm in the<br />

20 spring wheat lines (Li et al., in press), while the total alkylresorcinol<br />

content varies from 220 to 652 mg/g dm in these winter<br />

wheat lines and from 254 to 537 mg/g dm in the spring wheat lines<br />

(Andersson et al., in press). To try to evaluate the impact <strong>of</strong> genetic<br />

variability on grain tissue proportions using biochemical markers,<br />

tissue proportions <strong>of</strong> Tiger coarse bran were calculated with<br />

Crousty reference values (and vice versa). The mean relative error<br />

observed was 10%, and this variability was found to be less than the<br />

maximum analytical variability. This study should be carried out on<br />

other wheat cultivars. If this observation were confirmed, it could<br />

then be possible to use ‘‘standard reference values’’ (i.e. the means<br />

<strong>of</strong> the values obtained for the grain tissues <strong>of</strong> different cultivars) to<br />

quantify tissues proportion in fractions, to limit grain tissues<br />

Y. Hemery et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 55–64 63<br />

dissection, and to apply this quantification method to batches <strong>of</strong><br />

grains <strong>of</strong> various origins.<br />

Several biochemical analyses have to be carried out to assess the<br />

composition <strong>of</strong> fractions using the described method. Therefore, it<br />

would be difficult to use it for controls on the production line, but<br />

this method can be useful in quality control laboratories <strong>of</strong> firms,<br />

and can also be used as a reference for the calibration <strong>of</strong> equipment<br />

based on physical properties (i.e. fluorescence). This biochemical<br />

marker method has been proved to be an efficient tool for assessment<br />

<strong>of</strong> grain tissue proportions in various milling fractions<br />

exhibiting contrasting composition, such as flours, brans, or aleurone-rich<br />

fractions. The present study showed that this quantitative<br />

method allows, by calculation <strong>of</strong> the composition <strong>of</strong> the different<br />

products generated by grain or bran fractionation processes, an<br />

understanding <strong>of</strong> the effects <strong>of</strong> these processes and monitoring and<br />

adapting <strong>of</strong> them to reach the objective (i.e. efficient debranning,<br />

production <strong>of</strong> specific fractions).<br />

Acknowledgements<br />

The authors thank the Federal Research Centre for Nutrition and<br />

Food (BFEL), Germany, for the production <strong>of</strong> the conventional<br />

milling fractions used in this study. We also gratefully thank T.-M.<br />

Lasserre for her work on the biochemical analyses, and S. Chay and<br />

A. Putois for the manual dissection <strong>of</strong> grain tissues. This publication<br />

is financially supported by the European Commission in the<br />

Communities 6th Framework Program, Project HEALTHGRAIN<br />

(FOOD-CT-2005-514008). It reflects the author’s views and the<br />

Community is not liable for any use that may be made <strong>of</strong> the<br />

information contained in this publication.<br />

References<br />

AACC, 2000. American Association <strong>of</strong> <strong>Cereal</strong> Chemists Approved Methods, 10th ed.<br />

The Association, St Paul, Minnesota.<br />

Amrein, T.M., Gränicher, P., Arrigon, E., Amadò, R., 2003. In vitro digestibility and<br />

colonic fermentability <strong>of</strong> aleurone isolated from wheat bran. Lebensmittel-<br />

Wissenschaft und Technology 36, 451–460.<br />

Andersson, A.M.A., Kamal-Eldin, A., Fras, A., Boros, D., Åman, P. Alkylresocinols in<br />

wheat varieties in the HEALTHGRAIN diversity screen. <strong>Journal</strong> <strong>of</strong> Agricultural<br />

and Food Chemistry, in press.<br />

Antoine, C., Peyron, S., Mabille, F., Lapierre, C., Bouchet, B., Abecassis, J., Rouau, X.,<br />

2003. Individual contribution <strong>of</strong> grain outer layers and their cell wall structure<br />

to the mechanical properties <strong>of</strong> wheat bran. <strong>Journal</strong> <strong>of</strong> Agricultural and Food<br />

Chemistry 51, 2026–2033.<br />

Antoine, C., Peyron, S., Lullien-Pellerin, V., Abecassis, J., Rouau, X., 2004. Wheat bran<br />

tissue fractionation using biochemical markers. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 39,<br />

387–393.<br />

Aureli, G., D’Egidio, M.G., 2007. Efficacy <strong>of</strong> debranning on lowering <strong>of</strong> deoxynivalenol<br />

(DON) level in manufacturing processes <strong>of</strong> durum wheat. Tecnica<br />

Molitoria 58, 729–733.<br />

Bach Knudsen, K.E., Steenfeldt, S., Borsting, C.F., Eggum, B.O., 1995. The nutritive<br />

value <strong>of</strong> decorticated mill fractions <strong>of</strong> wheat. 1. Chemical composition <strong>of</strong> raw<br />

and enzyme treated fractions and balance experiments with rats. Animal Feed<br />

<strong>Science</strong> and Technology 52, 205–225.<br />

Bacic, A., Stone, B.A., 1981. Isolation and ultrastructure <strong>of</strong> aleurone cell walls from<br />

wheat and barley. Australian <strong>Journal</strong> <strong>of</strong> Plant Physiology 8, 453–474.<br />

Baldwin, P.M., Bertrand, D., Novales, B., Bouchet, B., Collobert, G., Gallant, D.J., 1997.<br />

Chemometric labeling <strong>of</strong> cereal tissues in multichannel fluorescence microscopy<br />

images using discriminant analysis. Analytical Chemistry 69, 4339–4348.<br />

Barrier-Guillot, B., Casado, P., Maupetit, P., Jondreville, C., Gatel, F., Larbier, M., 1996.<br />

Wheat phosphorus availability. 1. In vitro study; factors affecting endogenous<br />

phytasic activity and phytic phosphorus content. <strong>Journal</strong> <strong>of</strong> the <strong>Science</strong> <strong>of</strong> Food<br />

and Agriculture 70, 62–68.<br />

Barron, C., Surget, A., Rouau, X., 2007. Relative amounts <strong>of</strong> tissues in mature wheat<br />

(Triticum aestivum L.) grain and their carbohydrate and phenolic acid composition.<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 45, 88–96.<br />

Benoit, I., Navarro, D., Marnet, N., Rakotomanomana, N., Lesage-Meessen, L.,<br />

Sigoillot, J.C., Asther, M., Asther, M., 2006. Feruloyl esterases as a tool for the<br />

release <strong>of</strong> phenolic compounds from agro-industrial by-products. Carbohydrate<br />

Research 341, 1820–1827.<br />

Bohm, A., Bogoni, C., Behrens, R., Otto, T., 2003. Method for the extraction <strong>of</strong><br />

aleurone from bran. US Patent Application Publication US 2003/0175384 A1.<br />

Applicant: Buhler AG.<br />

Bohm, A., Kratzer, A., 2005. Method for isolating aleurone particles. US Patent<br />

Application Publication US 2005/0103907 A1. Applicant: Buhler AG.


64<br />

Brillouet, J.M., Rouau, X., Hoebler, C., Barry, J.L., Carre, B., Lorta, E., 1988. A new<br />

method for determination <strong>of</strong> insoluble cell-walls and soluble nonstarchy<br />

polysaccharides from plant materials. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry<br />

36, 969–979.<br />

Buri, R.C., von Reding, W., Gavin, M.H., 2004. Description and characterization <strong>of</strong><br />

wheat aleurone. <strong>Cereal</strong> Foods World 49, 274–282.<br />

Chen, Y., Ross, A.B., Aman, P., Kamal-Eldin, A., 2004. Alkylresorcinols as markers <strong>of</strong><br />

whole grain wheat and rye in cereal products. <strong>Journal</strong> <strong>of</strong> Agricultural and Food<br />

Chemistry 52, 8242–8246.<br />

Courcoux, P., Devaux, M.F., Bouchet, B., 2002. Simultaneous decomposition <strong>of</strong><br />

multivariate images using three-way data analysis. Application to the<br />

comparison <strong>of</strong> cereal grains by congocal laser scanning microscopy. Chemometrics<br />

and Intelligent Laboratory Systems 62, 103–113.<br />

Dexter, J.E., Wood, P.J., 1996. Recent applications <strong>of</strong> debranning <strong>of</strong> wheat before<br />

milling. Trends in Food <strong>Science</strong> & Technology 7, 35–41.<br />

Eugster, W., Gerschwiler, O., 2006. Method and installation for cleaning cereal. US<br />

Patent Application Publication US 2006/0147591 A1. Applicant: Bühler AG.<br />

Evers, A.D., Reed, M., 1988. Some novel observations by scanning electron microscopy<br />

on the seed coat and nucellus <strong>of</strong> the mature wheat grain. <strong>Cereal</strong> Chemistry<br />

65, 81–85.<br />

Evers, A.D., Millar, S., 2002. <strong>Cereal</strong> grain structure and development: some implications<br />

for quality. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 36, 261–284.<br />

Fincher, G.B., Stone, B.A., 1986. Cell walls and their components in cereal grain<br />

technology. In: Pomeranz, Y. (Ed.), Advances in <strong>Cereal</strong> <strong>Science</strong> and Technology,<br />

vol. 8. American Association <strong>of</strong> <strong>Cereal</strong> Chemists, St Paul, MN, pp. 207–295.<br />

Fleurat-Lessard, F., Chaurand, M., Marchegay, G., Abecassis, J., 2007. Effects <strong>of</strong> processing<br />

on the distribution <strong>of</strong> pirimiphos-methyl residues in milling fractions <strong>of</strong><br />

durum wheat. <strong>Journal</strong> <strong>of</strong> Stored Products Research 43, 384–395.<br />

Fulcher, R.G., O’Brien, T.P., Lee, J.W., 1972. Studies on the aleurone layer. I.<br />

Conventional and fluorescence microscopy <strong>of</strong> the cell wall with emphasis on<br />

phenol-carbohydrate complexes in wheat. Australian <strong>Journal</strong> <strong>of</strong> Biological<br />

<strong>Science</strong> 25, 23–34.<br />

Grabber, J.H., Ralph, J., Lapierre, C., Barriere, Y., 2004. Genetic and molecular basis <strong>of</strong><br />

grass cell-wall degradability. I. Lignin-cell wall matrix interactions. Comptes<br />

Rendus Biologies 327, 455–465.<br />

Greffeuille, V., Abecassis, J., Bar L’Helgouac’h, C., Lullien-Pellerin, V., 2005. Differences<br />

in the aleurone layer fate between hard and s<strong>of</strong>t common wheats at grain<br />

milling. <strong>Cereal</strong> Chemistry 82, 138–143.<br />

Hemery, Y., Rouau, X., Lullien-Pellerin, V., Barron, C., Abecassis, J., 2007. Dry<br />

processes to develop wheat fractions and products with enhanced nutritional<br />

quality. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 46, 327–347.<br />

Jacobs, D.R., Steffen, L.M., 2003. Nutrients, foods, and dietary patterns as exposures<br />

in research: a framework for food synergy. American <strong>Journal</strong> <strong>of</strong> Clinical Nutrition<br />

78, 508S–513S.<br />

Jensen, S.A., Munck, L., Martens, H., 1982. The botanical constituents <strong>of</strong> wheat and<br />

wheat milling fractions. 1. Quantification by aut<strong>of</strong>luorescence. <strong>Cereal</strong> Chemistry<br />

59, 477–484.<br />

Jensen, S.A., Martens, H.,1983. The botanical constituents <strong>of</strong> wheat and wheat milling<br />

fractions. 2. Quantification by amino-acids. <strong>Cereal</strong> Chemistry 60, 172–177.<br />

Laca, A., Mousia, Z., Diaz, M., Webb, C., Pandiella, S.S., 2006. Distribution <strong>of</strong><br />

microbial contamination within cereal grains. <strong>Journal</strong> <strong>of</strong> Food Engineering 72,<br />

332–338.<br />

Landberg, R., Kamal-Eldin, A., Salmenkallio-Marttila, M., Rouau, X., Aman, P., 2008.<br />

Localization <strong>of</strong> alkylresorcinols in barley, wheat and rye kernels. <strong>Journal</strong> <strong>of</strong><br />

<strong>Cereal</strong> <strong>Science</strong> 48, 401–406.<br />

Latta, M., Eskin, M., 1980. A simple and rapid colorimetric method for phytate<br />

determination. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 28, 1313–1315.<br />

Lempereur, I., Surget, A., Rouau, X., 1998. Variability in dehydrodiferulic acid<br />

composition <strong>of</strong> durum wheat (Triticum durum Desf.) and distribution in milling<br />

fractions. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 28, 251–258.<br />

Y. Hemery et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 55–64<br />

Li, L., Shewry, P.R., Ward, J.L. Phenolic acids in wheat varieties in the HEALTHGRAIN<br />

diversity screen. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry, in press.<br />

Liu, R.H., 2007. Whole grain phytochemicals and health. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

46, 207–219.<br />

Mateo Anson, N., Van den Berg, R., Havenaar, R., Bast, A., Haenen, G., 2008.<br />

Ferulic acid from aleurone determines the antioxidant potency <strong>of</strong> wheat<br />

grain (Triticum aestivum L.). <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 56,<br />

5589–5594.<br />

Miller, R.C., Bowles, D.J., 1982. A comparative-study <strong>of</strong> the localization <strong>of</strong> wheatgerm-agglutinin<br />

and its potential receptors in wheat grains. Biochemical<br />

<strong>Journal</strong> 206, 571–576.<br />

O’Dell, B.L., De Boland, A.R., Koirtyohann, S.R., 1972. Distribution <strong>of</strong> phytate and<br />

nutritionally important elements among the morphological components <strong>of</strong><br />

cereal grains. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 20, 718–723.<br />

Peyron, S., Surget, A., Mabille, F., Autran, J.C., Rouau, X., Abecassis, J., 2002.<br />

Evaluation <strong>of</strong> tissue dissociation <strong>of</strong> durum wheat grain (Triticum durum<br />

desf.) generated by the milling process. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 36,<br />

199–208.<br />

Pomeranz, Y., 1988. Chemical composition <strong>of</strong> kernel structures. In: Pomeranz, Y.<br />

(Ed.), Wheat: Chemistry and Technoloy, vol. 4. AACC, St Paul, pp. 97–158.<br />

Posner, E.S., Li, Y.Z., 1991. A technique for separation <strong>of</strong> wheat germ by impacting<br />

and subsequent grinding. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 13, 49–70.<br />

Poutanen, K., Shepherd, R., Shewry, P., Delcour, J.A., Björck, I., van der Kamp, J.W.,<br />

2008. Beyond whole grain: the European HEALTHGRAIN project aims at<br />

healthier cereal foods. <strong>Cereal</strong> Foods World 53, 32–35.<br />

Pussayanawin, V., Wetzel, D.L., Fulcher, R.G., 1988. Fluorescence detection and<br />

measurement <strong>of</strong> ferulic acid in wheat milling fractions by microscopy and HPLC.<br />

<strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 36, 515–520.<br />

Rhodes, D.I., Sadek, M., Stone, B.A., 2002. Hydroxycinnamic acids in walls <strong>of</strong> wheat<br />

Aleurone cells. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 36, 67–81.<br />

Ross, A.B., Shepherd, M.J., Schupphaus, M., Sinclair, V., Alfaro, B., Kamal-Eldin, A.,<br />

Aman, P., 2003. Alkylresorcinols in cereals and cereal products. <strong>Journal</strong> <strong>of</strong><br />

Agricultural and Food Chemistry 51, 4111–4118.<br />

Sun, R.C., Sun, X.F., Zhang, S.H., 2001. Quantitative determination <strong>of</strong> hydroxycinnamic<br />

acids in wheat, rice, rye, and barley straws, maize stems, oil palm frond<br />

fiber, and fast-growing poplar wood. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry<br />

49, 5122–5129.<br />

Surget, A., Barron, C., 2005. Histologie du grain de blé. Industrie des Céréales 145,<br />

3–7.<br />

Symons, S.J., Dexter, J.E., 1996. Aleurone and pericarp fluorescence as estimators <strong>of</strong><br />

mill stream refinement for various canadian wheat classes. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong><br />

<strong>Science</strong> 23, 73–83.<br />

Tapin, S., Sigoillot, J.C., Asther, M., Petit-Conil, M., 2006. Feruloyl esterase utilization<br />

for simultaneous processing <strong>of</strong> nonwood plants into phenolic compounds and<br />

pulp fibers. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 54, 3697–3703.<br />

Tluscik, F., Kozubek, A., Mejbaum-Katzenellenbogen, W., 1981. Alkylresorcinols in<br />

rye (Secale-<strong>Cereal</strong>e L) grains. 6. Colorimetric micromethod for the determination<br />

<strong>of</strong> alkylresorcinols with the use <strong>of</strong> diazonium salt, fast blue-B. Acta Societatis<br />

Botanicorum Poloniae 50, 645–651.<br />

Vaintraub, I.A., Lapteva, N.A., 1988. Colorimetric determination <strong>of</strong> phytate in<br />

unpurified extracts <strong>of</strong> seeds and the products <strong>of</strong> their processing. Analytical<br />

Biochemistry 175, 227–230.<br />

Vincenzi, S., Zoccatelli, G., Perbellini, F., Rizzi, C., Chignola, R., Curioni, A.,<br />

Peruffo, A.D.B., 2002. Quantitative determination <strong>of</strong> dietary lectin activities by<br />

enzyme-linked immunosorbent assay using specific glycoproteins immobilized<br />

on microtiter plates. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 50,<br />

6266–6270.<br />

Zhou, K., Laux, J.J., Yu, L., 2004. Comparison <strong>of</strong> swiss red wheat grain and fractions<br />

for their antioxidant properties. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 52,<br />

1118–1123.


Lateral growth <strong>of</strong> a wheat dough disk under various growth conditions<br />

Amy Penner a , Leaelaf Hailemariam b , Martin Okos a,b , Osvaldo Campanella a, *<br />

a School <strong>of</strong> Agricultural and Biological Engineering, Purdue University, West Lafayette, IN 47907-2093, USA<br />

b School <strong>of</strong> Chemical Engineering, Purdue University, West Lafayette, IN 47907-2100, USA<br />

article info<br />

Article history:<br />

Received 17 October 2007<br />

Received in revised form 24 June 2008<br />

Accepted 9 July 2008<br />

Keywords:<br />

Viscoelasticity<br />

Dough<br />

Pressure release<br />

Expansion<br />

1. Introduction<br />

abstract<br />

In the food industry, bubble formation is used to create structure<br />

and texture which are primarily responsible for the product attributes<br />

(Aguilera and Germain, 2007; Moraru and Kokini, 2003;<br />

Trater et al., 2005). Bubbles are generated through the evolution<br />

and expansion <strong>of</strong> gas, either through phase change (flash evaporation),<br />

or chemical reaction (yeast or chemical leavening). Gas that<br />

does not escape into the atmosphere is bound in nucleating sites<br />

which act as growth sites for bubbles in the expanding material<br />

(Baker and Mize, 1941). The quantity <strong>of</strong> bubbles created and<br />

sustained is the main factor in determining the final product<br />

quality, texture, and volume (Cauvain et al., 1999; Elmehdi et al.,<br />

2003; Fox,2004;Strybulevych et al., 2007; Whitworth and Alava,<br />

1999). Thus, an understanding <strong>of</strong> bubble growth mechanics is<br />

crucial for control and product optimization.<br />

Previous work on mathematical modeling <strong>of</strong> bubble growth has<br />

been done by Alavi et al. (2003), Chen and Rizvi (2006), de Cindio<br />

and Correra (1995), Fan et al. (1994, 1999), Schwartzberg et al.<br />

(1995), Shimiya and Yano (1988), Singh and Bhattacharya (2005),<br />

and Wang et al. (2005). Recently, Hailemariam et al. (2007)<br />

presented a model that included the effect <strong>of</strong> multi-scale mass<br />

* Corresponding author. Tel.: þ1 76 5496 6330; fax: þ1 76 5496 1115.<br />

E-mail address: campa@purdue.edu (O. Campanella).<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.07.007<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 65–72<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

Numerous studies have tried to understand and model bubble growth inside dough. Experimental<br />

studies are inconvenienced by the methods’ inability to capture the dynamic phenomena. In this paper,<br />

a versatile experimental method was developed to allow for macroscopic expansion <strong>of</strong> wheat dough. The<br />

study evaluates expansion <strong>of</strong> a dough disk under varying: moisture content (40, 41, 42, 43, and 44% wb),<br />

leavening acid concentration (30, 40, and 50% db), pressure schemas, pressurizing gas (compressed air<br />

and CO2), and lubrication (Teflon Ò film coating and Pam Ò aerosol lubricant). Dough expansion increased<br />

22.6% by increasing moisture content from 40 to 44%. Increased baking powder formulation (40% db) was<br />

used to enhance initial growth conditions and CO2 production. ‘Pressure pulse’ and ‘pressure vacuum<br />

methods’ added pressurization alternatively with full vacuum. The former method included a rest period<br />

before vacuum application, and increased expansion by 10.8%. Teflon Ò and Pam Ò reduced friction<br />

between the dough and acrylic plate and increased the final expansion by 14.7% compared to no<br />

lubricant following the ‘standard pressurization method’. ‘Pressure pulse’ and ‘pressure vacuum’<br />

experiments decreased expansion by 28.4 and 38.2%, respectively compared to ‘standard pressurization’<br />

while using Teflon Ò and Pam Ò .<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

transfer, surface tension, inertia and viscoelasticity in a multibubble<br />

environment. However, a major challenge with these<br />

models has been the experimental verification <strong>of</strong> bubble dynamics.<br />

Bell et al. (1975) described bubble growth in dough during baking<br />

using television microscopy <strong>of</strong> a thin film <strong>of</strong> dough baking on<br />

a small slide. Kumagai et al. (1991) used a similar approach,<br />

spreading a thin layer <strong>of</strong> dough on a slide heated by a thermostat.<br />

de Cindio and Correra (1995) microscopically analyzed bubble size<br />

distributions <strong>of</strong> dough samples which were immersed in liquid<br />

nitrogen and cut in thin slices. Similar analysis was done by<br />

Campbell et al. (1998) and Schwartzberg et al. (1995). Alavi et al.<br />

(1999) cut extrudate samples into thin sections which were dried in<br />

high temperature ovens before electron microscopy analysis. A<br />

similar approach was taken by Chen and Rizvi (2006) and Singh and<br />

Bhattacharya (2005).<br />

With the exception <strong>of</strong> the ‘dough smear’ methods <strong>of</strong> Bell et al.<br />

(1975), the experimental methods used thus far have yielded only<br />

the final bubble size distribution, which is <strong>of</strong> little help in studying<br />

bubble dynamics. The ‘dough smear’ methods do not allow for<br />

control <strong>of</strong> the surrounding environment and are difficult to scale<br />

up.<br />

Measuring the dough’s properties while it is undergoing physical<br />

and chemical changes is difficult due to the dynamic and fragile<br />

nature <strong>of</strong> dough (Aguilera and Germain, 2007). Invasion during the<br />

expansion process could cause gas cell leakage which would reduce<br />

expansion (Wagner et al., 2006). Older techniques used interval


66<br />

measurements, but this yielded inconsistent, time intensive results<br />

that were not fully accurate or representative <strong>of</strong> on-line changes<br />

(Fox et al., 2004). Modern microscopic measuring techniques may<br />

be applied for dynamic size measurements, but these are not<br />

without their own hurdles. These methods include electron<br />

microscopy (Aguilera and Germain, 2007; Trater et al. 2005;<br />

Whitmore and Alava, 1999), X-ray tomography (Babin et al., 2006,<br />

2007; Bellido et al., 2006; Trater et al., 2005; Whitmore and Alava,<br />

1999), Magnetic resonance imaging (MRI) (Wagner et al., 2006),<br />

and ultrasound (Elmhedi et al., 2003; Fox et al., 2004; Strybulevych<br />

et al., 2007). The reader is asked to consult these references for<br />

greater details.<br />

In summary, one may remark that no single experimental<br />

method is free <strong>of</strong> challenges. The ease <strong>of</strong> experimentation, the size<br />

range <strong>of</strong> the bubbles, and the versatility <strong>of</strong> the methods should be<br />

considered in using or developing an experimental method. The<br />

bubble size range in the food industry is 10–1000 mm(Fyrillas et al.,<br />

2000). Current detection methods are applied to determine gas<br />

bubble distribution, cell wall size, and cell wall thickness. The<br />

microscopic bubble size poses a challenge for the development <strong>of</strong><br />

a dynamic method <strong>of</strong> measuring bubble size because the material<br />

structure is fixed either through heat or freezing before analysis.<br />

This may destroy the structure, the entity to be measured, and is<br />

not fully representative <strong>of</strong> the dynamic system (Babin et al., 2007;<br />

Trader et al., 2005; Wagner et al., 2006).<br />

There is a need for a versatile experimental method that facilitates<br />

the dynamic observation <strong>of</strong> the bubble size history and<br />

distribution. To overcome the challenge <strong>of</strong> measuring bubble<br />

growth in situ, the authors developed an experimental method to<br />

study microstructure growth through macroscopic investigation. A<br />

sample <strong>of</strong> wheat dough was constrained to grow laterally in a thin<br />

gap between parallel plates which allowed the maximum change in<br />

density with time due to decreased surface area for gas escape.<br />

Some work has been done on lateral growth with a mass <strong>of</strong><br />

dough, including looking at cookie diameters (Lee and Inglett,<br />

2006; Swanson et al., 1999; Zoulias et al., 2002). Taki et al. (2003)<br />

developed an experimental method for analyzing the growth <strong>of</strong><br />

bubbles in polymers through examination <strong>of</strong> an expanding disk <strong>of</strong><br />

fluid. Recently, Elmehdi et al. (2007) described a method for<br />

measuring dough density between two polyacrylic plates at<br />

atmospheric conditions where radial growth was the only dimension<br />

encountered. Digital imaging was used to capture dough<br />

expansion every 2 min. In this paper, we extend that method and<br />

describe a method that allows for dynamic tracking <strong>of</strong> overall<br />

growth under different pressure schema with increased measurement<br />

recordings. This work concentrates on the study <strong>of</strong> rapid<br />

dough growth initiated by sudden pressure and vacuum changes.<br />

Analysis <strong>of</strong> the lateral growth in a dough disk may be imagined to<br />

be a cylindrical section <strong>of</strong> the expanding material. It was hypothesized<br />

that pressurization during the initial phase <strong>of</strong> growth would<br />

increase the final expansion due to prevention <strong>of</strong> early gas escape.<br />

To test this idea, different pressurization schemas were implemented<br />

to evaluate the effects <strong>of</strong> pressurization and vacuum<br />

application to aid in increasing expansion.<br />

The work describes a simple dough system consisting <strong>of</strong> flour,<br />

water, and baking powder to observe expansion in experimental<br />

testing. Moisture content and baking powder concentration work<br />

to drive dough expansion in this system and thus warranted further<br />

study to generate optimal testing conditions. Moisture content<br />

affects the release <strong>of</strong> CO2 from the baking powder. It also affects the<br />

ability <strong>of</strong> the dough matrix to expand and retain the gas that is<br />

produced. Baking powder concentration was varied to determine<br />

its effect on expansion. Use <strong>of</strong> external CO2 as the pressurizing gas<br />

was investigated for its potential effects in the final system equilibrium.<br />

Dough expansion is based on internal CO2 production and<br />

thus it was necessary to compare CO2 with compressed air as the<br />

A. Penner et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 65–72<br />

pressurizing gas. It was also hypothesized that expansion was<br />

limited by friction at the dough/acrylic plate interface, and thus<br />

a reduction in friction was investigated using Teflon Ò film and<br />

Pam Ò aerosol lubrication.<br />

2. Experiment description<br />

The pressure chamber consists <strong>of</strong> a clear acrylic cylinder with<br />

removable top and bottom clear acrylic plates. The plates sealed the<br />

chamber using a rubber gasket at each end and each end was<br />

securely bolted to maintain constant pressure once the sample was<br />

placed inside. The cylinder measured 10 inches vertically and had<br />

a 5 inch inner diameter. A schematic representation <strong>of</strong> the experimental<br />

setup is shown in Fig. 1. A sample <strong>of</strong> dough was placed<br />

between two 5-mm thick clear acrylic plates which were clamped in<br />

place with a fixed 2 mm gap. This ensured only lateral expansion<br />

which is shown in Fig. 2. The acrylic plate was marked with x and y<br />

major and minor axes in the horizontal plane using a mm length<br />

scale to determine the diameter <strong>of</strong> dough sample with time. All<br />

experiments were performed in triplicate and averaged results are<br />

reported. For all experiments temperature was held at 20 3 C.<br />

The following parameters were evaluated throughout this work:<br />

i. Moisture content: 40, 41, 42, 43, and 44% (wb).<br />

ii. Baking powder concentration: 30, 40, and 50% (db).<br />

iii. Pressure schema: combinations <strong>of</strong> increased pressure<br />

(2 bars), and/or full vacuum application; see Section 3.2 for<br />

full description.<br />

iv. Pressurizing gas: compressed air or CO2.<br />

v. Lubrication: Teflon Ò film coating and/or aerosol lubrication.<br />

Fig. 1. Expansion chamber: (a) top-schematic; and (b) bottom-laboratory setup.


Fig. 2. Parallel plate: (a) schematic; and (b) laboratory setup.<br />

3. Materials and methods<br />

3.1. Dough preparation<br />

Flour was determined to have 11.5% moisture content (AOAC<br />

method 925.10) and baking powders to have 6% moisture content<br />

(http://www.nutritiondata.com). The wheat dough was prepared<br />

from 10.00 0.02 g all purpose bleached flour (Gold Medal Ò<br />

General Mills, Minneapolis, MN, USA), 4.00 0.02 g double acting<br />

baking powder (Clabber Girl Ò , Terre Haute, IN, USA). A representative<br />

formula describes baking powder by mass as starch (40.07%),<br />

sodium bicarbonate (26.73%), sodium aluminum sulfate (19.92%)<br />

and monocalcium phosphate (13.28%) (Matz, 1992). Room<br />

temperature (22 2 C) distilled water was added to produce 40,<br />

41, 42, 43, and 44% moisture percentages on a wet basis. In this<br />

work, we used 40% baking powder (db) which is much higher than<br />

industrial practices such as 4–6.25% for baking powder biscuits<br />

(Matz, 1992). This was done to accelerate the production <strong>of</strong> CO2 and<br />

enhance expansion in experiments.<br />

A. Penner et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 65–72 67<br />

Flour and baking powder were kept in sealed containers in<br />

a freezer at 4 C to prevent moisture absorption from the air. They<br />

were taken out <strong>of</strong> the freezer immediately before weighing and<br />

stored in the freezer immediately after the sample was measured.<br />

Flour and baking powder were blended by stirring 20 times by<br />

hand. Water was added all at once and immediately mixed for<br />

2 min at 86 rpm in a 100-g bowl pin mixer (National Manufacturing<br />

Company, Lincoln, NE, USA).<br />

3.2. Measurement <strong>of</strong> lateral expansion<br />

Upon completion <strong>of</strong> mixing, a dough sample (2–3 g) was placed<br />

between the two acrylic plates. Clamps were added to the sides to<br />

ensure a constant 2 mm distance between the plates. This also<br />

eliminated the vertical expansion component as a variable in<br />

experiments. The plates holding the dough were placed inside the<br />

pressure chamber and sealed. Periodic photographs show the<br />

dough disk expanding as seen in Fig. 3. Measurements were taken<br />

<strong>of</strong> the horizontal major diameters (x and y radial diameters) every<br />

minute for 30 min starting 4 min after mixing was completed. The<br />

4 min window was used to mount the sample in the chamber<br />

between the clear acrylic plates.<br />

Density was determined from the mass <strong>of</strong> dough (assumed<br />

constant) divided by the changing volume <strong>of</strong> dough.<br />

rðtÞ ¼ m<br />

TA ¼<br />

Where:<br />

m<br />

Tp D hþDv<br />

4<br />

2<br />

A is the area <strong>of</strong> the dough (mm 2 ) r is density <strong>of</strong> dough as<br />

a function <strong>of</strong> time<br />

m is mass <strong>of</strong> dough in grams<br />

T is sample thickness, constant at 2 mm<br />

Dh is the horizontal distance occupied by the dough at each time<br />

measurement (mm)<br />

Dv is the vertical distance occupied by the dough at each time<br />

measurement (mm)<br />

This method <strong>of</strong> computing area was validated using Origin 6.1<br />

Scientific Graphing and Data Analysis S<strong>of</strong>tware (OriginLab Corporation,<br />

Northampton, MA, USA). Images <strong>of</strong> the dough disk were<br />

digitized, and the area was computed by using numerical integration<br />

through Origin 6.1 s<strong>of</strong>tware. Area measurements from on-line<br />

experiments were hand calculated and the results were in close<br />

agreement. For example, a hand calculated sample area was<br />

13.50 mm 2 while the corresponding s<strong>of</strong>tware calculated the area at<br />

13.52 mm 2 .<br />

The experiments were monitored, varying: moisture content,<br />

baking powder concentration, pressure and vacuum cycles in the<br />

Fig. 3. Change in dough disk dimensions through the pressure scheme: (a) before compression (1 bar); (b) pressurization (2 bar); (c) after release (1 bar). Dotted circle added to<br />

show changes in dough dimensions.


68<br />

chamber, pressurizing gas, and the application <strong>of</strong> lubricants on the<br />

acrylic plates. All the experiments, except for those done to study<br />

the effect <strong>of</strong> the pressure pr<strong>of</strong>ile (discussed later), were done with<br />

the ‘standard pressurization method’:<br />

Minutes 0–3: The chamber pressure was maintained at 1 bar to<br />

allow the dough to rest in the expansion chamber. This allowed<br />

for initial CO2 production and temperature equilibration.<br />

Minutes 4–8: Pressure was increased to 2 bars by CO2 or<br />

compressed air application to the chamber.<br />

Minutes 9–30: Pressure was immediately released and maintained<br />

at 1 bar.<br />

Such a pressure pr<strong>of</strong>ile is necessary because it allows the<br />

observer to simulate conditions during rapid expansion or<br />

compression <strong>of</strong> the dough. The diameter was recorded throughout<br />

the procedure at 1 min intervals and the dimensionless ratio <strong>of</strong><br />

density at any time to the initial density was compared. To study<br />

pressure effects, the ‘pressure pulse method’ was examined:<br />

Minutes 0–3: Chamber pressure was maintained at 1 bar to<br />

allow the dough to rest in the expansion chamber.<br />

Minutes 4–8: Pressure was increased to 2 bars by CO2 application<br />

to the chamber.<br />

Minutes 9–13: Pressure was released and the chamber pressure<br />

was maintained at 1 bar.<br />

Minutes 14–19: A full vacuum was applied to the chamber.<br />

Minutes 20–30: Vacuum was released and chamber pressure<br />

was maintained at 1 bar.<br />

The final pressurization schema used to study pressure effects<br />

was the ‘pressure vacuum method’:<br />

Minutes 0–3: Chamber pressure was maintained at 1 bar to<br />

allow the dough to rest in the expansion chamber.<br />

Minutes 4–8: Pressure was increased to 2 bars by CO2 application<br />

to the chamber.<br />

Minutes 9–13: Pressure was released into a full vacuum<br />

environment in the chamber and thereby eliminated the<br />

re-equilibration phase after pressurization.<br />

A. Penner et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 65–72<br />

Minutes 14–30: Vacuum was released and chamber pressure<br />

was maintained at 1 bar.<br />

4. Results<br />

4.1. Effect <strong>of</strong> moisture content (40, 41, 42, 43, and 44%)<br />

During experimentation, after pressure was applied the density<br />

dramatically increased due to compression <strong>of</strong> the dough. As<br />

moisture content increased from 40 to 44%, greater compression<br />

and total expansion were produced. Under the standard pressurization<br />

method the high moisture dough (44% wb) yielded a 26%<br />

greater expansion compared to the low moisture dough (40% wb)<br />

with only 4% expansion as seen in Fig. 4. It was noted that 40 and<br />

41% moisture content doughs were dry and non-cohesive. This was<br />

in strong contrast to the 44% dough which was highly viscous. The<br />

dough made using 43% moisture content was chosen as the balance<br />

between the above extremes and was used in all other<br />

experiments.<br />

4.2. Effect <strong>of</strong> baking powder concentration<br />

Formulation consisting <strong>of</strong> 40% baking powder is significantly<br />

higher than current industrial standards <strong>of</strong> 4.25–6% (db) for baking<br />

powder products such as biscuits (Matz, 1992). However, this value<br />

produced rapid CO2 generation for monitoring dough expansion.<br />

Using 6% baking powder (db) in the dough, produced low expansion<br />

that also produced poor repeatability when monitoring<br />

expansion. The growth effects were seen more easily with 40% (db)<br />

baking powder formulation. This concentration was varied to<br />

investigate 30, 40, and 50% (db) baking powder in the formula;<br />

however there was no significant difference in the final expansion.<br />

The authors were most concerned with initial conditions, and<br />

therefore used 40% baking powder concentration with all reactions<br />

because the author had used this with many previous experiments<br />

and sought to standardize the concentration.<br />

4.3. Effect <strong>of</strong> pressure change<br />

Lateral expansion was measured under three pressure schemas<br />

that were a variation <strong>of</strong> the ‘standard pressurization method’, as<br />

Fig. 4. Change in dough density with varying moisture content where dough followed the standard pressurization method: 1 bar in region A, 2 bar in region B, and 1 bar in region C.


described above. The dough was placed directly on the acrylic<br />

plates as the effects <strong>of</strong> friction were explored later. For the ‘atmospheric<br />

pressure method’, the dough sample was at 1 bar the entire<br />

measurement period. This showed the dough disk expanding for<br />

15 min. When the standard pressurization method was applied, the<br />

density quickly increased as the dough contracted, and upon<br />

release the dough showed gradual expansion. Pressure was introduced<br />

into the system to keep the CO2 generated within the<br />

compressed dough and demonstrate increased expansion after<br />

pressure release. With application <strong>of</strong> vacuum, the bubble volume<br />

was expected to be enhanced. To evaluate the effects <strong>of</strong> a reduced<br />

pressure environment, the ‘standard vacuum method’ was applied.<br />

This followed the standard pressurization method, but instead <strong>of</strong><br />

applying 2 bars <strong>of</strong> pressure during minutes 4–8, a full vacuum was<br />

applied. The dough expanded with the vacuum, but upon release, it<br />

collapsed and density increased. The use <strong>of</strong> vacuum produced the<br />

least overall expansion with 15.1%. Atmospheric pressure and<br />

standard pressurization method gave 17.7, and 19.6% expansion<br />

respectively, as represented in Fig. 5. The vertical lines in the figures<br />

denote regions <strong>of</strong> different pressure.<br />

4.3.1. Pressure pulse (with/without vacuum)<br />

The pressure pulse experiments were performed by subjecting<br />

the dough to large pressure differentials using a combination <strong>of</strong><br />

2 bars <strong>of</strong> pressure or a full vacuum. In all cases, 5 ml Fluoro-Grip Ò -F<br />

Optically Clear (HD-FEP) Teflon Ò coating (Integument Technologies,<br />

Inc., Tonawanda, NY, USA) was applied on the acrylic plates as<br />

a lubricant to reduce friction and enhance the effects <strong>of</strong> multiple<br />

changes in pressure. The effect <strong>of</strong> lubrication is discussed in a later<br />

section. The first and second test case followed the ‘pressure pulse<br />

method’ as described above with the addition <strong>of</strong> Pam Ò aerosol<br />

lubricant (Pam Ò , International Home FoodsÓ, Parsippany, NJ, USA)<br />

to the second case. The third test case followed the ‘pressure<br />

vacuum’ pr<strong>of</strong>ile where vacuum was applied immediately after<br />

pressurization. The intermediate rest period that occurred during<br />

minutes 9–13 in case 1 and 2 led to a greater expansion than case 3<br />

Fig. 5. The effect <strong>of</strong> sudden pressure release and sudden increase on dough expansion<br />

where dough followed the standard pressurization method: 1 bar in region A, 2 bar in<br />

region B, and 1 bar in region C; atmospheric pressure: 1 bar in region A, B and C;<br />

standard vacuum method: 1 bar in region A, full vacuum in region B, and 1 bar in<br />

region C.<br />

A. Penner et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 65–72 69<br />

with 20.7, 11.7, and 1.9%, respectively. Case 3 experienced a continuous<br />

drop in pressure which led to a significantly lower final<br />

expansion. The addition <strong>of</strong> a pressurization period increased<br />

expansion by the potential for distributing gas cells that had begun<br />

growth during the initial incubation phase. This is seen in Fig. 6,<br />

where vertical lines denote changes in pressure. The repeated<br />

change in pressure was performed to study the expansion behavior<br />

<strong>of</strong> the dough after significant deformation.<br />

4.4. Effect <strong>of</strong> pressurizing gas (air or CO2)<br />

Compressed air and CO2 gas were compared as the pressurizing<br />

media to determine if the added environmental CO-2 affected the<br />

chemical reaction which produced CO2. Each gas was used to<br />

pressurize the chamber to 2 bars following the standard<br />

pressurization<br />

4.4.1. Method<br />

Using CO2 for pressurization produced a 5.8% greater increase in<br />

the final dough density as seen in Fig. 7.<br />

The explanation for this enhancement is that CO2 has a higher<br />

solubility in the dough than air. CO2 is produced internally from the<br />

baking powder (sodium bicarbonate component) releasing the CO2<br />

from the chemical reaction. Environmental CO2 must be solubilized<br />

and transferred to the gaseous phase in the existing gas cells before<br />

it can aid in expansion (Eliasson and Larson, 1993). This presents<br />

the possibility <strong>of</strong> diffusion effects when multiple gases, e.g. air and<br />

CO2 are present.<br />

Fig. 6. The effect <strong>of</strong> different pressure schema on dough expansion through the<br />

pressure pulse method: 1 bar in region A, 2 bar in region B, 1 bar in region C, full<br />

vacuum in region D, and 1 bar in region E; and pressure vacuum method: 1 bar in<br />

region A, 2 bar in region B, full vacuum in region C, 1 bar in region D and E.


70<br />

Fig. 7. Effect <strong>of</strong> CO2 versus air as the pressurizing agent on dough expansion where<br />

dough followed the standard pressurization method: 1 bar in region A, 2 bar in region<br />

B, and 1 bar in region C.<br />

4.5. Effect <strong>of</strong> lubrication<br />

The authors were concerned that lateral expansion was inhibited<br />

by friction between the acrylic plate and the dough. Reducing<br />

this friction using lubricants (Teflon Ò and Pam Ò ) was done to<br />

evaluate the role <strong>of</strong> friction in expansion. The following four cases<br />

were studied with the standard pressurization method: expansion<br />

without lubrication, with Teflon Ò , with Pam Ò and with both. As<br />

expected, during the pressure increase to 2 bars the dough disk<br />

compressed the greatest with both Teflon Ò and Pam Ò at 30.9%,<br />

followed by Pam Ò with 19.0%, Teflon Ò with 9.9%, and without<br />

Fig. 8. Effect <strong>of</strong> lubrication on dough expansion where dough followed the standard<br />

pressurization method: 1 bar in region A, 2 bar in region B, and 1 bar in region C.<br />

A. Penner et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 65–72<br />

lubrication at 4.9%. Following the pressure release, the greatest<br />

expansion was with Teflon Ò and Pam Ò which expanded 40.1%<br />

compared to Pam Ò at 30.2% or Teflon Ò at 20.9%. It is interesting to<br />

note that no lubrication (adding neither Teflon Ò nor Pam) gave<br />

25.4% expansion which is greater than Teflon Ò alone as shown in<br />

Fig. 8. However, the difference in using Teflon Ò and no lubrication<br />

was not significant. It is <strong>of</strong> worth to note that in Fig. 6, Teflon Ò alone<br />

was compared to both Teflon and Pam Ò using the pressure pulse<br />

method. The combination <strong>of</strong> Teflon Ò and Pam Ò led to an overall<br />

expansion <strong>of</strong> 11.7% compared to Teflon Ò alone which was at 20.7%.<br />

The author would suggest that the friction between the dough and<br />

Teflon Ò surface restricted the dough’s movement after the full<br />

vacuum was released.<br />

The effect <strong>of</strong> lubrication was an indirect study <strong>of</strong> the viscosity <strong>of</strong><br />

the dough through the interaction with the stationary plates. The<br />

combination <strong>of</strong> dough and Pam Ò can be considered a two-liquid<br />

flow, while the Teflon Ò is considered stationary. The difference<br />

between the observed behaviors in the two cases would translate<br />

into the effect <strong>of</strong> wall roughness versus a liquid film.<br />

During expansion, bubbles were seen emerging at the Pam Ò -<br />

dough interface. Monitoring 1 s intervals at the interface, allows<br />

one to notice the emergence <strong>of</strong> bubbles that become entrapped<br />

within the Pam Ò liquid. This progression <strong>of</strong> bubble growth is<br />

highlighted with arrows in Fig. 9 which shows the same bubble<br />

increasing. The gas released from the dough matrix is believed to be<br />

the driving force for creation <strong>of</strong> the bubbles along the plates.<br />

5. Discussion<br />

A series <strong>of</strong> experiments was performed to study the lateral<br />

expansion <strong>of</strong> dough during rapid expansion and compression.<br />

Experiments under varying moisture contents (40, 41, 42, 43, and<br />

44%) showed a decrease in density with increase in the moisture<br />

content <strong>of</strong> the dough. This is consistent with the work <strong>of</strong> Elmhedi<br />

et al. (2007) who showed that dough density decreased as a function<br />

<strong>of</strong> fermentation time for yeast dough using digital imaging. Lai<br />

et al. (1989) reported that the addition <strong>of</strong> sodium bicarbonate<br />

contributed a weakened structure in wheat dough.<br />

One can observe the dramatic difference in the final expansion<br />

through the change in the pressure schema. It is expected that CO2<br />

is removed more quickly using a full vacuum which reduced the<br />

amount <strong>of</strong> retained gas which would have potentially formed<br />

bubbles. The presence <strong>of</strong> a short residence period <strong>of</strong> 1 bar pressure<br />

following pressurization in the pressure pulse experiments falls in<br />

line with this hypothesis: the rapid shift to vacuum would help<br />

deprive the dough <strong>of</strong> CO2. Application <strong>of</strong> vacuum after an initial<br />

growth period was used to understand dough behavior in<br />

a reduced pressure environment.<br />

The viscoelastic properties <strong>of</strong> dough may account for the<br />

expansion after removal <strong>of</strong> the deformation induced by pressure or<br />

vacuum. The elasticity <strong>of</strong> the dough is related to the release <strong>of</strong><br />

potential energy stored during the deformation by either extension<br />

or compression forces (Eliasson and Larsson, 1993). The growth<br />

phenomena under vacuum and pressure applications appear to<br />

qualitatively mirror each other. A comparison <strong>of</strong> pressurization by<br />

compressed air and CO2 showed very similar behavior, with CO2<br />

showing slightly more expansion and compression tendencies.<br />

These effects were minor and would not indicate a significant mass<br />

transfer. One possible occurrence for the small discrepancy could<br />

be the ability <strong>of</strong> increased CO2 in the pressurized atmosphere to be<br />

incorporated into the dough matrix.<br />

The authors were concerned about potential reduction in the<br />

dough’s final expansion due to friction at the dough-plate interface.<br />

To overcome this challenge, Elmhedi et al. (2007) applied mineral<br />

oil during expansion testing to reduce air bubble entrapment at the<br />

interface <strong>of</strong> the measuring vessel and the dough. Teflon Ò film


Fig. 9. Rapid growth <strong>of</strong> air bubbles through entrapment at the dough-Pam Ò interface:<br />

(a) top at 9:22.00 min; (b) middle at 9:22.30 min; (c) bottom 9:23.00 min.<br />

coating and Pam Ò aerosol spray were used to reduce friction for<br />

these experiments. These additions did not appear to affect the<br />

qualitative expansion behavior, but they did enhance the expansion<br />

and compression effects observed. The compression was significantly<br />

different for pressure pulse experiments when the<br />

A. Penner et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 65–72 71<br />

combination <strong>of</strong> Teflon Ò coating with Pam Ò lubrication was<br />

compared to only Teflon Ò coating as seen in Fig. 6. As expected,<br />

combining Teflon Ò and Pam Ò yielded the greatest expansion, but it<br />

was interesting to note that Teflon Ò had little effect by itself.<br />

6. Summary and future work<br />

Experimental studies for bubble growth are inconvenienced by<br />

the methods’ inability to capture the dynamic phenomena. In this<br />

paper, a versatile experimental method is developed to allow for<br />

macroscopic expansion <strong>of</strong> wheat dough. The study evaluated<br />

expansion <strong>of</strong> a dough disk under varying: moisture content (40, 41,<br />

42, 43, and 44% wb), leavening acid concentration (30, 40, and 50%<br />

db), pressure schemas, pressurizing gas composition (compressed<br />

air and CO2), and lubrication (Teflon Ò film coating and Pam Ò<br />

aerosol lubricant). Dough expansion increased 22.6% by increasing<br />

moisture content from 40 to 44% (wb). At 44% moisture (wb), this<br />

dough was highly viscous and 43% moisture content was chosen for<br />

all experiments. Baking powder was used at an increased level, 40%<br />

(db) to enhance initial growth conditions and facilitate rapid CO2<br />

production. ‘Pressure pulse’ and ‘pressure vacuum methods’ added<br />

pressurization alternatively with full vacuum. The former method<br />

included a rest period before vacuum application which increased<br />

expansion by 10.8%. The addition <strong>of</strong> a pressurization period aided<br />

expansion through potential for the promotion <strong>of</strong> distributing gas<br />

cells that had begun growth during the initial incubation phase.<br />

Teflon Ò and Pam Ò reduced friction between the dough and acrylic<br />

plate to increase the final expansion by 14.7% compared to no<br />

lubricant (‘standard pressurization method’). ‘Pressure pulse’ and<br />

‘pressure vacuum’ experiments decreased expansion by 28.4 and<br />

38.2%, respectively compared to ‘standard pressurization’ while<br />

using Teflon Ò and Pam Ò .<br />

Several extensions <strong>of</strong> this work are possible. Work is currently<br />

underway to model the reaction rate <strong>of</strong> CO2 production within<br />

a dough matrix. In addition to this, other upcoming extensions are<br />

to include the determination <strong>of</strong> whether the improved expansion<br />

for higher moisture content was a result <strong>of</strong> improved consistency or<br />

increased reaction rate. The incorporation <strong>of</strong> lateral mass transfer<br />

could be studied by drilling holes through the acrylic plates. Finally,<br />

work is to be done on the study <strong>of</strong> air entrapment and its effect on<br />

lubrication with an emphasis on possible creation <strong>of</strong> an air cushion<br />

between the dough and the acrylic surface.<br />

Acknowledgements<br />

The authors would like to thank Apri Melinda, Airine Airine,<br />

Reinaldo Kusliawan and Louis Nelson III for their invaluable help in<br />

developing the methodology. We would also like to thank the SURF<br />

program and NASA-SBIR grant for providing the funding.<br />

References<br />

Aguilera, J.M., Germain, J.C., 2007. Advances in image analysis for the study <strong>of</strong> food<br />

microstructure. In: McClements, D.J. (Ed.), Understanding and Controlling the<br />

Microstructure <strong>of</strong> Complex Food. Woodhead Publishing Limited, Cambridge, pp.<br />

261–287.<br />

Alavi, S.H., Gogoi, B.K., Khan, M., Bowman, B.J., Rizvi, S.S.H., 1999. Structural properties<br />

<strong>of</strong> protein-stabilized starch-based supercritical fluid extrudates. Food<br />

Research International 32, 107–118.<br />

Alavi, S.H., Rizvi, S.S.H., Harriot, P., 2003. Process dynamics <strong>of</strong> starch based microcellular<br />

foams produced by supercritical fluid extrusion I: model development<br />

and II. Numerical simulation and experimental evaluation. Food Research<br />

International 6, 309–330.<br />

AOAC International, 2000. Official Methods <strong>of</strong> Analysis, sixteenth ed. AOAC International,<br />

Gaithersburg, MD.<br />

Babin, P., Della Valle, G., Chiron, H., Cloetens, P., Hoszowska, J., Pernot, P.,<br />

Reguerre, A.L., Salvo, L., Dendievel, R., 2006. Fast X-ray tomography analysis <strong>of</strong><br />

bubble growth and foam setting during breadmaking. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

43 (3), 393–397.


72<br />

Babin, P., Della Valle, G., Dendievel, R., Lourdin, D., Salvo, L., 2007. X-ray tomography<br />

study <strong>of</strong> the cellular structure <strong>of</strong> extruded starches and its relation with<br />

expansion phenomenon and foam mechanical properties. Carbohydrate Polymers<br />

68 (2), 329–340.<br />

Baker, J.C., Mize, M.D., 1941. The origin <strong>of</strong> the gas cell in bread dough. <strong>Cereal</strong><br />

Chemistry 18, 19–34.<br />

Bell, A.V., Berger, K.G., Russo, J.V., White, G.W., Weathers, T.L., 1975. A study <strong>of</strong> the<br />

micro-baking <strong>of</strong> sponges and cakes using cine and television microscopy.<br />

<strong>Journal</strong> <strong>of</strong> Food Technology 10, 147–156.<br />

Bellido, G.G., Scanlon, M.G., Page, J.H., Hallgrimsson, B., 2006. The bubble size<br />

distribution in wheat flour dough. Food Research International 39 (10),<br />

1058–1066.<br />

Campbell, G.M., Reilly, C.D., Fryer, P.J., Sadd, P.A., 1998. Aeration <strong>of</strong> bread dough<br />

during mixing: effect <strong>of</strong> mixing dough at reduced temperature. <strong>Cereal</strong> Food<br />

World 433, 163–167.<br />

Cauvain, S.P., Whitworth, M.B., Alava, J.M., 1999. The evolution <strong>of</strong> bubble structure<br />

in bread doughs and its effect on bread structure. In: Campbell, G.M., Webb, C.,<br />

Pandiella, S.S., Niranjan, K. (Eds.), Bubbles in Foods. Eagan Press, St. Paul, MN,<br />

pp. 85–88.<br />

Chen, K.H., Rizvi, S.S.H., 2006. Rheology and expansion <strong>of</strong> starch-water-CO2<br />

mixtures with controlled gelatinization by supercritical fluid extrusion. International<br />

<strong>Journal</strong> <strong>of</strong> Food Properties 9, 863–876.<br />

de Cindio, B., Correra, S., 1995. Mathematical modeling <strong>of</strong> leavened cereal goods.<br />

<strong>Journal</strong> <strong>of</strong> Food Engineering 24, 379–403.<br />

Eliasson, A.C., Larson, K., 1993. <strong>Cereal</strong>s in Breadmaking – A Molecular Colloidal<br />

Approach. Marcel Dekker, Inc, New York.<br />

Elmehdi, H.M., Page, J.H., Scanlon, M.G., 2003. Using ultrasound to investigate the<br />

cellular structure <strong>of</strong> bread crumb. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 38, 33–42.<br />

Elmehdi, H.M., Page, J.H., Scanlon, M.G., 2007. Evaluating dough density changes<br />

during fermentation by different techniques. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> Chemistry 84,<br />

2007.<br />

Fan, J., Mitchell, J.R., Blanchard, J., 1994. A computer simulation <strong>of</strong> the dynamics <strong>of</strong><br />

bubble growth and shrinkage during extrudate expansion. <strong>Journal</strong> <strong>of</strong> Food<br />

Engineering 23, 337–356.<br />

Fan, J., Mitchell, J.R., Blanchard, J.M.V., 1999. A model for the oven rise <strong>of</strong> dough<br />

during baking. <strong>Journal</strong> <strong>of</strong> Food Engineering 41, 69–77.<br />

Fox, P., Smith, P.P., Sahi, S., 2004. Ultrasound measurements to monitor the specific<br />

gravity <strong>of</strong> food batters. <strong>Journal</strong> <strong>of</strong> Food Engineering 65 (3), 317–324.<br />

Fyrillas, M.M., Kloek, W., van Vliet, T., Mellema, J., 2000. Factors determining the<br />

stability <strong>of</strong> a gas cell in an elastic medium. Langmuir 16, 1014–1019.<br />

Hailemariam, L., Okos, M., Campanella, O., 2007. A mathematical model for the<br />

isothermal growth <strong>of</strong> bubbles in wheat dough. <strong>Journal</strong> <strong>of</strong> Food Engineering 82,<br />

466–477.<br />

http://www.nutritiondata.com/facts-C00001-01c21Ke.html Retrieved on 8/28/<br />

2007.<br />

A. Penner et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 65–72<br />

Kumagai, H., Lee, B.H., Kumagai, H., Yano, T., 1991. Critical radius and time course <strong>of</strong><br />

expansion <strong>of</strong> an isolated bubble in wheat flour dough under temperature rise.<br />

Agricultural and Biological Chemistry 55 (4), 1081–1087.<br />

Lai, C.S., Guetzlaff, J., Hoseney, R.C., 1989. Role <strong>of</strong> sodium bicarbonate and trapped<br />

air in extrusion. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> Chemistry 66, 69–73.<br />

Lee, S., Inglett, G.E., 2006. Rheological and physical evaluation <strong>of</strong> jet-cooked oat<br />

bran in low calorie cookies. International <strong>Journal</strong> <strong>of</strong> Food <strong>Science</strong> and Technology<br />

41, 553–559.<br />

Matz, S.A., 1992. Bakery Technology and Engineering, third ed. Van Nostrand<br />

Reinhold, New York.<br />

Moraru, C.I., Kokini, J.L., 2003. Nucleation and expansion during extrusion and<br />

microwave heating <strong>of</strong> cereal foods. Comprehensive Reviews in Food <strong>Science</strong><br />

and Food Safety 2, 120–138.<br />

Schwartzberg, H.G., Wu, J., Nussinovitch, A., Mugerwa, J., 1995. Modeling deformation<br />

and flow during vapor-induced puffing. <strong>Journal</strong> <strong>of</strong> Food Engineering 25,<br />

329–372.<br />

Shimiya, Y., Yano, T., 1988. Rates <strong>of</strong> shrinkage and growth <strong>of</strong> air bubbles entrained<br />

in wheat flour dough. Agricultural and Biological Chemistry 52 (11), 2879–<br />

2883.<br />

Singh, A.P., Bhattacharya, M., 2005. Development <strong>of</strong> dynamic modulus and cell<br />

opening <strong>of</strong> dough during baking. <strong>Journal</strong> <strong>of</strong> Texture Studies 36, 44–67.<br />

Strybulevych, A., Leroy, V., Scanlon, M.G., Page, J.H., 2007. Characterizing a model<br />

food gel containing bubbles and solid inclusions using ultrasound. S<strong>of</strong>t Matter 3<br />

(11), 1388–1394.<br />

Swanson, R.B., Carden, L.A., Parks, S.S., 1999. Effect <strong>of</strong> a carbohydrate-based fat<br />

substitute and emulsifying agents on reduced-fat peanut butter cookies. <strong>Journal</strong><br />

<strong>of</strong> Food Quality 22, 19–29.<br />

Taki, K., Nakayama, T., Yatsuzuka, T., Ohshima, M., 2003. Visual observations <strong>of</strong> batch<br />

and continuous foaming processes. <strong>Journal</strong> <strong>of</strong> Cellular Plastics 39, 155–169.<br />

Trater, A.M., Alavi, S., Rizvi, S.S.H., 2005. Use <strong>of</strong> non-invasive X-ray microtomography<br />

for characterizing microstructure <strong>of</strong> extruded biopolymer foams.<br />

Food Research International 38 (6), 709–719.<br />

Wagner, M., Zhang, L., Quellec, S., Doursat, C., Flick, D., Trystram, G., Lucas, T., 2006.<br />

Role <strong>of</strong> the crust formation on local expansion during bread baking. In:<br />

Campbell, G.M., Pyle, L., Scanlon, M. (Eds.), Proceedings <strong>of</strong> Bubbles in Food:<br />

Novelty, Health and Luxury, September 10–14, Windermere (UK).<br />

Wang, L., Ganjyal, G.M., Jones, D.D., Weller, C.L., Hanna, M.A., 2005. Modeling <strong>of</strong><br />

bubble growth dynamics and nonisothermal expansion in starch-based foams<br />

during extrusion. Advances in Polymer Technology 24 (1), 29–45.<br />

Whitworth, M.B., Alava, J.M., 1999. The imaging and measurement <strong>of</strong> bubbles in<br />

bread doughs. In: Campbell, G.M., Webb, C., Pandiella, S.S., Niranjan, K. (Eds.),<br />

Bubbles in Foods. Eagan Press, St. Paul, MN, pp. 221–231.<br />

Zoulias, E., Oreopoulou, V., Kounalaki, E., 2002. Effect <strong>of</strong> fat and sugar replacement<br />

on cookie properties. <strong>Journal</strong> <strong>of</strong> the <strong>Science</strong> <strong>of</strong> Food and Agriculture 82, 1637–<br />

1644.


Digestibility <strong>of</strong> protein and starch from sorghum (Sorghum bicolor)<br />

is linked to biochemical and structural features <strong>of</strong> grain endosperm<br />

Joshua H. Wong a , Tsang Lau a , Nick Cai a , Jaswinder Singh a,2 , Jeffrey F. Pedersen b , William H. Vensel c ,<br />

William J. Hurkman c , Jeff D. Wilson d , Peggy G. Lemaux a, *, Bob B. Buchanan a,1<br />

a Department <strong>of</strong> Plant and Microbial Biology, University <strong>of</strong> California, Berkeley, CA 94720, USA<br />

b United States Department <strong>of</strong> Agriculture, Agricultural Research Service, Lincoln, NE 68583, USA<br />

c United States Department <strong>of</strong> Agriculture, Agricultural Research Service, Western Regional Research Center, Albany, CA 94710, USA<br />

d United States Department <strong>of</strong> Agriculture, Agricultural Research Service, Manhattan, KS 66502, USA<br />

article info<br />

Article history:<br />

Received 21 December 2007<br />

Received in revised form 9 July 2008<br />

Accepted 9 July 2008<br />

Keywords:<br />

Disulfide proteins<br />

Starch–protein interface<br />

Granule-bound starch synthase I<br />

Sorghum<br />

1. Introduction<br />

abstract<br />

Grain sorghum (Sorghum bicolor L. Moench) ranks fifth in<br />

worldwide production among cereal crops, after wheat, rice, corn<br />

and barley. The U.S. is the number two producer and the number<br />

one exporter <strong>of</strong> sorghum, primarily to Mexico for use as animal feed<br />

(U.S. Grains Council, 2006). Popularity <strong>of</strong> sorghum is due in part to<br />

its ability to produce reasonable yields in warmer, drier regions.<br />

Because subsistence farmers in Africa and Asia cultivate sorghum<br />

widely as a staple food for home consumption, the crop is<br />

Abbreviations: DTT, dithiothreitol; GBSS, granule-bound starch synthase;<br />

IVDMD, in vitro dry matter disappearance; ME, 2-mercaptoethanol; Mr, relative<br />

molecular mass; PAGE, polyacrylamide gel electrophoresis; SDS, sodium dodecyl<br />

sulfate; MOPS, 3-(N-morpholino) propane sulfonic acid.<br />

* Corresponding author. Tel.: þ1 5106421589; fax: þ1 5106427356.<br />

E-mail addresses: viewmont@nature.berkeley.edu (J.H. Wong), shout_at_neko@<br />

hotmail.com (T. Lau), monto@nature.berkeley.edu (N. Cai), jaswinder.singh@<br />

mcgill.ca (J. Singh), jeff.pedersen@ars.usda.gov (J.F. Pedersen), vensel@pw.usda.gov<br />

(W.H. Vensel), bhurkman@pw.usda.gov (W.J. Hurkman), jeff.d.wilson@ars.usda.gov<br />

(J.D. Wilson), lemauxpg@nature.berkeley.edu (P.G. Lemaux), view@nature.<br />

berkeley.edu (B.B. Buchanan).<br />

1<br />

Submitting author.<br />

2<br />

Present address: McGill University, Macdonald Campus, 21,111 Lakeshore Road,<br />

Ste-Anne-de-Bellevue, Quebec H9X 3V9, Canada.<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.07.013<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 73–82<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

Although a principal source <strong>of</strong> energy and protein for millions <strong>of</strong> the world’s poorest people, the<br />

nutritional value <strong>of</strong> sorghum (Sorghum bicolor L. Moench) is diminished because <strong>of</strong> low digestibility <strong>of</strong><br />

grain protein and starch. To address this problem, we analyzed the properties <strong>of</strong> two sorghum lines that<br />

have a common pedigree but differ in digestibility. Consistent with results based on a ruminal fluid assay,<br />

the protein and starch <strong>of</strong> one line (KS48) was more thoroughly digested than that <strong>of</strong> the other (KS51)<br />

using in vitro assays based on pepsin and a-amylase. The indigestibility <strong>of</strong> KS51 relative to KS48 was<br />

shown to be due to (i) a greater abundance <strong>of</strong> disulfide-bonded proteins; (ii) presence in KS51 <strong>of</strong> nonwaxy<br />

starch and the accompanying granule-bound starch synthase; and (iii) the differing nature <strong>of</strong> the<br />

protein matrix and its interaction with starch. The current findings suggest that each <strong>of</strong> these factors<br />

should be considered in efforts to enhance the nutritional value <strong>of</strong> sorghum grain.<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

a principal source <strong>of</strong> energy and protein for millions <strong>of</strong> the world’s<br />

poorest (Klopfenstein and Hoseney, 1995).<br />

Seed proteins <strong>of</strong> sorghum are less digestible than those <strong>of</strong> other<br />

cereals and digestibility is exacerbated by wet cooking the meal or<br />

flour, which results in significant nutritional losses (Hamaker et al.,<br />

1986; MacLean et al., 1981; Mertz et al., 1984). Numerous factors<br />

contribute to the digestibility problem [for review, see (Duodu<br />

et al., 2003)]. Exogenous aspects include the interaction <strong>of</strong> protein<br />

with non-protein components, such as polyphenols, starch, nonstarch<br />

polysaccharides, phytates and lipids. Endogenous factors<br />

arise from the nature <strong>of</strong> the proteins themselves and their organization<br />

within the grain (Belton et al., 2006; Duodu et al., 2003;<br />

Ezeogu et al., 2005, 2008; Hamaker and Bugusu, 2003). As in other<br />

cereals, the low amount <strong>of</strong> protein relative to starch, i.e., approximately<br />

10% protein, vs. 70–80% starch, based on grain dry weight<br />

(Rooney and Pflugfelder, 1986), affects the functional properties <strong>of</strong><br />

starch, such as gelatinization and digestion rate, to a greater extent<br />

in sorghum than in other cereals (Chandrashekar and Kirleis, 1988;<br />

Duodu et al., 2002; Ezeogu et al., 2005, 2008).<br />

The proteins <strong>of</strong> the sorghum grain are classically divided, based on<br />

solubility in different solvents (Jambunathan et al., 1975; Landry and<br />

Moureaux, 1970): albumins (water-soluble), globulins (salt-soluble),<br />

kafirins (prolamins, aqueous alcohol-soluble), cross-linked kafirins<br />

(aqueous alcohol þ reducing agent-soluble), cross-linked glutelins<br />

(detergent þ reducing agent þ alkaline pH-soluble) and unextracted


74<br />

structural protein residue. A newer and more simplified classification<br />

scheme for sorghum proteins has been proposed that divides them<br />

into two groups, kafirins and non-kafirins. This scheme is based on<br />

the homogeneous nature and varied origin <strong>of</strong> the kafirin storage<br />

prolamins relative to the heterogeneous nature <strong>of</strong> the non-kafirin<br />

proteins (i.e., albumins, globulins and glutelins) that are involved in<br />

cellular functions (Hamaker and Bugusu, 2003; Hamaker et al.,1995).<br />

Despite a high degree <strong>of</strong> sequence homology to maize zeins, sorghum<br />

storage proteins contain a higher proportion <strong>of</strong> cross-linked fractions<br />

and are more hydrophobic, explaining their greater propensity to<br />

form intermolecular disulfide-cross linkages and possibly additional<br />

protein aggregates that could facilitate the formation <strong>of</strong> more covalent<br />

bonds compared to maize zeins (Belton et al., 2006; Hamaker<br />

and Bugusu, 2003).<br />

Kafirins, comprising 70–80% <strong>of</strong> the protein in whole grain<br />

sorghum flour (Hamaker et al., 1995), are synthesized and translocated<br />

into the lumen <strong>of</strong> the endoplasmic reticulum where they<br />

form protein bodies (Taylor et al., 1985). They are subclassified as:<br />

a- (23 and 25 kDa), b- (20 kDa) and g- (28 kDa) types based on<br />

molecular weight, extractability, structure and cross-reactivity with<br />

sera against analogous maize zeins (Belton et al., 2006; Mazhar<br />

et al.,1993; Shull et al., 1991). Comprising w80% <strong>of</strong> total kafirins, the<br />

a-type is considered the principal storage protein, followed by the<br />

g- (w15%) and b- (w5%) members. Recently, a DNA-derived<br />

sequence for a d-kafirin was shown to have high homology with the<br />

Mr 10,000 d-zein, except for absence <strong>of</strong> part <strong>of</strong> the methionine-rich<br />

region (Belton et al., 2006; Izquierdo and Godwin, 2005). a-Kafirins<br />

have low levels <strong>of</strong> cysteine relative to the b- and g-types (5 and<br />

7 mol%, respectively), whereas d-kafirins are rich in methionine<br />

(16–18 mol%) but lack cysteine. Electron microscopy <strong>of</strong> internal<br />

protein body structure reveals that g-kafirins, and to a lesser extent<br />

b-kafirins, encapsulate the more digestible a-kafirins in a disulfidebond<br />

polymer network (Oria et al., 2000; Shull et al., 1992), thereby<br />

impeding exposure to proteases.<br />

In the corneous endosperm, non-kafirins (albumins, globulins,<br />

glutelins) form around protein bodies, effectively ‘‘gluing’’ the<br />

bodies into a matrix surrounding the starch granules (Hamaker and<br />

Bugusu, 2003; Shull et al., 1990; Taylor et al., 1984). This protein<br />

matrix appears to act as a barrier to starch gelatinization and<br />

digestibility (Chandrashekar and Kirleis, 1988; Duodu et al., 2002;<br />

Ezeogu et al., 2005, 2008) due to cross-linking between g- and bkafirins<br />

and matrix proteins (Duodu et al., 2001; Hamaker and<br />

Bugusu, 2003; Oria et al., 1995a). Cooking reduces digestibility by<br />

effecting a conformational change in proteins that could facilitate<br />

formation <strong>of</strong> disulfide-linked polymers (Axtell et al., 1981; Duodu<br />

et al., 2002, 2003; Hamaker et al., 1987; Oria et al., 1995b). The<br />

negative impact <strong>of</strong> cooking on protein digestibility was mitigated<br />

by addition <strong>of</strong> 2-mercaptoethanol (ME) or other reducing agents<br />

(Elkhalifa et al., 1999; Hamaker et al., 1987). Sorghum grains rich in<br />

kafirin-containing protein bodies also have a lower capacity for<br />

starch gelatinization (Chandrashekar and Kirleis, 1988; Ezeogu<br />

et al., 2005, 2008)dan observation consistent with the finding that<br />

adding ME during cooking to cleave disulfide bonds within the<br />

protein matrix increased the degree <strong>of</strong> starch gelatinization and<br />

digestion (Elkhalifa et al., 1999; Ezeogu et al., 2005, 2008; Zhang<br />

and Hamaker, 1998).<br />

The protein barrier surrounding the starch granule may also<br />

reduce the hydrolysis <strong>of</strong> native and processed starch by amylolytic<br />

enzymes (Rooney and Pflugfelder, 1986). Thus, the addition <strong>of</strong><br />

pronase to hydrolyze the protein matrix significantly enhanced in<br />

vitro rates <strong>of</strong> starch hydrolysis by increasing surface area and<br />

enabling starch to interact with a-amylase and amyloglucosidase<br />

(Rooney and Pflugfelder, 1986). Similar effects were observed when<br />

sorghum flour was either treated with pepsin before cooking or<br />

cooked in the presence <strong>of</strong> dithiothreitol (DTT) or other reductants<br />

(Zhang and Hamaker, 1998). Abundance <strong>of</strong> starch granules may also<br />

J.H. Wong et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 73–82<br />

decrease proteolysis by limiting accessibility <strong>of</strong> proteolytic<br />

enzymes, especially when gelatinized during cooking (Duodu et al.,<br />

2002). Uniqueness <strong>of</strong> the protein matrix and its interaction with<br />

starch that affect the rate <strong>of</strong> starch digestion are key differences<br />

between the feed quality <strong>of</strong> sorghum and corn (Rooney and Miller,<br />

1982; Rooney and Pflugfelder, 1986). In addition, endosperm<br />

texture and cooking conditions have been shown to have a significant<br />

effect on in vitro digestibility <strong>of</strong> starch and protein in these two<br />

cereals (Duodu et al., 2002, 2003; Ezeogu et al., 2005, 2008).<br />

Sorghum displays significant variation in rates <strong>of</strong> starch disappearance<br />

(Wester et al., 1992). Based on a 12-h incubation in the in<br />

vitro dry matter disappearance (IVDMD) assay, loss <strong>of</strong> dry matter in<br />

sorghum correlated closely with rates <strong>of</strong> starch, but not necessarily<br />

protein digestion (Pedersen et al., 2000).<br />

Collectively, these findings suggest that properties <strong>of</strong> starch and<br />

protein in sorghum could affect their mutual digestibility. The<br />

majority <strong>of</strong> published references deal with digestibility <strong>of</strong> the<br />

proteins and their impact on starch, but evidence for the reverse is<br />

scant (Duodu et al., 2002). To pursue this issue and assess the<br />

importance <strong>of</strong> relevant factors in a single study, we investigated<br />

how protein and starch influence the breakdown <strong>of</strong> one another by<br />

comparing two sorghum lines having a common pedigree but<br />

differing in digestibility.<br />

2. Materials and methods<br />

2.1. Grain and preparation <strong>of</strong> materials<br />

The sorghum lines KS48 and KS51 were selected for these<br />

experiments because <strong>of</strong> previously determined differences in<br />

digestibility but the identical pedigree, Texioca-63 Short Kaura<br />

(Casady, 1972). Both have a clear pericarp and yellow endosperm.<br />

Field-grown seeds were produced at the University <strong>of</strong> Nebraska<br />

Agricultural Research Development Center, Ithaca, NE; additional<br />

seeds were generated from greenhouse-grown plants at the<br />

University <strong>of</strong> California, Berkeley. Seed size and hardness <strong>of</strong> KS48<br />

and KS51 were assessed using a Perten SKCS 4100 Single Kernel<br />

Characterization System instrument. The starch content <strong>of</strong> KS48<br />

and KS51 was 64.7% and 67.8%, respectively, on a dry weight basis.<br />

For 12-h IVDMD analysis, mature dry seeds were ground in a Wiley<br />

mill to pass a 20-mesh screen. For all other analyses, seeds were<br />

ground in a Wiley mill to pass a 40-mesh screen. Ground meal was<br />

stored at 25 C until used in digestion assays. Pepsin (porcine<br />

stomach mucosa, P-7000) and three types <strong>of</strong> a-amylases – bacterial<br />

(A-3403, Type XII-A), porcine pancreas (A-3176, Type VI-B) and<br />

human saliva (A-1031, Type XIII-A) – were used in digestibility<br />

assays (Sigma, St Louis, MO).<br />

2.2. Microscopy studies<br />

For microscopic observations, seeds were sectioned with a razor<br />

blade and stained with iodine (0.3 g I2, 1.0 g KI in 100 ml water).<br />

Sectioned seeds were used for electron microscopic analyses, using<br />

a Hitachi TM 1000 environmental scanning electron microscope<br />

(Hitachi High Technologies America, Inc., Pleasanton, CA). Prior to<br />

analysis, sections were coated with gold-palladium to 1 mm using<br />

a Tousimis sputter machine (Rockville, MD).<br />

2.3. Twelve-hour IVDMD <strong>of</strong> sorghum meal<br />

Sorghum meal was digested for 12 h in rumen fluid inoculum<br />

obtained from a ruminally fistulated steer using previously<br />

described methods (Pedersen et al., 2000).


2.4. Sequential protein extraction<br />

Sorghum meal (2 g) was sequentially extracted with the six<br />

solvents listed below in 20 ml lots at 25 C and centrifuged<br />

(18,900g, 10 min) at 4 C after each step [modified from (Hamaker<br />

et al., 1986)]. Initially, meal was extracted with 0.5 M NaCl by<br />

shaking for 60 min. Resulting NaCl-soluble albumins and globulins<br />

(fraction 1) were retained, and the pellet sequentially extracted<br />

yielding the following supernates: double distilled H2O (ddH2O) for<br />

20 min (water-wash fraction, fraction 2); 60% 2-propanol (v/v) for<br />

4 h (kafirins, fraction 3); 0.1 M borate buffer, pH 10.8, for 4 h (glutelin-like<br />

protein, fraction 4); 60% 2-propanol with 1% dithiothreitol<br />

(DTT) for 4 h (cross-linked kafirins, fraction 5); and 0.1 M borate<br />

buffer, pH 10.8, containing 1% DTT and 1% sodium dodecyl sulfate<br />

(SDS) for 18 h at 4 C (cross-linked glutelins, fraction 6). Protein<br />

fractions were stored at 20 C.<br />

2.5. Disulfide protein content <strong>of</strong> selected fractions<br />

Equal amounts <strong>of</strong> protein in fractions extracted without<br />

reducing agent were reduced with 1 mM DTT; protein sulfhydryl<br />

groups were labeled with monobromobimane (mBBr) and analyzed<br />

using SDS-PAGE (Wong et al., 2004). Relative intensity <strong>of</strong> fluorescence<br />

<strong>of</strong> labeled proteins was equated to relative amount <strong>of</strong> protein<br />

disulfides.<br />

2.6. Rapid in vitro pepsin digestion<br />

Time course <strong>of</strong> in vitro pepsin digestion <strong>of</strong> sorghum meal and<br />

processing <strong>of</strong> the resulting undigested residue were as described<br />

(Aboubacar et al., 2001; Nunes et al., 2004).<br />

2.7. Extraction <strong>of</strong> protein undigested by pepsin<br />

Protein in pellets, not digested by pepsin, was extracted with<br />

0.0125 M borate buffer, pH 10 containing 1% SDS and 2% 2-ME as<br />

described (Aboubacar et al., 2001). The extraction was modified<br />

further to include extraction <strong>of</strong> a granule-bound starch synthase I<br />

(GBSSI) from the borate–SDS–ME extracted residue (see Section 2.8).<br />

2.8. Extraction <strong>of</strong> granule-bound starch synthase<br />

To determine if unextracted protein remained in residues after<br />

extraction with borate–SDS–ME (0.0125 M borate, 1% SDS, 2% ME,<br />

pH 10), the washed pellets (in 0.5 ml <strong>of</strong> the same buffer) were<br />

mixed with 1 ml borate–SDS–ME buffer and extracted for 5–10 min<br />

in a boiling water bath and cooled in tap water. Gelatinized starch<br />

was pelleted by centrifugation (20,800g, 10 min) and the clear<br />

supernate was saved for SDS-PAGE analysis.<br />

2.9. SDS-PAGE<br />

A Criterion Pre-Cast Gel System using Tris–HCl gels (Bio-Rad,<br />

Hercules, CA) was used for SDS-PAGE analysis (Wong et al., 2004).<br />

Equal amounts <strong>of</strong> protein were used for protein distribution analyses<br />

and estimation <strong>of</strong> sulfhydryl content. For protein digestion<br />

analysis, equal volumes (w5 ml) with a specified amount <strong>of</strong><br />

extracted protein were used.<br />

2.10. SDS-PAGE <strong>of</strong> kafirins and glutelins undigested by pepsin<br />

NuPAGE Novex Bis–Tris gels (Invitrogen, Carlsbad, CA), with<br />

a neutral pH to minimize protein modifications, were used to<br />

obtain optimal resolution <strong>of</strong> small- to medium-sized proteins.<br />

Aliquots (w5 ml) <strong>of</strong> specific amounts <strong>of</strong> extracted protein were<br />

mixed with 1/4 volume <strong>of</strong> 4 NuPAGE sample buffer plus 0.4% ME,<br />

J.H. Wong et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 73–82 75<br />

boiled 5 min, centrifuged and run in 12% Bis–Tris gels with MOPS<br />

buffer for 1 h 20 min at 150 V. Fluorescent imaging and protein<br />

staining were as described (Section 2.11).<br />

2.11. Quantification <strong>of</strong> fluorescent proteins<br />

To determine redox status, fluorescent (unsaturated and saturated)<br />

images <strong>of</strong> reduced proteins were captured from gels on a Gel<br />

Doc 1000 imager using Quantity One s<strong>of</strong>tware (Bio-Rad, Hercules,<br />

CA) with 365 nm UV light either immediately or after a 15 min<br />

water-wash; images were saved in TIFF format. Gels were stained<br />

for protein with colloidal Coomassie Blue G-250 overnight at 25 C,<br />

destained in ddH2O and scanned with a UMAX PowerLook 1100<br />

scanner using UMAX MagicScan version 4.4 within Adobe Photoshop,<br />

v.6 (Adobe Systems, San Jose, CA). Amount <strong>of</strong> protein,<br />

expressed as volume (intensity area) <strong>of</strong> fluorescent bands and/or<br />

undigested kafirin bands, was quantified with Quantity One s<strong>of</strong>tware.<br />

Higher protein band volume indicates more protein in the<br />

gel, and thus lower digestibility (Aboubacar et al., 2001).<br />

2.12. 2-D gel electrophoresis<br />

Protein contained in equal volumes <strong>of</strong> extracts was precipitated<br />

by addition <strong>of</strong> 5 vol <strong>of</strong> cold ( 20 C) acetone and incubation<br />

at 20 C overnight. Proteins were recovered by centrifugation and<br />

separated by 2-D gel electrophoresis (Wong et al., 2004).<br />

2.13. Western blot analysis<br />

After electrophoresis, proteins were transferred to nitrocellulose<br />

membranes at 4 C for 75 min at 50 V constant voltage (Wong et al.,<br />

2004) and cross-reactivity analyzed with waxy protein antibody<br />

(Terada et al., 2000).<br />

2.14. In vitro starch digestion<br />

Digestibility <strong>of</strong> starch in uncooked sorghum meal was measured<br />

using a modified hydrolysis method (Zhang and Hamaker, 1998)<br />

with bacterial a-amylase (A-3403, Type XII-A, Sigma Chemical Co.,<br />

St. Louis, MO).<br />

2.15. Sequential addition <strong>of</strong> a-amylase/pepsin or pepsin/a-amylase<br />

2.15.1. a-Amylase followed by pepsin<br />

Meal (50 mg) was treated with 50 mg a-amylase (human saliva<br />

form, Sigma A-1031, Type XIII-A) in 0.3 ml 1 mM Na glycerophosphate–HCl<br />

buffer (25 mM NaCl, 5 mM CaCl2, 0.02% w/v azide),<br />

pH 6.9, at 37 C for 1 h with shaking. Pepsin solution (1 ml at 20 mg/<br />

ml, pH 2.0) was added, the mixture incubated for 2 h at 37 C and<br />

neutralized with 0.1 ml 2 N NaOH. Supernate containing a-amylaseand<br />

pepsin-digested material was collected by centrifugation<br />

(20,800g, 10 min) for reducing sugar determination. Pellet was<br />

washed with 1 ml 0.1 M KH2PO4 buffer, pH 7.0 and then 1 ml ddH2O<br />

and extracted for remaining kafirins and GBSSI as above. Controls<br />

were: (A) a-amylase alone, (B) pepsin alone and (C) buffer alone.<br />

2.15.2. Pepsin followed by a-amylase<br />

Meal (50 mg) was incubated with pepsin (1 ml at 20 mg/ml, pH<br />

2.0) for 2 h at 37 C with shaking and then neutralized (0.1 ml 2 N<br />

NaOH to pH 7.0). a-Amylase (50 mg) in 0.3 ml Na glycerophosphate<br />

buffer as above, was added and the mixture incubated for 1 h at 37 C.<br />

Reaction was stopped by acidification and the supernate containing<br />

pepsin- and a-amylase-digestible materials was saved for reducing<br />

sugar determination. Pellets were washed 2 with 1 ml ddH2O and<br />

extracted for kafirins and GBSSI (Sections 2.7 and 2.8). Controls were:<br />

(A) pepsin alone, (B) a-amylase alone and (C) buffer alone.


76<br />

2.16. Effect <strong>of</strong> DTT reduction on digestion <strong>of</strong> cross-linked kafirins<br />

and glutelins by pepsin<br />

Meal (50 mg) was incubated 5 mM DTT in 0.5 ml <strong>of</strong> 0.1 M<br />

KH2PO4, pH 7.0, for 1 h at 37 C and supernate removed by centrifugation<br />

(20,800g, 10 min). The DTT-treated pellet was washed<br />

with 1 ml (A) ddH2O and (B) 0.1 M KH2PO4 buffer, pH 2.0. The<br />

washed residue was incubated with pepsin (1 ml at 20 mg/ml in pH<br />

2.0 KH2PO4 buffer) at 37 C for 2 h with shaking and digestion was<br />

stopped with 0.1 ml 2 N NaOH. Pepsin-treated residue was<br />

extracted for remaining kafirins and glutelins and analyzed by<br />

NuPAGE as above.<br />

2.17. Granule-bound starch synthase (GBSSI) identification<br />

Coomassie blue-stained spots were excised from 2-D gels and<br />

placed in a 96-well reaction tray. After proteins were destained,<br />

reduced, alkylated and cleaved with trypsin by a robotic sample<br />

handler, they were analyzed by tandem mass spectrometry and<br />

identified as described (Balmer et al., 2006).<br />

2.18. Other analytical procedures<br />

Protein was estimated in buffer-soluble extracts by the dyebinding<br />

method and those in the presence <strong>of</strong> SDS and ME or DTT by<br />

the Non-Interfering Protein Assay (Geno Technology, Inc., St. Louis,<br />

MO) as described (Wong et al., 2004). The ratio <strong>of</strong> amylose and<br />

amylopectin was determined using a MegazymeÔ Amylose/<br />

Amylopectin Assay Kit (Megazyme, Bray, Ireland).<br />

3. Results and discussion<br />

The 12-h IVDMD assay estimates the in vitro rate <strong>of</strong> starch<br />

digestion, which is highly correlated to feed efficiency or gain/feed<br />

ratio (r ¼ 0.94) (Stock et al., 1987). Because it mimics biochemical<br />

conditions in cattle, use <strong>of</strong> ruminal fluid containing hydrolytic<br />

enzymes produced by the ruminant and its microbial flora is relevant<br />

to studies <strong>of</strong> this type. The 12-h IVDMD applied to the two<br />

sorghum lines, KS48 and KS51, gave digestibility values <strong>of</strong> 33.0%<br />

and 22.1%, respectively. The results revealed a difference in dry<br />

matter disappearance and, hence, in the rates <strong>of</strong> digestion <strong>of</strong> starch<br />

as well as protein. The difference revealed by IVDMD, plus the<br />

common pedigree <strong>of</strong> these two lines, led to their choice for further<br />

study <strong>of</strong> digestibility. In the present work, grain from the two lines<br />

are subjected to in vitro pepsin and a-amylase assays that are<br />

generally considered to reflect activities <strong>of</strong> the human stomach and<br />

small intestine, respectively (Astwood et al., 1996; Zhang and<br />

Hamaker, 1998).<br />

Table 1<br />

Protein distribution <strong>of</strong> two sorghum lines with a common pedigree differing in digestibility<br />

KS48 a<br />

3.1. Protein distribution <strong>of</strong> KS48 and KS51<br />

The distribution <strong>of</strong> protein in fractions extracted with the indicated<br />

solvents (Section 2.4) suggested that, in general, the two<br />

sorghum lines yielded approximately the same amount <strong>of</strong> total<br />

extractable protein (Table 1). However, differences in distribution<br />

<strong>of</strong> protein between the two lines suggested that the more digestible<br />

KS48 contains more protein than the less digestible KS51 in the first<br />

five fractions extracteddddH2O wash, albumin and globulin (saltsoluble),<br />

kafirin (60% 2-propanol), glutelin-like protein (borate<br />

soluble), and cross-linked kafirin (2-propanol þ DTT). By contrast,<br />

the least soluble (last-extracted) fraction, cross-linked glutelin<br />

(borate þ DTT þ SDS), represented a considerably greater fraction<br />

in KS51 than in KS48. To determine whether these differences in<br />

protein distribution affect digestibility, we examined the composition<br />

and redox status <strong>of</strong> protein in the six fractions (Fig. 1).<br />

Disulfide bonds between cysteine residues, a common feature <strong>of</strong><br />

most cereal seed proteins, have long been known to stabilize<br />

various molecular structures (Wall, 1971). Many proteins with<br />

disulfide bonds resist digestion by proteases (Astwood et al., 1996;<br />

Opstvedt et al., 1984; del Val et al., 1999). Since a large percentage <strong>of</strong><br />

sorghum kafirin storage proteins exist in polymeric forms linked by<br />

disulfide bonds in their native state (Duodu et al., 2002, 2003; El<br />

Nour et al., 1998; Hamaker et al., 1987; Oria et al., 1995b), differences<br />

in content <strong>of</strong> fractions rich in insoluble disulfide proteins, i.e.,<br />

2-propanol þ DTT (cross-linked kafirin) and borate þ DTT þ SDS<br />

(cross-linked glutelin) could contribute to protein digestibility<br />

differences. Accordingly, although KS48 had more 2-propanol<br />

þ DTT extractable proteins (cross-linked kafirins), the content<br />

<strong>of</strong> borate þ DTT þ SDS-extractable glutelins was appreciably<br />

greater in KS51 than in KS48 – a reflection <strong>of</strong> the relative abundance<br />

<strong>of</strong> intermolecular S–S groups in this fraction (Table 1). The latter<br />

could be a factor in the lower digestibility <strong>of</strong> KS51.<br />

3.2. Estimation <strong>of</strong> disulfide protein status<br />

Protein disulfide content was estimated by determining the<br />

sulfhydryl groups resulting from DTT reduction, using a thiolspecific<br />

fluorescent probe, mBBr (Wong et al., 2004). Four fractions<br />

(nos. 1–4 in Table 1), extracted without reducing agent, were<br />

monitored for DTT-induced sulfhydryl changes. KS51 showed more<br />

intense fluorescence relative to KS48 in fraction 1 salt (albumin and<br />

globulin) and fraction 4 borate (glutelin-like) (Fig. 1B). The more<br />

pronounced increase in proteins with disulfide bonds in fraction 6<br />

(borate þ DTT þ SDS) could also contribute to the lower digestibility<br />

<strong>of</strong> KS51, as could the markedly different distribution <strong>of</strong><br />

proteins in that fraction (Fig. 1A). For reasons not yet understood,<br />

the fluorescence intensity <strong>of</strong> the kafirin (propanol) fraction showed<br />

larger relative differences between the two lines in untreated vs.<br />

DTT-treated samples (Fig. 1B, lanes 9 vs. 11; 10 vs. 12). In summary,<br />

Fraction Extraction solvent Total mg protein extracted % Total protein extracted Total mg protein extracted % Total protein extracted D, %<br />

1 NaCl, 0.5 M 28.3 0.5 19.0 27.6 0.4 18.8 0.2<br />

2 H2O-wash 0.4 1.0 0.3 0.4 1.4 0.3 0.0<br />

3 2-Propanol, 60% 17.3 1.0 11.6 12.4 1.3 8.5 3.1<br />

4 Borate, 0.1 M pH 10.8 5.7 0.7 3.8 4.4 1.2 3.0 0.8<br />

5 2-Propanol, 60% þ 1% DTT 30.9 2.8 20.8 27.8 4.8 19.0 1.8<br />

6 Borate, 0.1 M pH 10.8 þ 1%<br />

DTT þ 1% SDS<br />

66.3 3.1 44.5 73.8 4.3 50.4 5.9<br />

Total 148.9 1.5 100 146.4 2.2 100 –<br />

a Mean <strong>of</strong> 5 experiments.<br />

J.H. Wong et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 73–82<br />

KS51 a<br />

KS48–KS51


the relative abundance <strong>of</strong> intermolecular S–S groups in the crosslinked<br />

glutelin fraction in KS51 suggests that the non-kafirins,<br />

which make up the protein matrix, have a strong influence on the<br />

digestibility <strong>of</strong> KS51 relative to KS48.<br />

3.3. Time course <strong>of</strong> in vitro pepsin digestion<br />

The gel-based system used to analyze protein digestion by<br />

pepsin (Aboubacar et al., 2001; Nunes et al., 2004) requires less<br />

material, is faster and can accommodate more samples than<br />

a commonly used batch procedure (Mertz et al., 1984). The gel<br />

procedure not only reveals the types <strong>of</strong> undigested kafirins, but also<br />

J.H. Wong et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 73–82 77<br />

Fig. 1. Effect <strong>of</strong> DTT on redox status <strong>of</strong> protein fractions extracted from grain <strong>of</strong> KS48 and KS51. (A): Protein stained with Coomassie blue; (B): protein labeled with mBBr visualized<br />

by UV absorption. Tris–HCl gel was used. Individual protein fractions are identified above (A).<br />

allows an estimate <strong>of</strong> the percentage digested over time. This gel<br />

system has been successfully used to separate sorghum proteins<br />

(Bean, 2003) and wheat glutenins (Kasarda et al., 1998). In our<br />

study, we observed that separation <strong>of</strong> different kafirin types is<br />

improved in the neutral pH environment <strong>of</strong> the NuPAGE Novex Bis–<br />

Tris gels relative to that seen with the conventional Tris–HCl gels<br />

(cf. Fig. 2Avs.Fig. 1A). A NuPAGE gel image (Fig. 2A) shows that high<br />

Mr components <strong>of</strong> the insoluble protein fraction, including glutelins<br />

(35–100 kDa), were readily digested, while the smaller kafirins (Mr<br />

18–27 kDa) were more resistant. Further, the time course <strong>of</strong> the in<br />

vitro digestion <strong>of</strong> this fraction demonstrated that kafirins in KS48<br />

were digested more rapidly by pepsin than those in KS51 (Fig. 2A).<br />

Fig. 2. Time course <strong>of</strong> in vitro digestion <strong>of</strong> storage protein and starch in KS48 and KS51. (A): BT-NuPAGE gel showing digestion <strong>of</strong> glutelins and kafirins over time; (B): estimation <strong>of</strong><br />

glutelin and kafirin digestion rate with the NI protein assay. Results were calculated by linear regression and expressed as % <strong>of</strong> control without pepsin; (C): starch digestion by<br />

bacterial a-amylase. Corn starch is shown as a control.


78<br />

Gel densitometry measurements revealed that various kafirins, a,<br />

b and g (El Nour et al., 1998), were digested at different rates:<br />

b > g > a (data not shown). Furthermore, data on the disappearance<br />

with time <strong>of</strong> insoluble proteins (kafirins and glutelins) (Fig. 2B)<br />

indicated that KS48 was digested more rapidly than KS51 – negative<br />

slope <strong>of</strong> 0.643 (SE 0.010) for KS48 vs. 0.530 (SE 0.023) for<br />

KS51, mean <strong>of</strong> three determinations. These values, which represent<br />

a 22% rate difference, positively correlate with the corresponding<br />

IVDMD difference <strong>of</strong> 36%. The lower value obtained with the in vitro<br />

pepsin method could be due to use <strong>of</strong> a single protease vs. a mixture<br />

<strong>of</strong> proteolytic and amylolytic enzymes present in ruminal fluid used<br />

in the IVDMD procedure.<br />

3.4. Time course <strong>of</strong> in vitro starch digestion by a-amylase<br />

The finding that the 12-h IVDMD value was higher for KS48 than<br />

KS51 suggested that starch digestion alone might follow the same<br />

trend, and this was confirmed (Fig. 2C). The average <strong>of</strong> three timecourse<br />

experiments for in vitro digestion <strong>of</strong> meal from KS48 and<br />

KS51 by bacterial a-amylase gave slopes <strong>of</strong> 1.74 10 3 (SE<br />

1.69 10 4 ) and 1.41 10 3 (SE 5.01 10 5 ), respectively. These<br />

results represent a 23% difference in starch digestion rate between<br />

the two lines – a value similar to that obtained with the in vitro<br />

pepsin (Fig. 2B) and IVDMD (Pedersen et al., 2000) procedures.<br />

More significantly, the three independent in vitro techniques used<br />

to assess digestibility showed a positive correlation. This is the first<br />

time the three approaches have been compared in a single study.<br />

Ezeogu et al. (2005) made significant contributions to the idea<br />

that in vitro digestibility <strong>of</strong> starch in sorghum and maize flours is<br />

related to endosperm texture and cooking conditions. These<br />

researchers found that starch digestion was significantly higher in<br />

floury sorghum endosperm than in vitreous endosperm, while<br />

their counterparts in maize were similar in this regard. Cooking<br />

with ME increased starch digestion in both sorghum and maize, but<br />

more so with sorghum, and with vitreous endosperm flours. This is<br />

consistent with the fact that more polymeric kafirins, formed by<br />

intermolecular disulfide bonds (Chandrashekar and Mazhar, 1999;<br />

Oria et al., 1995a), occurred in the vitreous endosperm fraction<br />

(Ezeogu et al., 2005; Kumari and Chandrashekar, 1994), probably<br />

because the cysteine-rich g- and b-kafirins abound in that part <strong>of</strong><br />

the sorghum grain (Chandrashekar and Mazhar, 1999; Ezeogu et al.,<br />

2005, 2008; Mazhar and Chandrashekar, 1995). Unlike their findings,<br />

which were based on similar types <strong>of</strong> starch – i.e., a 75:25<br />

amylopectin/amylose ratio from different parts <strong>of</strong> the sorghum and<br />

maize endosperm – our results are based on different types <strong>of</strong><br />

J.H. Wong et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 73–82<br />

starch, i.e., an amylopectin/amylose ratio <strong>of</strong> 97:3 and 70:30, for<br />

KS48 and KS51, respectively (see below).<br />

3.5. Starch digestibility differences in starch in KS48 and KS51<br />

In the original rapid protein digestion assay (Aboubacar et al.,<br />

2001), the pellet remaining after extraction <strong>of</strong> kafirins and glutelins<br />

with borate–SDS–ME was not further analyzed. In this study, we<br />

attempted to fill this gap by analyzing the residual pellet that still<br />

contained between 2.2–2.5% (pepsin-treated) and 4.7–5.0% (buffer<br />

control) <strong>of</strong> protein nitrogen (data not shown). After boiling the<br />

residual pellet with excess borate–SDS–ME buffer, we unexpectedly<br />

observed a marked difference in the gelatinization properties<br />

<strong>of</strong> starch in KS48 and KS51. The KS48 pellet yielded a transparent,<br />

s<strong>of</strong>t starch gel that swelled to a greater extent than the KS51<br />

counterpart. Further, when equal amounts <strong>of</strong> extraction buffer<br />

were added to the pellets before boiling, only a small amount <strong>of</strong><br />

supernate was recovered after centrifuging the cooled gel KS48<br />

preparation, whereas KS51 yielded a larger amount <strong>of</strong> clear<br />

supernate that separated easily from the harder opaque gel (data<br />

not shown). The gelatinization differences in KS48 and KS51 starch<br />

resemble those observed for waxy vs. non-waxy starch.<br />

In addition to obvious swelling differences in gelatinized starch,<br />

supernates from KS48 and KS51 contained different proteins. Onedimensional<br />

SDS-PAGE analysis revealed a major protein band <strong>of</strong><br />

Mr w 58 kDa in KS51, not detected in KS48 (Fig. 3A). Because it was<br />

released from starch granules only after gelatinization, the protein<br />

did not appear to change after treating the parent meal with pepsin<br />

(data not shown). Two-dimensional gel analysis gave information<br />

on the nature <strong>of</strong> the unknown band (Fig. 3C, D). Two major 61 kDa<br />

spots with acidic pI’s, uniquely observed in KS51 (note arrows,<br />

Fig. 3D), were identified as granule-bound starch synthase I by<br />

mass spectrometry [GBSSI, chloroplast (sorghum) accession No.<br />

Q43134]. Identity was further confirmed by Western blot analysis<br />

using an antibody against maize waxy protein (Fig. 3B). In KS51,<br />

a highly antibody-reactive band was seen at 58 kDa and a less<br />

reactive one at 61 kDa, corresponding to two is<strong>of</strong>orms <strong>of</strong> GBSSI:<br />

a major form in the endosperm (58 kDa) and a minor form in the<br />

pericarp (61 kDa) (Nakamura et al., 1998). While lower levels <strong>of</strong> the<br />

pericarp is<strong>of</strong>orm were seen in KS48 relative to KS51, the endosperm<br />

component was not detected. The granule-bound is<strong>of</strong>orm <strong>of</strong> starch<br />

synthase, GBSSI, also referred to as the waxy protein encoded in the<br />

waxy (Wx) locus <strong>of</strong> cereals, functions specifically to elongate amylose<br />

(Denyer et al., 2001) by adding glucose residues in a-1,4-glycosidic<br />

Fig. 3. Starch-bound proteins extracted from KS48 and KS51. The pellet remaining after removing glutelins and kafirins was boiled with borate–SDS–2-ME buffer and the proteins<br />

solubilized were analyzed by one-dimensional SDS-PAGE. (A): Protein; (B): Western blot analysis using waxy protein antibody; (C and D): two-dimensional gel analysis <strong>of</strong> starchbound<br />

proteins extracted in (A).


linkage. Recent genetic experiments have revealed that sorghum<br />

contains at least two GBSS genes (Pedersen et al., 2005).<br />

3.6. Starch <strong>of</strong> waxy and non-waxy sorghum<br />

Normal or non-waxy sorghum starch has approximately 75%<br />

amylopectin and 25% amylose, while waxy starch contains nearly<br />

100% amylopectin (Ezeogu et al., 2005; Rooney and Miller, 1982;<br />

Rooney and Pflugfelder, 1986). The apparent absence <strong>of</strong> GBSSI<br />

protein in the endosperm <strong>of</strong> KS48 (Fig. 3A, B) indicates that its<br />

starch should be low in amylose and contain mainly amylopectin.<br />

Waxy and non-waxy polymers <strong>of</strong> starch can be distinguished by<br />

iodine staining (Denyer et al., 2001; Pedersen et al., 2004). Starch <strong>of</strong><br />

non-waxy lines, containing amylose and amylopectin, forms blueblack<br />

complexes with iodine, while starch from waxy lines, lacking<br />

amylose, stains reddish brown. Since the waxy nature <strong>of</strong> KS48 and<br />

KS51 was not indicated in the USDA-ARS, Genetic Resource Information<br />

Network (Pedersen et al., 2004), grains from the two lines<br />

were stained with iodine. Results confirmed KS48 has a waxy<br />

(brownish red) and KS51 a non-waxy (dark blue) type <strong>of</strong> endosperm<br />

(see below). Biochemical analyses confirmed this observation:<br />

KS48 contains 3% and KS51 30% amylose.<br />

Numerous studies have shown that starch is more readily<br />

digested in waxy sorghum than in non-waxy types [for review, see<br />

(Rooney and Pflugfelder, 1986)]. Our results are consistent with this<br />

conclusion, not only for starch but also for protein (KS48 > KS51).<br />

Digestibility <strong>of</strong> protein, however, may not necessarily be consistent<br />

with waxy vs. non-waxy sorghum lines. The nature <strong>of</strong> the protein<br />

matrix and the extent <strong>of</strong> embedded starch in the endosperm are<br />

proposed to account for this inconsistency (Chandrashekar and<br />

Kirleis, 1988; Duodu et al., 2003; Rooney and Miller, 1982; Rooney<br />

and Pflugfelder, 1986; Shull et al., 1990). To gain insight into this<br />

point, we conducted experiments with our waxy and non-waxy<br />

lines as described below.<br />

3.7. Sequential addition <strong>of</strong> a-amylase and pepsin<br />

Digestion protocols, using sequential addition <strong>of</strong> pepsin followed<br />

by a-amylase, and vice versa, were designed to study the<br />

inter-relationship <strong>of</strong> starch and protein in KS48 and KS51. Prior<br />

digestion with pepsin, followed by treatment with human saliva aamylase<br />

(designated P/A in Table 2), enhanced to a greater extent<br />

starch hydrolysis, relative to treatment with a-amylase followed by<br />

pepsin (A/P), in KS48 compared to KS51 (1.8- vs. 1.5-fold, respectively).<br />

This observation suggests that KS51 starch granules were<br />

intrinsically harder to digest than those <strong>of</strong> KS48, even if the protein<br />

matrix was partially removed by prior pepsin digestion. This<br />

explanation agrees with the time course <strong>of</strong> in vitro starch digestion<br />

<strong>of</strong> meal (Fig. 2C), as well as with the relative ease <strong>of</strong> digestion <strong>of</strong><br />

isolated waxy vs. non-waxy starch (Rooney and Miller, 1982; Rooney<br />

and Pflugfelder, 1986).<br />

Reversal <strong>of</strong> this protocoldremoval <strong>of</strong> starch prior to pepsin<br />

digestiondshowed the opposite effect; proteolysis was enhanced<br />

to a greater extent with KS51 than KS48 (2.2- vs. 1.6-fold) (Table 3).<br />

Table 2<br />

Starch digestion with and without prior removal <strong>of</strong> protein by pepsin<br />

Sorghum line With pepsin<br />

pre-treatment<br />

(pepsin/a-amylase a )<br />

Without pepsin<br />

pre-treatment<br />

(a-amylase/pepsin a )<br />

Ratio<br />

P/A:A/P<br />

mg reducing sugar/h mg reducing sugar/h<br />

KS48 84.8 4.2 47.6 2.1 1.8<br />

KS51 59.1 2.7 39.3 1.3 1.5<br />

a Mean <strong>of</strong> 3 experiments.<br />

J.H. Wong et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 73–82 79<br />

Table 3<br />

Protein digestion by pepsin without and with prior removal <strong>of</strong> starch by a-amylase<br />

Sorghum<br />

line<br />

With a-amylase<br />

pre-treatment<br />

(a-amylase/pepsin a )<br />

Without a-amylase<br />

pre-treatment<br />

(pepsin/a-amylase a )<br />

This unexpected result suggests that, in addition to the intrinsic<br />

properties <strong>of</strong> the starch, spatial arrangement between the granules<br />

and protein bodies differs in the two lines. Further, starch in KS51<br />

appears not only to be harder to digest with a-amylase, but also<br />

more capable <strong>of</strong> impeding protein digestion by pepsin than in KS48.<br />

Once starch-imposed restrictions are removed, protein in KS51<br />

appears more exposed and thus more accessible to pepsin. A<br />

question arises as to whether the observed difference could simply<br />

be due to the physical properties <strong>of</strong> the two grains, i.e., variation in<br />

the hardness <strong>of</strong> the two grains, or different ratios <strong>of</strong> hard (closed) to<br />

s<strong>of</strong>t (open) endosperm, as reported by Ezeogu et al., 2005. Hardness<br />

data for KS48 and KS51 samples using the single kernel characterization<br />

system (SKCS) showed that they are practically and<br />

statistically the same, with a mean for seed hardness <strong>of</strong> 69.12 and<br />

69.83, respectively. Therefore, our interpretation is that the difference<br />

observed above is mostly due to the properties <strong>of</strong> the starch<br />

and proteins.<br />

3.8. Effect <strong>of</strong> DTT reduction on digestion <strong>of</strong> kafirins and<br />

glutelins by pepsin<br />

Ratio<br />

A/P:P/A<br />

Protein remaining, volume, % Protein remaining, volume, %<br />

KS48 65.4 2.0 40.9 2.7 1.6<br />

KS51 75.3 1.5 34.0 3.8 2.2<br />

a Mean <strong>of</strong> 3 experiments.<br />

Reducing agents, such as 2-ME (Elkhalifa et al., 1999; Ezeogu<br />

et al., 2005, 2008; Hamaker et al., 1987) and DTT or thioredoxin (del<br />

Val et al., 1999), were shown to improve protein digestibility, thus<br />

highlighting the importance <strong>of</strong> components with disulfide bonds<br />

(Hamaker et al., 1987). In assessing DTT effects, we observed that<br />

42% and 50% <strong>of</strong> the kafirins <strong>of</strong> KS48 and KS51, respectively, were not<br />

digested by pepsin in control samples. Reduction with DTT<br />

Fig. 4. Effect <strong>of</strong> redox status on digestion <strong>of</strong> the combined glutelin and kafirin fractions<br />

<strong>of</strong> KS48 and KS51 by pepsin. D þ P ¼ DTT þ pepsin; P ¼ pepsin; D ¼ DTT; C ¼ control<br />

(minus DTT and pepsin). * ¼ Kafirin region.


80<br />

enhanced digestibility by 3.5-fold and 2.0-fold, with undigested<br />

protein decreasing from 42% to 12% for KS48 and from 50% to 25%<br />

for KS51 (Fig. 4, * designates kafirin region). To achieve a level <strong>of</strong><br />

kafirin digestion with KS51 equivalent to that <strong>of</strong> KS48, either<br />

a longer incubation period or more reductant was required (data<br />

not shown).<br />

3.9. Microscopy studies<br />

Our results thus far show that a greater abundance <strong>of</strong> disulfide<br />

proteins and the non-waxy type <strong>of</strong> starch contribute to the low<br />

digestibility <strong>of</strong> KS51 vs. KS48. We have performed experiments<br />

with additional waxy and non-waxy lines <strong>of</strong> sorghum and corn<br />

(3 pairs <strong>of</strong> sorghum lines, 1 pair <strong>of</strong> corn lines). Digestion <strong>of</strong> waxy<br />

starch was consistently faster than that <strong>of</strong> non-waxy starch in all<br />

samples tested but protein digestibility was mixed (data not<br />

shown). Thus, this data also show the importance <strong>of</strong> the interactions<br />

between starch and protein to the digestibility <strong>of</strong> both<br />

J.H. Wong et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 73–82<br />

components. However, they provide no information on the nature<br />

<strong>of</strong> the differences observed in these interactions between the two<br />

lines.<br />

To study this point, we examined iodine-stained sections <strong>of</strong><br />

grains microscopically and found that the endosperm in KS51 was<br />

more highly organized than that <strong>of</strong> KS48di.e., there was a distinct<br />

demarcation between floury and corneous endosperm in KS51, not<br />

present in KS48 (Fig. 5A). Taken together, the analytical data presented<br />

above and the iodine staining results confirmed the earlier<br />

classification <strong>of</strong> the lines as waxy (KS48) and non-waxy (KS51)<br />

(Pedersen et al., 2004, 2005). Electron micrographs revealed<br />

organizational differences between the two. While floury endosperm<br />

was similar (Fig. 5C), corneous endosperm differed markedly<br />

in the two lines (Fig. 5B). Protein bodies were more numerous and<br />

more tightly associated with starch granules in KS51 than in KS48<br />

(Fig. 5B). This type <strong>of</strong> interaction between the two components<br />

could contribute to the low digestibility <strong>of</strong> protein and starch in<br />

KS51.<br />

Fig. 5. Light and electron microscopy <strong>of</strong> KS48 and KS51 seeds. A. Cross section <strong>of</strong> seeds stained with iodine showing floury (F) and corneous (C) endosperm. Magnification: 32 .B.<br />

Electron micrograph <strong>of</strong> corneous endosperm <strong>of</strong> KS48 and KS51, 5000 . C. Electron micrograph <strong>of</strong> floury endosperm <strong>of</strong> KS48 and KS51, 5000 .SG¼ starch granule; PB ¼ protein<br />

body; M ¼ protein matrix.


In summary, our results highlight the importance <strong>of</strong> disulfide<br />

proteins and GBSSI in conferring the chemical and structural<br />

properties that influence the digestibility <strong>of</strong> stored reserves, both<br />

protein and starch. These findings also add insight into the role <strong>of</strong><br />

the protein matrix and its interaction with starch. Comparison <strong>of</strong><br />

two sorghum lines with a common pedigree that differed in<br />

digestibility made it possible to demonstrate how protein–starch<br />

interactions in the seed influence the digestibility <strong>of</strong> each. Attempts<br />

to improve the digestibility <strong>of</strong> grain by classical breeding and<br />

genetic engineering should further our understanding <strong>of</strong> these<br />

factors and their relative contribution to digestibility. These efforts<br />

may also provide answers to related fundamental questions, such<br />

as what role the key redox protein, thioredoxin, plays in forming<br />

and maintaining the protein matrix (Buchanan and Balmer, 2005)<br />

and the importance <strong>of</strong> this unique packaging <strong>of</strong> stored reserves to<br />

the structure and physiology <strong>of</strong> seed. Answers to these questions<br />

will in all likelihood be relevant to human and animal nutrition and<br />

to the use <strong>of</strong> sorghum as a bi<strong>of</strong>uel.<br />

Acknowledgements<br />

This study was supported by the U.S. Agency for International<br />

Development and the Bill and Melinda Gates Foundation. PGL was<br />

supported by the USDA Cooperative Extension Service and BBB by<br />

the Agricultural Experiment Station through UC Berkeley. We thank<br />

Drs. K. McDonald and G. Min (Electron Microscope Laboratory) and<br />

D. Schichnes (Biological Imaging Laboratory) for assistance with the<br />

microscopy studies and Dr. S.R. Wessler for the gift <strong>of</strong> waxy protein<br />

antibody. The late Dr. K. Kobrehel helped launch this study during<br />

a sabbatical visit to Berkeley in the early 1990s. Numerous undergraduate<br />

students also made contributions, especially D. Bergmann,<br />

P.-H. Ren and P. Yu-Man Kei.<br />

References<br />

Aboubacar, A., Axtell, J.D., Huang, C.-P., Hamaker, B.R., 2001. A rapid protein<br />

digestibility assay for identifying highly digestible sorghum lines. <strong>Cereal</strong><br />

Chemistry 78.<br />

Astwood, J.D., Leach, J.N., Fuchs, R.L., 1996. Stability <strong>of</strong> food allergens to digestion in<br />

vitro. Nature Biotechnology 14, 1269–1273.<br />

Axtell, J.D., Kirleis, A., Hassen, M.M., D’Croz-Mason, N., Mertz, E.T., Munck, L., 1981.<br />

Digestibility <strong>of</strong> sorghum proteins. Proceedings <strong>of</strong> the National Academy <strong>of</strong><br />

<strong>Science</strong>s <strong>of</strong> the United States <strong>of</strong> America 78, 1333–1335.<br />

Balmer, Y., Vensel, W.H., DuPont, F.M., Buchanan, B.B., Hurkman, W.J., 2006. Proteome<br />

<strong>of</strong> amyloplasts isolated from developing wheat endosperm presents<br />

evidence <strong>of</strong> broad metabolic capability. <strong>Journal</strong> <strong>of</strong> Experimental Botany 57,<br />

1591–1602.<br />

Bean, S.R., 2003. Evaluation <strong>of</strong> novel pre-cast gradient gels for the separation <strong>of</strong><br />

sorghum proteins. <strong>Cereal</strong> Chemistry 80, 500–504.<br />

Belton, P.S., Delgadillo, I., Halford, N.G., Shewry, P.R., 2006. Kafirin structure and<br />

functionality. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 44, 272–286.<br />

Buchanan, B.B., Balmer, Y., 2005. Redox regulation: a broadening horizon. Annual<br />

Reviews <strong>of</strong> Plant Biology 56, 187–220.<br />

Casady, A.J., 1972. 10 Sorghum parental lines. Crop <strong>Science</strong> 12, 722.<br />

Chandrashekar, A., Kirleis, A.W., 1988. Influence <strong>of</strong> protein on starch gelatinization<br />

in sorghum. <strong>Cereal</strong> Chemistry 65, 457–462.<br />

Chandrashekar, A., Mazhar, H., 1999. The biochemical basis and implications <strong>of</strong><br />

grain strength in sorghum and maize. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 30, 193–207.<br />

Denyer, K., Johnson, P., Zeeman, S., Smith, A.M., 2001. The control <strong>of</strong> amylose<br />

synthesis. <strong>Journal</strong> <strong>of</strong> Plant Physiology 158, 479–487.<br />

Duodu, K.G., Nunes, A., Delgadillo, I., Parker, M.L., Mills, E.N.C., Belton, P.S.,<br />

Taylor, J.R.N., 2002. Effect <strong>of</strong> grain structure and cooking on sorghum and maize<br />

in vitro protein digestibility. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 35, 161–174.<br />

Duodu, K.G., Tang, H., Grant, A., Wellner, N., Belton, P.S., Taylor, J.R.N., 2001. FTIR and<br />

solid state 13 C spectroscopy <strong>of</strong> proteins <strong>of</strong> wet cooked and popped sorghum and<br />

maize. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 33, 261–269.<br />

Duodu, K.G., Taylor, J.R.N., Belton, P.S., Hamaker, B.R., 2003. Factors affecting<br />

sorghum protein digestibility. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 38, 117–131.<br />

El Nour, I.N.A., Peruffo, A.D.B., Curioni, A., 1998. Characterisation <strong>of</strong> sorghum kafirins<br />

in relation to their cross-linking behaviour. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 28,<br />

197–207.<br />

Elkhalifa, A.E.O., Chandrashekar, A., Mohamed, B.E., El Tinay, A.H., 1999. Effect <strong>of</strong><br />

reducing agents on the in vitro protein and starch digestibilities <strong>of</strong> cooked<br />

sorghum. Food Chemistry 66, 323–326.<br />

J.H. Wong et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 73–82 81<br />

Ezeogu, L.I., Duodu, K.G., Taylor, J.R.N., 2005. Effects <strong>of</strong> endosperm texture and<br />

cooking conditions on the in vitro starch digestibility <strong>of</strong> sorghum and maize<br />

flours. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 42, 33–44.<br />

Ezeogu, L.I., Duodu, K.G., Emmanbux, M.N., Taylor, J.R.N., 2008. Influence <strong>of</strong> cooking<br />

conditions on the protein matrix <strong>of</strong> sorghum and maize endosperm flours.<br />

<strong>Cereal</strong> Chemistry 85, 397–402.<br />

Hamaker, B.R., Bugusu, B.A., 2003. Overview: sorghum proteins and food quality.<br />

Paper presented at: Workshop on the proteins <strong>of</strong> sorghum and millets:<br />

enhancing nutritional and functional properties for Africa [CD] (Pretoria: South<br />

Africa).<br />

Hamaker, B.R., Kirleis, A.W., Butler, L.G., Axtell, J.D., Mertz, E.T., 1987. Improving the<br />

in vitro protein digestibility <strong>of</strong> sorghum with reducing agents. Proceedings <strong>of</strong><br />

the National Academy <strong>of</strong> <strong>Science</strong>s <strong>of</strong> the United States <strong>of</strong> America 84, 626–628.<br />

Hamaker, B.R., Kirleis, A.W., Mertz, E.T., Axtell, J.D., 1986. Effect <strong>of</strong> cooking on the<br />

protein pr<strong>of</strong>iles and in vitro digestibility <strong>of</strong> sorghum and maize. <strong>Journal</strong> <strong>of</strong><br />

Agricultural and Food Chemistry 34, 647.<br />

Hamaker, B.R., Mohamed, A.A., Habben, J.E., Huang, C.-P., Larkins, B.A., 1995. An efficient<br />

procedure for extracting maize and sorghum kernel proteins reveals higher<br />

prolamin contents than the conventional method. <strong>Cereal</strong> Chemistry 72, 583–588.<br />

Izquierdo, L., Godwin, I.D., 2005. Molecular characterization <strong>of</strong> novel methioninerich<br />

delta-kafirin seed storage protein gene in sorghum (Sorghum bicolor L.).<br />

<strong>Cereal</strong> Chemistry 82, 706.<br />

Jambunathan, R., Mertz, E.T., Axtell, J.D., 1975. Fractionation <strong>of</strong> soluble proteins <strong>of</strong><br />

high-lysine and normal sorghum grain. <strong>Cereal</strong> Chemistry 52, 119–121.<br />

Kasarda, D.D., Woodard, K.M., Elva, A.A., 1998. Resolution <strong>of</strong> high molecular weight<br />

glutenin subunits by a new SDS-PAGE system incorporating a neutral pH buffer.<br />

<strong>Cereal</strong> Chemistry 75, 70–71.<br />

Klopfenstein, C.F., Hoseney, R.C., 1995. Nutritional properties <strong>of</strong> sorghum and the<br />

millets. In: Dendy, D.A.V. (Ed.), Sorghum and Millets: Chemistry and Technology.<br />

American Association <strong>of</strong> <strong>Cereal</strong> Chemists, St. Paul, MN, pp. 125–168.<br />

Kumari, S.R., Chandrashekar, A., 1994. Relationships between grain hardness and<br />

the contents <strong>of</strong> prolamins and three anti-fungal proteins in sorghum. <strong>Journal</strong> <strong>of</strong><br />

<strong>Cereal</strong> <strong>Science</strong> 20, 93–99.<br />

Landry, J., Moureaux, T., 1970. Heterogeneite des gluenines du grain de mais:<br />

Extraction slelective et composition en acides amines des trios fractions isolees.<br />

Bulletin de la Societe de Chimie Bioologique 52, 1021–1031.<br />

MacLean, W.C., de Romana, G.L., Placko, R.P., Graham, G.G., 1981. Protein quality and<br />

digestibility <strong>of</strong> sorghum in preschool children: balance studies and plasma free<br />

amino acids. <strong>Journal</strong> <strong>of</strong> Nutrition 111, 1928–1936.<br />

Mazhar, H., Chandrashekar, A., 1995. Quantitation and distribution <strong>of</strong> kafirins in the<br />

kernels <strong>of</strong> sorghum cultivars varying in endosperm vitreousness. <strong>Journal</strong> <strong>of</strong><br />

<strong>Cereal</strong> <strong>Science</strong> 21, 155–162.<br />

Mazhar, H., Chandrashekar, A., Shetty, H.S., 1993. Isolation and immunochemical<br />

characterization <strong>of</strong> alcohol-extractable proteins (kafirins) <strong>of</strong> Sorghum bicolor (L.)<br />

Moench. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 17, 83–93.<br />

Mertz, E.T., Hassen, M.M., Cairn-Whittern, C., Kirleis, A.W., Tu, L., Axtell, J.D., 1984.<br />

Pepsin digestibility <strong>of</strong> proteins in sorghum and other major cereals. Proceedings<br />

<strong>of</strong> the National Academy <strong>of</strong> <strong>Science</strong>s <strong>of</strong> the United States <strong>of</strong> America 81, 1–2.<br />

Nakamura, T., Vrinten, P., Hayakawa, K., Ikeda, J., 1998. Characterization <strong>of</strong><br />

a granule-bound starch synthase is<strong>of</strong>orm found in pericarp <strong>of</strong> wheat. Plant<br />

Physiology 118, 451–459.<br />

Nunes, A., Correia, I., Barros, A., Delgadillo, I., 2004. Sequential in vitro pepsin<br />

digestion <strong>of</strong> uncooked and cooked sorghum and maize samples. <strong>Journal</strong> <strong>of</strong><br />

Agricultural and Food Chemistry 52, 2052–2058.<br />

Opstvedt, J., Miller, R., Hardy, R.W., Spinelli, J., 1984. Heat-induced changes in<br />

sulfhydryl groups and disulfide bonds in fish protein and their effect on protein<br />

and amino acid digestibility in rainbow trout (Salmo gairdneri). <strong>Journal</strong> <strong>of</strong><br />

Agricultural and Food Chemistry 32, 924–935.<br />

Oria, M.P., Hamaker, B.R., Axtell, J.D., Huang, C.-P., 2000. A highly digestible sorghum<br />

mutant cultivar exhibits a unique folded structure <strong>of</strong> endosperm protein bodies.<br />

Proceedings <strong>of</strong> the National Academy <strong>of</strong> <strong>Science</strong>s <strong>of</strong> the United States <strong>of</strong><br />

America 97, 5065–5070.<br />

Oria, M.P., Hamaker, B.R., Shull, J.M., 1995a. In vitro protein digestibility <strong>of</strong> developing<br />

and mature sorghum grain in relation to a-, b- and g-kafirin disulfide<br />

crosslinking. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 22, 85–93.<br />

Oria, M.P., Hamaker, B.R., Shull, J.M., 1995b. Resistance <strong>of</strong> sorghum a-, b- and gkafirins<br />

to pepsin digestion. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 43,<br />

2148–2152.<br />

Pedersen, J.F., Bean, S.R., Funnell, D.L., Graybosch, R.A., 2004. Rapid iodine staining<br />

techniques for identifying the waxy phenotype in sorghum grain and waxy<br />

genotype in sorghum pollen. Crop <strong>Science</strong> 44, 764–767.<br />

Pedersen, J.F., Bean, S.R., Graybosch, R.A., Park, S.H., Tilley, M., 2005. Characterization<br />

<strong>of</strong> waxy grain sorghum lines in relation to granule-bound starch synthase.<br />

Euphytica 144, 151–156.<br />

Pedersen, J.F., Milton, T., Mass, R.A., 2000. A twelve-hour in vitro procedure for<br />

sorghum grain feed quality assessment. Crop <strong>Science</strong> 40, 204–208.<br />

Rooney, L.W., Miller, F.R., 1982. Variation in the structure and kernel characteristics<br />

<strong>of</strong> sorghum. Paper presented at: International Symposium on Sorghum Grain<br />

Quality (ICRISAT Center, Patancheru, India).<br />

Rooney, L.W., Pflugfelder, R.L., 1986. Factors affecting starch digestibility with special<br />

emphasis on sorghum and corn. <strong>Journal</strong> <strong>of</strong> Animal <strong>Science</strong> 63, 1607–1623.<br />

Shull, J.M., Chandrashekar, A., Kirleis, A.W., Ejeta, G., 1990. Development <strong>of</strong> sorghum<br />

[Sorghum bicolor (L.) Moench] endosperm in varieties <strong>of</strong> varying hardness. Food<br />

Structure 9, 253–267.<br />

Shull, J.M., Watterson, J.J., Kirleis, A.W., 1991. Proposed nomenclature for the<br />

alcohol-soluble proteins (kafirins) <strong>of</strong> Sorghum bicolor (L. Moench) based on


82<br />

molecular weight, solubility, and structure. <strong>Journal</strong> <strong>of</strong> Agricultural and Food<br />

Chemistry 39, 83–87.<br />

Shull, J.M., Watterson, J.J., Kirleis, A.W., 1992. Purification and immunocytochemical<br />

localization <strong>of</strong> kafirins in Sorghum bicolor (L. Moench) endosperm. Protoplasma<br />

171, 64–74.<br />

Stock, R.A., Brink, D.R., Kreikemeier, K.K., Smith, K.K., 1987. Evaluation <strong>of</strong> earlyharvested<br />

and reconstituted grain sorghum in finishing steers. <strong>Journal</strong> <strong>of</strong><br />

Animal <strong>Science</strong> 65, 548–556.<br />

Taylor, J.R.N., Novellie, L., Liebenberg, N.V.D.W., 1984. Sorghum protein body<br />

composition and ultrastructure. <strong>Cereal</strong> Chemistry 61, 69.<br />

Taylor, J.R.N., Schussler, L., Liebenberg, N.V.D.W., 1985. Protein body formation in the<br />

starchy endosperm <strong>of</strong> developing Sorghum bicolor (L.) Moench seeds. S-Afr<br />

Tydskr Plantk 51, 81–86.<br />

Terada, R., Nakajima, M., Isshiki, M., Okaagaki, R.J., Wessler, S.R., Shimamoto, K.,<br />

2000. Antisense waxy genes with highly active promoters effectively suppress<br />

waxy gene expression in transgenic rice. Plant and Cell Physiology 41, 881–888.<br />

J.H. Wong et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 73–82<br />

U.S. Grains Council, 2006. Sorghum, grain sorghum: any <strong>of</strong> various plants <strong>of</strong> the<br />

genus sorghum family Poaceae, a cereal grain.<br />

del Val, G., Yee, B.C., Lozano, R.M., Buchanan, B.B., Ermel, R.W., Lee, Y.M., Frick, O.L.,<br />

1999. Thioredoxin treatment increases digestibility and lowers allergenicity <strong>of</strong><br />

milk. <strong>Journal</strong> <strong>of</strong> Allergy and Clinical Immunology 1103, 690–697.<br />

Wall, J.S., 1971. Disulfide bonds: determination, location and influence on molecular<br />

properties <strong>of</strong> proteins. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 19, 619.<br />

Wester, T.J., Gramlich, S.M., Britton, R.A., Stock, R.A., 1992. Effect <strong>of</strong> grain sorghum<br />

hybrid on in vitro rate <strong>of</strong> starch disappearance and finishing performance <strong>of</strong><br />

ruminants. <strong>Journal</strong> <strong>of</strong> Animal <strong>Science</strong> 70, 2866–2876.<br />

Wong, J.H., Cai, N., Tanaka, C.K., Vensel, W.H., Hurkman, W.J., Buchanan, B.B., 2004.<br />

Thioredoxin reduction alters the solubility <strong>of</strong> proteins <strong>of</strong> wheat starchy endosperm:<br />

an early event in cereal germination. Plant and Cell Physiology 45,<br />

407–415.<br />

Zhang, G., Hamaker, B.R., 1998. Low a-amylase starch digestibility <strong>of</strong> cooked<br />

sorghum flours and the effect <strong>of</strong> protein. <strong>Cereal</strong> Chemistry 75, 710–713.


Fundamental study on protein changes taking place during malting <strong>of</strong> oats<br />

Christina Klose, Beatus D. Schehl, Elke K. Arendt *<br />

Department <strong>of</strong> Food and Nutritional <strong>Science</strong>s, National University <strong>of</strong> Ireland, University College Cork, College Road, Cork, Ireland<br />

article info<br />

Article history:<br />

Received 28 February 2008<br />

Received in revised form 8 July 2008<br />

Accepted 9 July 2008<br />

Keywords:<br />

Oats<br />

Malting<br />

Proteins<br />

Lab-on-a-Chip analysis<br />

1. Introduction<br />

abstract<br />

Oats (Avena sativa) is one <strong>of</strong> the most popular cereals for human<br />

consumption and they have received increased interest because <strong>of</strong><br />

their excellent health-related properties, such as high contents <strong>of</strong><br />

dietary fibre, especially b-glucan, as well as minerals and antioxidants.<br />

The most abundant antioxidants are vitamin E (tocols),<br />

phytic acid, phenolic compounds and avenanthramides; these are<br />

concentrated in the outer layers <strong>of</strong> the kernel (Peterson, 2001). The<br />

major component <strong>of</strong> the endosperm cell walls <strong>of</strong> the oat grain or<br />

caryopsis is a (1 / 3), (1 / 4)-b-D-glucan, known as oat b-glucan<br />

(Ren et al., 2003). The early interest in oat b-glucan arose from<br />

knowledge <strong>of</strong> its highly viscous properties and therefore potential<br />

commercial value as a thickening agent in food formulations and<br />

other industrial applications, which is interesting for healthy<br />

gluten-free products. Whether oats is gluten-free is not yet clear.<br />

However, oats can be tolerated by most but not all people with<br />

coeliac disease and in Finland 69% <strong>of</strong> patients were reported to use<br />

oat products (Peraaho et al., 2004). For this purpose, a gluten-free<br />

beer out <strong>of</strong> malted oats could also be attractive for coeliacs.<br />

Malted oats has been widely used as an ingredient for beer<br />

production in medieval times and before. Nowadays oat malt is<br />

used in the brewing industry mainly as a flavour adjunct for the<br />

* Corresponding author. Tel.: þ353 21 490 2064; fax: þ353 21 427 0213.<br />

E-mail address: e.arendt@ucc.ie (E.K. Arendt).<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.07.014<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 83–91<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

During the malting process, storage proteins are degraded by proteolytic enzymes into small peptides<br />

and amino acids. The activity <strong>of</strong> these enzymes was measured during malting <strong>of</strong> oats and was found to be<br />

increased. To quantify proteolytic degradation, proteins <strong>of</strong> unmalted, germinating and malted grains<br />

were fractionated. After extracting the oat proteins (Osborne fractionation), protein fractions were<br />

analysed using a Lab-on-a-Chip technique, which separates the proteins – based on their molecular<br />

weight – by capillary electrophoresis. This new technique for the analysis <strong>of</strong> proteins was supported by<br />

using two-dimensional gel electrophoresis. In addition, amino acid analysis was carried out. In general<br />

a degradation <strong>of</strong> proteins to small peptides and amino acids could be observed in the globulin, prolamin<br />

and glutelin fractions. In the albumin fraction a protein increase was observed, which is due to the fact<br />

that this fraction contains the majority <strong>of</strong> the metabolically active proteins. Amino acid analysis supported<br />

the observation <strong>of</strong> increased protein amount in the albumin fraction and decreased protein<br />

amounts in the other fractions. Some proteins, which have not been described in the literature, were<br />

detected in the albumin and glutelin fraction, since Lab-on-a-Chip technique allows detection <strong>of</strong> proteins<br />

with low molecular weights <strong>of</strong> 4.5 kDa.<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

production <strong>of</strong> special lagers, ales and stouts. Malting is the initial<br />

step in beer production and strongly defines type and quality <strong>of</strong> the<br />

beer. The main purpose <strong>of</strong> malting is to produce enzymes and to<br />

breakdown cell walls surrounding starch granules. One <strong>of</strong> the most<br />

important physical–chemical changes that occur during malting is<br />

the degradation <strong>of</strong> the proteinaceous matrix that surrounds the<br />

starch granules within the cells <strong>of</strong> the endosperm and their<br />

conversion into soluble peptides and amino acids to provide<br />

substrates for the synthesis <strong>of</strong> proteins in the growing embryo<br />

(Briggs et al., 1981). Malt proteins have a high impact on the<br />

brewing process and the subsequent quality <strong>of</strong> beer. Protein<br />

content and size distribution are <strong>of</strong> particular interest in terms <strong>of</strong><br />

filtering, fermentability, foam and haze stability. Compared to the<br />

commonly used grain for malting, barley and oats has not only<br />

a unique protein composition, but also a high protein content <strong>of</strong><br />

11–15 %, <strong>of</strong> which about a third passes into the final beer. A number<br />

<strong>of</strong> authors (Lapvetelainen et al., 1995; Ma and Harwalkar, 1984;<br />

Robbins et al., 1971; Robert et al., 1985, 1983a; Shewry, 1995; Wu,<br />

1983; Zarkadas, 1982) have reported on the protein composition <strong>of</strong><br />

oats. In general cereal proteins have been classified based on their<br />

solubility (Osborne fractionation), albumins (water soluble), globulins<br />

(salt water soluble), prolamins (soluble in dilute alcohol<br />

solutions) and glutelins (soluble in acids or bases). The difference<br />

between oats and other cereal grains in the structure <strong>of</strong> the<br />

proteinaceous components is consistent with the differences in the<br />

distribution <strong>of</strong> the protein fractions. Oats lacks the protein matrix,<br />

which is characteristic <strong>of</strong> wheat and some other cereals. In wheat


84<br />

and some other cereals, the storage proteins are insoluble in salt<br />

solutions. In terms <strong>of</strong> oats, a large portion <strong>of</strong> the salt water-soluble<br />

globulins also belong to the storage proteins <strong>of</strong> the endosperm.<br />

Although the literary data varies widely, oats contains a relatively<br />

low quantity <strong>of</strong> prolamins and a high amount <strong>of</strong> globulins, up to<br />

80% <strong>of</strong> the total oat protein (Lásztity, 1996). The prolamins (avenins)<br />

<strong>of</strong> oats account for about 15% <strong>of</strong> the total protein and have, similar<br />

to other cereal prolamins, mainly a storage function (Robert et al.,<br />

1983b). Glutelins have typically been difficult to completely solubilise<br />

and therefore values from 5 to 66% have been reported,<br />

depending on the extraction solvent and concentration (Robert<br />

et al., 1985). Most <strong>of</strong> the metabolically active proteins <strong>of</strong> oats are in<br />

the water-soluble albumin fraction. Nevertheless, it cannot be<br />

excluded that the globulin fraction and probably the glutelin fraction<br />

contain this type <strong>of</strong> protein. Oat albumins account for 1–12% <strong>of</strong><br />

the total protein (Lásztity, 1996). Oat protein distribution affects<br />

also the amino acid composition <strong>of</strong> oats. The generally higher lysine<br />

content <strong>of</strong> albumins and globulins causes the relatively high lysine<br />

content in oats, compared to other cereals. In addition, glutamic<br />

acid and proline contents are relatively lower (Lásztity, 1996).<br />

The aim <strong>of</strong> this study was to evaluate the changes in protein and<br />

amino acid composition from the raw oats through germination to<br />

the final malt. For this purpose, a novel analysis technique, the Labon-a-Chip<br />

capillary electrophoresis, was used and was supported<br />

by using two-dimensional gel electrophoresis. This is the first<br />

report on a detailed analysis <strong>of</strong> the protein changes during malting<br />

<strong>of</strong> oats.<br />

2. Experimental<br />

2.1. Materials<br />

The oat grain was harvested in 2005 and obtained from Raisio,<br />

Finland. Malting was carried out using a micro malting machine<br />

(Joe White Malting Systems, Perth, Australia). For this purpose, 8 kg<br />

<strong>of</strong> oat seeds were steeped three times, each stage containing a wet<br />

stage at 13 C (6 h, 4 h, 3 h) and an air rest stage at 18 C (10 h, 7 h,<br />

1 h). After a total <strong>of</strong> 31 h <strong>of</strong> steeping, seeds were germinated for 5<br />

days and 6 h at three stages (17 C, 15 C and 13 C) and kilned in six<br />

stages for 23 h using a schedule that started at 55 C and increased<br />

to 85 C.<br />

2.2. Total protein analysis <strong>of</strong> oats<br />

Samples were taken daily throughout the malting process.<br />

Preliminary analysis <strong>of</strong> all samples with the Lab-on-a-Chip system<br />

revealed degradation <strong>of</strong> all proteins during the process, whereas<br />

largest changes could be observed during the germination process<br />

and only minor changes after germination (before kilning). Thus,<br />

unmalted, 3-day-germinating and kilned (malted) samples were<br />

taken and freeze-dried (VirTus Benchtop 4k, Biopharma Process<br />

Systems, Winchester, UK). Roots were removed and samples were<br />

milled with a Bühler Miag laboratory-scale disc mill (Bühler GmbH,<br />

Braunschweig, Germany) set at a fine setting <strong>of</strong> 0.05 mm and stored<br />

at 80 C.<br />

The proteolytic enzyme activity during the malting process was<br />

measured according to the method <strong>of</strong> Brijs et al. (2002), where<br />

haemoglobin was used as substrate. After incubation, the reaction<br />

was stopped by the addition <strong>of</strong> 10% (w/v) trichloroacetic acid. The<br />

free a-amino nitrogen levels <strong>of</strong> the supernatants were assayed with<br />

trinitrobenzene-sulfonic acid reagent (0.3%, v/v, in 0.2 M sodium<br />

phosphate buffer, pH 8.0) using L-leucine as standard. For this<br />

purpose, the supernatant and trinitrobenzene-sulfonic acid reagent<br />

were incubated and reaction was stopped with 0.2 M HCl. The<br />

absorbance was measured at 340 nm. One unit <strong>of</strong> activity corresponds<br />

to the enzyme activity that liberated 1 mg <strong>of</strong> L-leucine/h<br />

C. Klose et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 83–91<br />

under the assay conditions. Total nitrogen contents <strong>of</strong> the unmalted,<br />

germinating and malted samples were determined by<br />

combustion (according to Analytica-EBC, method 4.3.2) using<br />

a Leco FP-528 nitrogen determinator (St Joseph, MI).<br />

2.3. Modified Osborne fractionation <strong>of</strong> oat proteins<br />

The extraction procedure was carried out once, where first,<br />

unmalted, germinating and malted oats (10 g <strong>of</strong> each sample) were<br />

defatted with 100 mL hexane using a Soxhlet extractor. The defatted<br />

samples were then extracted twice with 100 mL <strong>of</strong> distilled<br />

water (albumin fraction). After water extraction the residue was<br />

extracted twice with 100 mL <strong>of</strong> 5.0% NaCl (globulin fraction). The<br />

remaining flour was then extracted three times with 150 ml <strong>of</strong> 55%<br />

1-propanol þ 1% DTT (prolamin fraction) and the glutelin fraction<br />

together with the residue was extracted three times with a solvent<br />

containing 6 M urea, 2% SDS and 1% DTT. Each extraction was<br />

carried out at room temperature for 10 min and centrifuged afterwards<br />

at 10,000 rpm for 10 min. All supernatants were collected,<br />

further dialyzed against distilled water for 24 h, freeze-dried and<br />

stored at 80 C.<br />

2.4. Lab-on-a-Chip analysis <strong>of</strong> total protein composition<br />

To evaluate protein changes during the malting process, a novel<br />

analysis technique, the Lab-on-a-Chip capillary electrophoresis,<br />

was used. The principles <strong>of</strong> these electrophoretic assays are based<br />

on traditional gel electrophoresis principles that have been transferred<br />

to a chip format. The chip accommodates sample wells, gel<br />

wells and a well for an external standard (ladder). Micro-channels<br />

are fabricated in glass to create interconnected networks among<br />

these wells. During chip preparation, the micro-channels are filled<br />

with a sieving polymer and fluorescence dye. Once the wells and<br />

channels are filled, the chip becomes an integrated electrical circuit.<br />

Charged biomolecules like proteins are electrophoretically driven<br />

by a voltage gradient, similar to slab gel electrophoresis. Because <strong>of</strong><br />

a constant mass-to-charge ratio and the presence <strong>of</strong> a sieving<br />

polymer matrix, the molecules are separated by size. Smaller<br />

fragments are migrating faster than larger ones. Dye molecules<br />

intercalate into protein micelles. These complexes are detected by<br />

laser-induced fluorescence. Data is translated into gel-like images<br />

(bands) and electropherograms (peaks). With the help <strong>of</strong> a ladder<br />

that contains components <strong>of</strong> known sizes, a standard curve <strong>of</strong><br />

migration time versus fragments size is plotted. From the migration<br />

times measured for each fragment in the sample, the size is<br />

calculated. Two markers are run with each <strong>of</strong> the samples bracketing<br />

the overall sizing range. The ‘‘lower’’ and ‘‘upper’’ markers are<br />

internal standards used to align the ladder data with data from the<br />

sample wells. This is necessary to compensate for drift effects that<br />

may occur during the course <strong>of</strong> a chip run (Agilent Technologies,<br />

2005).<br />

For analysis <strong>of</strong> the total protein content, 40 mg <strong>of</strong> milled oat<br />

samples were extracted at room temperature for 5 min with 400 mL<br />

2 M urea solution containing glycerol, Tris–HCl and 0.1 M DTT and<br />

centrifuged for 15 min. Four microliters <strong>of</strong> each supernatant were<br />

denatured using 2 mL <strong>of</strong> Agilent denaturing solution and heated for<br />

5 min at 100 C. After dilution with deionised water, 6 mL were<br />

applied to the Protein 230 þ LabChip and the Protein 80 þ LabChip<br />

for analysis in the Agilent 2100 Bioanalyzer (Agilent Technologies,<br />

Palo Alto, CA).<br />

2.5. Lab-on-a-Chip analysis <strong>of</strong> extracted protein fractions<br />

Ten micrograms <strong>of</strong> each extracted protein fraction were dissolved<br />

in 1 mL <strong>of</strong> their extraction solvents and applied to the Agilent<br />

2100 Bioanalyzer, as described above. For the 230 þ LabChip,


each run included a ladder comprising reference proteins <strong>of</strong> 7, 15,<br />

28, 46, 63, 95, 150 kDa plus an upper marker <strong>of</strong> 240 kDa and a lower<br />

marker <strong>of</strong> 4.5 kDa. Each sample contained an internal standard<br />

comprising the upper and lower marker as well. Any peak detected<br />

below 14 kDa is termed a system peak and is not included in<br />

analysis. The detection performance was also carried out in<br />

a molecular weight range between 4.5 and 95 kDa using the Protein<br />

80 þ LabChip. For this analysis the ladder consisted <strong>of</strong> reference<br />

proteins <strong>of</strong> 3.5, 6.5, 15, 28, 46, 63 kDa plus the upper and the lower<br />

markers <strong>of</strong> 95 and 1.6 kDa. According to the Agilent manual any<br />

peak detected below 5 kDa is named a system peak and is not<br />

included in analysis. Analyses were carried out in triplicate on each<br />

LabChip and a relative standard deviation <strong>of</strong> 8% was considered.<br />

2.6. Two-dimensional gel electrophoresis <strong>of</strong> extracted protein<br />

fractions<br />

Initially oat protein fractions <strong>of</strong> unmalted, germinating and<br />

malted samples were defrosted. All samples were then dissolved in<br />

a solubilisation buffer containing 9 M urea, 4% CHAPS, 0.05% Triton<br />

X100 and 65 mM DTT. Protein solution (125 mL) was applied to each<br />

strip. Isoelectr<strong>of</strong>ocusing (IEF) was carried out using 7-cm IPG 3–10<br />

strips (ReadyStripÔ, BioRad) and a BioRad PROTEAN IEF cell with<br />

controlled cell temperature <strong>of</strong> 20 C. The running conditions were<br />

as follows: passive rehydration: 8 h, active rehydration (50 V): 8 h,<br />

rapid (300 V): 30 min, linear (4000 V): 20,000 V-h, rapid (300 V):<br />

99 h. After IEF was completed, the IPG strips were equilibrated for<br />

15 min in a buffer containing 50 mM Tris–HCl pH 8.8, 6 M urea, 30%<br />

glycerol, 2% SDS and130 mM DTT. After that, the strips were<br />

equilibrated for another 15 min in the same solution except that<br />

DTT was replaced with 130 mM iodoacetamide and traces <strong>of</strong> bromophenol<br />

blue. SDS-PAGE was performed in a BioRad Criterion<br />

Dodeca cell with gels <strong>of</strong> total acrylamide concentration <strong>of</strong> 12.5% at<br />

20 C, sealed with Tris/glycine/SDS running buffer. The gels were<br />

run at 100 V until the tracking dye reached the bottom <strong>of</strong> the gel. To<br />

visualize the proteins, the gels were first incubated in a fixation<br />

solution, containing ethanol and phosphoric acid and then in an<br />

incubation solution containing methanol, phosphoric acid and<br />

ammonium sulphate dissolved in distilled water. The staining was<br />

carried out with the incubation solution containing Coomassie<br />

brilliant blue for 4 days. Analysis was carried out in triplicate.<br />

2.7. Amino acid analysis<br />

The free amino acid pr<strong>of</strong>ile <strong>of</strong> the milled samples was measured<br />

in triplicate using a Jeol JLC-500/V amino acid analyser (Jeol (UK)<br />

Ltd., Garden City, Herts, UK) fitted with a Jeol Na þ high performance<br />

cation exchange column. Prior to analysis, samples were extracted<br />

with a 12% (w/v) trichloracetic acid solution. This procedure was<br />

carried out three times. For the total amino acid pr<strong>of</strong>ile, milled<br />

samples as well as each protein fraction were analysed three times<br />

according to the method <strong>of</strong> Moore and Stein (1963). Due to this<br />

method, tryptophan, which is not stable during hydrolysis, could<br />

not be identified, and glutamine and asparagine were converted to<br />

their acids during hydrolysis, so that they were detected with their<br />

acid.<br />

3. Results<br />

3.1. Proteolytic activity<br />

Total proteolytic enzyme activity was analysed using haemoglobin<br />

as substrate and absorption was measured against L-leucine<br />

as standard. During the malting process an 8.6-fold increase in<br />

proteolytic activity could be observed. The results revealed an<br />

increase in the proteolytic activity level also during steeping (from<br />

C. Klose et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 83–91 85<br />

2.4 to 4.5 mg <strong>of</strong> L-leucine/h/g), during 3 days <strong>of</strong> germination (from<br />

4.5 to 16.3 mg <strong>of</strong> leucine/h/g) and during the last steps <strong>of</strong> malting<br />

(from 16.3 to 20.3 mg <strong>of</strong> leucine/h/g). The highest increase could be<br />

detected during the first 3 days <strong>of</strong> germination.<br />

3.2. Total protein analysis <strong>of</strong> oats<br />

The results <strong>of</strong> total protein analysis using the Protein 230 þ<br />

LabChip showed several protein peaks in the area <strong>of</strong> 23–30 kDa<br />

(area 1), as well as in the area <strong>of</strong> 38–52 kDa (area 2) in all three<br />

samples (results not shown). Peak heights in area 1 demonstrate<br />

that these proteins were degraded by 47%, from about 500 FU<br />

(Fluorescence Units) to about 200 FU during malting. In area 2, peak<br />

heights decreased by 74% after malting. Identical protein patterns<br />

were obtained with the Protein 80 þ LabChip (results not shown).<br />

Both peak areas decreased by 44% (1) or by 71% (2) during the<br />

malting process. However, the Protein 80 þ LabChip reveals more<br />

protein peaks <strong>of</strong> low molecular sizes between 6 and 19 kDa, where<br />

not only a decrease, but also an increase <strong>of</strong> a protein peak with the<br />

molecular size <strong>of</strong> 10 kDa can be seen.<br />

Total nitrogen analysis revealed nitrogen levels <strong>of</strong> 1.734% for the<br />

unmalted, 1.725% for the germinating and 1.750% for the malted<br />

oats. According to Wu (1983), the total nitrogen content <strong>of</strong> oat<br />

kernels decreased at first during germination, which is probably<br />

due to the loss <strong>of</strong> water-soluble nitrogen. After kilning the<br />

percentage <strong>of</strong> total nitrogen was increased.<br />

3.3. Protein fractions<br />

Total protein analysis alone did not give a detailed picture <strong>of</strong> the<br />

changes occurring during the malting process. Proteins were fractioned<br />

according to Osborne to further investigate how oat protein<br />

alters during malting.<br />

3.3.1. Albumin fraction<br />

The electropherogram <strong>of</strong> the water-soluble fraction <strong>of</strong> unmalted<br />

and malted oat samples using the Protein 80 þ LabChip is shown in<br />

Fig. 1A. Results indicate important changes among the overall<br />

malting process. After malting, new or increased protein peaks can<br />

be seen. The peak area reaching from 6 to 9 kDa and the peak at<br />

61 kDa were not present before, but after malting. Increased<br />

protein peaks could be found at 13, 20 and 50 kDa. Proteins with<br />

molecular weights <strong>of</strong> 35–44, 15 and 76 kDa were decreased during<br />

the process.<br />

Results from the Lab-on-a-Chip capillary electrophoresis agreed<br />

with the effect, found when two-dimensional gel electrophoresis<br />

(Fig. 1B) was applied. An increase in intensity <strong>of</strong> protein spots as<br />

well as new spots could be detected in the malted sample. Nevertheless,<br />

results <strong>of</strong> Lab-on-a-Chip analysis appear clearer and more<br />

precisely.<br />

3.3.2. Globulin fraction<br />

Fig. 2A shows the patterns <strong>of</strong> the unmalted, germinating and<br />

malted oat globulin samples on the Protein 230 þ LabChip. Results<br />

reveal a reduction <strong>of</strong> the intensity <strong>of</strong> the group <strong>of</strong> bands located<br />

between 42 and 51 kDa (2) over the malting process. In the area <strong>of</strong><br />

the molecular weights between 20 and 28 kDa (1), the band with<br />

the size <strong>of</strong> 25 kDa disappears. These patterns are clearly depicted in<br />

Fig. 2B, where the molecular weight distribution <strong>of</strong> the unmalted,<br />

germinating and malted globulin fraction can be seen in the electropherogram.<br />

A decrease in the protein peaks reaching from 42 to<br />

51 kDa (2) was also detected. In this molecular weight range,<br />

protein peaks were degraded by 67% during the first 3 days <strong>of</strong><br />

germination and by 76% during the overall malting process,<br />

calculated based on the peak area. The electropherogram shows<br />

that the 25 kDa protein peak decreases in peak height by 30%, but


86<br />

a decrease <strong>of</strong> the concentration <strong>of</strong> all the 20–28 kDa proteins (1),<br />

calculated based on the peak area, could not be observed.<br />

Results <strong>of</strong> two-dimensional gel electrophoresis are shown in<br />

Fig. 2C. A large number <strong>of</strong> overlaying spots in the area <strong>of</strong> 40–50 kDa<br />

in the unmalted sample (picture a) make the identification difficult.<br />

However, a decrease <strong>of</strong> these proteins after the malting process<br />

(picture c) can be seen. This fact concurs with the results from the<br />

Lab-on-a-Chip analysis, where a decrease in peak area 2 (42–<br />

51 kDa) could be detected. The lower molecular weight protein<br />

spots ranging from 20 to 30 kDa seem to be changed only slightly<br />

during the malting process, whereas the protein spots with pH <strong>of</strong><br />

5–6 decrease and the spot with pH 7 seems not to be most affected.<br />

This is in accordance with results <strong>of</strong> Lab-on-a-Chip analysis.<br />

3.3.3. Prolamin (avenin) fraction<br />

The results from the Protein 230 þ LabChip <strong>of</strong> prolamin degradation<br />

during the malting process are illustrated in Fig. 3A. The<br />

electropherogram displays a low molecular weight peak area<br />

extending from 15 to 23 kDa (1), a protein peak with the size <strong>of</strong><br />

31 kDa (2) and an area between the molecular weights from 44 to<br />

51 kDa (3). All prolamins appear to be decreased during the<br />

process. In area 1 a general decrease by about 90% could be<br />

observed. The peak height <strong>of</strong> the 31 kDa peak (2) was reduced<br />

during 3 days <strong>of</strong> germination from 188 to 107 FU and to 52 FU after<br />

kilning, what resembles a decline <strong>of</strong> 72%. The 44–51 kDa part (3)<br />

was reduced from a peak area <strong>of</strong> 68–37 FU during germination and<br />

to 13 FU after the entire malting process, which equals a decrease<br />

by 46 and 81%, respectively.<br />

Fig. 3B shows the gels <strong>of</strong> unmalted (a), germinating (b) and<br />

malted (c) prolamin samples. Protein spot areas <strong>of</strong> 15, 30–35 and<br />

40–45 kDa could be observed. Results correspond approximately<br />

with the peak areas <strong>of</strong> the electropherogram. During germination<br />

the spot intensity decreases and, after kilning, almost all spots<br />

disappeared.<br />

3.3.4. Glutelin fraction<br />

Glutelin proteins decreased during the malting process (Fig. 4A).<br />

A protein peak with molecular weight <strong>of</strong> 9 kDa could be detected,<br />

C. Klose et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 83–91<br />

Unmalted albumins<br />

Malted albumins<br />

Fig. 1. (A) Electropherogram <strong>of</strong> the albumin fraction <strong>of</strong> unmalted and malted samples (Protein 80 þ LabChip). (B) 2D gels <strong>of</strong> unmalted (a) and malted (b) albumin fraction.<br />

A<br />

90<br />

kDa<br />

10<br />

kDa<br />

90<br />

kDa<br />

10<br />

kDa<br />

pH 3<br />

which seems to increase slightly by 10% during malting. Peak areas<br />

ranging from 22 to 27 kDa and from 46 to 51 kDa were detected as<br />

well. A moderate decrease by 40% during the process could be<br />

measured at the protein peaks ranging from 20 to 27 kDa. The<br />

peaks between 46 and 51 kDa appear to be entirely degraded<br />

during the malting procedure.<br />

The acrylamide gels <strong>of</strong> the two-dimensional gel electrophoresis<br />

showed an overall decrease <strong>of</strong> intensity <strong>of</strong> spots after the malting<br />

process (Fig. 4B).<br />

3.4. Amino acid analysis<br />

pH 10<br />

3.4.1. Free and total amino acid composition <strong>of</strong> unmalted and<br />

malted oats<br />

In Table 1 the free amino acid composition <strong>of</strong> unmalted and<br />

malted oats is specified. Glutamic acid (254.1 mg/g oats) was found<br />

to be the major free amino acid in unmalted oats. In addition,<br />

aspartic acid (103.7 mg/g oats) and arginine (94.1 mg/g oats) were<br />

also present in high amounts in the unmalted oat kernel. Furthermore,<br />

no free proline could be detected. Results clearly show an<br />

increase in all free amino acids in oats after the malting process.<br />

Arginine (735.5 mg/g oats) was present in the highest concentration<br />

<strong>of</strong> malted oats, followed by proline (664.4 mg/g oats), threonine<br />

(654.3 mg/g oats), glutamic acid (481.6 mg/g oats) and phenylalanine<br />

(477.5 mg/g oats). Especially the free proline content increased from<br />

0 to 664.4 mg/g oats during malting. And also tyrosine, threonine<br />

and phenylalanine increased significantly by 98, 97, and 95%.<br />

Table 2 shows the total amino acid composition <strong>of</strong> unmalted and<br />

malted oats. Glutamic acid is by far the amino acid with the highest<br />

amount (20.61 mg/g) in unmalted oats. Asparitc acid, followed by<br />

arginine and leucine are also present in relatively high amounts <strong>of</strong><br />

8.27, 7.81 and 7.72 mg/g, respectively. Also the essential amino<br />

acids valine, phenylalanine and lysine could be found in high<br />

amounts <strong>of</strong> 5.86, 5.53 and 4.38 mg/g. Considering a maximum 5%<br />

error in the method used, the following amino acid levels significantly<br />

decrease during malting: tyrosine (by 57%), arginine (by<br />

16%), proline (by 14%), phenylalanine (by 14%) and glycine (by 13%).<br />

Only the amino acid histidin appears to increase during the process<br />

B<br />

a<br />

b


1<br />

by 20%. However, in total, a value <strong>of</strong> 102.23 in the unmalted and<br />

a slightly lower level <strong>of</strong> 94.39 mg/g in the malted sample could be<br />

detected.<br />

3.4.2. Total amino acid composition <strong>of</strong> unmalted, germinating and<br />

malted oat protein fractions<br />

The total amino acid composition <strong>of</strong> oat protein fractions shows<br />

significant differences (Table 3). On the one hand, the amino acid<br />

composition changes during the malting process and on the other<br />

hand, amino acids occur in different amounts in the four fractions.<br />

In the albumin fraction, a total <strong>of</strong> 289.65 g amino acids/kg albumin<br />

could be found in the unmalted sample, whereas glutamic acid has<br />

the highest concentration (39.00 g/kg albumin), followed by<br />

glycine (24.96 g/kg albumin) and aspartic acid (24.09 g/kg<br />

albumin). All these amino acids were increased after malting by 12,<br />

27 and 35%, respectively. These increases led to a total amino acid<br />

content in the malted sample <strong>of</strong> 357.60 g amino acids/kg albumin.<br />

For the unmalted globulin sample, a value <strong>of</strong> 587.18, for the<br />

germinating sample <strong>of</strong> 554.38 and, for the malted sample, 503.11 g<br />

amino acids/kg globulin was found. In the globulin fraction as well,<br />

glutamic acid has the highest, but in a 3.3-fold increased concentration<br />

(127.37 g/kg globulin). Aspartic acid (55.52 g/kg globulin),<br />

2<br />

A<br />

C. Klose et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 83–91 87<br />

2<br />

1<br />

Unmalted globulins<br />

Germinating globulins<br />

Malted globulins<br />

90<br />

kDa<br />

10<br />

kDa<br />

90<br />

kDa<br />

10<br />

kDa<br />

90<br />

kDa<br />

10<br />

kDa<br />

pH<br />

3<br />

40-50 kDa<br />

20-30 kDa<br />

Fig. 2. (A) SDS-PAGE image <strong>of</strong> the globulin fraction <strong>of</strong> unmalted, germinating and malted samples. (B) Electropherogram <strong>of</strong> the globulin fraction <strong>of</strong> unmalted, germinating and<br />

malted samples (Protein 230 þ LabChip). (C) 2D gels <strong>of</strong> unmalted (a), germinating (b) and malted (c) globulin fraction.<br />

B<br />

pH<br />

10<br />

arginine (54.27 g/kg globulin), leucine (44.03 g/kg globulin) and<br />

phenylalanine (34.30 g/kg globulin) appeared also in high<br />

concentrations. Except for lysine, which increased by 14%, all other<br />

amino acids decreased over the process. Glutamic acid was<br />

decreased by 28%, as well as tyrosine (by 22%), valine (by 20%) and<br />

serine (by 19%). For the prolamin fraction, a total <strong>of</strong> 354.07 g/kg<br />

prolamin was determined in the unmalted sample. The total<br />

concentration was decreased during germination to 311.69 g/kg<br />

prolamin and further decreased after kilning to 224.83 g/kg<br />

prolamin. Again, glutamic acid was found in the highest concentration<br />

(125.23 g/kg prolamin), but was decreased markedly during<br />

malting by 43%. Values <strong>of</strong> the other amino acids were a lot lower<br />

than glutamic acid concentration, but were degraded as well;<br />

leucine by 37%, valine by 38%, phenylalanine by 34% and tyrosine by<br />

37%. The amino acid concentration <strong>of</strong> the glutelin fraction was<br />

(192.38 g/kg glutelin) in the unmalted sample and it was decreased<br />

during the process to 169.42 g/kg glutelin and further to 134.67 g/<br />

kg glutelin. As already found in the other fractions, glutamic acid<br />

(37.76 g/kg glutelin) and aspartic acid (18.07 g/kg glutelin) were<br />

apparent in the highest concentrations, followed by arginine<br />

(16.46 g/kg glutelin), leucine (13.31 g/kg glutelin), valine (12.14 g/<br />

kg glutelin) and phenylalanine (11.4 g/kg glutelin). All amino acids<br />

C<br />

a<br />

b<br />

c


88<br />

were degraded markedly, especially proline (by 56%), tyrosine (by<br />

41%) and valine (by 41%).<br />

4. Discussion<br />

4.1. Total protein analysis <strong>of</strong> oats<br />

1<br />

2<br />

The total nitrogen content <strong>of</strong> oat samples during malting was<br />

analysed using the combustion method. In this study a total<br />

nitrogen content <strong>of</strong> 1.734, 1.725 and 1.750% was found for the<br />

unmalted, germinating and malted sample. The protein content<br />

was calculated by multiplying the nitrogen content by the<br />

conversion factor 5.4 according to Mariotti et al. (2008). The<br />

calculated protein content is 9.36, 9.32 and 9.45% for the three<br />

samples. According to the literature (Lásztity, 1996) the nitrogen<br />

content <strong>of</strong> oats ranges from 11 to 15 %, probably determined by<br />

using conversion factors <strong>of</strong> 6.25 or 5.83. According to Mossé (1990)<br />

and Mariotti et al. (2008) this factor leads to overestimated protein<br />

content. Thus, the corrected conversion factor for oats <strong>of</strong> 5.4 was<br />

used for calculating the protein content <strong>of</strong> the oats used, which<br />

leads to the relatively low protein content <strong>of</strong> the analysed oat grain.<br />

A comparison <strong>of</strong> the free amino acid content in unmalted and<br />

malted oats (Table 1) revealed an increase in all amino acids during<br />

malting. This is due to hydrolysis <strong>of</strong> the native protein, which gives<br />

both high molecular weight and low molecular weight protein<br />

breakdown products (peptides and free amino acids) (Kunze, 1996).<br />

Quantitatively, glutamic acid and aspartic acid were most present<br />

in the unmalted sample. In the malted sample, instead arginine,<br />

3<br />

C. Klose et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 83–91<br />

Unmalted prolamins<br />

Germinating prolamins<br />

Malted prolamins<br />

A<br />

90<br />

kDa<br />

10<br />

kDa<br />

90<br />

kDa<br />

10<br />

kDa<br />

90<br />

kDa<br />

10<br />

kDa<br />

proline and threonine had the highest concentrations, which was<br />

also reported by Chittenden et al. (1987), where free proline was<br />

found to increase during germination <strong>of</strong> wheat. The largest increase<br />

could be observed in the concentration <strong>of</strong> free proline, which was<br />

not present in the unmalted sample and thus was increased by<br />

100%. In particular, the essential amino acids isoleucine, leucine,<br />

lysine, threonine, valine and phenylalanine were highly increased<br />

by about 90% each.<br />

Total amino acid composition <strong>of</strong> unmalted oats (Table 2)<br />

corresponds to that reported by Lásztity (1996). There are no great<br />

differences during malting. Total amino acid concentration <strong>of</strong><br />

unmalted oats is 102.23 and <strong>of</strong> malted grain is 94.39 mg/g.<br />

According to Briggs et al. (1981), there is no net loss or gain <strong>of</strong><br />

nitrogen in barley grain during malting, apart from substances<br />

leached during steeping. However, tyrosine was decreased by 57%<br />

and histidine is the only amino acid, which was increased by 20%<br />

during malting. Before and after malting, glutamic acid and aspartic<br />

acid had the highest concentrations, which can be attributed to the<br />

fact that during hydrolysis glutamine and asparagine were converted<br />

into their acids.<br />

4.2. Protein fractions<br />

pH 3<br />

pH 10<br />

4.2.1. Albumin fraction<br />

The albumin fraction <strong>of</strong> other cereals, such as barley, contains<br />

metabolically active proteins (Lásztity, 1996); thus, most <strong>of</strong> the<br />

enzymes <strong>of</strong> oats are probably found in the water-soluble fraction as<br />

well. Since the main purpose <strong>of</strong> malting is the production <strong>of</strong><br />

B<br />

a<br />

ca. 40-45 kDa<br />

ca. 30-35 kDa<br />

ca. 15 kDa<br />

Fig. 3. (A) Electropherogram <strong>of</strong> the prolamin fraction <strong>of</strong> unmalted, germinating and malted samples (Protein 230 þ LabChip). (B) 2D gels <strong>of</strong> unmalted (a), germinating (b) and malted<br />

(c) prolamin fraction.<br />

b<br />

c


enzymes, an increase in the albumin fraction was expected. As<br />

shown in Fig. 1 and Table 3, an increase in several proteins could be<br />

observed. Proteins with molecular weights <strong>of</strong> 14–17, 20–27 and 36–<br />

47 kDa have been described in the literature (Lásztity, 1996) and<br />

could be detected with the Protein 80 þ LabChip, whereas the 36–<br />

47 kDa protein is decreasing during malting. The other proteins (13,<br />

20 and 23–26 kDa) were increased during the process, which can<br />

be explained by the increasing enzyme activity. In addition,<br />

a smaller protein peak area (6–9 kDa) was found with LabChip,<br />

which cannot be detected with gel electrophoresis, because the<br />

commonly used protein analysis technique detects proteins above<br />

14 kDa only. These small proteins and the protein with the size <strong>of</strong><br />

61 kDa are probably protein breakdown products, which are<br />

formed during malting and are water soluble. The formation <strong>of</strong><br />

water-soluble protein breakdown products is well known for barley<br />

proteins (Kunze, 1996). The isoelectric points <strong>of</strong> oat albumins,<br />

detected with two-dimensional gel electrophoresis (Fig. 1B), range<br />

from pH 5 to 8. Similar values have been reported by Ma and<br />

Harwalkar (1984). Amino acid analysis <strong>of</strong> oat albumins revealed an<br />

Table 1<br />

Free amino acid composition <strong>of</strong> the unmalted and malted oat grain (mg/g [dry wt])<br />

Amino Acid<br />

Essential<br />

Unmalted Malted<br />

Isoleucine 15.0 0.7 245.3 5.0<br />

Leucine 18.5 0.9 316.5 9.0<br />

Lysine 42.3 2.0 356.5 9.5<br />

Threonine 21.2 1.0 654.3 12.0<br />

Valine 46.2 2.1 430.3 10.1<br />

Phenylalanine 22.1 1.1 477.5 9.9<br />

Histidin 58.0 2.8 322.5 8.9<br />

Arginine 94.1 4.0 735.5 13.2<br />

Nonessential<br />

Alanine 55.0 2.2 364.5 9.3<br />

Aspartic acid 103.7 5.0 189.1 4.5<br />

Glutamic acid 254.1 5.2 481.6 10.5<br />

Glycine 21.4 1.0 67.2 2.7<br />

Proline 0 664.4 12.1<br />

Serine 25.4 1.2 277.3 6.4<br />

Tyrosine 8.0 0.3 336.9 6.9<br />

1<br />

2<br />

C. Klose et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 83–91 89<br />

Unmalted glutelins<br />

Malted glutelins<br />

3<br />

A<br />

90<br />

kDa<br />

10<br />

kDa<br />

90<br />

kDa<br />

10<br />

kDa<br />

increase in all amino acids, with the exception <strong>of</strong> arginine and<br />

histidine. The increase <strong>of</strong> proline by 48% is notable. The high<br />

increase in albumin–proline probably contributes to the strong<br />

increase in free proline as mentioned in Section 4.1. But also<br />

phenylalanine, aspartic acid and tyrosine were increased markedly.<br />

This effect concurs with results reported by Wu (1983), where an<br />

increase in phenylalanine plus tyrosine was reported during<br />

germination.<br />

4.2.2. Globulin fraction<br />

The salt water-soluble protein <strong>of</strong> oats is known to be the major<br />

protein fraction. This fraction functions as storage proteins and<br />

consists <strong>of</strong> 3S, 7S and 12S globulins (Lásztity, 1996; Shewry and<br />

Halford, 2002). Peterson (1978) showed that the 12S globulin has<br />

a molecular weight <strong>of</strong> 322 kDa and consists <strong>of</strong> two subunits with<br />

molecular weights <strong>of</strong> about 32 and 22 kDa, called the a- and bsubunits.<br />

As shown in Fig. 2, these subunits could be detected with<br />

the Protein 230 þ LabChip, covered in area 1. This type <strong>of</strong> protein<br />

seems to be affected only slightly by the degradation process<br />

B<br />

pH 3 pH 10<br />

Fig. 4. (A) Electropherogram <strong>of</strong> the glutelin fraction <strong>of</strong> unmalted and malted samples (Protein 80 þ LabChip). (B) 2D gels <strong>of</strong> unmalted (a) and malted (b) glutelin fraction.<br />

Table 2<br />

Total amino acid composition <strong>of</strong> the unmalted and malted oat grain (mg/g [dry wt])<br />

Amino Acid<br />

Essential<br />

Unmalted Malted<br />

Isoleucine 3.88 0.2 3.79 0.1<br />

Leucine 7.72 0.4 7.16 0.3<br />

Lysine 4.38 0.2 4.36 0.2<br />

Threonine 3.22 0.1 3.34 0.1<br />

Valine 5.86 0.2 5.17 0.2<br />

Phenylalanine 5.53 0.2 4.77 0.2<br />

Histidin 3.00 0.1 3.59 0.1<br />

Arginine 7.81 0.4 6.57 0.3<br />

Nonessential<br />

Alanine 5.31 0.2 4.79 0.2<br />

Aspartic acid 8.27 0.4 8.89 0.4<br />

Glutamic acid 20.61 1.0 19.38 0.9<br />

Glycine 5.56 0.2 4.85 0.2<br />

Proline 5.94 0.3 5.09 0.2<br />

Serine 4.43 0.2 4.44 0.2<br />

Tyrosine 2.87 0.1 1.25 0.0<br />

a<br />

b


90<br />

Table 3<br />

Average amino acid compositions <strong>of</strong> the different protein fractions extracted from oats grown in Finland (mg/g protein fraction)<br />

Albumin Globulin<br />

1a 2a 3a 1b 2b 3b<br />

Isoleucine 8.33 0.4 7.80 0.3 10.03 0.5 26.49 1.3 26.25 1.3 23.21 1.1<br />

Leucine 17.45 0.8 16.45 0.8 20.74 1.0 44.03 2.2 42.09 2.1 38.97 1.9<br />

Lysine 13.91 0.6 11.88 0.5 15.41 0.7 18.89 0.9 19.40 0.9 21.53 1.0<br />

Threonine 11.54 0.5 11.11 0.5 14.59 0.7 20.85 1.0 19.66 0.9 19.04 0.9<br />

Valine 12.13 0.6 11.18 0.5 14.61 0.7 33.34 1.6 29.71 1.4 26.68 1.3<br />

Phenylalanine 9.73 0.4 11.68 0.5 13.18 0.6 34.30 1.7 31.50 1.5 29.25 1.4<br />

Histidin 8.15 0.4 6.92 0.3 8.00 0.4 18.16 0.9 17.92 0.8 17.34 0.8<br />

Arginine 16.88 0.8 13.33 0.6 15.73 0.7 54.27 2.7 52.31 2.6 44.96 2.2<br />

Alanine 17.52 0.8 15.41 0.7 20.82 1.0 27.64 1.3 27.08 1.3 27.53 1.3<br />

Aspartic acid 24.09 1.2 24.07 1.2 32.46 1.6 55.52 2.7 52.93 2.6 50.55 2.5<br />

Glutamic acid 39.00 1.9 35.07 1.7 43.84 2.1 127.37 6.3 114.37 5.7 91.51 4.5<br />

Glycine 24.96 1.2 23.90 1.1 31.57 1.5 28.81 1.4 28.63 1.4 26.83 1.3<br />

Proline 13.45 0.6 14.77 0.7 19.92 0.9 23.09 1.1 24.46 1.2 21.39 1.0<br />

Serine 20.75 1.0 20.60 1.0 26.74 1.3 29.97 1.4 26.93 1.3 24.32 1.2<br />

Tyrosine 12.44 0.6 12.65 0.6 16.64 0.8 18.84 0.9 16.85 0.8 14.72 0.7<br />

Total 289.65 276.88 357.60 587.18 554.38 503.11<br />

Prolamin Glutelin<br />

1c 2c 3c 1d 2d 3d<br />

Isoleucine 11.44 0.5 10.48 0.5 7.48 0.3 8.96 0.4 8.63 0.4 6.21 0.3<br />

Leucine 33.34 1.6 29.87 1.4 21.15 1.0 13.31 0.6 13.43 0.6 10.37 0.5<br />

Lysine 4.84 0.2 5.01 0.2 4.03 0.2 7.50 0.3 3.74 0.1 4.79 0.2<br />

Threonine 7.51 0.3 7.02 0.3 5.21 0.2 6.55 0.3 5.85 0.2 4.65 0.2<br />

Valine 18.66 0.9 16.69 0.8 11.54 0.5 12.14 0.6 11.00 0.5 7.21 0.3<br />

Phenylalanine 21.73 1.0 18.53 0.9 14.33 0.7 11.40 0.5 10.53 0.5 9.81 0.4<br />

Histidin 9.51 0.4 8.77 0.4 7.54 0.3 7.93 0.3 6.67 0.3 7.37 0.3<br />

Arginine 17.13 0.8 16.36 0.8 11.23 0.5 16.46 0.8 15.00 0.7 12.08 0.6<br />

Alanine 14.19 0.7 13.04 0.6 9.81 0.4 9.05 0.4 8.55 0.4 7.36 0.3<br />

Aspartic acid 12.72 0.6 12.17 0.6 10.91 0.5 18.07 0.9 17.14 0.8 12.31 0.6<br />

Glutamic acid 125.23 4.2 104.82 5.0 71.79 3.5 37.76 1.8 32.00 1.5 27.48 1.3<br />

Glycine 8.26 0.4 7.93 3.9 6.24 0.3 8.82 0.4 8.03 0.4 6.17 0.3<br />

Proline 22.51 1.1 19.97 0.9 15.35 0.7 6.76 0.3 5.81 0.2 2.98 0.1<br />

Serine 9.18 0.4 8.31 0.4 6.30 0.3 8.89 0.4 7.48 0.3 5.97 0.2<br />

Tyrosine 4.60 0.2 4.06 0.2 2.89 0.1 6.58 0.3 5.79 0.2 3.85 0.1<br />

Total 354.07 311.69 224.83 192.38 169.42 134.67<br />

Sample 1a–1d ¼ unmalted; sample 2a–2d ¼ germinating; sample 3a–3d ¼ malted.<br />

taking place during malting. In contrast to the 12S globulin, the 7S<br />

globulin, which mainly consists <strong>of</strong> a 55 kDa polypeptide and is<br />

found in area 2 in Fig. 2, is decreasing during the malting process.<br />

This area probably represents only the 7S globulin, although the<br />

disulfide-bonded dimer, formed by the a- and b-subunit has<br />

a molecular weight <strong>of</strong> 54 kDa. Due to the fact that analysis was<br />

performed under reducing conditions, no disulfide-bond<br />

remained. The 3S fraction was described in the literature as<br />

a protein consisting <strong>of</strong> a 15 and 21 kDa subunit (Lásztity, 1996).<br />

The 15 kDa polypeptide could not clearly be detected with the<br />

LabChip, but the 21 kDa component was probably found in area 1<br />

(Fig 2B). Amino acid analysis revealed decreasing proteins as well,<br />

since the total amino acid amount <strong>of</strong> the globulin fraction is<br />

decreasing, which concurs with the degradation in the globulin<br />

fraction during malting. In particular the lysine content was found<br />

to be increased during the process, which is remarkable from<br />

a nutritional point <strong>of</strong> view and was also found by Wu (1983), who<br />

found an increasing lysine content in germinating oats. This can be<br />

attributed to the fact that the globulin fraction, which is high in<br />

lysine, was degraded during the malting process. This probably<br />

revealed free lysine, which is water soluble and thus, was<br />

extracted with the albumin and globulin fraction. The amino acid<br />

composition <strong>of</strong> globulins corresponds to that reported from<br />

Lásztity (1996). Glutamic acid is the most frequently found amino<br />

acid and a decrease by 28% <strong>of</strong> this amino acid was detected, which<br />

is probably due to decreases in the 12S globulin subunit, since the<br />

12S subunit has been described to contain a higher amount <strong>of</strong><br />

glutamic acid than the 7S globulin.<br />

C. Klose et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 83–91<br />

4.2.3. Prolamin fraction<br />

Since oats main storage protein fraction is the globulin fraction,<br />

oats prolamin fraction only accounts for 15% <strong>of</strong> total oat protein<br />

(Lásztity, 1996), which is different compared to other cereals, such<br />

as wheat, barley or rye. However, two-dimensional gel electrophoresis<br />

revealed more than 20 components (Fig. 3B, a). The Labon-a-Chip<br />

analysis disclosed three main peak areas in the same<br />

molecular weight range (Fig. 3A). The pH <strong>of</strong> isoelectric focusing was<br />

found to range from pH 6 to 9. These results concur with those<br />

reported by Lásztity (1996). During the malting process the<br />

prolamin fraction was degraded almost entirely, which is in<br />

agreement with other germinating cereals such as barley, reported<br />

by Celus et al. (2006). The degradation by endoproteases is due to<br />

the fact that storage proteins are supporting the embryo during the<br />

first stages <strong>of</strong> development. The oats prolamin fraction is like many<br />

other cereals also rich in the amino acids proline and glutamine<br />

(Simpson, 2001). Thus, prolamin proteins most likely function as<br />

storage proteins, because the aim <strong>of</strong> the glutamine- and prolinerich<br />

polypeptides is to supply the embryo with amino acids and<br />

nitrogen during germination.<br />

4.2.4. Glutelin fraction<br />

The proteins remaining after removing the water-soluble albumins,<br />

the salt water-soluble globulins and the alcohol soluble<br />

prolamins are called glutelins. Unfortunately, these extractions are<br />

generally incomplete and some nitrogen remains in the residue.<br />

The quantity <strong>of</strong> glutelins observed is <strong>of</strong>ten directly related to the<br />

efficiency <strong>of</strong> the preceding extractions <strong>of</strong> the albumins, globulins


and prolamins (Shewry et al., 1978). Thus, it is likely to extract not<br />

only glutelins with the last extraction step, but also remaining<br />

proteins from the water, salt water and alcohol soluble fractions,<br />

especially the globulin fraction (Lásztity, 1996). Robert et al. (1985)<br />

reported two-dimensional gel electrophoresis <strong>of</strong> oats proteins,<br />

where they found a few proteins spots, which probably are true<br />

candidates for the glutelin fraction, along with proteins spots,<br />

which are most likely a- and b-globulin subunits. A similar effect<br />

can be seen in our results <strong>of</strong> the glutelin fraction (Fig. 4). Compared<br />

to the gels from the globulin fraction (Fig. 2C) it seems likely that<br />

not all globulins had been extracted with the 5.0% NaCl solution and<br />

remained until the extraction with urea, SDS and DTT, since the<br />

molecular weights <strong>of</strong> 22–27 and 42–51 kDa could be found in both<br />

fractions. The proteins ranging from 22 to 27 kDa are probably bsubunit<br />

proteins from the 12S globulin. However, the LabChip<br />

revealed a so far unknown polypeptide with molecular weight <strong>of</strong><br />

9 kDa, which is most likely a glutelin protein, that seems to be<br />

unaffected by the malting process.<br />

In conclusion, this study reveals an understanding about the<br />

protein changes taking place during malting <strong>of</strong> oats. In general<br />

a degradation <strong>of</strong> the proteins to small peptides and amino acids could<br />

be observed in any fraction except in the albumin fraction, in which<br />

some proteins increased in amount. This is due to the fact that these<br />

represent most likely metabolically active proteins. Amino acid<br />

analysis supported the observation <strong>of</strong> increased protein amount in<br />

the albumin fraction and decreased protein amounts in the other<br />

fractions. In addition, so far unknown proteins could be detected in<br />

the albumin and glutelin fraction with the Lab-on-a-Chip analysis.<br />

This technique was found to be appropriate for analysis <strong>of</strong> degrading<br />

or increasing proteins, as it revealed repeatable and reliable results,<br />

which could be validated by using common protein analysis techniques<br />

such as two-dimensional gel electrophoresis.<br />

Acknowledgements<br />

This project was supported by the Irish Government under the<br />

National Development Plan, 2006–2013. The authors would like to<br />

thank Mrs. Paula O’Connor for her support in amino acid analysis.<br />

References<br />

Agilent Technologies, 2005. Agilent 2100 Bioanalyzer - 2100 Expert User’s Guide.<br />

Agilent Technologies, Waldbronn, Germany.<br />

Briggs, D.E., Hough, J.S., Stevens, R., Young, T.W., 1981. Malting and Brewing <strong>Science</strong>,<br />

second ed. Chapman & Hall, London, U.K.<br />

Brijs, K., Trogh, I., Jones, B.L., Delcour, J.A., 2002. Proteolytic enzymes in germinating<br />

rye grains. <strong>Cereal</strong> Chemistry 79, 423–428.<br />

Celus, I., Brijs, K., Delcour, J.A., 2006. The effects <strong>of</strong> malting and mashing on barley<br />

protein extractability. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 44, 203–211.<br />

C. Klose et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 83–91 91<br />

Chittenden, C.G., Laidman, D.L., Ahmad, N., Wyn Jones, R.G., 1987. Amino acids and<br />

quaternary nitrogen compounds in the germinating wheat grain. Phytochemistry<br />

17, 1209–1216.<br />

Kunze, W., 1996. Technology Brewing and Malting, seventh ed. Versuchs- und<br />

Lehranstalt für Brauerei, Berlin, Germany.<br />

Lapvetelainen, A., Bietz, J.A., Huebner, F.R., 1995. Reversed-phase high-performance<br />

liquid chromatography <strong>of</strong> oat proteins: application to cultivar comparison and<br />

analysis <strong>of</strong> the effect <strong>of</strong> wet processing. <strong>Cereal</strong> Chemistry 72, 259–264.<br />

Lásztity, R., 1996. The Chemistry <strong>of</strong> <strong>Cereal</strong> Proteins, second ed. CRC Press, Boca<br />

Raton, Florida.<br />

Ma, C.-Y., Harwalkar, V.R., 1984. Chemical characterisation and functionality<br />

assessment <strong>of</strong> oat protein fractions. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry<br />

32, 144–149.<br />

Mariotti, F., Tome, D., Mirand, P.P., 2008. Converting nitrogen into protein – beyond<br />

6.25 and Jones’ factors. Critical Reviews in Food <strong>Science</strong> and Nutrition 48,<br />

177–184.<br />

Moore, S., Stein, W.H., 1963. Chromatographic determination <strong>of</strong> amino acids by<br />

the use <strong>of</strong> automatic recording equipment. Methods in Enzymology 6,<br />

819–831.<br />

Mossé, J., 1990. Nitrogen to protein conversion factor for ten cerals and six legumes<br />

or oilseeds. A reappraisal <strong>of</strong> its definition and determination. Variation<br />

according to species and to seed protein. <strong>Journal</strong> <strong>of</strong> Agricultural and Food<br />

Chemistry 38, 18–24.<br />

Peraaho, M., Kaukinen, K., Mustalahti, K., Vuolteenaho, N., Maki, M., Laippala, P.,<br />

Collin, P., 2004. Effect <strong>of</strong> an oats-containing gluten-free diet on symptoms and<br />

quality <strong>of</strong> life in coeliac disease. A randomized study. Scandinavian <strong>Journal</strong> <strong>of</strong><br />

Gastroenterology 39, 27–31.<br />

Peterson, D.M., 1978. Subunit structure and composition <strong>of</strong> oat seed globulin. Plant<br />

Physiology 62, 506–509.<br />

Peterson, D.M., 2001. Oat antioxidants – mini review. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 33,<br />

115–129.<br />

Ren, Y., Ellis, P.R., Ross-Murphy, S.B., Wang, Q., Wood, P.J., 2003. Dilute and semidilute<br />

solution properties <strong>of</strong> (1 / 3), (1 / 4)-b-D-glucan, the endosperm cell<br />

wall polysaccharide <strong>of</strong> oats (Avena sativa L.). Carbohydrate Polymers 53,<br />

401–408.<br />

Robbins, G.S., Pomeranz, Y., Briggle, L.W., 1971. Amino acid composition <strong>of</strong> oat<br />

groats. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 19, 536.<br />

Robert, L.S., Nozzolillo, C., Altosaar, I., 1985. Characterization <strong>of</strong> oat (Avena sativa L.)<br />

residual proteins. <strong>Cereal</strong> Chemistry 62, 276–279.<br />

Robert, L.S., Matlashewski, G.J., Adeli, K., Nozzolillo, C., Altosaar, I., 1983a. Electrophoretic<br />

and developmental characterization <strong>of</strong> oat (Avena sativa L.)<br />

globulins in cultivars <strong>of</strong> different protein content. <strong>Cereal</strong> Chemistry 60,<br />

231–234.<br />

Robert, L.S., Nozzolillo, C., Altosaar, I., 1983b. Molecular weight and charge<br />

heterogeneity <strong>of</strong> prolamins (avenins) from nine oat (Avena sativa L.) cultivars <strong>of</strong><br />

different protein content and from developing seeds. <strong>Cereal</strong> Chemistry 60,<br />

438–442.<br />

Shewry, P.R., 1995. Plant storage proteins. Biological Reviews 70, 375–426.<br />

Shewry, P.R., Halford, N.G., 2002. <strong>Cereal</strong> seed storage proteins: structures, properties<br />

and role in grain utilization. <strong>Journal</strong> <strong>of</strong> Experimental Botany 53,<br />

947–958.<br />

Shewry, P.R., Hill, J.M., Pratt, H.M., Leggatt, M.M., Miflin, B.J., 1978. An evaluation <strong>of</strong><br />

techniques for the extraction <strong>of</strong> hordein and glutelin from barley seed and<br />

a comparison <strong>of</strong> the protein composition <strong>of</strong> Bomi and Risø 1580. <strong>Journal</strong> <strong>of</strong><br />

Experimental Botany 29, 677–692.<br />

Simpson, D.J., 2001. Proteolytic degradation <strong>of</strong> cereal prolamins – the problem with<br />

proline. Plant <strong>Science</strong> 161, 825–838.<br />

Wu, Y.V., 1983. Effect <strong>of</strong> germination on oats and oat protein. <strong>Cereal</strong> Chemistry 60,<br />

418–420.<br />

Zarkadas, C.G., 1982. A comparison <strong>of</strong> the amino acid composition <strong>of</strong> two<br />

commercial oats groats. <strong>Cereal</strong> Chemistry 59, 323–327.


Effect <strong>of</strong> milling, pasta making and cooking on minerals in durum wheat<br />

Francesco Cubadda a, *, Federica Aureli a , Andrea Raggi a , Marina Carcea b<br />

a Istituto Superiore di Sanità, Viale Regina Elena 299, 00161 Rome, Italy<br />

b Istituto Nazionale di Ricerca per gli Alimenti e la Nutrizione, Via Ardeatina 546, 00178 Rome, Italy<br />

article info<br />

Article history:<br />

Received 14 April 2008<br />

Received in revised form 8 July 2008<br />

Accepted 9 July 2008<br />

Keywords:<br />

Durum wheat<br />

Pasta<br />

Minerals<br />

Trace elements<br />

Milling<br />

Processing<br />

Cooking<br />

1. Introduction<br />

abstract<br />

<strong>Cereal</strong>s and derived products are among the major dietary<br />

sources <strong>of</strong> essential elements for humans. The contribution <strong>of</strong><br />

cereal products to the estimated dietary intake <strong>of</strong> several minerals<br />

and nutritionally beneficial trace elements is about 20–30% <strong>of</strong> the<br />

total intake in Western countries (Carcea et al., 2007). In the case <strong>of</strong><br />

iron and manganese the contribution is as high as 40–50% (Carcea<br />

et al., 2007). These figures are even higher in other regions <strong>of</strong> the<br />

world (Choi and Kim, 2007; Hattori et al., 2004) and especially in<br />

developing countries (Hussein and Brüggemann, 1997).<br />

Wheat is the main cereal crop used for human consumption in<br />

many areas worldwide. Common wheat (Triticum aestivum L.) is<br />

widely used for breadmaking, whereas durum wheat, i.e., Triticum<br />

turgidum L. subsp. durum (Desf.) Husn., is mainly employed in the<br />

production <strong>of</strong> other food items, pasta being the most popular. In<br />

Italy, the main pasta producer, pasta is a staple food. However pasta<br />

Abbreviations: RDA, daily recommended dietary allowance.<br />

* Corresponding author. Tel.: þ39 06 4990 3643; fax: þ39 06 4990 2540.<br />

E-mail address: francesco.cubadda@iss.it (F. Cubadda).<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.07.008<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 92–97<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

The effect <strong>of</strong> technological processing on the contents <strong>of</strong> eight minerals – i.e., calcium, copper, iron,<br />

magnesium, phosphorous, potassium, selenium, and zinc – was investigated in pasta making. Milling <strong>of</strong><br />

durum wheat as well as pasta making were carried out in a pilot plant by using three different grain<br />

samples. Pasta samples purchased on the market were also surveyed to gain information on the mineral<br />

content <strong>of</strong> commercial products. The effect <strong>of</strong> cooking was also investigated in order to determine the<br />

retention <strong>of</strong> the selected elements in the final ‘ready-to-eat’ product. Analyte concentrations in whole<br />

grains, semolina, pasta and cooked pasta were determined by inductively coupled plasma-mass<br />

spectrometry.<br />

Conventional roller milling significantly reduced the content <strong>of</strong> each mineral in durum wheat grains.<br />

However concentration losses as a consequence <strong>of</strong> milling widely differed among elements, from 16%<br />

for Se to 66% for Mg and Zn on a dry weight basis. Retention <strong>of</strong> elements after milling followed the order<br />

Se > Ca > Cu > P z K > Fe > Mg z Zn. Pasta making had little effect on element concentrations in<br />

semolina. Cooking caused an increase in the calcium content <strong>of</strong> pasta whereas the concentrations <strong>of</strong> the<br />

other elements were either unchanged or slightly reduced (0–18% on a dry weight basis) except<br />

potassium, which showed a decrease <strong>of</strong> 74%.<br />

Commercial pasta samples showed concentrations <strong>of</strong> minerals similar to those <strong>of</strong> the experimental<br />

samples, except selenium which was higher due to the use <strong>of</strong> imported wheat with higher levels <strong>of</strong><br />

selenium in industrial semolina production. Overall, pasta appears to be a valuable source <strong>of</strong> several<br />

minerals, especially selenium, copper, magnesium, and zinc.<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

consumption has become widespread in several countries (UNIPI,<br />

2007). Notwithstanding its popularity, the mineral pr<strong>of</strong>ile <strong>of</strong> pasta<br />

as influenced by processing (i.e., milling and pasta making) and<br />

cooking has been the subject <strong>of</strong> systematic investigations mainly in<br />

the 1980s (Albrecht et al., 1987; Brondi et al., 1984; Ranhotra et al.,<br />

1984, 1985; Toepfer et al., 1972; Yaseen, 1993). The introduction <strong>of</strong><br />

new cultivars as well as alterations in agronomic practices and<br />

changes in the processing conditions do require updated studies.<br />

These studies should employ state-<strong>of</strong>-the-art techniques and<br />

include analytical quality control schemes allowing, for instance,<br />

precise and accurate determinations <strong>of</strong> essential microelements<br />

such as selenium.<br />

Milling is the critical process affecting the concentrations <strong>of</strong><br />

inorganic elements in wheat-derived food products. As the outer<br />

parts <strong>of</strong> the kernel, especially the aleurone layer and the germ, are<br />

richer in minerals when compared to the starchy endosperm,<br />

conventional milling reduces their content in flour (semolina, in the<br />

case <strong>of</strong> durum wheat) and concentrates them in the milling residues<br />

(Brondi et al., 1984). However, it has been reported that<br />

differences in the mineral content likely exist even between the<br />

inner endosperm and the outer endosperm (Pomeranz, 1988). The<br />

grain shape and texture (which both depend also on cultivar) and


the technical conditions <strong>of</strong> milling, primarily the extraction rate, are<br />

important in determining the extent <strong>of</strong> mineral loss. However,<br />

when all these variables are fixed, the distribution <strong>of</strong> the mineral in<br />

the various milling fractions ultimately depends on how the<br />

element is unevenly distributed within the kernel. Therefore, it will<br />

vary on an element-specific basis.<br />

There is little information on the effect <strong>of</strong> milling on durum<br />

wheat minerals (Brondi et al., 1984; Toepfer et al., 1972). This<br />

information is needed to assess the nutritional significance <strong>of</strong><br />

mineral loss as a consequence <strong>of</strong> conventional roller milling and to<br />

provide basic knowledge in order to establish, for instance, whether<br />

there is a need for fortification, which minerals should be supplemented<br />

(in bioaccessible form), and at what levels. Limited and<br />

sometimes contradictory data are available on the effect <strong>of</strong> further<br />

processing <strong>of</strong> semolina into pasta as well as pasta cooking (Albrecht<br />

et al., 1987; Ranhotra et al., 1984, 1985; Yaseen, 1993).<br />

The present study was undertaken to determine the effect <strong>of</strong><br />

conventional roller milling, pasta making and cooking on eight<br />

nutritionally important minerals (namely calcium, copper, iron,<br />

magnesium, phosphorous, potassium, selenium, and zinc) in durum<br />

wheat. Milling and pasta making pilot plants were employed for the<br />

manufacture <strong>of</strong> semolina and long (‘spaghetti’ type) pasta in order to<br />

be able to control the whole processing chain. Commercial pasta<br />

samples from major Italian brands were purchased on the market<br />

and included in this study for comparison. The effect <strong>of</strong> cooking was<br />

investigated in order to gain information on the retention <strong>of</strong> the<br />

selected elements in the final ‘ready-to-eat’ product.<br />

2. Experimental<br />

2.1. Milling and pasta making<br />

Three samples <strong>of</strong> durum wheat grain <strong>of</strong> about 10 kg each were<br />

included in the study. The samples belonged to three different<br />

cultivars, namely, Appio, Duilio, and Simeto (hereafter indicated as<br />

samples 1, 2, and 3) grown in Southern Italy and widely used in Italy<br />

for industrial pasta production.<br />

Semolina was obtained from each sample by cleaning and<br />

tempering the grains for 36–40 h to 16.5% moisture and then<br />

milling them in a pilot mill (Model MLU 202, Bühler, Uzwill,<br />

Switzerland) equipped with three breaks, three reduction rolls, six<br />

steel screens and coupled with a purifier to give standard grade<br />

semolina, following Approved Methods 26-10 A and 26-41 (AACC,<br />

2000). Semolina yield (ash


94<br />

(model 510) and an ADX-500 autodilutor (both from CETAC Technologies,<br />

Omaha, NE, USA), was used for quantification <strong>of</strong> the<br />

analytes. Details <strong>of</strong> the instrumentation and the operating conditions<br />

are reported elsewhere (Cubadda et al., 2002). The analytical<br />

masses were 43 Ca, 63 Cu, 57 Fe, 39 K, 24 Mg, 31 P, 77þ82 Se, and 64 Zn.<br />

Multielemental calibration standards were prepared from<br />

1000 mg L 1 stock solutions <strong>of</strong> individual elements (BDH, Poole,<br />

England) by dilution with 3% v/v ultrapure concentrated HNO3. In<br />

all measurements, rhodium (20 mg L 1 ) was selected as internal<br />

standard for correction <strong>of</strong> matrix effects and instrumental drift.<br />

Correction <strong>of</strong> 44 Ca 16 OH interference on 57 Fe was accomplished by<br />

application <strong>of</strong> a mathematical equation calculated daily as reported<br />

previously (Cubadda et al., 2002).<br />

Randomly selected samples were analysed on different days to<br />

verify the precision <strong>of</strong> results. Samples <strong>of</strong> the two reference<br />

materials were analysed in each analytical run to check accuracy <strong>of</strong><br />

measurements. Digestion blanks were run together with samples<br />

belonging to the same analytical batch and their signal was subtracted<br />

from that <strong>of</strong> the sample when calculating analyte<br />

concentrations.<br />

2.5. Statistical analysis<br />

The existence <strong>of</strong> significant differences in the element concentrations<br />

(on a dry weight basis) among the studied matrixes on<br />

account <strong>of</strong> the various treatments (i.e., milling, pasta making,<br />

cooking) was determined by analysis <strong>of</strong> variance. The test resulted<br />

significant for all metals (p ¼ 0.05) and thus a multiple comparison<br />

test (Scheffé test) was performed to identify which treatments<br />

caused significant variations (p ¼ 0.05).<br />

3. Results and discussion<br />

The results obtained for the reference materials analysed for<br />

quality control purposes are summarized in Table 1. Good agreement<br />

was observed with the certified or best estimated values <strong>of</strong><br />

each element, indicating effective recovery <strong>of</strong> analytes after<br />

digestion and subsequent accurate detection. Precision expressed<br />

as the coefficient <strong>of</strong> variation (CV) was, on average, 6% for calcium,<br />

3% for copper, 2% for iron, 1% for potassium, 1% for magnesium, 2%<br />

for phosphorous, 2% for selenium, and 4% for zinc.<br />

Tables 2–4 show the element concentrations measured in the<br />

three grain samples selected for this study and in their derived<br />

products, including cooked pasta. Each result is the average <strong>of</strong> four<br />

experimental replicates, on which duplicate subsampling was<br />

carried out for analyses. Data are reported both on a dry and on<br />

a fresh weight basis. Average moisture content <strong>of</strong> samples was 9.8%,<br />

Table 1<br />

Results obtained for the reference materials analysed (N ¼ 6)<br />

Element CRM 189 (wholemeal flour) RM 8436 (durum wheat flour)<br />

Ca b<br />

Cu b<br />

Fe b<br />

K d<br />

Mg d<br />

P d<br />

Se b<br />

Zn b<br />

Found Certified Found Best estimated<br />

Mean (c.i.) a<br />

Mean (c.i.) a<br />

Mean (c.i.) a Mean (c.i.) a<br />

540 (22) [520] c<br />

276 (14) 278 (26)<br />

6.36 (0.13) 6.4 (0.2) 4.28 (0.10) 4.30 (0.69)<br />

69.2 (0.6) 68.3 (1.9) 41.1 (1.2) 41.5 (4.0)<br />

6.23 (0.09) [6.3] c<br />

3.20 (0.02) 3.18 (0.14)<br />

1.92 (0.01) [1.9] c<br />

1.12 (0.02) 1.07 (0.08)<br />

5.49 (0.03) [5.3] c<br />

2.91 (0.06) 2.90 (0.22)<br />

0.137 (0.002) 0.132 (0.010) 1.25 (0.02) 1.23 (0.09)<br />

56.6 (1.1) 56.5 (1.7) 22.3 (1.1) 22.2 (1.7)<br />

a<br />

Uncertainty as half-width <strong>of</strong> the 95% confidence interval <strong>of</strong> the mean.<br />

b 1<br />

Concentrations in mg g .<br />

c<br />

Indicative value.<br />

d 1<br />

Concentrations in mg g .<br />

F. Cubadda et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 92–97<br />

12.6%, 10.1% and 61.1% for grain, semolina, pasta and cooked pasta,<br />

respectively.<br />

In order to allow easy recognition <strong>of</strong> the effects <strong>of</strong> processing<br />

and cooking, the overall percent variation in element concentrations<br />

caused by the different treatments in relation to their original<br />

levels in grain (¼100%) is shown in Tables 2–4, together with the<br />

percent variation <strong>of</strong> the element concentrations in each type <strong>of</strong><br />

sample as a consequence <strong>of</strong> a specific treatment.<br />

3.1. Effect <strong>of</strong> milling<br />

As expected, milling significantly reduced the concentrations <strong>of</strong><br />

all elements (p < 0.01), with average losses ranging from 16% for<br />

selenium to 66% for magnesium and zinc on a dry weight basis<br />

(Tables 2–4).<br />

A modest reduction <strong>of</strong> selenium concentration upon milling<br />

(8%–20%) has also been found in studies on common wheat (Lyons<br />

et al., 2005; Toepfer et al., 1972). Similar to sulphur, selenium,<br />

which occurs predominantly as selenomethionine in wheat grains,<br />

is mostly protein-bound and more evenly distributed throughout<br />

the kernel when compared to other minerals (Lyons et al., 2005).<br />

Therefore, due to a higher proportion being stored in the endosperm,<br />

lower proportions <strong>of</strong> Se are removed in the milling process.<br />

Calcium showed an average loss <strong>of</strong> 41% in concentration terms,<br />

which is in agreement with the results obtained in earlier work on<br />

durum wheat (Toepfer et al., 1972). Studies on the distribution <strong>of</strong><br />

minerals within the kernel <strong>of</strong> common wheat reported that the<br />

endosperm contains about 50% <strong>of</strong> calcium whereas about 25% is<br />

found in the aleurone layer (Pomeranz, 1988). For other elements,<br />

such as magnesium and zinc, only a minor proportion is found in<br />

the endosperm, whereas 70% and w50%, respectively, is in the<br />

aleurone (Bock, 2000; Pieczonka and Rosopulo, 1985; Pomeranz,<br />

1988). Accordingly, calcium was found to follow an entirely<br />

different pattern from that <strong>of</strong> magnesium, zinc and iron in industrially<br />

milled wheat flours (Lorenz et al., 1980). The level <strong>of</strong> calcium<br />

depended mainly on the calcium content <strong>of</strong> the wheat, whereas<br />

magnesium, zinc and iron depended more on milling variables<br />

(Lorenz et al., 1980).<br />

Next to calcium, copper showed a concentration decrease <strong>of</strong> 47%<br />

in this study. This is similar to the results obtained in earlier work<br />

on durum wheat (Brondi et al., 1984; Toepfer et al., 1972) and<br />

matches studies showing that about 45% <strong>of</strong> copper is in the endosperm<br />

<strong>of</strong> common wheat kernels (O’Dell et al., 1972; Pieczonka and<br />

Rosopulo, 1985).<br />

The concentration <strong>of</strong> phosphorous and potassium in semolina<br />

was 56% lower than that in parent grains. Phosphorous mainly<br />

occurs in wheat kernels as phytic acid and its salts. Potassium is<br />

associated with phosphorous as it forms, together with magnesium,<br />

a major part <strong>of</strong> the phytates found in the kernels (Bock, 2000).<br />

A sizeable reduction <strong>of</strong> phytic acid in semolina is achieved upon<br />

milling <strong>of</strong> durum wheat grains so that a major proportion <strong>of</strong><br />

phosphorous is present as nucleoprotein, lipid, and inorganic<br />

phosphorous (Pomeranz, 1988). Since phytic acid is a metalchelating<br />

agent which can lower the absorption <strong>of</strong> several essential<br />

metals (including iron, zinc, calcium and magnesium) its removal<br />

into milling by-products is nutritionally beneficial (Bock, 2000;<br />

O’Dell et al., 1972). Recently, low phytic acid mutants <strong>of</strong> wheat have<br />

been isolated with the view <strong>of</strong> improving the nutritional quality <strong>of</strong><br />

wheat by reducing the major storage form <strong>of</strong> phosphorous and<br />

increasing the level <strong>of</strong> inorganic phosphorous, which is more<br />

readily absorbed by humans and other monogastric animals (Guttieri<br />

et al., 2006).<br />

A group <strong>of</strong> three elements showed a major decrease in<br />

concentration following milling, i.e., iron (63%), magnesium and<br />

zinc (66%). These metals are particularly abundant in the aleurone<br />

(O’Dell et al., 1972; Pieczonka and Rosopulo, 1985). Zinc is present


Table 2<br />

Variations <strong>of</strong> potassium, phosphorous, and magnesium concentration in durum wheat as a consequence <strong>of</strong> processing and cooking<br />

Sample On a fresh weight basis On a dry weight basis<br />

Grain Semolina Pasta Cooked pasta Grain Semolina Pasta Cooked pasta<br />

Potassium<br />

1 a<br />

4.37 0.04 1.99 0.02 2.06 0.02 0.22 0.01 4.85 0.04 2.28 0.02 2.30 0.02 0.58 0.01<br />

2 a<br />

4.32 0.04 1.80 0.01 1.87 0.01 0.20 0.01 4.79 0.04 2.06 0.02 2.08 0.02 0.52 0.01<br />

3 a<br />

5.08 0.04 2.07 0.02 2.14 0.02 0.25 0.01 5.62 0.05 2.37 0.02 2.37 0.02 0.64 0.01<br />

% vs prec b<br />

100 43 (41–46) 104 (103–104) 11 (11–12) 100 44 (42–47) 101 (100–101) 26 (25–27)<br />

% vs grain c<br />

100 43 (41–46) 44 (42–47) 5 (5–5) 100 44 (42–47) 44 (42–47) 11 (11–12)<br />

Phosphorous<br />

1 a<br />

2 a<br />

3 a<br />

% vs prec b<br />

% vs grain c<br />

Magnesium<br />

1 a<br />

2 a<br />

3 a<br />

% vs prec b<br />

% vs grain c<br />

4.04 0.07 1.70 0.03 1.76 0.03 0.65 0.01 4.48 0.07 1.95 0.03 1.96 0.03 1.68 0.03<br />

3.69 0.06 1.68 0.03 1.73 0.03 0.61 0.01 4.10 0.07 1.92 0.03 1.92 0.03 1.57 0.03<br />

4.07 0.07 1.69 0.03 1.74 0.03 0.59 0.01 4.50 0.07 1.93 0.03 1.94 0.03 1.52 0.02<br />

100 43 (42–45) 103 (103–103) 36 (34–37) 100 44 (43–47) 100 (100–100) 82 (79–86)<br />

100 43 (42–45) 44 (43–47) 16 (15–17) 100 44 (43–47) 45 (43–47) 37 (34–38)<br />

1.20 0.01 0.39 0.01 0.40 0.01 0.18 0.01 1.33 0.01 0.44 0.01 0.45 0.01 0.45 0.01<br />

1.04 0.01 0.36 0.01 0.38 0.01 0.16 0.01 1.15 0.01 0.42 0.01 0.43 0.01 0.42 0.01<br />

1.20 0.01 0.40 0.01 0.41 0.01 0.18 0.01 1.33 0.01 0.45 0.01 0.46 0.01 0.46 0.01<br />

100 33 (32–35) 104 (104–105) 43 (43–44) 100 34 (33–36) 101 (101–102) 100 (99–101)<br />

100 33 (32–35) 35 (34–37) 15 (15–16) 100 34 (33–36) 35 (34–37) 35 (34–37)<br />

a 1<br />

Concentration in mg g 95% confidence interval <strong>of</strong> the mean.<br />

b<br />

Percent variation versus preceding item: average <strong>of</strong> the three samples (range).<br />

c<br />

Percent variation versus grain: average <strong>of</strong> the three samples (range).<br />

at a relatively high concentration in the germ as well, this part<br />

accounting for over 10% <strong>of</strong> the zinc burden in the caryopsis<br />

compared to, e.g., 5% <strong>of</strong> copper (Pieczonka and Rosopulo, 1985). As<br />

a consequence, a major proportion <strong>of</strong> iron, magnesium and zinc in<br />

the grain is removed in the conventional roller milling process.<br />

Overall, the retention <strong>of</strong> elements after milling followed the<br />

order Se > Ca > Cu > P z K > Fe > Mg z Zn. The sequence<br />

Ca > Cu > (P, K, Fe, Mg, Zn) was obtained in previous investigations<br />

on durum wheat by Toepfer et al. (1972) (in this case with an<br />

inversion between phosphorous and copper) and Brondi et al.<br />

Table 3<br />

Variations <strong>of</strong> calcium, iron, zinc, and copper concentration in durum wheat as a consequence <strong>of</strong> processing and cooking<br />

(1984). The same order results from studies on common wheat<br />

(Lorenz et al., 1980; Pomeranz and Dikeman, 1983; Toepfer et al.,<br />

1972), even though some investigations found a higher proportion<br />

<strong>of</strong> either potassium (Brondi et al., 1984; Lyons et al., 2005; Zhang<br />

et al., 1997) oriron(Brondi et al., 1984; Brüggemann and Kumpulainen,<br />

1995; Rao and Deosthale, 1981; Zhang et al., 1997) in flour. In<br />

most studies the percentage retentions <strong>of</strong> copper and phosphorous<br />

in flour were found to be similar (Pomeranz and Dikeman, 1983;<br />

Rao and Deosthale, 1981; Toepfer et al., 1972; Zhang et al., 1997),<br />

although generally slightly higher for copper. Deviations from the<br />

Sample On a fresh weight basis On a dry weight basis<br />

Grain Semolina Pasta Cooked pasta Grain Semolina Pasta Cooked pasta<br />

Calcium<br />

1 a<br />

373 18 210 10 228 11 169 8 413 20 241 12 254 12 436 21<br />

2 a<br />

412 20 224 11 248 12 193 9 456 22 256 12 276 13 496 24<br />

3 a<br />

346 17 206 10 229 11 180 9 383 18 235 11 254 12 462 22<br />

% vs prec b<br />

100 57 (54–60) 110 (108–111) 77 (74–79) 100 59 (56–61) 107 (105–108) 178 (172–182)<br />

% vs grain c<br />

100 57 (54–60) 63 (60–66) 48 (45–52) 100 59 (56–61) 63 (61–66) 112 (105–121)<br />

Iron<br />

1 a<br />

2 a<br />

3 a<br />

% vs prec b<br />

% vs grain c<br />

Zinc<br />

1 a<br />

2 a<br />

3 a<br />

% vs prec b<br />

% vs grain c<br />

Copper<br />

1 a<br />

2 a<br />

3 a<br />

% vs prec b<br />

% vs grain c<br />

42.0 0.7 14.7 0.2 15.7 0.3 6.3 0.1 46.6 0.7 16.9 0.3 17.5 0.3 16.3 0.3<br />

33.7 0.5 12.3 0.2 13.4 0.2 5.1 0.1 37.4 0.6 14.1 0.2 14.9 0.2 13.0 0.2<br />

37.0 0.6 13.2 0.2 14.5 0.2 5.4 0.1 41.0 0.7 15.1 0.2 16.1 0.3 13.9 0.2<br />

100 36 (35–36) 109 (107–110) 38 (37–40) 100 37 (36–38) 106 (104–107) 89 (86–93)<br />

100 36 (35–36) 39 (37–40) 15 (15–15) 100 37 (36–38) 39 (38–40) 35 (34–35)<br />

39.5 1.3 13.3 0.4 13.7 0.4 5.2 0.2 43.8 1.4 15.3 0.5 15.3 0.5 13.3 0.4<br />

33.6 1.1 10.3 0.3 10.7 0.3 4.5 0.1 37.3 1.2 11.8 0.4 11.9 0.4 11.6 0.4<br />

29.8 1.0 10.3 0.3 10.8 0.3 4.4 0.1 33.0 1.1 11.8 0.4 12.0 0.4 11.3 0.4<br />

100 33 (31–35) 104 (103–104) 40 (38–42) 100 34 (32–36) 101 (100–101) 93 (87–97)<br />

100 33 (31–35) 34 (32–36) 14 (13–15) 100 34 (32–36) 34 (32–36) 32 (30–34)<br />

5.54 0.13 2.92 0.07 3.06 0.07 1.13 0.03 6.15 0.15 3.35 0.08 3.41 0.08 2.92 0.07<br />

5.30 0.13 2.62 0.06 2.71 0.07 0.98 0.02 5.88 0.14 3.00 0.07 3.02 0.07 2.52 0.06<br />

5.47 0.13 2.87 0.07 3.09 0.07 1.08 0.03 6.06 0.15 3.28 0.08 3.43 0.08 2.76 0.07<br />

100 52 (49–53) 105 (104–107) 36 (35–37) 100 53 (51–55) 102 (101–104) 83 (80–86)<br />

100 52 (49–53) 54 (51–56) 20 (18–20) 100 53 (51–55) 54 (51–57) 45 (43–47)<br />

a 1<br />

Concentration in mg g 95% confidence interval <strong>of</strong> the mean.<br />

b<br />

Percent variation versus preceding item: average <strong>of</strong> the three samples (range).<br />

c<br />

Percent variation versus grain: average <strong>of</strong> the three samples (range).<br />

F. Cubadda et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 92–97 95


96<br />

Table 4<br />

Variations <strong>of</strong> selenium concentration in durum wheat as a consequence <strong>of</strong> processing and cooking<br />

Sample On a fresh weight basis On a dry weight basis<br />

Grain Semolina Pasta Cooked pasta Grain Semolina Pasta Cooked pasta<br />

1 a<br />

85.8 1.4 72.8 1.2 74.5 1.2 31.2 0.5 95.2 1.5 83.5 1.3 83.1 1.3 80.3 1.3<br />

2 a<br />

99.1 1.6 81.2 1.3 83.5 1.3 35.8 0.6 109.8 1.8 93.0 1.5 93.0 1.5 92.0 1.5<br />

3 a<br />

110.6 1.8 85.2 1.4 87.5 1.4 36.2 0.6 122.5 2.0 97.3 1.6 97.2 1.6 92.8 1.5<br />

% vs prec b<br />

100 81 (77–85) 103 (102–103) 42 (41–43) 100 84 (79–88) 100 (99–100) 97 (95–99)<br />

% vs grain c<br />

100 81 (77–85) 83 (79–87) 35 (33–36) 100 84 (79–88) 84 (79–87) 81 (76–84)<br />

a 1<br />

Concentration in ng g 95% confidence interval <strong>of</strong> the mean.<br />

b<br />

Percent variation versus preceding item: average <strong>of</strong> the three samples (range).<br />

c<br />

Percent variation versus grain: average <strong>of</strong> the three samples (range).<br />

common pattern occasionally observed for potassium, iron, and<br />

phosphorous may be due to varying milling conditions among<br />

studies, primarily extraction rate (Rao and Deosthale, 1981), and to<br />

an uneven release <strong>of</strong> some elements from milling equipment,<br />

especially iron (Cubadda et al., 2005). Furthermore, the proportion<br />

<strong>of</strong> each element in the morphological sections <strong>of</strong> the wheat kernel<br />

is dependent on the genotype and varies among cultivars (Lyons<br />

et al., 2005), and this can further explain the slightly different<br />

patterns from one study to another.<br />

3.2. Effect <strong>of</strong> pasta making<br />

Overall, pasta making had little effect on element concentrations<br />

in parent semolina. Slight, non-statistically significant<br />

enrichments <strong>of</strong> calcium (107%), iron (106%) and copper (102%) were<br />

observed. The concentrations <strong>of</strong> these elements in the water used<br />

for dough preparation were 52 1, 0.10 0.01, and<br />

0.041 0.02 mg L 1 , respectively, and explained the major part <strong>of</strong><br />

the calcium increase, a low proportion <strong>of</strong> that <strong>of</strong> copper, and 1% <strong>of</strong><br />

that <strong>of</strong> iron. Therefore, the release from pieces <strong>of</strong> equipment used in<br />

the pasta making process appeared to be the cause <strong>of</strong> the<br />

concentration increase <strong>of</strong> iron and, to a lesser extent, <strong>of</strong> copper.<br />

The elemental concentrations found in the commercial pasta<br />

samples are summarized in Table 5. These concentrations closely<br />

match those <strong>of</strong> the experimental samples. The only differences<br />

were a slightly higher magnesium concentration (p ¼ 0.014) and an<br />

almost double concentration <strong>of</strong> selenium (p < 0.001). Lower<br />

concentrations <strong>of</strong> selenium in the experimental samples compared<br />

to commercial samples are the consequence <strong>of</strong> the relatively low<br />

levels <strong>of</strong> plant-available selenium in the areas <strong>of</strong> Southern Italy<br />

from which experimental grain samples originated (Spadoni et al.,<br />

2007). The widespread use <strong>of</strong> wheat imported from selenium-rich<br />

areas in industrial semolina production results in a higher level <strong>of</strong><br />

selenium in commercial pasta.<br />

It was investigated as to whether differences in elemental levels<br />

between long pasta (spaghetti and related types) and short pasta<br />

(macaroni and related types) existed in commercial samples.<br />

Differences, indeed, turned out to be negligible, even though<br />

a slightly significantly higher content <strong>of</strong> copper and zinc in long<br />

pasta was detected (0.01 < p < 0.05).<br />

Table 5<br />

Element concentrations in commercial pasta samples (N ¼ 12) a<br />

Ca<br />

(mg g 1 Cu<br />

) (mg g 1 Fe<br />

) (mg g 1 K<br />

) (mg g 1 Mg<br />

) (mg g 1 P<br />

) (mg g 1 Se<br />

) (ng g 1 Zn<br />

) (mg g 1 )<br />

Median 266 3.30 12.9 2.07 430 1.66 141 12.7<br />

Mean 260 3.21 13.0 2.07 430 1.67 154 12.9<br />

Min 200 2.76 11.3 1.87 398 1.51 84 10.3<br />

Max 314 3.56 14.7 2.29 457 1.82 227 15.4<br />

CV(%) 11 9 9 6 5 6 34 14<br />

a<br />

Fresh weight basis (average water content for conversion to dry wt basis is<br />

11.0%).<br />

F. Cubadda et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 92–97<br />

Perhaps <strong>of</strong> more importance is the variability <strong>of</strong> concentrations<br />

in commercial pasta, which differed widely among minerals.<br />

Potassium, magnesium and phosphorous exhibited the lowest CV<br />

(5–6%), followed by copper and iron (9%), calcium (11%), zinc (14%)<br />

and selenium (34%). The large spread <strong>of</strong> selenium levels is not<br />

surprising since this element is not essential for plants and its<br />

concentration in grains and derived cereal products correlates with<br />

plant-available selenium in the soil (Spadoni et al., 2007).<br />

The mineral concentrations in pasta in this study are generally<br />

similar to those found in recent surveys elsewhere (USDA, 2007),<br />

even though more confidence can be attributed to the data presented<br />

here due to a more representative sampling. However, in<br />

earlier studies, substantially lower levels <strong>of</strong> calcium (Albrecht et al.,<br />

1987; Ranhotra et al., 1984, 1985; Yaseen, 1993), copper (Albrecht<br />

et al., 1987), potassium (Yaseen, 1993), magnesium (Yaseen, 1993),<br />

and phosphorous (Ranhotra et al., 1984; Yaseen, 1993) were reported.<br />

It is unclear whether the lower concentrations measured in<br />

these earlier studies are the result <strong>of</strong> analytical shortcomings or<br />

reflect actual differences due to, e.g., the introduction <strong>of</strong> new cultivars,<br />

modifications in agronomic practices, or changes in processing.<br />

3.3. Effect <strong>of</strong> cooking<br />

The effect <strong>of</strong> cooking can be best assessed considering data<br />

expressed on a dry weight basis in Tables 2–4. Cooking markedly<br />

increased calcium concentration <strong>of</strong> pasta (p < 0.001), which can be<br />

ascribed to the calcium content <strong>of</strong> the cooking water (26 1mg<br />

L 1 ). Except for potassium, the concentrations <strong>of</strong> the other<br />

elements remained unchanged (magnesium) or showed a slight<br />

reduction in the range 79–97%, which turned out to be significant<br />

only for copper (p < 0.001). Potassium showed a sizeable decrease<br />

(p < 0.001), its concentration after cooking being reduced to about<br />

1/4 <strong>of</strong> that in dry pasta.<br />

The results obtained in this study can not be easily compared to<br />

those <strong>of</strong> previous investigations. When expressed on a dry weight<br />

basis, the retention values calculated in this study correspond to the<br />

‘apparent retention’ <strong>of</strong> nutrients in cooked foods as defined by<br />

Murphy et al. (1975). Apparent retentions after pasta cooking were<br />

determined in an earlier study, where slightly lower values were<br />

found for all minerals (70–82%), including calcium, whereas<br />

potassium again displayed the highest loss (Yaseen, 1993).<br />

However, in this latter study, pasta was cooked in distilled deionized<br />

water without addition <strong>of</strong> salt, which does not allow easy<br />

comparison with the results <strong>of</strong> the present study. A similar<br />

approach was used by Ranhotra et al. (1984), but in this case no<br />

distilled water was used and calcium turned out to have the highest<br />

retention. Other studies (Albrecht et al., 1987; Ranhotra et al., 1985)<br />

took into account the loss <strong>of</strong> solids during cooking and calculated<br />

the so-called ‘true retention’ (Murphy et al., 1975). Albrecht et al.<br />

(1987) compared the effect <strong>of</strong> salt addition when pasta is cooked<br />

either in distilled or tap water. Calcium true retention increased<br />

from 89% to about 100% when salt was added to distilled water.


Calcium retention in unsalted tap water was about 129% (average <strong>of</strong><br />

macaroni and spaghetti) and increased to 158% upon salt addition,<br />

with negligible differences following subsequent rinsing. In<br />

agreement with the other studies, potassium was found to have the<br />

lowest retention and salt addition decreased it, which probably<br />

explains the low retention value for this element obtained in the<br />

present study.<br />

Studies on different cooking approaches may give insight into<br />

the effect <strong>of</strong> each specific medium on mineral retention and suggest<br />

explanations for different retention patterns. However, in this study<br />

pasta was cooked in the customary way, i.e., in salted tap water, in<br />

order to identify the actual mineral content <strong>of</strong> the final product<br />

when prepared according to common household practice.<br />

Using the retention factors obtained in this study, the amount <strong>of</strong><br />

each mineral provided by a standard serving <strong>of</strong> pasta (80 g <strong>of</strong> the<br />

uncooked product) containing each mineral at the average<br />

concentration found in the commercial samples (Table 5) was<br />

calculated. This amount was compared with the Italian daily recommended<br />

dietary allowance (RDA) for the adult population (SINU,<br />

1996). It resulted that the proportion <strong>of</strong> the RDA provided by a daily<br />

serving <strong>of</strong> pasta is on average 22% for selenium, 18% for copper, 10%–<br />

14% for zinc (for males and females, respectively), 11% for phosphorous,<br />

9%–5% for iron (for males and females, respectively), 4%<br />

for calcium, and 1% for potassium. For magnesium, only a range <strong>of</strong><br />

intakes is established instead <strong>of</strong> a RDA; if a value one-third above<br />

the lower level <strong>of</strong> this range is chosen as a reference, the proportion<br />

<strong>of</strong> such an amount provided by a daily serving <strong>of</strong> pasta turns out to<br />

be 17%. Overall, pasta appears to be a valuable source <strong>of</strong> several<br />

minerals <strong>of</strong> importance in human nutrition and well-being.<br />

4. Conclusions<br />

This study led to a better understanding <strong>of</strong> the effect <strong>of</strong> durum<br />

wheat processing on the levels <strong>of</strong> eight minerals, namely, calcium,<br />

copper, iron, magnesium, phosphorous, potassium, selenium, and<br />

zinc. For selenium, no data on the changes induced by durum wheat<br />

processing were available so far, notwithstanding the ever<br />

increasing awareness <strong>of</strong> the importance <strong>of</strong> this element to human<br />

health. Furthermore, the effect <strong>of</strong> cooking was investigated in order<br />

to determine the retention factors to be used for the estimation <strong>of</strong><br />

the content <strong>of</strong> each element in the final product (as consumed)<br />

when the element concentration in the uncooked product is<br />

known.<br />

Milling was the most important processing step in the production<br />

<strong>of</strong> conventional pasta in regard to the change in content <strong>of</strong><br />

minerals originally present in the durum wheat grains. At least six<br />

groups <strong>of</strong> elements could be distinguished on the basis <strong>of</strong> their<br />

concentration decrease upon milling. Selenium had the highest<br />

retention with concentrations in semolina equal to 77%–85% <strong>of</strong> that<br />

in grain (dry weight basis), followed by calcium (54%–60%), copper<br />

(49%–53%), potassium and phosphorous (42%–47%), iron (36%–<br />

38%), magnesium and zinc (32%–36%).<br />

Pasta making had little effect on element concentrations in<br />

semolina whereas cooking caused negligible to small losses <strong>of</strong><br />

elements, except for calcium and potassium which greatly<br />

increased and decreased their concentration, respectively. Using<br />

the retention factors determined by the cooking experiments and<br />

the average concentrations ascertained in the commercial pasta<br />

samples it was assessed that pasta can provide nutritionally<br />

important amounts <strong>of</strong> several minerals, especially selenium,<br />

copper, magnesium, and zinc.<br />

Acknowledgements<br />

The skilled technical help <strong>of</strong> Mr. L. Bartoli in the milling <strong>of</strong> grains<br />

and in the pasta making is acknowledged.<br />

F. Cubadda et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 92–97 97<br />

References<br />

AACC International, 2000. Approved Methods <strong>of</strong> the American Association <strong>of</strong> <strong>Cereal</strong><br />

Chemists, 10th Ed. The Association, St. Paul, MN, USA.<br />

Albrecht, J.A., Asp, E.H., Buzzard, I.M., 1987. Contents and retentions <strong>of</strong> sodium and<br />

other minerals in pasta cooked in unsalted or salted water. <strong>Cereal</strong> Chemistry 64,<br />

106–109.<br />

Bock, M.A., 2000. Minor constituents <strong>of</strong> cereals. In: Kulp, K., Ponte, J.G. (Eds.),<br />

Handbook <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> and Technology, second ed. Marcel Dekker, New<br />

York, pp. 479–504.<br />

Brondi, M., Ciardi, A.M., Cubadda, R., Dall’Aglio, M., Mastino, G.G., Michetti, I.,<br />

Santaroni, G.P., 1984. Trasferimento di elementi traccia dall’ambiente alla catena<br />

alimentare: livelli in grani e loro prodotti. La Rivista della Società Italiana di<br />

Scienza dell’Alimentazione 13, 27–38.<br />

Brüggemann, J., Kumpulainen, J., 1995. The status <strong>of</strong> trace elements in staple foods.<br />

II. Some effects <strong>of</strong> cereal and potato processing. Zeitschrift für Lebensmittel-<br />

Untersuchung und-Forschung 201, 7–11.<br />

Carcea, M., Aureli, F., Cubadda, F., 2007. Minerals and trace elements in the Italian<br />

wheat and products. Tecnica Molitoria International 58, 129–139.<br />

Choi, M.K., Kim, E.Y., 2007. Evaluation <strong>of</strong> dietary manganese intake in Korean men<br />

and women over 20 years old. The <strong>Journal</strong> <strong>of</strong> the Korean Society <strong>of</strong> Food <strong>Science</strong><br />

and Nutrition 36, 447–452.<br />

Cubadda, F., Baldini, M., Carcea, M., Pasqui, L.A., Raggi, A., Stacchini, P., 2001.<br />

Influence <strong>of</strong> laboratory homogenization procedures on trace element content <strong>of</strong><br />

food samples: an ICP-MS study on s<strong>of</strong>t and durum wheat. Food Additives and<br />

Contaminants 18, 778–787.<br />

Cubadda, F., Raggi, A., Testoni, A., Zanasi, F., 2002. Multielemental analysis <strong>of</strong> food<br />

and agricultural matrixes by inductively coupled plasma-mass spectrometry.<br />

<strong>Journal</strong> <strong>of</strong> AOAC International 85, 113–121.<br />

Cubadda, F., Raggi, A., Marconi, E., 2005. Effects <strong>of</strong> processing on five selected metals<br />

in the durum wheat food chain. Microchemical <strong>Journal</strong> 79, 97–102.<br />

Guttieri, M.J., Peterson, K.M., Souza, E.J., 2006. Mineral distributions in milling<br />

fractions <strong>of</strong> low phytic acid wheat. Crop <strong>Science</strong> 46, 2692–2698.<br />

Hattori, H., Ashida, A., Ito, C., Yoshida, M., 2004. Determination <strong>of</strong> molybdenum in<br />

foods and human milk, and an estimate <strong>of</strong> average molybdenum intake in the<br />

Japanese population. <strong>Journal</strong> <strong>of</strong> Nutritional <strong>Science</strong> & Vitaminology 50, 404–409.<br />

Hussein, L., Brüggemann, J., 1997. Zinc analysis <strong>of</strong> Egyptian foods and estimated<br />

daily intakes among an urban population group. Food Chemistry 58, 391–398.<br />

Lorenz, K., Loewe, R., Weadon, D., Wolf, W., 1980. Natural Levels <strong>of</strong> nutrients in<br />

commercially milled wheat flours. III. Mineral Analysis. <strong>Cereal</strong> Chemistry 57,<br />

65–69.<br />

Lyons, G.H., Genc, Y., Stangoulis, J.C.R., Palmer, L.T., Graham, R.D., 2005. Selenium<br />

distribution in wheat grain, and the effect <strong>of</strong> postharvest processing on wheat<br />

selenium content. Biological Trace Element Research 103, 155–168.<br />

Murphy, E.W., Criner, P.E., Gray, B.C., 1975. Comparisons <strong>of</strong> methods for calculating<br />

retentions <strong>of</strong> nutrients in cooked foods. <strong>Journal</strong> <strong>of</strong> Agricultural and Food<br />

Chemistry 23, 1153–1157.<br />

O’Dell, B.L., de Boland, A., Koirtyohann, S.R., 1972. Distribution <strong>of</strong> phytate and<br />

nutritionally important elements among morphological components <strong>of</strong> cereal<br />

grains. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 20, 718–721.<br />

Pieczonka, K., Rosopulo, A., 1985. Distribution <strong>of</strong> cadmium, copper, and zinc in the<br />

caryopsis <strong>of</strong> wheat (Triticum aestivum L.). Fresenius’ <strong>Journal</strong> <strong>of</strong> Analytical<br />

Chemistry 322, 697–699.<br />

Pomeranz, Y., 1988. In: Pomeranz, Y. (Ed.), Wheat Chemistry and Technology.<br />

Chemical composition <strong>of</strong> kernel structures, Vol. I. American Association <strong>of</strong><br />

<strong>Cereal</strong> Chemists, St. Paul, MN, USA, pp. 97–158.<br />

Pomeranz, Y., Dikeman, E., 1983. Minerals and protein contents in hard red winter<br />

wheat flours. <strong>Cereal</strong> Chemistry 60, 80–82.<br />

Ranhotra, G.S., Gelroth, J.A., Novak, F.A., Bock, M.A., Winterringer, G.L.,<br />

Matthews, R.H., 1984. Nutritive value <strong>of</strong> selected variety breads and pastas.<br />

<strong>Journal</strong> <strong>of</strong> the American Dietetic Association 84, 322–327.<br />

Ranhotra, G.S., Gelroth, J.A., Novak, F.A., Bock, M.A., Matthews, R.H., 1985. Retention<br />

<strong>of</strong> selected minerals in enriched pasta products during cooking. <strong>Cereal</strong> Chemistry<br />

62, 117–119.<br />

Rao, D.S.S., Deosthale, Y.G., 1981. Mineral and trace element composition <strong>of</strong> wheat<br />

and wheat flours <strong>of</strong> different extraction rates. <strong>Journal</strong> <strong>of</strong> Plant Foods 3, 251–257.<br />

SINU, 1996. Livelli di Assunzione Raccomandati di Energia e Nutrienti per la<br />

Popolazione Italiana (Recommended dietary allowances for energy and nutrients<br />

for the Italian population). Società Italiana di Nutrizione Umana, Rome.<br />

Spadoni, M., Voltaggio, M., Carcea, M., Coni, E., Raggi, A., Cubadda, F., 2007. Bioaccesible<br />

selenium in Italian agricultural soils: comparison <strong>of</strong> the biogeochemical<br />

approach with a regression model based on geochemical and<br />

pedoclimatic variables. The <strong>Science</strong> <strong>of</strong> the Total Environment 376, 160–177.<br />

Toepfer, E.W., Polansky, M.M., Heart, J.F., Slover, H.T., Morris, E.R., Hepburn, F.N.,<br />

Quackenbush, F.W., 1972. Nutrient composition <strong>of</strong> selected wheats and wheat<br />

products. XI. Summary. <strong>Cereal</strong> Chemistry 49, 173–186.<br />

UNIPI, 2007. Consumo di pasta alimentare nei diversi paesi (Consumption <strong>of</strong><br />

alimentary pasta in the different countries). Web site: http://www.unipi-pasta.<br />

it/dati/PDF/pdf%202006/TAB31.pdf Tab. 31.<br />

USDA, 2007. National Nutrient Database for Standard Reference, Release 20, NBD<br />

No. 20420. Web site: http://nal.usda.gov/fnic/foodcomp/cgi-bin/list_nut_edit.pl.<br />

Yaseen, A.A.E., 1993. Effect <strong>of</strong> processing conditions and cooking on retention <strong>of</strong><br />

minerals in macaroni. Die Nahrung 37, 449–455.<br />

Zhang, Z.W., Moon, C.S., Watanabe, T., Shimbo, S., Ikeda, M., 1997. Contents <strong>of</strong><br />

pollutant and nutrient elements in rice and wheat grown on the neighbouring<br />

fields. Biological Trace Element Research 57, 39–50.


Starch granule size distribution <strong>of</strong> hard red winter and hard red spring wheat:<br />

Its effects on mixing and breadmaking quality<br />

Seok-Ho Park, Jeff D. Wilson *, Bradford W. Seabourn<br />

Grain Marketing and Production Research Center, Agricultural Research Service, Department <strong>of</strong> Agriculture (USDA), 1515 College Avenue, Manhattan, KS 66502, USA<br />

article info<br />

Article history:<br />

Received 24 April 2008<br />

Received in revised form 2 July 2008<br />

Accepted 10 July 2008<br />

Keywords:<br />

Starch granule size distribution<br />

HRW<br />

HRS<br />

Mixing property<br />

Breadmaking quality<br />

Crumb grain score<br />

Optimum range <strong>of</strong> B-granules<br />

1. Introduction<br />

abstract<br />

Starch is an important part <strong>of</strong> wheat endosperm, not only<br />

because starch accounts for 65–73% <strong>of</strong> dry flour mass when milling<br />

extraction is


Nomenclature<br />

A-granules larger than 10 mm in diameter<br />

B-granules smaller than 10 mm in diameter<br />

HRS hard red spring<br />

HRW hard red winter<br />

r simple correlation coefficient<br />

the Federal Grain Inspection Service (FGIS) Technical Center,<br />

(Kansas City, MO), Grain Inspection, Packers, and Stockyards<br />

Administration (GIPSA), U.S. Department <strong>of</strong> Agriculture. Detailed<br />

information about the samples has been previously reported<br />

(Maghirang et al., 2006).<br />

2.2. Starch isolation and determination <strong>of</strong> granule size distribution<br />

Starch was isolated by enzymatic digestion using pepsin A<br />

(P7012, Sigma, St. Louis, MO), hemicellulase 90 (90,000 U/g activity,<br />

Amano Enzyme U.S.A., Lombard, IL), and cleaned further by<br />

a detergent mix (5% SDS, 5% Triton X-100, 5% Tween 40, and 5%<br />

Triton X-15) (Bechtel and Wilson, 2000).<br />

The size distribution <strong>of</strong> isolated starch granules was measured<br />

using a single wavelength Beckman Coulter LS 13 320 Particle Size<br />

Analyzer (Beckman Coulter, Miami, FL) with the Universal Liquid<br />

Module for liquid-based measurements. Each starch sample was<br />

slurried with 1.0 mL <strong>of</strong> water and vortexed before analysis. The<br />

standard refractive indices used were 1.31 for water and 1.52 for<br />

starch, which is within the sample concentration range <strong>of</strong> the<br />

instrument’s specifications. Volumes <strong>of</strong> all starch granules were<br />

calculated on the assumption that all granules were spherical in<br />

shape.<br />

2.3. Evaluation <strong>of</strong> wheat and flour properties<br />

The following properties <strong>of</strong> wheat were analyzed: test weight<br />

(lb/bu) by American Association <strong>of</strong> <strong>Cereal</strong> Chemists (AACC, 2000)<br />

Approved Method 55-10; protein content using near-infrared<br />

reflectance (AACC Approved Method 39-25); single kernel hardness<br />

(AACC Approved Method 55-31), weight, and size using the<br />

SKCS 4100 (Perten, Springfield, IL); and ash content (AACC<br />

Approved Method 08-01). Wheat was milled using a Brabender<br />

Quadrumat Sr. experimental mill (AACC Approved Method 26-10A).<br />

Flour protein and ash content were measured using AACC<br />

Approved Method 39-11 and 08-01, respectively. Flour color (L * , a * ,<br />

and b * ) was determined using a colorimeter (CR-300, Minolta,<br />

Osaka, Japan). Mixing characteristics <strong>of</strong> flour were evaluated using<br />

Mixograph (AACC Approved Method 54-40A) and Farinograph<br />

(AACC Approved Method 54-21). A modified optimized straightdough<br />

breadmaking method (AACC Approved Method 10-10B) was<br />

used for evaluation <strong>of</strong> experimental breadmaking properties <strong>of</strong><br />

flours. The detailed baking method and description <strong>of</strong> crumb grain<br />

score were reported previously (Park et al., 2004). All tests were<br />

conducted with at least 2 duplicates.<br />

2.4. Statistical analysis<br />

A complete randomized experimental design was used. The<br />

difference between starch granule size and volume distributions <strong>of</strong><br />

HRW and HRS was analyzed using the General Linear Models<br />

procedure <strong>of</strong> the Statistical Analysis System (SAS Institute, Cary,<br />

NC). Statistical abbreviations were simple correlation coefficient (r),<br />

coefficient <strong>of</strong> determinant (R 2 ), P < 0.01 (*), P < 0.001 (**),<br />

P < 0.0001 (***), and standard error <strong>of</strong> the mean (SEM).<br />

S.-H. Park et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 98–105 99<br />

*<br />

coefficient <strong>of</strong> determinant<br />

P < 0.01<br />

** P < 0.001<br />

*** P < 0.0001<br />

SEM standard error <strong>of</strong> the mean<br />

SKCS single kernel characterization system<br />

R 2<br />

3. Results and discussion<br />

3.1. Starch granule size distributions <strong>of</strong> HRW and HRS<br />

Fig. 1 shows a typical bimodal size distribution <strong>of</strong> wheat starch<br />

granules, plotting proportions by volume % <strong>of</strong> granules in equal<br />

diameter intervals against diameter. Point 1 and 3 represent the<br />

most frequent differential volume (%) <strong>of</strong> the small and large granules,<br />

and Point 2 is differential volume (%) between small and large<br />

granules that sets apart small and large starch granules. The<br />

respective integrating areas (%) (A, B, and C) represent proportion <strong>of</strong><br />

volume distribution (%) split by each Point.<br />

Starch granule size distributions <strong>of</strong> HRW (n ¼ 98) and HRS<br />

(n ¼ 99) showed significant differences in mean values <strong>of</strong> different<br />

aspects <strong>of</strong> granule size distribution including differential volume,<br />

diameter, and volume distribution (Table 1). Table 1 shows that<br />

HRW has less B-granules as indicated by lower differential volume<br />

compared to HRS wheat (2.25% vs. 2.80%, respectively). Differential<br />

volumes <strong>of</strong> HRW and HRS wheats at Point 2, however, were not<br />

significantly different. At Point 3, differential volume for HRW<br />

wheat was higher (6.55%) than that <strong>of</strong> HRS (6.03%.), and is due to<br />

the fact that differential volume (%) represents proportionality.<br />

HRW and HRS wheats have similar starch granule size distribution<br />

range from less than 1 mm to about 40 mm, thus, a lower % in one<br />

proportion results in higher % proportion elsewhere. Fig. 2 shows<br />

typical starch granule size distribution curves <strong>of</strong> HRW and HRS<br />

wheats, demonstrating that HRS wheat had higher and lower<br />

differential volumes at Point 1 and 3, respectively, compared with<br />

HRW wheat. It should be noted that even though these curves<br />

(differential volume) were obtained from calculations using real<br />

numbers <strong>of</strong> starch granules at a specific size, the differential volume<br />

represents the proportion <strong>of</strong> starch granule size distribution in the<br />

kernel and not the actual volume.<br />

The mean values <strong>of</strong> starch granule diameter at each point,<br />

compared to differential volume, represent actual size. HRW had<br />

significantly smaller size <strong>of</strong> B-granules (4.32 mm) compared with<br />

Differential volume (%)<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

Point 1 Point 2<br />

0<br />

0 1 2 6 16 40<br />

Size (diameter, µm)<br />

Point 3<br />

Fig. 1. Starch granule size distribution indicating high and low points (Point 1, 2, and 3)<br />

<strong>of</strong> differential volume and their volume percents (Area A, B, C).


100<br />

Table 1<br />

Starch granule size distribution in differential volume (%), diameter (mm), and volume percent distribution (%) <strong>of</strong> HRW and HRS a<br />

HRW HRS<br />

Size Distribution b<br />

Mean Minimum Maximum SEM c<br />

Mean Minimum Maximum SEM c<br />

Differential volume (%)<br />

Point 1 2.25*** 1.30 3.15 0.025 2.80*** 1.91 4.12 0.029<br />

Point 2 1.53 0.91 1.96 0.015 1.56 0.73 2.14 0.022<br />

Point 3 6.55*** 5.17 7.91 0.029 6.03*** 4.74 7.20 0.034<br />

Diameter (mm)<br />

Point 1 4.32*** 3.52 5.11 0.030 4.49*** 3.86 5.11 0.029<br />

Point 2 8.72*** 5.61 10.78 0.063 9.46*** 8.15 10.78 0.043<br />

Point 3 21.49 18.86 24.95 0.087 21.46 18.86 24.95 0.088<br />

Volume distribution (%)<br />

A 23.53*** 14.94 30.87 0.245 29.35*** 20.06 38.24 0.254<br />

A þ B 38.01*** 12.57 49.07 0.402 46.88*** 34.03 57.81 0.308<br />

A þ BþC 75.78*** 57.59 89.00 0.300 80.05*** 73.34 87.30 0.180<br />

a<br />

*** ¼ Mean values <strong>of</strong> HRW and HRS in the same row are significantly different at P ¼ 0.0001.<br />

b<br />

Explanation <strong>of</strong> specific points and area are given in Fig. 1.<br />

c<br />

SEM ¼ standard error mean.<br />

HRS (4.49 mm) at Point 1 (Table 1). The observation that HRW<br />

wheat had a smaller size <strong>of</strong> B-granules compared with HRS wheat<br />

was confirmed by the fact that the threshold diameters at Point 2<br />

were 8.72 mm for HRW wheat and 9.46 mm for HRS wheat. The<br />

mean diameter <strong>of</strong> A-granules, however, was not significantly<br />

different at Point 3 (21.49 and 21.46 mm, respectively).<br />

The HRW and HRS wheats showed a wide range <strong>of</strong> volume<br />

distributions in terms <strong>of</strong> A- and B-granules. The volume distribution<br />

<strong>of</strong> HRW wheat B-granules ranged from 12.57% to 49.07%, while<br />

HRS wheat B-granules ranged from 34.03% to 57.81% (Table 1,<br />

Fig. 1). The mean volume distributions <strong>of</strong> HRW wheat B-granules<br />

were significantly lower (23.53% and 38.01% for area A and A þ B,<br />

respectively) than that <strong>of</strong> HRS (29.35% and 46.88% for area A and<br />

A þ B, respectively). It is obvious that HRW wheat had smaller<br />

B-granules in size and number.<br />

The volume distributions <strong>of</strong> subdivided ranges <strong>of</strong> starch granule<br />

size distribution are shown in Table 2. This data confirms that HRW<br />

wheat has significantly smaller proportions <strong>of</strong> B-granules (less than<br />

10 mm) compared with HRS wheat in our study.<br />

Specific volume ranges for A- and B-granules derived from HRW<br />

and HRS wheats have not been reported in the literature, and most<br />

Differential volume (%)<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

HRS<br />

HRW<br />

0<br />

0 1 2 6 16 40<br />

Size (diameter, µm)<br />

Fig. 2. Starch granule size distribution <strong>of</strong> HRW (GIPSA identification number:<br />

03027680) and HRS (GIPSA identification number: 03036243).<br />

S.-H. Park et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 98–105<br />

previous reports on the proportion <strong>of</strong> B-granules in total starch<br />

were on a weight basis (Brocklehurst and Evers, 1977; D’Appolonia<br />

and Gilles, 1971; Dengate and Meredith, 1984; Evers, 1973; Evers<br />

and Lindley, 1977; Hughes and Briarty, 1976; Meredith, 1981; Park<br />

et al., 2004; Soulaka and Morrison, 1985a). Our data, however, can<br />

still be compared with previous reports because the difference in<br />

density <strong>of</strong> A- and B-starch granules is negligible (Dengate et al.,<br />

1979; Meredith, 1981). Our volume data on the range <strong>of</strong> B-granules<br />

was higher than most previous reports. D’Appolonia and Gilles<br />

(1971) reported less than 10% B-granules by weight in total starch<br />

from 12 HRS flours, whereas others reported 30% B-granules (Evers,<br />

1973; Evers and Lindley, 1977; Hughes and Briarty, 1976). Park et al.<br />

(2004) observed 15.7–27.0% B-granules from 12 HRW wheat flours,<br />

and Soulaka and Morrison (1985a) found 13–35%. Stoddard (1999)<br />

reported a similar range <strong>of</strong> B-granule volume (%), showing 17–50%<br />

<strong>of</strong> B-granules using laser diffraction sizing methods from 130<br />

Australian hexaploid wheat cultivars and Landraces from Asia.<br />

Bechtel et al. (1990) found 48% <strong>of</strong> B- and C-type granules (less than<br />

15.9 mm) from mature HRW wheat, but this discrimination was<br />

larger than the commonly used 10-mm range. These different<br />

observations on the proportion <strong>of</strong> B-granules in total starch were<br />

most likely due to different cultivars, growing conditions, and<br />

starch extraction and analysis methods.<br />

The smaller size and number <strong>of</strong> B-granules in HRW wheat<br />

compared with HRS wheat could be a characteristic <strong>of</strong> wheat class,<br />

or it could be caused by higher temperature during the grain-filing<br />

period. Dengate and Meredith (1984) found that drought affected<br />

granule size distribution (weight %), decreasing B-granules<br />

(4–10 mm) and increasing smaller A-granules (10–20 mm). The<br />

volume % <strong>of</strong> B-granules decreased when growth temperature was<br />

increased from 15 Cto40 C during grain-filling period (Shi et al.,<br />

1994). High temperature also seemed to be associated with<br />

a reduction in the number <strong>of</strong> B-granules (Bhullar and Jenner, 1985).<br />

Barley, which also has a bimodal starch granule size distribution,<br />

showed a similar low ratio in number <strong>of</strong> B-granules induced by high<br />

temperature (MacLeod and Duffus, 1988; Tester et al., 1991). These<br />

environmental effects could cause a difference in the grain-filling<br />

pattern between A- and B-granules. The A-granule starts to form in<br />

the amyloplast at about 4–5 days after anthesis (Bechtel et al., 1990;<br />

Parker, 1985), and continues to increase in size until reaching<br />

a maximum diameter <strong>of</strong> 25–50 mm at physiological maturity<br />

(Bechtel et al., 1990; Dengate and Meredith, 1984; Simmonds and<br />

O’Brien, 1981). The final number <strong>of</strong> A-amyloplasts, however, is<br />

achieved much earlier at approximately seven days post-anthesis<br />

when cell division ceases (Briarty et al., 1979). On the other hand,


Table 2<br />

Volume percent distribution (%) on different ranges <strong>of</strong> granule size <strong>of</strong> HRW and HRS a<br />

HRW HRS<br />

Range (mm) Mean Minimum Maximum SEM b<br />

Mean Minimum Maximum SEM b<br />

30 9.6*** 2.2 22.9 0.24 6.5*** 3.3 13.3 0.13<br />


102<br />

Table 3<br />

Correlations between starch granule size distribution <strong>of</strong> HRW and HRS and wheat and flour properties<br />

Differential volume (%) Diameter (mm) Area range (%)<br />

Point 1 Point 2 Point 3 Point 1 Point 2 Point 3 A A þ B AþBþC HRW<br />

Wheat protein (%) 0.39 *** 0.40 *** 0.49 *** 0.47 ***<br />

Test weight (bu/lb) 0.36 ** 0.32 *<br />

Kernel size distribution (%)<br />

Large kernel 0.56 *** 0.32 * 0.40 *** 0.33 ** 0.61 *** 0.28 * 0.40 *** 0.41 ***<br />

Medium kernel 0.56 *** 0.33 ** 0.39 *** 0.31 * 0.60 *** 0.28 * 0.39 *** 0.41 ***<br />

Small kernel 0.28 * 0.28 * 0.36 ** 0.27 *<br />

Single kernel character<br />

Weight (mg) 0.40 *** 0.38 *** 0.48 *** 0.34 ** 0.29 *<br />

Size (mm) 0.49 *** 0.45 *** 0.30 * 0.53 *** 0.30 * 0.33 ** 0.36 **<br />

Hardness 0.29 * 0.30 * 0.31 *<br />

Flour protein (%) 0.35 ** 0.41 *** 0.44 *** 0.45 ***<br />

Flour color: L 0.32 * 0.41 *** 0.32 *<br />

HRS<br />

Wheat protein (%) 0.72 *** 0.48 *** 0.62 *** 0.52 *** 0.67 *** 0.68 *** 0.66 *** 0.30 *<br />

Test weight (bu/lb) 0.72 *** 0.41 *** 0.67 *** 0.45 *** 0.64 *** 0.70 *** 0.71 *** 0.26 *<br />

Kernel size distribution (%)<br />

Large kernel 0.68 *** 0.50 *** 0.60 *** 0.53 *** 0.70 *** 0.30 * 0.67 *** 0.63 *** 0.29 *<br />

Medium kernel 0.68 *** 0.50 *** 0.60 *** 0.53 *** 0.70 *** 0.30 * 0.67 *** 0.64 *** 0.29 *<br />

Small kernel 0.29 * 0.28 * 0.29 * 0.28 *<br />

Single kernel character<br />

Weight (mg) 0.59 *** 0.49 *** 0.53 *** 0.56 *** 0.48 *** 0.27 * 0.59 *** 0.45 ***<br />

Size (mm) 0.63 *** 0.44 *** 0.58 *** 0.57 *** 0.54 *** 0.63 *** 0.53 ***<br />

Hardness 0.29 * 0.27 *<br />

Flour protein (%) 0.72 *** 0.51 *** 0.59 *** 0.54 *** 0.70 *** 0.29 * 0.68 *** 0.64 *** 0.28 *<br />

Flour color: L 0.53 *** 0.33 ** 0.50 *** 0.44 *** 0.52 *** 0.54 *** 0.50 *** 0.31 *<br />

*, **, and *** ¼ significant correlations at P < 0.01, P < 0.001, and P < 0.0001, respectively.<br />

at Points 1 and 2 (r ¼ 0.53***, 0.65***, respectively), and area<br />

range A and A þ B(r ¼ 0.64***, 0.59***, respectively) showed<br />

significant inverse correlations with mixing absorption. This<br />

observation is most likely due to mixing absorption being generally<br />

positively correlated to protein content (Ohm and Chung, 1999;<br />

Park et al., 2006). The B-granules were reported to have higher<br />

Table 4<br />

Correlations between starch granule size distribution and mixing properties<br />

S.-H. Park et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 98–105<br />

swelling power (intragranular plus entrained intergranular water),<br />

which was due to higher water absorption capability, and was<br />

triggered only when temperature was much higher than normal<br />

mixing temperature, above 90 C(Kulp, 1973; Park et al., 2004;<br />

Seib, 1994; Wong and Lelievre, 1982). Haraszi et al. (2004) observed<br />

a decrease in water absorption when protein content decreased<br />

Differential volume (%) Diameter (mm) Area range (%)<br />

Point 1 Point 2 Point 3 Point 1 Point 2 Point 3 A A þ B AþBþC HRW<br />

Mixograph<br />

Absorption<br />

Mix time<br />

0.30 * 0.42 *** 0.40 *** 0.42 ***<br />

Mix tolerance 0.26 *<br />

Farinograph<br />

Absorption 0.35 ** 0.30 * 0.36 **<br />

Development time 0.32 * 0.32 * 0.28 * 0.31 * 0.39 ***<br />

Stability<br />

Tolerance 0.34 **<br />

Breakdown 0.29 * 0.36 ** 0.34 **<br />

HRS<br />

Mixograph<br />

Absorption 0.67 *** 0.47 *** 0.56 *** 0.53 *** 0.65 *** 0.30 * 0.64 *** 0.59 *** 0.27 *<br />

Mix time 0.59 *** 0.29 * 0.53 *** 0.36 ** 0.45 *** 0.58 *** 0.57 *** 0.35 **<br />

Mix tolerance 0.55 *** 0.35 ** 0.47 *** 0.41 *** 0.39 *** 0.55 *** 0.49 *** 0.32 *<br />

Farinograph<br />

Absorption 0.29 * 0.38 *** 0.29 * 0.35 **<br />

Development time 0.59 *** 0.33 ** 0.49 *** 0.37 ** 0.54 *** 0.59 *** 0.61 *** 0.31 *<br />

Stability 0.48 *** 0.38 *** 0.41 *** 0.46 *** 0.48 *** 0.48 *** 0.40 ***<br />

Tolerance 0.33 ** 0.39 *** 0.34 ** 0.27 * 0.29 **<br />

Breakdown 0.62 *** 0.37 ** 0.52 *** 0.43 *** 0.60 *** 0.61 *** 0.61 *** 0.29 *<br />

*, **, and *** ¼ significant correlations at P < 0.01, P < 0.001, and P < 0.0001, respectively.


with starch addition. Also, mixing absorption is highly influenced<br />

by protein subclass, 50% 1-propanol insoluble polymeric protein<br />

content in flour (Park et al., 2006). HRW wheat showed similar<br />

trends, but with rather weak and insignificant relationships.<br />

Mixograph mixing time in HRS wheat showed negative correlations<br />

with B-granules, whereas in HRW wheat, a positive correlation<br />

was observed between mix time and differential volume at<br />

Point 1 (Table 4). Park et al. (2004) found that mix time did not<br />

show significant correlations to protein content with 49 HRW<br />

wheat flours due to contrasting effects among protein subclass on<br />

mix time. However, they suggested that mix time would be shorter<br />

as protein content increases because they also found that 50%<br />

1-propanol insoluble polymeric protein based on total protein was<br />

positively correlated with mix time and tended to decrease with<br />

increasing total protein content. Consequently, considering significant<br />

inverse correlations between protein content and parameters<br />

<strong>of</strong> B-granules, the relationship between mix time and parameters<br />

<strong>of</strong> B-granules was expected to be positive. The positive correlation<br />

(r ¼ 0.33**), however, was obtained only from HRW wheat. The<br />

inverse correlations between HRS wheat and parameters <strong>of</strong><br />

B-granules were probably due in part to a positive correlation<br />

between protein content and mix time (r ¼ 0.44***, data not<br />

shown) in HRS wheat. In HRW wheat, there was no significant<br />

correlation between protein content and mix time. It has also been<br />

reported that B-granules require shorter optimum mix time<br />

compared with A-granules (D’Appolonia and Gilles, 1971; Petr<strong>of</strong>sky<br />

and Hoseney, 1995). Chiotelli and Le Meste (2002) also reported<br />

that B-granules had a higher affinity for water than the A-granules<br />

at room temperature, resulting in faster hydration. Therefore, the<br />

present work suggests that a higher proportion <strong>of</strong> B-granules in<br />

HRS wheat may partially account for the inverse relationship<br />

between mix time and parameters <strong>of</strong> B-granules.<br />

Farinograph absorption showed weak but significant inverse<br />

correlations to parameters <strong>of</strong> B-granules in HRS wheat, whereas in<br />

HRW wheat, diameter at Point 2 showed significant inverse<br />

correlation (Table 4). A similar reason, as given for Mixograph mix<br />

absorption, could explain these relationships. Farinograph<br />

absorption was positively correlated with wheat protein<br />

(r ¼ 0.71***, data not shown), and parameters <strong>of</strong> B-granules were<br />

inversely correlated to protein content, resulting in negative<br />

correlations between Farinograph absorption and parameters <strong>of</strong><br />

B-granules. Soh et al. (2006) observed a significant increase in<br />

Farinograph absorption at constant protein content (17.4–17.7%) as<br />

% B-granules increased from 17% to 32.4%. Therefore, it appears that<br />

starch granule size distribution may affect Farinograph absorption<br />

when protein content and quality are invariable. Our data suggests<br />

that starch granule size distribution may not be an important<br />

variable for mix absorption when protein content varies. The<br />

protein content in the HRS wheat flour varied widely from 10.6% to<br />

17.8% (data not shown), and the protein composition (quality) was<br />

altered as the protein content increased (Park et al., 2006).<br />

The relationship between Farinograph development time and<br />

starch granule size distribution was similar to the relationship<br />

between Mixograph mix time and starch granule size distribution.<br />

The same explanation would be applicable for both relationships.<br />

Farinograph stability in HRS wheat also showed inverse<br />

correlation with parameters <strong>of</strong> B-granules. Several authors have<br />

reported the effects <strong>of</strong> starch on rheological properties. B- and nonwheat<br />

starch granules have been reported to result in large<br />

rheological differences using a constant vital gluten ratio in the<br />

reconstituted dough (Petr<strong>of</strong>sky and Hoseney, 1995). They suggested<br />

that there were strong interactions between vital gluten and nonwheat<br />

starches and wheat B-granules, resulting in less extensibility.<br />

Miller and Hoseney (1999) found that starch isolated from Kansas<br />

grown wheat gave significantly lower elastic modulus (G 0 ) and<br />

viscous modulus (G 00 ) than starches from other Kansas and strong<br />

S.-H. Park et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 98–105 103<br />

Canadian wheats when dough was prepared from the same gluten<br />

source. Thus, it appears that starch from different cultivars and<br />

different size distributions makes a difference in determining the<br />

rheological properties <strong>of</strong> dough when a constant gluten source is<br />

used. The inverse relationships between Farinograph stability and<br />

parameters <strong>of</strong> B-granules, however, are hard to explain from<br />

previous reports (Miller and Hoseney, 1999; Petr<strong>of</strong>sky and Hoseney,<br />

1995) for two reasons. First, the present study used actual HRS<br />

wheat flours, and not blends <strong>of</strong> starches and constant gluten, as did<br />

previous researches, which had various components that could<br />

affect mixing properties. Second, there is limited research to<br />

connect rheological parameters <strong>of</strong> B-granules to Farinograph<br />

stability. It is generally recognized that rheological development in<br />

dough is triggered by the formation <strong>of</strong> a continuous gluten network<br />

(MacRitchie, 1992), with starch granules embedded in the network.<br />

In addition, previous studies reported that breakdown (loss <strong>of</strong><br />

stability) during over-mixing was caused by ‘‘depolymerization’’ <strong>of</strong><br />

the protein network (Danno and Hoseney,1982; Skerritt et al.,1999;<br />

Tanaka and Bushuk, 1973; Weegels et al., 1996). Therefore, with<br />

large variation in flour quality including protein content (10.6–<br />

17.8%, 14% moist base), Farinograph absorption (59.8–73.1%),<br />

development time (5.2–44.5 min), and loaf volume (803–1238 cm 3 )<br />

(data not shown), inverse relationships between Farinograph<br />

stability and parameters <strong>of</strong> B-granules could be explained by the<br />

positive correlations between protein content and Farinograph<br />

stability (r ¼ 0.46***, data not shown). As previously discussed,<br />

protein content was inversely correlated to parameters <strong>of</strong> B-granules,<br />

possibly giving the inverse correlation between Farinograph<br />

stability and parameters <strong>of</strong> B-granules. Farinograph breakdown<br />

showed a similar relationship with stability, and a similar explanation<br />

could be applicable. Protein content and breakdown were<br />

highly positively correlated (r ¼ 0.79***, data not shown). The HRW<br />

wheat showed a similar trend in these relationships, but was<br />

weaker with many non-significant correlations.<br />

3.4. Relationships with breadmaking parameters<br />

The breadmaking parameters showed significant correlations to<br />

starch granule size distribution (Table 5). A greater number <strong>of</strong> the<br />

parameters <strong>of</strong> HRS wheat, again, were significantly correlated. The<br />

trends in relationships were similar for HRW and HRS wheat, so our<br />

discussion will focus on HRS wheat.<br />

Baking absorption and mix time were positively correlated to<br />

Mixograph absorption (r ¼ 0.85***), mix time (r ¼ 0.93***), Farinograph<br />

absorption (r ¼ 0.73***) and development time (r ¼ 0.77***)<br />

(data not shown).<br />

Loaf volume was inversely correlated with parameters <strong>of</strong><br />

B-granules. Differential volume at Point 1 showed the greatest<br />

inverse correlation (r ¼ 0.68***), followed by area range A<br />

(r ¼ 0.66***), A þ B (r ¼ 0.62***) and diameter at Point 1<br />

(r ¼ 0.52***). Park et al. (2005) summarized previous studies<br />

using four different concepts concerning the effects <strong>of</strong> small starch<br />

granule size on breadmaking properties: ‘‘beneficial’’ (Hayman<br />

et al., 1998; Sahlstrom et al., 1998; Van Vliet et al., 1992); ‘‘detrimental’’(D’Appolonia<br />

and Gilles, 1971; Kulp, 1973); ‘‘little effect’’<br />

(Hoseney et al., 1971); and ‘‘optimum ratio’’ <strong>of</strong> A- and B-granules<br />

(Lelievre et al., 1987; Park et al., 2004; Soulaka and Morrison,<br />

1985b). More recently, Park et al. (2005) confirmed their previous<br />

observation that there seemed to be an optimum weight ratio <strong>of</strong> Aand<br />

B-granules for crumb grain score, and there was no benefit <strong>of</strong><br />

small starch granules to loaf volume and crumb grain score when<br />

used with a constant source <strong>of</strong> gluten. Considering previous<br />

controversial results, it is difficult to state that B-granules affect loaf<br />

volume negatively. It is more appropriate to admit that the relationships<br />

were obtained due to a positive correlation between<br />

protein content and loaf volume (r ¼ 0.91***, data not shown), and


104<br />

Table 5<br />

Correlations between starch granule size distribution and straight-dough breadmaking parameters<br />

Differential volume (%) Diameter (mm) Area range (%)<br />

Point 1 Point 2 Point 3 Point 1 Point 2 Point 3 A A þ B AþBþC HRW<br />

Breadmaking<br />

Absorption<br />

Mix time<br />

0.35 ** 0.28 * 0.26 *<br />

Pro<strong>of</strong> height 0.46 *** 0.38 *** 0.31 * 0.45 *** 0.51 *** 0.26 * 0.28 * 0.27 *<br />

Crumb grain score 0.35 **<br />

Loaf volume 0.35 ** 0.43 *** 0.28 * 0.42 *** 0.40 ***<br />

HRS<br />

Breadmaking<br />

Absorption 0.57 *** 0.30 * 0.52 *** 0.40 *** 0.57 *** 0.28 * 0.56 *** 0.58 *** 0.37 **<br />

Mix time 0.52 *** 0.46 *** 0.34 ** 0.43 *** 0.52 *** 0.53 *** 0.36 **<br />

Pro<strong>of</strong> height 0.56 *** 0.34 ** 0.51 *** 0.37 ** 0.47 *** 0.54 *** 0.52 ***<br />

Crumb grain score 0.28 * 0.27 * 0.41 *** 0.27 *<br />

Loaf volume 0.68 *** 0.45 *** 0.60 *** 0.52 *** 0.67 *** 0.28 * 0.66 *** 0.62 *** 0.28 *<br />

*, **, and *** ¼ significant correlations at P < 0.01, P < 0.001, and P < 0.0001, respectively.<br />

All correlations are significant in this table at P ¼ 0.01 (>r ¼j 0.261j), P ¼ 0.001 (>r ¼j 0.326j), and P ¼ 0.0001 (>r ¼j 0.374j).<br />

protein content was inversely correlated with parameters <strong>of</strong><br />

B-granules.<br />

Crumb grain score showed a generally weak but significant<br />

inverse linear correlation to differential volume at Point 1<br />

(r ¼ 0.28*), diameter at Point 1 (r ¼ 0.41***), and area range <strong>of</strong> A<br />

(r ¼ 0.27*). These inverse relationships agreed with results from<br />

Park et al. (2005) where the authors found the lowest value <strong>of</strong><br />

crumb grain score and fineness from the bread baked with 100%<br />

B-granules. In this study, crumb grain score was not significantly<br />

correlated to protein, consequently relationships were independent<br />

<strong>of</strong> protein content. It should be pointed out that there were no<br />

other significant correlations to crumb grain score from 68 wheat<br />

and flour quality parameters, except wheat kernel weight and size<br />

(r ¼ 0.36***, respectively, data not shown), which were positively<br />

correlated to parameters <strong>of</strong> B-granules. So, even though those<br />

correlation values were low, we believe it has significance.<br />

3.5. Polynomial relationships between B-granules<br />

and crumb grain score<br />

Several authors reported that there seemed to be an optimum<br />

weight % ratio <strong>of</strong> A- and B-granules for breadmaking (Lelievre et al.,<br />

1987; Park et al., 2004, 2005; Soulaka and Morrison, 1985b) and<br />

spaghetti production (Soh et al., 2006). A polynomial relationship<br />

was applied between B-granules and crumb grain score, resulting<br />

in improved correlations for both HRW and HRS wheats (Table 6).<br />

Correlation values <strong>of</strong> HRS wheat increased from 0.073* to 0.154***<br />

for area A, and 0.013 to 0.088* for area A þ B. In addition, we found<br />

that the linear relationships improved when the ratio <strong>of</strong> volume %<br />

<strong>of</strong> B-granules to flour protein content was used. The correlation<br />

Table 6<br />

Linear and polynomial correlations <strong>of</strong> B-type granule volume % and ratio with<br />

protein contents to crumb grain score<br />

Crumb grain score vs. Linear<br />

Polynomial<br />

correlation<br />

correlation<br />

R 2<br />

R 2<br />

HRW Area A 0.004 0.018<br />

Area A þ B 0.005 0.039<br />

Area A/flour protein 0.130 ** 0.141 ***<br />

Area A þ B/flour protein 0.136 ** 0.158 ***<br />

HRS Area A 0.073 * 0.154 ***<br />

Area A þ B 0.013 0.088 *<br />

Area A/flour protein 0.089 * 0.222 ***<br />

Area A þ B/flour protein 0.049 0.164 ***<br />

*, **, and *** ¼ significant correlations at P < 0.01, P < 0.001, and P < 0.0001,<br />

respectively.<br />

S.-H. Park et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 98–105<br />

improvements were more obvious for HRW wheat, which<br />

improved from 0.004 to 0.130** for area range A, and from 0.005 to<br />

0.136** for area A þ B. The linear correlations between volume % <strong>of</strong><br />

B-granules and crumb grain score were improved from 0.004 to<br />

0.141*** and from 0.005 to 0.158*** for HRW wheat, and from<br />

0.073* to 0.222*** and from 0.013 to 0.164*** for HRS wheat after<br />

obtaining the ratios with protein content and applying polynomial<br />

correlation. Therefore, it appears that there is an optimum range <strong>of</strong><br />

volume % B-granules, and the optimum range could vary depending<br />

on protein content. Lelievre et al. (1987) found different optimum<br />

starch size fractions for different protein concentrations when used<br />

for breadmaking. Park et al. (2005) suggested that high water<br />

absorption during baking and/or high surface area <strong>of</strong> B-granules<br />

could be responsible for gas cell stabilization and the resultant<br />

crumb grain score. It has been shown that protein–starch interactions<br />

produce different rheological properties depending on variety<br />

and granule size (Miller and Hoseney, 1999; Petr<strong>of</strong>sky and Hoseney,<br />

1995). Also, bulk rheological properties <strong>of</strong> dough could affect<br />

overall gas cell stability (Van Vliet et al., 1992). Consequently, the<br />

reason for different optimum weight % <strong>of</strong> B-starch granules for<br />

different protein contents may be due to rheological properties<br />

imparted by protein–starch interactions. The higher water<br />

absorbing properties <strong>of</strong> B-granules could pull water from the<br />

attached protein matrix and liquid film in the dough system during<br />

baking. Gan et al. (1995) proposed that gas cells are stabilized by<br />

a continuous liquid film on the protein–starch matrix. Depending<br />

on how extreme the situations are, e.g. low protein content with<br />

high content <strong>of</strong> B-granules (too stiff) vs. high protein content with<br />

low content <strong>of</strong> B-granules (too viscous), the overall viscoelastic<br />

properties and gas cell stability could be changed, resulting in<br />

different crumb structures.<br />

4. Conclusion<br />

Starch granule size distributions <strong>of</strong> HRW and HRS wheats<br />

showed significant differences in differential volume, diameter, and<br />

volume distribution. HRW has smaller size and proportion <strong>of</strong><br />

B-granules than HRS wheat. The reason is not clear in this study,<br />

but an explanation could be due to hot and dry growing conditions<br />

during grain filling. Parameters <strong>of</strong> B-granules showed many<br />

significant correlations with wheat and flour properties, partly due<br />

to the inverse correlation between protein content and parameters<br />

<strong>of</strong> B-granules. There appears to be different optimum ranges <strong>of</strong> Bgranule<br />

weight % for flours with different protein contents. It seems<br />

that starch granule size distribution is a unique property that


affects physicochemical properties <strong>of</strong> wheat, flour, and breadmaking<br />

properties in conjunction with its counterpart, protein.<br />

References<br />

AACC, 2000. Approved Methods <strong>of</strong> the American Association <strong>of</strong> <strong>Cereal</strong> Chemists,<br />

10th ed. AACC, St. Paul, MN, USA.<br />

Altenbach, S.B., DuPont, F.M., Kothari, K.M., Chan, R., Johnson, E.L., Lieu, D., 2003.<br />

Temperature, water and fertilizer influence the timing <strong>of</strong> key events during<br />

grain development in a U.S. spring wheat. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 37, 9–20.<br />

Bechtel, D.B., Wilson, J.D., 2000. Variability in a starch isolation method and automated<br />

digital image analysis system used for the study <strong>of</strong> starch size distributions<br />

in wheat flour. <strong>Cereal</strong> Chemistry 77, 401–405.<br />

Bechtel, D.B., Zayas, I., Kaleikau, L., Pomeranz, Y., 1990. Size-distribution <strong>of</strong> wheat<br />

starch granules during endosperm development. <strong>Cereal</strong> Chemistry 67, 59–63.<br />

Bhullar, S.S., Jenner, C.F., 1985. Differential responses to high temperatures <strong>of</strong> starch<br />

and nitrogen accumulation in the grain <strong>of</strong> four cultivars <strong>of</strong> wheat. Australian<br />

<strong>Journal</strong> <strong>of</strong> Plant Physiology 12, 363–375.<br />

Briarty, L.G., Hughes, C.E., Evers, A.D., 1979. The developing endosperm <strong>of</strong> wheat –<br />

a stereological analysis. Annals <strong>of</strong> Botany 44, 641–658.<br />

Brocklehurst, P.A., Evers, A.D., 1977. The size distribution <strong>of</strong> starch granules in<br />

endosperm <strong>of</strong> different sized kernels <strong>of</strong> the wheat cultivar Maris Huntsman.<br />

<strong>Journal</strong> <strong>of</strong> the <strong>Science</strong> <strong>of</strong> Food and Agriculture 28, 1084–1089.<br />

Chiotelli, E., Le Meste, M., 2002. Effect <strong>of</strong> small and large wheat starch granules on<br />

thermomechanical behavior <strong>of</strong> starch. <strong>Cereal</strong> Chemistry 79, 286–293.<br />

Danno, G., Hoseney, R.C., 1982. Effects <strong>of</strong> dough mixing and rheologically active<br />

compounds on relative viscosity <strong>of</strong> wheat proteins. <strong>Cereal</strong> Chemistry 59,196–198.<br />

D’Appolonia, B.L., Gilles, K.A., 1971. Effect <strong>of</strong> various starches in baking. <strong>Cereal</strong><br />

Chemistry 48, 625–636.<br />

Dengate, H., Meredith, P., 1984. Variation in size distribution <strong>of</strong> starch granules from<br />

wheat grain. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 3, 83–90.<br />

Dengate, H.N., Baruch, D.W., McKenzie, J., Simmons, L.D., Meredith, P.W.,<br />

Morrison, R., 1979. Fractionation <strong>of</strong> wheat starch on the basis <strong>of</strong> hydrated<br />

density. Starch 31, 40–44.<br />

Dronzek, B.L., Hwang, P., Bushuk, W., 1972. Scanning electron microscopy <strong>of</strong> starch<br />

from sprouted wheat. <strong>Cereal</strong> Chemistry 49, 232–239.<br />

Eliasson, A.C., 1989. Characterization <strong>of</strong> wheat starch and gluten as related to enduse<br />

properties. In: Pomeranz, Y. (Ed.), Wheat is Unique. AACC International, St.<br />

Paul, MN, pp. 177–192.<br />

Evers, A.D., 1973. The size distribution among starch granules in wheat endosperm.<br />

Starch 25, 303–304.<br />

Evers, A.D., Lindley, J., 1977. The particle-size distribution in wheat endosperm<br />

starch. <strong>Journal</strong> <strong>of</strong> <strong>Science</strong> and Food Agriculture 28, 98–102.<br />

Gan, Z., Ellis, P.R., Sch<strong>of</strong>ield, J.D., 1995. Mini review. Gas cell stabilization and gas<br />

retention in wheat bread dough. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 21, 215–230.<br />

Gibson, L.R., McCluskey, P.J., Tilley, K.A., Paulsen, G.M., 1998. Quality <strong>of</strong> hard red<br />

winter wheat grown under high temperature conditions during maturation and<br />

ripening. <strong>Cereal</strong> Chemistry 75, 421–427.<br />

Gibson, L.R., Paulsen, G.M., 1999. Yield components <strong>of</strong> wheat grown under high<br />

temperature stress during reproductive growth. Crop <strong>Science</strong> 39, 1841–1846.<br />

Haraszi, R., Gras, P.W., Tömösközi, S., Salgó, A., Békés, F., 2004. Application <strong>of</strong> a micro<br />

Z-arm mixer to characterize mixing properties and water absorption <strong>of</strong> wheat<br />

flour. <strong>Cereal</strong> Chemistry 81, 555–560.<br />

Hayman, D., Sipes, K., Hoseney, R.C., Faubion, J.M., 1998. Factors controlling gas cell<br />

failure in bread dough. <strong>Cereal</strong> Chemistry 75, 585–589.<br />

Hoseney, R.C., Finney, K.F., Pomeranz, Y., Shogren, M.D., 1971. Functional (breadmaking)<br />

and biochemical properties <strong>of</strong> wheat flour components. VII. Starch.<br />

<strong>Cereal</strong> Chemistry 48, 191–201.<br />

Hughes, C.E., Briarty, L.G., 1976. Stereological analysis <strong>of</strong> the contribution to mature<br />

wheat endosperm starch by large and small granules. Starch 28, 336–337.<br />

Karlsson, R., Olered, R., Eliasson, A.-C., 1983. Changes in starch granule size distribution<br />

and starch gelatinization properties during development and maturation<br />

<strong>of</strong> wheat, barley and rye. Starch 35, 335–340.<br />

Kulp, K., 1973. Characteristics <strong>of</strong> small-granule starch <strong>of</strong> flour and wheat. <strong>Cereal</strong><br />

Chemistry 50, 666–679.<br />

Lelievre, J., Lorenz, K., Meredith, P., Baruch, D.W., 1987. Effects <strong>of</strong> starch particle size<br />

and protein concentration on breadmaking performance. Starch 39, 347–352.<br />

Lineback, D.R., 1984. The starch granule organization and properties. Bakers Digest<br />

58 (2), 16. 18–21.<br />

MacLeod, L.C., Duffus, C.M., 1988. Temperature effects on starch granules in<br />

developing barley grain. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 8, 29–37.<br />

MacRitchie, F., 1992. Physicochemical properties <strong>of</strong> wheat proteins in relation to<br />

functionality. Advances in Food Research 36, 1–85.<br />

Maghirang, E.B., Lookhart, G.L., Bean, S.R., Pierce, R.O., Xie, F., Caley, M.S.,<br />

Wilson, J.D., Seabourn, B.W., Ram, M.S., Park, S.H., Chung, O.K., Dowell, F.E.,<br />

2006. Comparison <strong>of</strong> quality characteristics and breadmaking functionality <strong>of</strong><br />

hard red winter and hard red spring wheat. <strong>Cereal</strong> Chemistry 83, 520–528.<br />

Meredith, P., 1981. Large and small starch granules in wheat – are they really<br />

different? Starch 33, 40–44.<br />

Miller, K.A., Hoseney, R.C., 1999. Dynamic rheological properties <strong>of</strong> wheat starch–<br />

gluten doughs. <strong>Cereal</strong> Chemistry 76, 105–109.<br />

Morrison, W.R., Gadan, H., 1987. The amylose and lipid contents <strong>of</strong> starch granule in<br />

developing wheat endosperm. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 5, 263–275.<br />

S.-H. Park et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 98–105 105<br />

Morrison, W.R., Scott, D.C., 1986. Measurements <strong>of</strong> the dimensions <strong>of</strong> wheat starch<br />

granule populations using a Coulter Counter with 100-channel analyzer. <strong>Journal</strong><br />

<strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 4, 13–21.<br />

Nikuni, Z., 1978. Studies on starch granules. Starch 30, 105–111.<br />

Ohm, J.B., Chung, O.K., 1999. Gluten, pasting, and mixograph parameters <strong>of</strong> hard<br />

winter wheat flours in relation to breadmaking. <strong>Cereal</strong> Chemistry 76,<br />

606–613.<br />

Ohm, J.B., Chung, O.K., Deyoe, C.W., 1998. Single kernel characteristics <strong>of</strong> hard<br />

winter wheats in relation to milling and baking quality. <strong>Cereal</strong> Chemistry 75,<br />

156–161.<br />

Park, S.H., Chung, O.K., Seib, P.A., 2004. Size distribution and properties <strong>of</strong> wheat<br />

starch granules in relation to crumb grain score <strong>of</strong> pup-loaf bread. <strong>Cereal</strong><br />

Chemistry 81, 699–704.<br />

Park, S.H., Chung, O.K., Seib, P.A., 2005. Effects <strong>of</strong> varying weight ratios <strong>of</strong> large and<br />

small wheat starch granules on experimental straight-dough bread. <strong>Cereal</strong><br />

Chemistry 82, 166–172.<br />

Park, S.H., Chung, O.K., Seib, P.A., 2006. Levels <strong>of</strong> protein and protein composition in<br />

hard winter wheat flours and the relationship to breadmaking. <strong>Cereal</strong> Chemistry<br />

83, 418–423.<br />

Parker, M.L., 1985. The relationship between A-type and B-type starch granules in<br />

the developing endosperm <strong>of</strong> wheat. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 3, 271–278.<br />

Peng, M., Gao, M., Abdel-Aal, E.-S.M., Hucl, P., Chibbar, R.N., 1999. Separation and<br />

characterization <strong>of</strong> A- and B-type starch granules in wheat endosperm. <strong>Cereal</strong><br />

Chemistry 76, 375–379.<br />

Petr<strong>of</strong>sky, K.E., Hoseney, R.C., 1995. Rheological properties <strong>of</strong> dough made with<br />

starch and gluten from several sources. <strong>Cereal</strong> Chemistry 72, 53–58.<br />

Pomeranz, Y., 1988. Chemical composition <strong>of</strong> kernel structures. In: Pomeranz, Y.<br />

(Ed.), Wheat Chemistry and Technology, vol. 1. AACC International, St. Paul, MN,<br />

pp. 97–158.<br />

Raeker, M.O., Gaines, C.S., Finney, P.L., Donelson, T., 1998. Granule size distribution<br />

and chemical composition <strong>of</strong> starches from 12 s<strong>of</strong>t wheat cultivars. <strong>Cereal</strong><br />

Chemistry 75, 721–728.<br />

Sandstedt, R.M., 1961. The function <strong>of</strong> starch in the baking <strong>of</strong> bread. Bakers Digest 35<br />

(3), 36–42. 44.<br />

Sahlstrom, S., Brathen, E., Lea, P., Autio, K., 1998. Influence <strong>of</strong> starch granule size<br />

distribution on bread characteristics. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 28, 157–164.<br />

Seib, P.A., 1994. Wheat starch: isolation, structure, and properties. Oyo Toshitsu<br />

Kagaku 41, 49–69.<br />

Shi, Y.C., Seib, P.A., Bernardin, J.E., 1994. Effects <strong>of</strong> temperature during grain-filling<br />

on starches from six wheat cultivars. <strong>Cereal</strong> Chemistry 71, 369–383.<br />

Skerritt, J.H., Hac, L., Lindsay, M.P., Bekes, F., 1999. Depolymerization <strong>of</strong> the gluten<br />

macropolymer during mixing: II. Differences in retention <strong>of</strong> specific glutenin<br />

subunits. <strong>Cereal</strong> Chemistry 76, 402–409.<br />

Simmonds, D.H., O’Brien, T.P., 1981. Morphological and biochemical development <strong>of</strong><br />

wheat endosperm. Advances in <strong>Cereal</strong> <strong>Science</strong> and Technology 4, 5–14.<br />

Soh, H.N., Sissons, M.J., Turner, M.A., 2006. Effects <strong>of</strong> starch granule size distribution<br />

and elevated amylose content on durum dough rheology and spaghetti cooking<br />

quality. <strong>Cereal</strong> Chemistry 83, 513–519.<br />

Sollars, W.F., Rubenthaler, G.L., 1971. Performance <strong>of</strong> wheat and other starches in<br />

reconstituted flours. <strong>Cereal</strong> Chemistry 48, 397–410.<br />

Soulaka, A.B., Morrison, W.R., 1985a. The amylose and lipid contents, dimensions,<br />

and gelatinization characteristics <strong>of</strong> some wheat starches and their A- and Bgranule<br />

fractions. <strong>Journal</strong> <strong>of</strong> the <strong>Science</strong> <strong>of</strong> Food and Agriculture 36, 709–718.<br />

Soulaka, A.B., Morrison, W.R., 1985b. The bread baking quality <strong>of</strong> six wheat starches<br />

differing in composition and physical properties. <strong>Journal</strong> <strong>of</strong> the <strong>Science</strong> <strong>of</strong> Food<br />

and Agriculture 36, 719–727.<br />

Stoddard, F.L., 1999. Survey <strong>of</strong> starch particle-size distribution in wheat and related<br />

species. <strong>Cereal</strong> Chemistry 76, 145–149.<br />

Stone, P.J., Nicolas, M.E., 1994. Wheat cultivars vary widely in their responses <strong>of</strong><br />

grain yield and quality to short periods <strong>of</strong> post-anthesis heat stress. Australian<br />

<strong>Journal</strong> <strong>of</strong> Plant Physiology 21, 887–900.<br />

Tanaka, K., Bushuk, W., 1973. Changes in flour proteins during dough mixing. II. Gel<br />

filtration and electrophoresis results. <strong>Cereal</strong> Chemistry 50, 597–605.<br />

Tashiro, T., Wardlaw, I.F., 1990. The effects <strong>of</strong> high temperature at different stages <strong>of</strong><br />

ripening on grain set, grain weight, and grain dimension in the semi-dwarf<br />

wheat ‘Banks’. Annals <strong>of</strong> Botany 65, 51–61.<br />

Tester, R.F., South, J.B., Morrison, W.R., Ellis, R.P., 1991. The effects <strong>of</strong> ambient<br />

temperature during the grain-filling period on the composition and properties<br />

<strong>of</strong> starch from four barley genotypes. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 13, 113–127.<br />

U.S. Wheat Associate, 2002. Harvest Report – October 8, 2002 – final. http://www.<br />

uswheat.org/harvestReports/doc/3DCEBA32BAAE4ACF85256F8100572BE1?Open<br />

Document# (accessed April 24, 2008).<br />

Van Vliet, T., Janssen, A.M., Bloksma, A.H., Walstra, P., 1992. Strain hardening <strong>of</strong><br />

dough as a requirement for gas retention. <strong>Journal</strong> <strong>of</strong> Texture Studies 23,<br />

439–460.<br />

Weegels, P.L., van de Pijpekamp, A.M., Graveland, A., Hamer, R.J., Sch<strong>of</strong>ield, J.D.,<br />

1996. Depolymerisation and repolymerisation <strong>of</strong> wheat glutenin during dough<br />

processing. I. Relationships between glutenin macropolymer content and<br />

quality parameters. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 23, 103–111.<br />

Wilson, J.D., Bechtel, T.C., Seib, P.A., 2006. Measurement <strong>of</strong> wheat starch granule size<br />

distribution using image analysis and laser diffraction technology. <strong>Cereal</strong><br />

Chemistry 83, 259–268.<br />

Wong, R.B.K., Lelievre, J., 1982. Comparison <strong>of</strong> the crystallinities <strong>of</strong> wheat starches<br />

with different swelling capacities. Starch 34, 159–161.


Total phenolics, flavonoids, antioxidant capacity in rice grain and<br />

their relations to grain color, size and weight<br />

Yun Shen, Liang Jin, Peng Xiao, Yan Lu, Jinsong Bao *<br />

Institute <strong>of</strong> Nuclear Agricultural <strong>Science</strong>s, Key Laboratory <strong>of</strong> Chinese Ministry <strong>of</strong> Agriculture and Zhejiang Province for Nuclear-Agricultural <strong>Science</strong>s, College <strong>of</strong> Agriculture and<br />

Biotechnology, Zhejiang University, Hua Jiachi Campus, Hangzhou 310029, People’s Republic <strong>of</strong> China<br />

article info<br />

Article history:<br />

Received 2 March 2008<br />

Received in revised form 14 June 2008<br />

Accepted 7 July 2008<br />

Keywords:<br />

Antioxidant capacity<br />

Flavonoid<br />

Phenolics<br />

Rice<br />

1. Introduction<br />

abstract<br />

Rice is a staple food being consumed by nearly half <strong>of</strong> the world<br />

population. Nutritional quality <strong>of</strong> rice has received more attention<br />

in the developing countries, where monotonous consumption <strong>of</strong><br />

rice may lead to deficiencies <strong>of</strong> essential minerals, vitamins, and<br />

other nutritional compositions (Bouis et al., 2003). This is not<br />

caused by nutritional deficiency in rice grain itself, but due to it<br />

being traditionally eaten in the form <strong>of</strong> the milled white kernel.<br />

Milling <strong>of</strong> the brown rice to obtain milled rice removes bran layers<br />

Abbreviations: ABTS, 2,2-azino-bis-(3-ehylbenzothiazoline-6-sulphonic acid)<br />

diammonium salt; GAE, gallic acid equivalent; RE, rutin equivalent; TEAC, trolox<br />

equivalent antioxidant capacity.<br />

* Corresponding author. Tel.: þ86 571 8697 1932; fax: þ86 571 8697 1421.<br />

E-mail address: jsbao@zju.edu.cn (J. Bao).<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.07.010<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 106–111<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

Total phenolics, flavonoid contents and antioxidant capacity from a wide collection <strong>of</strong> rice germplasm<br />

were measured, and their relations to grain color, grain size and 100-grain weight were investigated.<br />

Highly significant genotypic differences were observed in total phenolics, flavonoid contents and 2,2azino-bis-(3-ehylbenzothiazoline-6-sulphonic<br />

acid) diammonium salt (ABTS) radical cation antioxidant<br />

capacity. They displayed an increasing order in the white rice, red rice and black rice, yet several white<br />

rice had higher phenolics and flavonoids contents than the red rice. Significant positive pair-wise<br />

correlations were found among the phenolics, flavonoid contents and antioxidant capacity, and the<br />

coefficient between the phenolic contents and antioxidant capacity was extremely high (r ¼ 0.96).<br />

Among all rice accessions, the grain color parameters had negative correlations with the phenolics,<br />

flavonoid contents and antioxidant capacity (p < 0.001). The negative correlation between a* and antioxidant<br />

capacity, and the positive correlation between H and antioxidant capacity were consistent<br />

within the respective white rice and red rice groups. Flavonoid contents had positive correlation with<br />

grain length and length to width ratio, and had negative correlation with the 100-grain weight among all<br />

rice accessions. It was also found that 100-grain weight still had negative correlations with phenolics,<br />

flavonoid contents and antioxidant capacity within the white rice genotypes. These relationships may<br />

serve as indexes to indirectly select breeding lines high in the phenolics, flavonoids and antioxidant<br />

capacity. Principal component analysis including the information for phenolics, flavonoids, antioxidant<br />

capacity, grain color parameters, grain size and 100-grain weight extracted five principal components<br />

that explained 83.7% <strong>of</strong> the total variances. The results <strong>of</strong> this study may provide new opportunities for<br />

rice breeders and eventually commercial rice growers to promote the production <strong>of</strong> rice with enhanced<br />

nutritional quality.<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

that are rich in protein, fiber, oil, minerals, vitamins, and other<br />

phytochemicals (Orthoefer and Eastman, 2004; Yokoyama, 2004),<br />

leading to loss <strong>of</strong> most <strong>of</strong> the nutritional components <strong>of</strong> the rice<br />

grain. Numerous studies have shown that the essential phytochemicals<br />

in fruits, vegetables and cereal grains, including rice, are<br />

significantly associated with reduced risk <strong>of</strong> developing chronic<br />

diseases such as cardiovascular disease, type 2 diabetes, and some<br />

cancers (Liu, 2007; Yawadio et al., 2007; Yokoyama, 2004).<br />

Bi<strong>of</strong>ortification <strong>of</strong> rice to improve nutritional quality and to<br />

combat nutritional deficiency by a transgenic engineering<br />

approach has several successful examples (Bouis et al., 2003). Ye<br />

et al. (2000) introduced the b-carotene synthesis pathway to<br />

rice endosperm by genetic engineering to obtain the golden rice<br />

that produced 1.6 mg/g <strong>of</strong> b-carotene in the grain. Storozhenko<br />

et al. (2007) reported bi<strong>of</strong>ortification <strong>of</strong> folate content in rice<br />

grain by over expression <strong>of</strong> two Arabidopsis genes encoding GTP<br />

cyclohydrolase I and aminodeoxychorismate synthase under the


control <strong>of</strong> strong endosperm-specific promoters. The transgenic<br />

rice grain contained up to 100-fold higher folate levels<br />

compared to the wild type. It should be noted that conventional<br />

breeding is still possible to improve the nutritional components<br />

in grains. For example, Harjes et al. (2008) reported that variations<br />

at the lycopene epsilon cyclase (lcyE) locus alter flux down<br />

a-carotene versus b-carotene branches <strong>of</strong> the carotenoid pathways<br />

in the maize grain, and four lcyE natural polymorphism<br />

explained 58% <strong>of</strong> the variation in these two branches and<br />

a threefold difference in provitamin A compounds. Thus, selection<br />

<strong>of</strong> favorable lcyE alleles with inexpensive molecular markers<br />

enables developing-country breeders to more effectively<br />

improve provitamin A levels. However, effort towards bi<strong>of</strong>ortification<br />

<strong>of</strong> rice grain to improve nutritional quality by<br />

conventional breeding has been scarcely reported (Bouis et al.,<br />

2003). As a primary step to achieve this goal, it is necessary to<br />

investigate the genotypic diversity in the phytochemicals among<br />

diverse rice accessions, so as to find a way to enrich these<br />

compositions by breeding.<br />

The genotypic diversity <strong>of</strong> some phytochemicals in rice bran<br />

layers has been widely characterized (Dykes and Rooney, 2007; Liu,<br />

2007). For example, Bergman and Xu (2003) reported genotypic<br />

and environmental effects on tocopherol, tocotrienol, and g-oryzanol<br />

contents <strong>of</strong> rice, and Miller and Engel (2006) also reported the<br />

contents <strong>of</strong> g-oryzanol in brown rice. Jiang et al. (2007) reported<br />

minerals contents and their correlation with other quality traits <strong>of</strong><br />

rice. However, some phytochemicals, including phenolics and<br />

flavonoids, have not received as much attention as other compositions<br />

in rice grains and the phytochemicals in other cereals, fruits<br />

and vegetables (Liu, 2007).<br />

Phenolics are compounds possessing one or more aromatic<br />

rings with one or more hydroxyl groups (Liu, 2007). Phenolic<br />

compounds in diet may provide health benefits associated with<br />

reduced risk <strong>of</strong> chronic disease (Liu, 2007). Chinese medicinal<br />

plants have high levels <strong>of</strong> phenolics and potent antioxidant<br />

capacity, which might contribute to the protective effects against<br />

cancer (Cai et al., 2004). In rice, G<strong>of</strong>fman and Bergman (2004)<br />

studied the genotypic and environmental effects <strong>of</strong> the kernel<br />

phenolic content, and found that bran color was highly statistically<br />

significant for bran phenolic contents. Flavonoids are one<br />

group <strong>of</strong> phenolics, which consists <strong>of</strong> two aromatic rings linked<br />

by 3 carbons that are usually in an oxygenated heterocycle ring<br />

(Liu, 2004). Anthocyanins are a group <strong>of</strong> reddish to purple watersoluble<br />

flavonoids that are the primary pigments in the red and<br />

black grains, and have been widely identified and characterized<br />

in cereal grains (Abdel-Aal et al., 2006). The major components<br />

<strong>of</strong> anthocyanidins in colored rice are cyaniding-3-O-b-glucoside<br />

and peonidin-3-O-b-glucoside (Abdel-Aal et al., 2006; Yawadio<br />

et al., 2007). There have been few reports on characterization <strong>of</strong><br />

other flavonoids such as flavonols, flavones, flavanols, and<br />

flavanones.<br />

The phenolic compounds are also known as antioxidants<br />

(Abdel-Aal et al., 2006; Adom and Liu, 2002; Hu et al., 2003).<br />

Antioxidants have long been recognized to have protective functions<br />

against oxidative damage, and are associated with reduced<br />

risk <strong>of</strong> chronic diseases (Adom and Liu, 2002; Liu, 2007). Other<br />

phytochemicals such as carotenoids, tocols and g-oryzanols are also<br />

antioxidants (Aguilar-Garcia et al., 2007; Choi et al., 2007; Xu et al.,<br />

2001).<br />

The objective <strong>of</strong> this study was to evaluate total phenolics,<br />

flavonoids and antioxidant capacity <strong>of</strong> a large number <strong>of</strong> rice<br />

genotypes (481 accessions) and to analyze their relationships with<br />

grain color, size and 100-grain weight. The results <strong>of</strong> this study<br />

could provide rice breeders and eventually commercial rice<br />

growers new opportunities to promote the production <strong>of</strong> rice with<br />

enhanced levels <strong>of</strong> the bioactive compounds.<br />

Y. Shen et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 106–111 107<br />

2. Experimental<br />

2.1. Materials<br />

A total <strong>of</strong> 481 rice accessions including 423 white rice, 52 red<br />

rice and 6 black rice, were employed in this study. All the rice was<br />

grown in the Zhejiang University farm in 2006. They were sown in<br />

late May, transplanted on June 20, and harvested in October, and<br />

the field management followed conventional practices. Rice grains<br />

were air-dried and stored at room temperature for three months.<br />

Then they were dehusked on a Satake Rice Machine (Satake Co.,<br />

Japan), and ground to pass through a 100-mesh sieve on a Cyclone<br />

Sample Mill (UDY Corporation, Fort Collins, Colorado, USA). 2,2-<br />

Azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) diammonium<br />

salt (ABTS), potassium persulfate and gallic acid were purchased<br />

from BBI (Ontario, Canada), 6-hydroxy-2,5,7,8-tetramethylchromane-2-carboxylic<br />

acid (Trolox) from Sigma/Aldrich (St. Louis,<br />

MO), Folin–Ciocalteu reagent from Fluka Chemie AG (Buchs, Switzerland),<br />

and rutin from Oukang Company (Chengdu, China).<br />

2.2. Extraction<br />

Wholemeal flours (1 g) <strong>of</strong> each accession were extracted with<br />

25 mL <strong>of</strong> methanol containing 1% HCl for 24 h at 24 C. The<br />

procedure was repeated twice. The methanolic extracts were<br />

centrifuged at w4000g for 15 min and the supernatants were<br />

pooled and stored at 4 C.<br />

2.3. Color <strong>of</strong> rice grain<br />

The color <strong>of</strong> rice grain sample was measured with a TC-PIIG<br />

automatic color difference meter (Beijing Optical Instrument<br />

Factory, Beijing, China). Color measurements were expressed as<br />

tristimulus parameters, L*, a*, and b*. L* indicates lightness<br />

(100 ¼ white and 0 ¼ black). a* indicates redness–greenness and b*<br />

indicates yellowness–blueness. In addition, the chroma (C) value<br />

indicates color intensity or saturation, calculated as<br />

C ¼ða* 2 þ b* 2 Þ 1=2 , and Hue angle was calculated as H ¼ tan 1 (b*/<br />

a*) (Bao et al., 2005).<br />

2.4. Total phenolics<br />

Total phenolic content was assayed by the Folin–Ciocalteu<br />

colorimetric method with slight modification (Bao et al., 2005; Cai<br />

Table 1<br />

Variations in phenolics, flavonoids contents and antioxidant capacity among white<br />

(n ¼ 423), red (n ¼ 52) and black (n ¼ 6) rice genotypes<br />

Phenolics a<br />

Flavonoids a<br />

Antioxidant capacity a<br />

Total rice<br />

Mean SD 197.5 144.8 134.7 19.8 0.413 0.696<br />

CV (%) 73.3 14.7 168.63<br />

Range 108.1–1244.9 88.6–286.3 0.012–5.533<br />

White rice<br />

Mean SD 151.8 19.5 131.6 14.2 0.196 0.073<br />

CV (%) 12.9 10.8 37.33<br />

Range 108.1–251.4 88.6–170.7 0.012–0.413<br />

Red rice<br />

Mean SD 470.1 107.2 147.2 18.0 1.705 0.600<br />

CV (%) 22.8 12.3 35.22<br />

Range 165.8–731.8 108.7–190.3 0.291–2.963<br />

Black rice<br />

Mean SD 1055.7 176.2 240.6 38.1 4.484 1.095<br />

CV (%) 16.7 15.8 24.41<br />

Range 841.0–1244.9 187.6–286.3 2.527–5.533<br />

a Phenolics content was expressed as mg GAE/100 g, flavonoids content was<br />

expressed as mg RE/100 g, and antioxidant capacity was expressed as mM TAEC.


108<br />

No. <strong>of</strong> accessions<br />

No. <strong>of</strong> accessions<br />

No. <strong>of</strong> accessions<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

450<br />

400<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

600<br />

Phenolics content (mg GAE / 100g)<br />

180<br />

Flavonoids content (mg RE / 100g)<br />

3.0<br />

Antioxidant capacity (mM TAEC)<br />

black<br />

red<br />

white<br />

black<br />

red<br />

white<br />

black<br />

red<br />

white<br />

Fig. 1. Mean distributions <strong>of</strong> phenolics, flavonoids contents and antioxidant capacity<br />

among white rice (423), red rice (52) and black rice (6) accessions.<br />

et al., 2004). Briefly, aliquots (1.0 mL) <strong>of</strong> appropriately diluted<br />

extracts or standard solutions were mixed with 0.5 mL 0.5 N<br />

Folin–Ciocalteu reagent, then the reaction was neutralized with<br />

saturated sodium carbonate (75 g/L). The absorbance <strong>of</strong> the<br />

resulting blue color was recorded using a spectrophotometer after<br />

incubation for 2 h at 23 C. A calibration curve was prepared using<br />

gallic acid solution. Total phenolics contents were expressed as<br />

milligrams <strong>of</strong> gallic acid equivalent (mg GAE) per 100 g <strong>of</strong> dry<br />

weight.<br />

Y. Shen et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 106–111<br />

2.5. Total flavonoids<br />

Total flavonoid content was determined by a colorimetric<br />

method (Bao et al., 2005) with minor modification. Aliquots<br />

(0.5 mL) <strong>of</strong> appropriately diluted extracts or standard solutions<br />

were pipetted into 15-mL polypropylene conical tubes containing<br />

2 mL double distilled H2O and mixed with 0.15 mL 5% NaNO2. After<br />

5 min, 0.15 mL 10% AlCl3$6H2O solution was added, and the mixture<br />

was allowed to stand for another 5 min, and then 1 mL 1 M NaOH<br />

was added. The reaction solution was well mixed, kept for 15 min,<br />

and the absorbance was determined at 415 nm. Total flavonoid<br />

content was calculated using the standard rutin curve, and<br />

expressed as mg rutin equivalent (mg RE) per 100 g <strong>of</strong> dry weight.<br />

2.6. Radical cation ABTS þ scavenging activity<br />

Total antioxidant capacity <strong>of</strong> rice extracts was carried out<br />

using a spectrophotometer by the improved 2,2-azino-bis-<br />

(3-ehylbenzothiazoline-6-sulphonic acid) diammonium salt<br />

(ABTS) radical cation method as described (Bao et al., 2005; Cai<br />

et al., 2004). ABTS þ solution (3.9 mL, absorbance <strong>of</strong> 0.700) was<br />

added to 0.1 mL <strong>of</strong> the extracts and mixed thoroughly. The<br />

reaction mixture was kept at room temperature for 6 min and the<br />

absorbance was immediately recorded at 734 nm. Trolox standard<br />

solution in 80% ethanol was prepared and assayed under the<br />

same conditions. Results were expressed in terms <strong>of</strong> Trolox<br />

equivalent antioxidant capacity (TEAC, mM Trolox equivalents per<br />

100 g dry weight).<br />

2.7. Statistical analysis<br />

All the analyses were carried out at least in duplicate and in<br />

randomized order with mean values being reported. Analysis <strong>of</strong><br />

variance (ANOVA), correlation analysis and principal component<br />

analysis <strong>of</strong> the results were performed in SAS (S<strong>of</strong>tware Version 9.1.<br />

SAS Institute Inc., Cary, NC).<br />

3. Results<br />

3.1. Total phenolics, flavonoid contents and antioxidant capacity<br />

There were wide range <strong>of</strong> variations in the total phenolics in rice<br />

grain (Table 1, Fig. 1). Among all the rice accessions, total phenolic<br />

content ranged from 108.1 to 1244.9 mg GAE/100 g, with the lower<br />

values coming from the white rice, while the higher values were<br />

from red and black rice (Fig. 1). Variations were still found within<br />

the white rice and red rice, ranging from 108 to 251 mg GAE/100 g<br />

and from 165.8 to 731.8 mg GAE/100 g for white and red rice,<br />

respectively (Table 1). However, several red rice accessions still had<br />

lower total phenolic contents than the white rice (Fig. 1).<br />

Flavonoid contents in all the rice ranged from 88.6 to<br />

286.3 mg RE/100 g. The mean flavonoid contents among the white,<br />

red and black rice were 131.6, 147.2 and 240.6 mg RE/100 g,<br />

respectively. Even though red rice had average higher levels <strong>of</strong><br />

flavonoids than the white rice, some red rice accessions still had<br />

Table 2<br />

Pair-wise correlations among phenolics, flavonoids contents and antioxidant<br />

capacity (ABTS) among the total rice, white and red rice genotypes<br />

Total rice White rice Red rice<br />

Flavonoids ABTS Flavonoids ABTS Flavonoids ABTS<br />

Phenolics 0.681*** 0.962*** 0.703*** 0.231*** 0.461*** 0.777***<br />

Flavonoids 0.612*** 0.101* 0.342*<br />

* , ** and *** were significant at 0.05, 0.01 and 0.001 probability level, respectively.


70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

-10<br />

-20<br />

-30<br />

white rice red rice black rice<br />

Fig. 2. Mean color parameters <strong>of</strong> white, red and black rice.<br />

lower contents than white rice, whereas most <strong>of</strong> red rice had lower<br />

contents than the black rice (Fig. 1).<br />

The total antioxidant capacity was measured using the ABTS<br />

assay. It was varied to a great extent, averaged 0.413 mM TEAC,<br />

ranging from 0.012 to 5.533 mM TEAC among the total rice<br />

accessions (Table 1, Fig. 1). Among the white rice, the mean<br />

ABTS was 0.196 mM TEAC, ranging from 0.012 to 0.413 mM TEAC,<br />

whereas among the red rice, it averaged 1.705 mM TEAC, ranging<br />

from 0.291 to 2.963 mM TEAC. The six black rice samples had<br />

average antioxidant capacity <strong>of</strong> 4.484 mM TEAC, around three<br />

times <strong>of</strong> that <strong>of</strong> the red rice (Table 1).<br />

3.2. Correlations among total phenolics, flavonoids and<br />

antioxidant capacity<br />

Pair-wise correlations between the three parameters were<br />

positive (P < 0.001) among all rice groups, which were still true<br />

within the respective white rice and red rice accessions (Table 2).<br />

The phenolic contents were highly positively correlated with the<br />

antioxidant capacity (r ¼ 0.962) among all the rice. However, the<br />

correlation coefficient within the white rice accessions (r ¼ 0.231)<br />

was much smaller than that within the red rice accessions<br />

(r ¼ 0.777). The correlation coefficient between phenolic contents<br />

and flavonoid contents was higher within white rice groups<br />

(r ¼ 0.703) than that within the red rice accessions (r ¼ 0.461). The<br />

correlation coefficient between flavonoid content and antioxidant<br />

capacity was much smaller within the white and red rice accessions<br />

(Table 2) than that within the total rice groups (r ¼ 0.612), which<br />

might be due to the smaller variation in each group when<br />

compared to the total rice accessions (Table 2).<br />

3.3. Relationships between phenolics, flavonoids, antioxidant<br />

capacity and grain color<br />

Apparently, the red rice had L*, a*, and b* values <strong>of</strong> color<br />

parameters smaller than the white rice, but larger than the black<br />

rice, though the color parameters still differed within the white and<br />

L*<br />

a*<br />

b*<br />

red rice groups (Fig. 2). Among total rice accessions, the five color<br />

parameters (L*, a*, b*, C and H ) were significantly negatively<br />

correlated with phenolics, flavonoid contents, and antioxidant<br />

capacity (Table 3). Within the white rice accessions, some correlations<br />

were still significant (Table 3), but only the L* vs phenolic<br />

contents, a* vs antioxidant capacity, H vs phenolic contents and<br />

flavonoid contents were negatively correlated, whereas antioxidant<br />

capacity vs L*, b* and H , and a* vs phenolics and flavonoid contents<br />

were positive, which implied that the correlation from total rice<br />

accessions could not be applied to the individual rice groups.<br />

Within the red rice group, a* was still negatively correlated with<br />

phenolic contents and antioxidant capacity, C was also negatively<br />

correlated with antioxidant capacity, but the H was positively<br />

correlated with phenolic contents and antioxidant capacity<br />

(Table 3). It was worthy to note that the negative correlation<br />

between a* and antioxidant capacity was consistent in white and<br />

red rice groups, and so was the positive correlation between H and<br />

antioxidant capacity (Table 3), suggesting that these correlations<br />

could be used as indirect indexes to select rice breeding lines with<br />

high antioxidant capacity.<br />

3.4. Relationships between phenolics, flavonoids, antioxidant<br />

capacity and grain size and 100-grain weight<br />

Among all rice accessions, only flavonoid contents were poorly<br />

positively correlated with grain length, length/width ratio and 100grain<br />

weight (Table 4). However, it seems that the antioxidant<br />

capacity was negatively correlated with, whereas the flavonoids<br />

contents were positively correlated with grain length and length/<br />

width ratio within the white rice group (Table 4). The 100-grain<br />

weight was negatively correlated with phenolics, flavonoid<br />

contents and antioxidant capacity. Interestingly, the flavonoid<br />

contents still had positive correlation with grain length within the<br />

red rice group (r ¼ 0.325). Contrast to the relationship within the<br />

total rice and white rice, the 100-grain weight was positively<br />

correlated with the flavonoid contents within the red rice accessions<br />

(Table 4).<br />

3.5. Principal component analysis<br />

Principal component analysis was performed on the ten variables<br />

including phenolic contents, flavonoid content, antioxidant<br />

capacity, color parameters, grain size and 100-grain weight (Tables<br />

5 and 6). All the twelve principal components and their corresponding<br />

eigenvalues and variances are listed in Table 5. The results<br />

indicated that the first five principal components could explain<br />

83.7% <strong>of</strong> total variance. The first principal component (PC1) was the<br />

most important one, explaining 26.1% <strong>of</strong> the total variance (Table 5).<br />

The PC1 represented the phenolics content, C and H <strong>of</strong> color<br />

parameters (Table 6). The second principal component (PC2)<br />

accounted for an additional 19.2% <strong>of</strong> the total variances, which was<br />

mainly attributed to flavonoid contents, L* and a* <strong>of</strong> color parameters.<br />

The third principal component (PC3) accounted for 16.2% <strong>of</strong><br />

the total variance; the variation was mainly contributed by<br />

Table 3<br />

Correlation coefficients between phenolics, flavonoids contents, antioxidant capacity (ABTS) and grain color parameters among the total rice, white and red rice genotypes<br />

Total rice White rice Red rice<br />

Phenolics Flavonoids ABTS Phenolics Flavonoids ABTS Phenolics Flavonoids ABTS<br />

L* 0.774*** 0.463*** 0.738*** 0.103* 0.065 0.333*** 0.166 0.014 0.128<br />

a* 0.349*** 0.277*** 0.398*** 0.170*** 0.136** 0.460*** 0.292* 0.165 0.335*<br />

b* 0.592*** 0.481*** 0.579*** 0.066 0.018 0.212*** 0.161 0.101 0.261<br />

C 0.271*** 0.145** 0.311*** 0.156** 0.089 0.001 0.191 0.089 0.307*<br />

H 0.589*** 0.529*** 0.582*** 0.152** 0.126* 0.477*** 0.326* 0.195 0.332*<br />

* , ** and *** were significant at 0.05, 0.01 and 0.001 probability level, respectively.<br />

Y. Shen et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 106–111 109


110<br />

Table 4<br />

Correlation coefficients between phenolics, flavonoids contents, antioxidant capacity (ABTS) and grain size and 100-grain weight among the total rice, white and red rice<br />

genotypes<br />

Total rice White rice Red rice<br />

Phenolics Flavonoids ABTS Phenolics Flavonoids ABTS Phenolics Flavonoids ABTS<br />

Length 0.005 0.169*** 0.036 0.064 0.177*** 0.178*** 0.072 0.325* 0.106<br />

Width 0.011 0.030 0.007 0.042 0.052 0.054 0.021 0.155 0.068<br />

Length/Width ratio 0.031 0.106* 0.042 0.014 0.132** 0.097* 0.064 0.083 0.113<br />

100 grain weight 0.067 0.110* 0.065 0.146** 0.117* 0.160** 0.126 0.322* 0.168<br />

* , ** and *** were significant at 0.05, 0.01 and 0.001 probability level, respectively.<br />

antioxidant capacity, grain length and grain width. The fourth<br />

principal component mainly representing the grain length to width<br />

ratio and 100-grain weight explained 13.2% <strong>of</strong> the total variance.<br />

The fifth principal component representing the b* <strong>of</strong> color parameter<br />

with eigenvalue <strong>of</strong> 1.095, explained an additional 9.1% <strong>of</strong> the<br />

total variance (Tables 5 and 6).<br />

4. Discussion<br />

The role <strong>of</strong> phenolics as natural antioxidants has attracted<br />

considerable interest due to their pharmacological functions.<br />

Increased consumption <strong>of</strong> phenolic compounds has been associated<br />

with the reduced risk <strong>of</strong> cardiovascular diseases and certain<br />

cancers (Liu, 2004, 2007; Dykes and Rooney, 2007). The whole rice<br />

grain had phenolic contents ranging from 108.1 to 1244.9 mg GAE/<br />

100 g (Table 1), depending on the color <strong>of</strong> grain (Choi et al., 2007;<br />

G<strong>of</strong>fman and Bergman, 2004). G<strong>of</strong>fman and Bergman (2004)<br />

reported that the phenolic contents in the white, red and purple<br />

rice ranged from 25 to 246, 34 to 424, 69 to 535 mg GAE/100 g,<br />

which was a little lower than this study (Table 1). Three reasons<br />

may explain the differences. First, they used rice materials grown in<br />

two years, which means that material harvested from the first year<br />

had to be stored. Storage <strong>of</strong> rice grain results in a decrease <strong>of</strong><br />

phenolic content (Zhou et al., 2004), thus their data was lower than<br />

ours. Second, rice phenolic compounds exist in free, esterified and<br />

insoluble-bound forms, and insoluble-bound phenolics may be<br />

released by base, acid or enzymatic treatment <strong>of</strong> samples prior to<br />

extraction (Adom and Liu, 2002; Choi et al., 2007; Zhou et al., 2004).<br />

Our extraction solution including 1% HCL may cause release <strong>of</strong> at<br />

least part <strong>of</strong> the bound phenolics, thus leading to higher phenolic<br />

contents. Third, more rice accessions were used in this study may<br />

provide more wide diversity in the phenolic contents. Flavonoids<br />

have potent antioxidant and anticancer activities (Adom and Liu,<br />

2002; Dykes and Rooney, 2007; Hu et al., 2003). Coloration <strong>of</strong> rice<br />

may be derived from accumulation <strong>of</strong> anthocyanins (Furukawa<br />

et al., 2007). Thus, it could be expected that the white rice had<br />

mean flavonoid content (131.6 mg RE/100 g) lower than those <strong>of</strong><br />

red rice (147.2 mg RE/100 g) and black rice (240 mg RE/100 g)<br />

Table 5<br />

Principal component analysis for all the ten parameters<br />

Component Eigenvalue Variance (%) Cumulative variance (%)<br />

1 3.126 26.05 26.05<br />

2 2.300 19.17 45.22<br />

3 1.941 16.18 61.40<br />

4 1.579 13.16 74.56<br />

5 1.095 9.12 83.68<br />

6 0.726 6.05 89.73<br />

7 0.691 5.76 95.49<br />

8 0.263 2.19 97.68<br />

9 0.215 1.79 99.47<br />

10 0.037 0.30 99.78<br />

11 0.023 0.19 99.97<br />

12 0.004 0.03 100.00<br />

Y. Shen et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 106–111<br />

(Table 1). It should be noted that some red rice accessions were<br />

lower in the flavonoid contents than white rice (Fig. 1). It was<br />

possible that other flavonoid compositions (e.g. flavonols, flavanones)<br />

rather than anthocyanins in the white rice were higher than<br />

those in the red rice. The antioxidant capacity <strong>of</strong> rice grain was<br />

contributed not only by phenolic compounds, but also by other<br />

phytochemicals, such as carotenoids, tocols and g-oryzanols (Choi<br />

et al., 2007; Xu et al., 2001). However, phenolic contents were<br />

reportedly positively correlated with TABS antioxidant capacity<br />

(r ¼ 0.962), which was consistent with nearly all other studies<br />

(Adom and Liu, 2002; Choi et al., 2007) using samples from<br />

different plant origins (Cai et al., 2004).<br />

Bi<strong>of</strong>ortification by conventional breeding is one <strong>of</strong> the strategies<br />

to improve nutritional quality <strong>of</strong> rice grain. It is apparent that<br />

phenolics, flavonoids and antioxidant capacity differ from the grain<br />

color (Table 1), yet these components and antioxidant capacity still<br />

differ among the white color rice (Table 1). Therefore, breeding<br />

efforts for better nutritional quality may be applied for white rice, if<br />

the whole grain is used for food product development. For ease <strong>of</strong><br />

breeding, indirect selection approaches are <strong>of</strong>ten used in the<br />

breeding program. One <strong>of</strong> indirect selection methods is to use the<br />

color parameters which can be easily measured in an automatic<br />

color difference meter. The color parameters L*, b* and H are<br />

positively, while a* is negatively correlated with antioxidant<br />

capacity (Table 3). Thus, breeding lines high in antioxidant could be<br />

indirectly selected by selecting grain with larger L*, b* and H , but<br />

smaller a*.<br />

It is well known that most <strong>of</strong> the phytochemicals are rich in the<br />

bran layers which make up approximately 8–10% <strong>of</strong> the rough rice<br />

weight (Yokoyama, 2004), though some <strong>of</strong> them are still present in<br />

the milled rice. It is general knowledge that, in a given weight <strong>of</strong><br />

rice, the larger the grain size, the smaller surface area per weight <strong>of</strong><br />

grain. Smaller surface area might imply lower phytochemicals in<br />

the bran layers. However, whether the grain size negatively<br />

correlated with phenolics, flavonoids and antioxidant capacity was<br />

not tested before. It is found that the 100-grain weight was truly<br />

negatively correlated with flavonoid contents among the total rice<br />

accessions, and the correlations were still true among white rice,<br />

but not true among red rice accessions (Table 4). The grain size was<br />

Table 6<br />

Sources <strong>of</strong> variation for the first five principal components (PC)<br />

PC1 PC2 PC3 PC4 PC5<br />

Phenolics 0.481 0.065 0.092 0.075 0.215<br />

Flavonoids 0.143 0.591 0.187 0.136 0.164<br />

ABTS 0.323 0.237 0.433 0.029 0.333<br />

L* 0.224 0.479 0.320 0.094 0.240<br />

a* 0.173 0.548 0.255 0.135 0.210<br />

b* 0.274 0.098 0.087 0.018 0.564<br />

C 0.485 0.011 0.297 0.111 0.109<br />

H 0.464 0.062 0.313 0.095 0.062<br />

Length 0.076 0.079 0.492 0.305 0.455<br />

Width 0.128 0.128 0.394 0.364 0.413<br />

Length/width ratio 0.106 0.101 0.095 0.522 0.030<br />

100 grain weight 0.073 0.123 0.049 0.655 0.053


not necessarily negatively correlated with flavonoids, but inversely<br />

the relationship was positive. The antioxidant capacity was negatively<br />

correlated with grain length, length/width ratio and 100grain<br />

weight (Table 4), so in the white rice breeding practice, high<br />

antioxidant capacity <strong>of</strong> breeding lines could be indirectly selected<br />

with short grain length and smaller 100-grain weight in breeding<br />

programs.<br />

In conclusion, this study found wide diversity in the phenolics,<br />

flavonoid contents and antioxidant capacity in the whole rice grain.<br />

These data provide opportunities to further increase the content <strong>of</strong><br />

phenolics, flavonoids and antioxidant capacity by breeding, especially<br />

in white rice. Their relationships with grain color, grain size<br />

and 100-grain weight could serve as indexes to indirectly select rice<br />

breeding lines high in phenolics, flavonoids and antioxidant<br />

capacity. The results could provide rice breeders and eventually<br />

commercial rice producers, with new opportunities to promote the<br />

production <strong>of</strong> rice with enhanced levels <strong>of</strong> the phytochemicals. Rice<br />

genotypes rich in phytochemicals may be incorporated into functional<br />

foods.<br />

Acknowledgement<br />

The authors thank Mr. Junquan Yu and Dr. Jianliang Lu for their<br />

assistance in chemical analysis and the color parameters test,<br />

respectively. We also thank the anonymous reviewers for their<br />

constructive comments. Financial support for this work was<br />

provided in part by National High Technology Development Project,<br />

National Natural <strong>Science</strong> Foundation <strong>of</strong> China and the <strong>Science</strong> and<br />

Technology Department <strong>of</strong> Zhejiang Province.<br />

References<br />

Abdel-Aal, E.M., Young, J.C., Rabalski, I., 2006. Anthocyanin composition in black,<br />

blue, pink, purple, and red cereal grains. <strong>Journal</strong> <strong>of</strong> Agricultural and Food<br />

Chemistry 54, 4696–4704.<br />

Adom, K.K., Liu, R.H., 2002. Antioxidant activity <strong>of</strong> grains. <strong>Journal</strong> <strong>of</strong> Agricultural and<br />

Food Chemistry 50, 6182–6187.<br />

Aguilar-Garcia, C., Gavino, G., Baragano-Mosqueda, M., Hevia, P., Gavino, V.C., 2007.<br />

Correlation <strong>of</strong> tocopherol, tocotrienol, g-oryzanol and total polyphenol content<br />

in rice bran with different antioxidant capacity assays. Food Chemistry 102,<br />

1228–1232.<br />

Bao, J.S., Cai, Y., Sun, M., Wang, G.Y., Corke, H., 2005. Anthocyanins, flavonols, and<br />

free radical scavenging activity <strong>of</strong> Chinese bayberry (Myrica rubra) extracts and<br />

their color properties and stability. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry<br />

53, 2327–2332.<br />

Y. Shen et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 106–111 111<br />

Bergman, C.J., Xu, Z., 2003. Genotype and environment effects on tocopherol,<br />

tocotrienol, and g-oryzanol contents <strong>of</strong> Southern U.S. rice. <strong>Cereal</strong> Chemistry 80,<br />

446–449.<br />

Bouis, H.E., Chassy, B.M., Ochanda, J.O., 2003. Genetically modified food crops and<br />

their contribution to human nutrition and food quality. Trends in Food <strong>Science</strong><br />

and Technology 14, 191–209.<br />

Cai, Y.Z., Luo, Q., Sun, M., Corke, H., 2004. Antioxidant activity and phenolic<br />

compounds <strong>of</strong> 112 traditional Chinese medicinal plants associated with anticancer.<br />

Life <strong>Science</strong> 74, 2157–2184.<br />

Choi, Y., Jeong, H.S., Lee, J., 2007. Antioxidant activity <strong>of</strong> methanolic extracts from<br />

some grains consumed in Korea. Food Chemistry 103, 130–138.<br />

Dykes, L., Rooney, L.W., 2007. Phenolic compounds in cereal grains and their health<br />

benefits. <strong>Cereal</strong> Foods World 52, 105–111.<br />

Furukawa, T., Maekawa, M., Oki, T., Suda, I., Lida, S., Shimada, H., Takamure, I.,<br />

Kadowaki, K.I., 2007. The Rc and Rd genes are involved in proanthocyanidin<br />

synthesis in rice pericarp. Plant <strong>Journal</strong> 49, 91–102.<br />

G<strong>of</strong>fman, F.D., Bergman, C.J., 2004. Rice kernel phenolic content and its relationship<br />

with antiradical efficiency. <strong>Journal</strong> <strong>of</strong> the <strong>Science</strong> <strong>of</strong> Food and Agriculture 84,<br />

1235–1240.<br />

Harjes, C.E., Rocheford, T.R., Bai, L., Brutnell, T.P., Kandianis, C.B., Sowinski, S.G., et al.,<br />

2008. Natural genetic variation in Lycopene Epsilon Cyclase tapped for maize<br />

bi<strong>of</strong>ortification. <strong>Science</strong> 319, 330–333.<br />

Hu, C., Zawistowski, J., Ling, W., Kitts, D.D., 2003. Black rice (Oryza sativa L. indica)<br />

pigmented fraction suppresses both reactive oxygen species and nitric oxide in<br />

chemical and biological model systems. <strong>Journal</strong> <strong>of</strong> Agricultural and Food<br />

Chemistry 51, 5271–5277.<br />

Jiang, S.L., Wu, J.G., Feng, Y., Yang, X.E., Shi, C.H., 2007. Correlation analysis <strong>of</strong><br />

mineral element contents and quality traits in milled rice (Oryza sativa L.).<br />

<strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 55, 9608–9613.<br />

Liu, R.H., 2004. Potential synergy <strong>of</strong> phytochemicals in cancer prevention: mechanism<br />

<strong>of</strong> action. <strong>Journal</strong> <strong>of</strong> Nutrition 134, 3479S–3485S.<br />

Liu, R.H., 2007. Whole grain phytochemicals and health. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

46, 207–219.<br />

Miller, A., Engel, K.H., 2006. Content <strong>of</strong> g-oryzanol and composition <strong>of</strong> steryl ferulates<br />

in brown rice (Oryza sativa L.) <strong>of</strong> European origin. <strong>Journal</strong> <strong>of</strong> Agricultural<br />

and Food Chemistry 54, 8127–8133.<br />

Orthoefer, F.T., Eastman, J., 2004. Rice bran and oil. In: Champagne, E.T. (Ed.), Rice<br />

Chemistry and Technology. AACC, St. Paul, Minnesota, U.S.A., pp. 569–593.<br />

Storozhenko, S., Brouwer, V.E., Volckaert, M., Navarrete, O., Blancquaert, D.,<br />

Zhang, G.F., et al., 2007. Folate fortification <strong>of</strong> rice by metabolic engineering.<br />

Nature Biotechnology 25, 1277–1279.<br />

Xu, Z., Hua, N., Godber, J.S., 2001. Antioxidant activity <strong>of</strong> tocopherols, tocotrienols,<br />

and g-oryzanol components from rice bran against cholesterol oxidation<br />

accelerated by 2,2 0 -azobis(2-methylpropionamidine) dihydrochloride. <strong>Journal</strong><br />

<strong>of</strong> Agricultural and Food Chemistry 49, 2077–2081.<br />

Yawadio, R., Tanimori, S., Morita, N., 2007. Identification <strong>of</strong> phenolic compounds<br />

isolated from pigmented rices and their aldose reductase inhibitory activities.<br />

Food Chemistry 101, 1644–1653.<br />

Ye, X., Al-Babili, S., Kloti, A., Zhang, J., Lucca, P., Beyer, P., Potrokus, I., 2000. Engineering<br />

the pro-vitamin A (beta-carotene) biosynthetic pathway into (carotenoid-free)<br />

rice endosperm. <strong>Science</strong> 287, 303–305.<br />

Yokoyama, W., 2004. Nutritional properties <strong>of</strong> rice and rice bran. In:<br />

Champagne, E.T. (Ed.), Rice Chemistry and Technology. AACC, St. Paul, Minnesota,<br />

U.S.A., pp. 595–609.<br />

Zhou, Z.K., Robards, K., Helliwell, S., Blanchard, C., 2004. The distribution <strong>of</strong> phenolic<br />

acids in rice. Food Chemistry 87, 401–406.


Nucleotide polymorphisms in the waxy gene <strong>of</strong> NaN3-induced<br />

waxy rice mutants<br />

Toong Long Jeng a , Chang Sheng Wang b , Tung Hai Tseng a , Min Tze Wu a , Jih Min Sung c, *<br />

a Division <strong>of</strong> Biotechnology, Taiwan Agricultural Research Institute, Wu Fong, Taichung 413, Taiwan, ROC<br />

b Department <strong>of</strong> Agronomy, National Chung Hsing University, Taichung 402, Taiwan, ROC<br />

c Department <strong>of</strong> Food <strong>Science</strong> & Nutrition, Hungkuang University, Shalu, Taichung 433, Taiwan, ROC<br />

article info<br />

Article history:<br />

Received 21 February 2008<br />

Received in revised form 10 July 2008<br />

Accepted 14 July 2008<br />

Keywords:<br />

bp duplication<br />

G-to-T substitution<br />

Microsatellite<br />

Polymorphism<br />

Rice<br />

Waxy mutation<br />

1. Introduction<br />

abstract<br />

Starch is the most abundant storage reserves in rice (Oryza<br />

sativa L.) grains. The starch reserves <strong>of</strong> rice grains generally contain<br />

17–30% amylose, which plays an important role in palatability and<br />

processing qualities (Ramesh et al., 1999; Zhou et al., 2002).<br />

Granule bound starch synthase (GBSS) is the key enzyme responsible<br />

for amylose synthesis in rice grain (James et al., 2003; Vandeputte<br />

and Delcour, 2004). Six loci are involved in the amylose<br />

synthesis <strong>of</strong> rice grains, with the dominant Wx gene that encodes<br />

GBSS being the major locus (Itoch et al., 2003). Mutation <strong>of</strong> the<br />

dominant Wx locus into recessive wx causes inactivation or absence<br />

<strong>of</strong> GBSS (Han et al., 2004), and subsequently results in rice grain <strong>of</strong><br />

opaque endosperm that contains little or no amylose (Wang and<br />

Wang, 2002).<br />

Waxy (wx) mutants have been observed in both japonica and<br />

indica rice genotypes by spontaneous and induced mutation (Han<br />

et al., 2004; Hori et al., 2007; Sato and Nishio, 2003). Several<br />

functional molecular markers associated with Wx gene are<br />

detectable in waxy rice genotypes. An alternative splicing at the 5 0<br />

* Correspondence author. Tel.: þ886 4 26318652x5015.<br />

E-mail address: sungjm@sunrise.hk.edu.tw (J.M. Sung).<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.07.009<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 112–116<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

Spontaneous and induced waxy phenotype, associated with endosperm containing little or no amylose,<br />

has been recognized in rice (Oryza sativa L.). Mutation <strong>of</strong> a dominant gene Wx into a recessive gene wx,<br />

which causes the inactivation or absence <strong>of</strong> granule bound starch synthase, is believed to be responsible<br />

for the change in endosperm starch leading to the waxy grain. In the present study, the nucleotide<br />

polymorphism in the Wx gene <strong>of</strong> rice genotype Tainung 67 (wild type) and its 35 NaN3-induced wx<br />

mutants were examined. Iodine staining confirmed that all the mutants had waxy grain trait. The G-to-T<br />

single base substitution analysis indicated that the wild type genotype Tainung 67 and its waxy mutants<br />

carried Wx b allele. Moreover, 23-bp duplication in exon 2 was detected in all the waxy mutants.<br />

Microsatellite polymorphism (CT)n was also detectable on the Wx gene <strong>of</strong> the tested genotypes and<br />

mutants, with at least 5 classes <strong>of</strong> (CT)n microsatellites identified at the Wx locus. Electrophoretic<br />

analyses also confirmed the observed nucleotide polymorphsim. Thus, nucleotide polymorphsim exist<br />

among NaN3-induced waxy mutants in rice. However, only the 23-bp duplication in exon 2 may be used<br />

as a molecular marker to characterize waxy grain trait in rice genotypes.<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

splice site <strong>of</strong> intron 1 caused by a single base substitution (AGGT to<br />

AGTT) generally exists in waxy rice genotypes (Han et al., 2004;<br />

Isshiki et al., 1998; Itoch et al., 2003; Yamanaka et al., 2004). A<br />

duplication <strong>of</strong> 23-bp in the second exon, which causes the inactivation<br />

or loss <strong>of</strong> GBSS, is detectable only in waxy rice genotypes<br />

(Hori et al., 2007; Jeng et al., 2007; Wanchana et al., 2003). Polymorphic<br />

cytosine and thymidine (CT)n microsatellites in the 5 0 -<br />

untranslated region are also noted in waxy rice genotypes, but with<br />

the number <strong>of</strong> alleles (4 alleles) consistently identified much less<br />

than non-waxy rice genotypes (10 alleles) (Bao et al., 2006).<br />

Many gamma ray- or chemical mutagen-induced waxy mutant<br />

lines have been obtained in rice, and the point mutations in Wx<br />

gene have been identified (Isshiki et al., 2001; Sato and Nishio,<br />

2003). In this study, DNA polymorphisms <strong>of</strong> the Wx gene among<br />

Tainung 67 and its NaN3-induced wx mutant lines (Wang et al.,<br />

2002) were investigated. The G-to-T single base substitution and<br />

23-bp duplication, which were generally used to differentiate waxy<br />

and non-waxy rice varieties on a molecular level, were used to<br />

identify the possible variations in the Wx locus <strong>of</strong> the tested<br />

genotypes and mutants. CT-microsatellite polymorphism was also<br />

employed to compare the possible differences among the tested<br />

mutants. The results <strong>of</strong> this study are useful for further study in<br />

identifying and breeding the waxy rice genotype with desirable<br />

agronomic traits.


2. Materials and methods<br />

A total <strong>of</strong> 38 rice (O. sativa L.) accessions, including wild type<br />

genotype Tainung 67 and its 35 NaN3-induced waxy mutants, and 2<br />

naturally mutated waxy genotypes TKW 1 and TCSW 1, were<br />

obtained from the Taiwan Agricultural Research Institute (Table 1).<br />

Additionally, a NaN3-induced non-waxy recovery <strong>of</strong> SA 419 (a NaN3induced<br />

waxy mutant <strong>of</strong> genotype Tainung 67) was also used to<br />

identify the possible effect <strong>of</strong> 23-bp duplication on the expression <strong>of</strong><br />

wx gene. Seeds were sown in the nursery plots. The 3-leaf stage<br />

seedlings were transplanted to experimental plots (3 6 m) at a hill<br />

spacing <strong>of</strong> 30 15 cm with 3 seedlings per hill. Each plot received<br />

a basal application <strong>of</strong> fertilizer before transplanting (24 kg N ha 1 ,<br />

36 kg P2O5 ha 1 and 24 kg K2Oha 1 ) and 3 top-dressings <strong>of</strong> fertilizer<br />

at 20 d (6 kg N ha 1 , 9 kg P2O5 ha 1 and 6 kg K2Oha 1 ), 40 d<br />

(9 kg N ha 1 , 13.5 kg P2O5 ha 1 and 9 kg K2Oha 1 ) and 60 d<br />

(9 kg N ha 1 , 13.5 kg P2O5 ha 1 and 9 kg K2Oha 1 ) after transplanting<br />

(Jeng et al., 2003). All plants were grown to maturity.<br />

Mature dry seeds were hand-cut with a razor blade, and the halfseeds<br />

without embryo were soaked in distilled water at 4 Covernight.<br />

After the treatment, the half-seeds were sliced and stained<br />

with potassium iodide for visual examination (Itoch et al., 2003).<br />

Genomic DNA was extracted from leaves <strong>of</strong> 30-day-old seedlings<br />

<strong>of</strong> each genotype or line (Doyle and Doyle, 1990). For sequence<br />

analysis <strong>of</strong> the Wx alleles, a fragment was amplified by PCR (Invitrogen<br />

Co., USA) using two primer pairs (i.e., 5 0 -CAAGCTGGAAA<br />

AGCAAAAG-3 0 and 5 0 -TTGACCAACTCGGCTACTAA, 5 0 -ATGTCTCTCG<br />

CCACTGGA-3 0 and 5 0 -CTCAGCCACAACGCTGGTAT-3 0 ). PCR amplification<br />

was carried out by using 10 ng genomic DNA, 2 mM <strong>of</strong> each<br />

primer and 1 U rTag DNA polymerase (TaKaRa Co., Japan) at 56.4 C<br />

with 32 cycles. DNA sequencing was performed on an ABI 373<br />

automated sequencer following the manufacturer’s instruction<br />

(Applied Biosystems, Inc., USA) using Dye Terminator Cycle<br />

Sequence Ready Reaction Kit (Perkin Elmer Co., USA).<br />

The G-to-T substitution at the 5 0 leader intron splice donor site<br />

<strong>of</strong> the Wx alleles was examined using a derived cleaved amplified<br />

polymorphic sequence (dCAPS) technique developed by Yamanaka<br />

et al. (2004). The forward 5 0 -TGTTGTTCATCAGGAAGAACATCTC<br />

CAAG-3 0 and reverse primers 5 0 -TTAATTTCCAGCCCAACACC-3 0<br />

generate a unique EcoT14I restriction site characteristic <strong>of</strong> the Wx a<br />

allele. Total DNA was isolated from leaf, and the DNA fragments at<br />

the splice donor site <strong>of</strong> the first intron were amplified by PCR<br />

(Invitrogen Co., USA). PCR amplification was carried out by using<br />

100 ng genomic DNA, 0.4 mM <strong>of</strong> each primer and 1 U rTag DNA<br />

polymerase at 56 C with 32 cycles. Five microliters <strong>of</strong> each PCR<br />

product was digested with EcoT14I in a total volume <strong>of</strong> 20 mL at<br />

37 C overnight. After digestion, each digest was electrophoresed in<br />

a30gkg 1 Nusieve 3:1 agarose gel with ethidium bromide staining.<br />

The 23-bp duplication was detected using the technique<br />

detailed by Wanchana et al. (2003). The forward 5 0 -TGCAGA<br />

GATCTTCCACAGCA-3 0 and reverse primers 5 0 -GCTGGTCGTCAC<br />

GCTGAG-3 0 were used to generate amplicons specific for Wx alleles.<br />

Table 1<br />

The 35 NaN3-induced rice mutants <strong>of</strong> genotype Tainung 67<br />

Code Mutant Code Mutant Code Mutant Code Mutant<br />

1 SA 419 10 SA 1472 19 SA 572 28 SA 967<br />

2 SA 402 11 SA 491 20 SA 573 29 SA 1047<br />

3 SA 420 12 SA 492 21 SA 580 30 SA 1200<br />

4 SA 472 13 SA 495 22 SA 588 31 SA 1220<br />

5 SA 893 14 SA 497 23 SA 649 32 SA 1314<br />

6 SA 897 15 SA 502 24 SA 674 33 SA 1324<br />

7 SA 1917 16 SA 523 25 SA 860 34 SA 1398<br />

8 SA 1170 17 SA 526 26 SA 861 35 SA 1399<br />

9 SA 1445 18 SA 543 27 SA 908<br />

The code numbers are used to identify the tested mutants presented on Figs. 1–3.<br />

T.L. Jeng et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 112–116 113<br />

The forward primer was marked by IRDye700 (MWG-Biotech Co.,<br />

Germany). A reaction mixture (20 mL) containing 20 ng <strong>of</strong> genomic<br />

DNA as the template, 0.2 mM <strong>of</strong> dNTPs, 0.2 mM <strong>of</strong> each primer, 0.5 U<br />

<strong>of</strong> Taq DNA polymerase, 50 mM HCl, 2.0 mM MgCl2, and 10 mM<br />

Tris-HCl (pH 8.3), and deionized H2O were added to make 20 mL<br />

final volume. The amplification reactions were carried out by the<br />

following pr<strong>of</strong>ile: 94 C pre-denaturation for 2 min followed by 32<br />

cycles <strong>of</strong> (94 C denaturation for 30 s, 60 C annealing for 30 s and<br />

72 C for 1 min) and the final extension at 72 C for 5 min. The PCR<br />

products were then electrophoresed in 70 g kg 1 polyacrylamide<br />

gel with LICOR 4300 DNA Analyzer (Licor Co., Germany). Similar<br />

PCR conditions were also used to detect CT-microsatellite polymorphism<br />

by using the forward 5 0 -CTTTGTCTATCTCAAGACAC-3 0<br />

and reverse primers 5 0 -TTGCAGATGTTCTTCCTGATG-3 0 to generate<br />

amplicons specific for Wx alleles.<br />

3. Results and discussion<br />

Notable differences in appearance <strong>of</strong> de-hulled grains were<br />

observed among the tested genotypes and NaN3-induced mutants<br />

(Fig. 1). All the tested mutants had chalky endosperm. Of these<br />

samples, the majority <strong>of</strong> mutants had white pericarp. Only 2<br />

mutants had dark blue color pericarp (SA 526 and SA 572) and 3<br />

mutants had brown color pericarp (SA 897, SA 674 and SA 1399). In<br />

Southeast Asia, waxy rice is grown mainly as a staple food (Wanchana<br />

et al., 2003). In Taiwan, waxy rice, particularly the waxy rice<br />

with colored pericarp, is generally reserved for use in festival foods<br />

and desserts (Jeng et al., 2007). Jeng et al. (2006) previously<br />

reported that, under adverse environmental conditions, the grain<br />

yield <strong>of</strong> NaN3-induced low amylose mutant SA 419 (code number 1<br />

in Table 1 and Figs. 1–3) was better than its wild type genotype<br />

Tainung 67, and therefore was recommended to waxy rice growers<br />

and food manufacturers. In this regard, these 5 mutants with dark-<br />

Fig. 1. The de-hulled grains <strong>of</strong> rice genotype Tainung 67 (TNG67) and its 35 NaN3induced<br />

mutants as well as two naturally occurred waxy mutants TKW 1 (japonica<br />

type) and TCSW 1 (indica type). The numbers marked in the figure match the code<br />

numbers listed on Table 1.


114<br />

Fig. 2. The iodine-stained endosperms <strong>of</strong> rice genotype Tainung 67 (TNG67) and its 35<br />

NaN3-induced mutants as well as two naturally occurred waxy mutants TKW 1<br />

(japonica type) and TCSW 1 (indica type). The numbers marked in the figure match the<br />

code numbers listed on Table 1.<br />

colored pericarp could also be recommended to growers and food<br />

manufacturers, if their palatability, processing qualities and grain<br />

yield were proved promising.<br />

Rice genotype is considered as waxy with 0–20 g kg 1 <strong>of</strong> grain<br />

amylose (Jeng et al., 2006). The opaque endosperm is generally<br />

regarded as a phenotypic marker for waxy rice (Bao et al., 2006;<br />

Han et al., 2004). This grain trait can be easily detected using iodine<br />

staining, which is a very sensitive technique to detect amylose in<br />

various tissues (Itoch et al., 2003). In the present study, the dark<br />

blue color <strong>of</strong> iodine staining suggested that the Wx gene was<br />

expressed in cultivar Tainung 67 (200 g kg 1 amylose) (Jeng et al.,<br />

2006) (Fig. 2). On the other hand, all the waxy mutants (genotype<br />

TCSW 1, genotype TKW 1 and 35 NaN3-induced mutants) had<br />

orange color <strong>of</strong> stained endosperm, suggesting that they either<br />

have no or severely reduced amylose (Fig. 2).<br />

Several nucleotide polymorphisms are associated with the Wx<br />

gene, namely single nucleotide polymorphism (G/T) in the intron 1,<br />

T.L. Jeng et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 112–116<br />

23-bp duplication in the exon 2 and CT-microsatellites (Bao et al.,<br />

2002; Wanchana et al., 2003; Yamanaka et al., 2004). Extensive<br />

research has indicated that two functional Wx alleles exist in<br />

cultivated rice genotypes, with Wx a having AGGT sequence and Wx b<br />

having an alternative AGTT sequence at 5 0 splice site in intron 1<br />

(Han et al., 2004; Isshiki et al., 1998; Itoch et al., 2003; Yamanaka<br />

et al., 2004). A mutation from AGGT sequence to AGTT sequence at<br />

5 0 splice site <strong>of</strong> Wx intron 1 leads to incomplete post-transcriptional<br />

processing <strong>of</strong> Wx pre-mRNA. Thus, the expression <strong>of</strong> Wx a is higher<br />

than Wx b at RNA, GBSS and amylose levels (Isshiki et al., 1998).<br />

Waxy rice genotypes generally have no detectable levels <strong>of</strong> spliced<br />

mRNA as a result <strong>of</strong> this mutation (Wang et al., 1995). They are<br />

characterized as the Wx b allele based on the G-to-T base substitution<br />

(Wang et al., 1995) and have very low or no amylose in mature<br />

grains (Wang and Wang, 2002). Nevertheless, Yamanaka et al.<br />

(2004) recently examined 353 waxy rice strains collected from<br />

various regions <strong>of</strong> Asia, and found that the collected waxy strains<br />

had both Wx a (AGGT)- and Wx b (AGTT)-derived alleles. They further<br />

indicated that the Wx b (AGTT)-derived allele was predominant, and<br />

Wx a (AGGT) allele was detected in only a minority <strong>of</strong> samples. Thus,<br />

G-to-T base substitution alone might not be a reliable marker for<br />

differentiating waxy and non-waxy rice varieties at a molecular<br />

level. In the present study, nucleotide sequence was also analyzed<br />

in genotype Tainung 67 and its 9 NaN3-induced waxy mutants as<br />

well as 2 waxy genotypes TKW 1 (japonica type waxy rice) and<br />

TCSW 1 (indica type waxy rice) (Table 2). Two genotypes Nipponbare<br />

(japonica genotype with intermediate amylose level) and 93-<br />

11 (indica genotype with high amylose level), in which their<br />

nucleotide sequence had been analyzed and released on internet<br />

http://www.ncbi.nlm.nih.gov/, were also added for comparison<br />

(Table 2). The results showed that only indica genotype 93-11 had<br />

AGGT sequence at putative 5 0 leader intron. All the other tested<br />

accessions had AGTT sequence, regardless <strong>of</strong> their origin (indica or<br />

japonica). These results suggest that all the tested waxy mutants<br />

carry the Wx b allele. On the other hand, both japonica type<br />

genotypes Tainung 67 and Nipponbare (both with intermediate<br />

amylose level) have AGTT sequence. These results seem to indicate<br />

that the G-to-T polymorphism might be a useful molecular marker<br />

for identifying the rice genotype with grain amylose below<br />

intermediate level, but it is not sufficient to identify the waxy<br />

starch grain trait.<br />

Fig. 3. Nucleotide polymorphisms <strong>of</strong> (A) 23-bp duplication, (B) G-to-T substitution and (C) (CT)n repeats on Wx alleles <strong>of</strong> rice genotype Tainung 67 (TNG67) and its 35 NaN3-induced<br />

mutants as well as two naturally occurred waxy mutants TKW 1 (japonica type) and TCSW 1 (indica type). A high amylose genotype TN 1 is added for comparison (a is wild type<br />

genotype Tainung 67, b is high amylose genotype TN 1, c is genotype TCSW 1, d is genotype TKW 1, M is marker). The numbers marked in figures match the code numbers listed on<br />

Table 1.


Table 2<br />

Comparisons <strong>of</strong> nucleotide sequences <strong>of</strong> Wx alleles <strong>of</strong> rice genotype Tainung 67<br />

(TNG67) and its NaN3-induced mutants<br />

.exon 1.intron 1.exon 2.<br />

TNG 67<br />

1.50.142.1389.<br />

.(CT)18.T.(ACGGGTTCCAGGGCCTCAAGCCC)1<br />

Nipponbare .(CT)18.T.(ACGGGTTCCAGGGCCTCAAGCCC)1<br />

93-11 .(CT)18.G.(ACGGGTTCCAGGGCCTCAAGCCC)1<br />

TKW 1 .(CT)18.T.(ACGGGTTCCAGGGCCTCAAGCCC)2<br />

TCSW 1 .(CT)16.T.(ACGGGTTCCAGGGCCTCAAGCCC)2<br />

SA 419 .(CT)15.T.(ACGGGTTCCAGGGCCTCAAGCCC)2<br />

SA 402 .(CT)15.T.(ACGGGTTCCAGGGCCTCAAGCCC)2<br />

SA 420 .(CT)19.T.(ACGGGTTCCAGGGCCTCAAGCCC)2<br />

SA 491 .(CT)17.T.(ACGGGTTCCAGGGCCTCAAGCCC)2<br />

SA 492 .(CT)17.T.(ACGGGTTCCAGGGCCTCAAGCCC)2<br />

SA 893 .(CT)17.T.(ACGGGTTCCAGGGCCTCAAGCCC)2<br />

SA 1917 .(CT)17.T.(ACGGGTTCCAGGGCCTCAAGCCC)2<br />

SA 1170 .(CT)16.T.(ACGGGTTCCAGGGCCTCAAGCCC)2<br />

SA 1445 .(CT)17.T.(ACGGGTTCCAGGGCCTCAAGCCC)2<br />

Two naturally mutated waxy genotypes TKW 1 (japonica type) and TCSW 1 (indica<br />

type) were added for comparison. Additionally, the nucleotide sequences <strong>of</strong> genotypes<br />

Nipponbare (japonica type) and 93-11 (indica type) are collected from internet<br />

http://www.ncbi.nlm.nih.gov/.<br />

While apparently required for the mutation, the presence <strong>of</strong> the<br />

single base substitution (G-to-T) at the 5 0 splice site <strong>of</strong> intron 1<br />

alone does not ensure that waxy mutation is expressed (Prathepha<br />

and Baimai, 2004; Yamanaka et al., 2004). Wanchana et al. (2003)<br />

identified a unique duplication <strong>of</strong> 23-bp in the second exon, which<br />

was present only in the waxy rice genotypes. The 23-bp duplication<br />

in exon 2 would result in a frame shift and subsequently cause<br />

a premature stop codon in the second exon and bring about<br />

degradation <strong>of</strong> mRNA by nonsense-mediated decay (Isshiki et al.,<br />

2001). Hori et al. (2007) examined 29 waxy genotypes and a 23-bp<br />

duplication at the second exon was detected in 27 genotypes<br />

(including paddy and upland waxy genotypes), and the two other<br />

waxy genotypes were considered to have a mutation (insertion <strong>of</strong><br />

7764 bp in exon 9) different from the 23-bp duplication. We also<br />

analyzed TKW 1, TCSW 1 and 9 NaN3-induced waxy mutants<br />

derived from Tainung 67 and detected 23-bp duplication in all the<br />

tested waxy mutants (Table 2). As expected, no 23-bp insertion was<br />

detected on non-waxy japonica genotypes Tainung 67, Nipponbare<br />

and indica genotype 93-11. Thus, the findings <strong>of</strong> Wanchana et al.<br />

(2003) are still valid even though the waxy mutation is induced by<br />

chemical mutagen NaN3, at least for these NaN3-induced waxy<br />

mutants derived from genotype Tainung 67.<br />

Polymorphic cytosine and thymidine (CT)n microsatellites are<br />

noted in the 5 0 -untranslated region <strong>of</strong> the Wx gene (Bergmaan et al.,<br />

2001; Prathepha and Baimai, 2004), which seems to correlate well<br />

with the various amylose contents in non-waxy genotypes (Bao<br />

et al., 2002). Bao et al. (2002) examined a set <strong>of</strong> waxy rice genotypes,<br />

and found 4 microsatellite alleles, (CT)16, (CT)17, (CT)18, (CT)19<br />

at the wx locus, <strong>of</strong> which (CT)17 was the most frequent. A similar<br />

conclusion was also reported from other studies (Han et al., 2004;<br />

Prathepha and Baimai, 2004). In the present study, a total <strong>of</strong> 5<br />

classes <strong>of</strong> (CT)n microsatellites, located at 55-bp upstream <strong>of</strong> the<br />

putative 5 0 leader intron splice-junction were detected. Among the<br />

tested waxy mutants, 5 waxy mutants carried (CT)17 allele, 2<br />

mutants contained (CT)15 allele, 2 mutants contained (CT)16 allele,<br />

a mutant had a (CT)18 allele and a mutant had (CT)19 allele (Table 2).<br />

Similar (CT)18 allele were also detected on non-waxy genotypes<br />

TNG 67, Nipponbare and 93-11. It should be pointed out that the<br />

(CT)15 allele was the only allele not detected in other studies (Bao<br />

et al., 2002; Han et al., 2004; Prathepha and Baimai, 2004). CT<br />

repeats are present in the 5 0 -untranslated region and variations <strong>of</strong><br />

the number <strong>of</strong> the CT repeats from 8 to 20 have been report in nonwaxy<br />

genotype (Olsen and Purugganan, 2002). Therefore, CT<br />

T.L. Jeng et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 112–116 115<br />

repeats is not considered to be the cause <strong>of</strong> the waxy trait (Hori<br />

et al., 2007). However, CT repeats might potentially serve as<br />

a molecular marker for other starch qualities <strong>of</strong> waxy rice, such as<br />

G/T-related thermal properties as well as some <strong>of</strong> the pasting<br />

viscosity parameters (Bao et al., 2002).<br />

Direct nucleotide sequencing is labor intensive and costly to<br />

analyze a large sample set. Electrophoretic analysis is known to be<br />

an accurate technique to analyze large numbers <strong>of</strong> samples quickly,<br />

without the need for direct sequencing determination (Yamanaka<br />

et al., 2004). In the present study, all the 35 waxy mutants<br />

(including the accessions used for nucleotide sequence analysis)<br />

and their wild type genotype Tainung 67 plus 2 waxy genotypes<br />

TKW 1 and TCSW 1 that were differentiated by iodine staining<br />

(Fig. 2) were used to identify 23-bp duplication polymorphism<br />

through electrophoresis. A well-known non-waxy indica type<br />

genotype TN 1 (Taichung Native 1) was also added to serve as<br />

a comparison. The electrophoretic results (Fig. 3A) showed that no<br />

23-bp duplication was detected in genotypes Tainung 67 and TN1,<br />

which contain intermediate and high grain amylose, respectively.<br />

As expected, all the waxy accessions (including cultivars TKW 1 and<br />

TCSW 1) had the 23-bp insertion in the second exon (Fig. 3A). This<br />

premature stop codon due to the 23-bp insersion would cause the<br />

loss <strong>of</strong> function <strong>of</strong> GBSS encoded by the Wx gene (Sato and Nishio,<br />

2003). For further verifying the functional role <strong>of</strong> the 23-bp<br />

duplication, an electrophoresis on a NaN3-induced non-waxy<br />

recovery line <strong>of</strong> SA 419 was conducted and the results indicated<br />

that no 23-bp duplication at the second exon was found for this<br />

non-waxy recovery line <strong>of</strong> SA 419 (Fig. 4). This result clearly indicates<br />

that the 23-bp duplication at the second exon is essential for<br />

NaN3-induced waxy mutation, at least for these mutants derived<br />

from genotype Tainung 67.<br />

Electrophoresis <strong>of</strong> G-to-T polymorphism indicated that only<br />

indica genotype TN 1 having high grain amylose content had AGGT<br />

sequence on intron 1 splice site (Fig. 3B). On the other hand, all the<br />

tested waxy mutants, including TKW 1 and TCS 1, as well as their<br />

wild type TNG 67 contained the AGTT sequence at the 5 0 splice site<br />

<strong>of</strong> the first intron (Fig. 3B). Moreover, variations in the length <strong>of</strong> CTmicrosatellite<br />

in the 5 0 -untranslated region <strong>of</strong> the wx gene (Bao<br />

et al., 2002; Prathepha and Baimai, 2004) was also obtainable in the<br />

tested accessions (Fig. 3C).<br />

In conclusion, the present results indicate that significant<br />

differences in grain appearance are observed among the tested<br />

genotypes and NaN3-induced waxy mutants. Nucleotide polymorphisms<br />

also exist among the tested mutants. G/T polymorphism<br />

analysis indicates that all the NaN3-induced waxy<br />

mutants carry Wx b allele. However, the presence <strong>of</strong> the single base<br />

substitution (G-to-T) at the 5 0 splice site <strong>of</strong> intron 1 alone does not<br />

ensure that waxy mutation is expressed, because both wild type<br />

genotype Tainung 67 and non-waxy genotype Nipponbare (both<br />

are genotypes with intermediate grain amylose level) also shows<br />

same G/T substitution at 5 0 splice site <strong>of</strong> intron 1. Additionally, at<br />

least 5 classes <strong>of</strong> (CT)n microsatellites at the Wx locus are identified<br />

among the tested waxy mutants. However, the CT repeats are also<br />

not the cause <strong>of</strong> the waxy trait. On the other hand, a 23-bp<br />

Fig. 4. Nucleotide polymorphisms <strong>of</strong> 23-bp duplication on Wx alleles <strong>of</strong> (1) mutant SA<br />

419 and its (2) non-waxy recovery.


116<br />

duplication in the second exon is present in all the NaN3-induced<br />

waxy mutants. This 23-bp duplication appears to be necessary for<br />

the waxy trait in genotype Tainung 67. Thus, the 23-bp duplication<br />

in exon 2 may be used as a molecular marker to characterize waxy<br />

grain trait in rice breeding program.<br />

Acknowledgements<br />

We gratefully acknowledge financial support from the National<br />

<strong>Science</strong> Council <strong>of</strong> ROC.<br />

References<br />

Bao, J.S., Corke, H., Sun, M., 2002. Microsatellites in starch-synthesizing genes in<br />

relation to starch physicochemical properties in waxy rice (Oryza sativa L.).<br />

Theoretical and Applied Genetics 105, 898–905.<br />

Bao, J.S., Corke, H., Sun, M., 2006. Microsatellites, single nucleotide polymorphisms<br />

and a sequence tagged site in starch-synthesizing genes in relation to starch<br />

physicochemical properties in non-waxy rice (Oryza sativa L.). Theoretical and<br />

Applied Genetics 113, 1185–1196.<br />

Bergmaan, C.J., Delgado, J.T., McClung, A.M., Fjellstrom, R.G., 2001. An improved<br />

method for using a microsatellite sequence closely linked to the waxy gene to<br />

determine amylose class. <strong>Cereal</strong> Chemistry 78, 257–260.<br />

Doyle, J.J., Doyle, J.L., 1990. Isolation <strong>of</strong> plant DNA from fresh tissue. Focus 12, 13–15.<br />

Han, Y., Xu, M., Liu, X., Yan, C., Korban, S.S., Chen, X., Gu, M., 2004. Genes coding for<br />

starch branching enzymes are major contributors to starch viscosity characteristics<br />

in waxy rice (Oryza sativa L.). Plant <strong>Science</strong> 166, 357–364.<br />

Hori, Y., Fujimoto, R., Sato, Y., Nishio, T., 2007. A novel wx mutation caused by<br />

insertion <strong>of</strong> a retrotransposon-like sequence in a glutinous cultivar <strong>of</strong> rice<br />

(Oryza sativa). Theoretical and Applied Genetics 115, 217–224.<br />

Isshiki, M., Yamamoto, Y., Satoh, H., Shimamoto, K., 2001. Nonsense-mediated decay<br />

<strong>of</strong> mutant waxy mRNA inrice. Plant Physiology 125, 1388–1395.<br />

Isshiki, M., Morino, K., Nakajima, M., Okagaki, R.J., Wessier, S.R., Izawa, T.,<br />

Shimamoto, K., 1998. A naturally occurring functional allele <strong>of</strong> the rice waxy<br />

locus has a GT to TT mutation at the 5 0 splice site <strong>of</strong> the first intron. Plant <strong>Journal</strong><br />

15, 133–138.<br />

T.L. Jeng et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 112–116<br />

Itoch, K., Ozaki, H., Okada, K., Hori, H., Takeda, Y., Mitsui, Y., 2003. Introduction <strong>of</strong><br />

Wx transgene into rice wx mutants lead to both high- and low-amylose rice.<br />

Plant and Cell Physiology 44, 473–480.<br />

James, M.G., Denyer, K., Meyers, A.M., 2003. Starch synthesis in the cereal<br />

endosperm. Current Opinions in Plant Biology 6, 215–222.<br />

Jeng, T.L., Wang, C.S., Yseng, T.H., Sung, J.M., 2007. Expression <strong>of</strong> granule-bound<br />

starch synthase in developing rice grain. <strong>Journal</strong> <strong>of</strong> the <strong>Science</strong> <strong>of</strong> Food and<br />

Agriculture 87, 2456–2463.<br />

Jeng, T.L., Tseng, T.H., Wang, C.S., Chen, C.L., Sung, J.M., 2003. Starch biosynthesizing<br />

enzymes in developing grains <strong>of</strong> rice cultivar Tainung 67 and its sodium-azideinduced<br />

rice mutant. Field Crops Research 84, 261–266.<br />

Jeng, T.L., Tseng, T.H., Wang, C.S., Chen, C.L., Sung, J.M., 2006. Yield and grain<br />

uniformity in contrasting rice genotypes suitable for different growth environment.<br />

Field Crops Research 99, 59–66.<br />

Olsen, K.M., Purugganan, M.D., 2002. Molecular evidence on the origin and<br />

evolution <strong>of</strong> glutinuous rice. Genetics 162, 941–950.<br />

Prathepha, P., Baimai, V., 2004. Variation <strong>of</strong> Wx microsatellite allele, waxy allele<br />

distribution and differentiation <strong>of</strong> chloroplast DNA i n a collection <strong>of</strong> Thai rice<br />

(Oryza sativa L. Euphytica 140, 231–237.<br />

Ramesh, M., Ali, S.Z., Bhattacharya, K.R., 1999. Structure <strong>of</strong> rice starch and its<br />

relation to cooked-rice texture. Carbohydrate Polymers 38, 337–347.<br />

Sato, Y., Nishio, T., 2003. Mutation detection in rice waxy mutants by PCR-RF-SSCP.<br />

Theoretical and Applied Genetics 107, 560–567.<br />

Vandeputte, G.E., Delcour, J.A., 2004. From sucrose to starch granule to starch<br />

physical behaviour: a focus on rice starch. Carbohydrate Polymers 58, 245–266.<br />

Wanchana, S., Toojinda, T., Tragoonrung, S., Vanavichit, A., 2003. Duplicated coding<br />

sequence in the waxy allele <strong>of</strong> tropical glutinuous rice (Oryza sativa L.). Plant<br />

<strong>Science</strong> 165, 1193–1199.<br />

Wang, Y.J., Wang, L., 2002. Structures <strong>of</strong> four waxy rice starches in relation to<br />

thermal, pasting and textural properties. <strong>Cereal</strong> Chemistry 79, 252–256.<br />

Wang, C.-T., Tseng, T.-H., Lin, C.-Y., 2002. Rice biotech research at the Taiwan<br />

Agricultural Research Institute. Asia Pacific Biotechnology News 6, 950–956.<br />

Wang, Z.-Y., Zheng, F.-Q., Shen, G.-Z., Gao, J.-P., Snutad, D.P., Li, M.-G., Zhang, J.-L.,<br />

Hong, M.-M., 1995. The amylose content in rice endosperm is related to the<br />

post-transcriptional regulation <strong>of</strong> the waxy gene. Plant <strong>Journal</strong> 7, 613–622.<br />

Yamanaka, S., Nakamura, I., Watanabe, K.N., Sato, Y.-I., 2004. Identification <strong>of</strong> SNPs<br />

in the waxy gene among glutinuous rice cultivars and their evolutionary<br />

significance during the domestication process <strong>of</strong> rice. Theoretical and Applied<br />

Genetics 108, 1200–1204.<br />

Zhou, Z., Robards, K., Helliwell, S., Blanchard, C., 2002. Ageing <strong>of</strong> stored rice: changes<br />

in chemical and physical attributes. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 35, 65–78.


Amaranth (Amaranthus hypochondriacus) as an alternative crop for sustainable<br />

food production: Phenolic acids and flavonoids with potential impact on its<br />

nutraceutical quality<br />

A.P. Barba de la Rosa a, *, Inge S. Fomsgaard b , Bente Laursen b , Anne G. Mortensen b , L. Olvera-Martínez c ,<br />

C. Silva-Sánchez a , A. Mendoza-Herrera d , J. González-Castañeda e ,A.DeLeón-Rodríguez a<br />

a<br />

Institute for Scientific and Technological Research at San Luis Potosi, Molecular Biology Division, Camino a La Presa San Jose No. 2055,<br />

Lomas 4 a sección, San Luis Potosí, SLP, Mexico<br />

b<br />

Department <strong>of</strong> Integrated Pest Management, Faculty <strong>of</strong> Agricultural <strong>Science</strong>s, Aarhus University, Denmark<br />

c<br />

CBETa 196 Villa de Pozos, San Luis Potosí, SLP, Mexico<br />

d<br />

Horticultural <strong>Science</strong> Department, HFSB Room 202, 2133 TAMU, College Station, United States<br />

e<br />

Instituto de Ciencias Agrícolas, Universidad de Guanajuato, Irapuato, Gto, Mexico<br />

article info<br />

Article history:<br />

Received 7 May 2008<br />

Received in revised form 4 July 2008<br />

Accepted 10 July 2008<br />

Keywords:<br />

Crop yield<br />

LC/MS/MS<br />

Protein content<br />

Phytochemicals<br />

RAPD<br />

1. Introduction<br />

abstract<br />

The challenge for agricultural practices to increase food<br />

production to obtain food security still persists after 40 years <strong>of</strong> the<br />

Green Revolution (Hobbs, 2007). The first Millennium Development<br />

Goal is to reduce hunger and poverty by 2015 (Dixon et al.,<br />

2006). The demand for food is increasing, not only because <strong>of</strong> the<br />

growing population, but also to provide more nutritious food with<br />

high protein quality and nutraceutical compounds.<br />

Water resources, especially surface and ground water will be<br />

more limited as domestic and industrial needs increase just as it is<br />

limited in semi-desert zones with low precipitation, and so future<br />

crops must be more suited to low water use. Amaranth (Amaranthus<br />

hypochondriacus) is a crop naturally resistant to water deficit<br />

and is a good source <strong>of</strong> nutritious seeds; the seeds have high<br />

* Corresponding author. Tel.: þ52 4448342000; fax: þ52 4448342010.<br />

E-mail address: apbarba@ipicyt.edu.mx (A.P. Barba de la Rosa).<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.07.012<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 117–121<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

The demand for food is increasing, not only to meet food security for growing populations, but also to<br />

provide more nutritious food, rich in good quality proteins and nutraceutical compounds. The amaranth<br />

(Amaranthus hypochondriacus) plant, in addition to its high nutritive and nutraceutical characteristics,<br />

has excellent agronomic features. The objective <strong>of</strong> the present study was to analyze some physical and<br />

proximal-nutritional properties <strong>of</strong> amaranth seeds obtained from different varieties grown in arid zones<br />

and characterize their phenolic acids and flavonoids. Two commercial (Tulyehualco and Nutrisol) and<br />

two new (DGETA and Gabriela) varieties <strong>of</strong> A. hypochondriacus were grown at the Mexican Highlands<br />

zone. Tulyehualco and DGETA varieties had higher seed yield <strong>of</strong> 1475 and 1422 kg ha 1 , respectively,<br />

comparable to corn and soybean production in agricultural areas. Gabriela had the highest protein<br />

content <strong>of</strong> 17.3%, but all varieties had an adequate balance <strong>of</strong> essential amino acids. Polyphenols as rutin<br />

(4.0–10.2 mgg 1 flour) and nicotiflorin (7.2–4.8 mgg 1 flour) were detected. Amaranth can be cultivated<br />

in arid zones where commercial crops cannot be grown; the seeds besides their well known nutritive<br />

characteristics could be a source <strong>of</strong> phenolic compounds <strong>of</strong> high antioxidant properties.<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

amounts <strong>of</strong> protein containing essential amino acid such as lysine<br />

and methionine; also significant levels <strong>of</strong> squalene, an important<br />

precursor for all steroids (He et al., 2002), were reported in<br />

amaranth oil. Since the beginning <strong>of</strong> the 1980s, amaranth has been<br />

rediscovered and several reports have tried to promote it as a basic<br />

crop (Kauffman, 1992). In addition to nutritional characteristics,<br />

amaranth plants have agronomic features identifying it as an<br />

alternative crop where cereals and vegetables cannot be grown (dry<br />

soils, high altitudes and high temperatures) (Omami et al., 2006). In<br />

general, the selection <strong>of</strong> promising genotype in a breeding program<br />

is based on various criteria, with the most important being final<br />

crop yield and seed quality (Kozak et al., 2008).<br />

The aim <strong>of</strong> the research presented in this paper was to analyze<br />

four amaranth varieties and select the most promising for growing<br />

in arid zones like the Mexican Highlands or arid zones <strong>of</strong> Southern<br />

Europe. Variety selection was carried out on the basis <strong>of</strong> seed yield<br />

and proximal quality. The presence <strong>of</strong> a range <strong>of</strong> phytochemical<br />

compounds that contribute to nutraceutical properties <strong>of</strong> the seeds<br />

was determined as well.


118<br />

2. Materials and methods<br />

2.1. Plant material<br />

Two commercial varieties (Tulyehualco and Nutrisol) and two<br />

new varieties (DGETA and Gabriela) <strong>of</strong> A. hypochondriacus were<br />

cultivated at the CBTa 196 Villa de Pozos in San Luis Potosi, under<br />

low watering conditions, precipitations between 25 and 35 mm.<br />

The experiment was in randomized complete blocks with five<br />

blocks per variety. Seeds were drilled manually in four rows, each<br />

6 m long, 80 cm apart and with 30 cm spacing in the row<br />

(Henderson et al., 2000). Plants were watered at 35% <strong>of</strong> field<br />

capacity, at the beginning <strong>of</strong> soil preparation (April 2002) and in<br />

August 2002. After stem elongation was complete (November<br />

2002), hand-harvest was carried out at maturity by cutting plants<br />

at soil level.<br />

2.2. Physical analysis<br />

Measurements <strong>of</strong> morphologic parameters were taken from 10<br />

randomly chosen plants. Plant height was measured from the<br />

ground to the top <strong>of</strong> the inflorescence head. Spike length was<br />

measured from the base to the top <strong>of</strong> inflorescence. For seed yield,<br />

all plants from each block were hand-harvest, seeds were cleaned<br />

and weighted, and yield was calculated and extrapolated to kg ha 1 .<br />

A sample <strong>of</strong> 100 seeds per triplicate was weighted and extrapolated<br />

to obtain the weight <strong>of</strong> 1000 seeds. The seeds were milled to obtain<br />

the flour able to pass through a 100-mesh screen and flour yield<br />

was determined as the percentage <strong>of</strong> flour recovery. The flours<br />

were defatted with hexane at a flour/hexane ratio <strong>of</strong> 1:10 (w/v)<br />

under continuous stirring for 4 h at 4 C, the slurry was centrifuged<br />

at 9000 g for 20 min and the flour was air-dried at room temperature<br />

and stored at 4 C(Barba de la Rosa et al., 1992).<br />

2.3. Chemical analysis<br />

Total nitrogen (method 954.01), fat (920.39), ash (923.03), crude<br />

fiber (962.09), and moisture (925.09) contents <strong>of</strong> the flours were<br />

determined according to AOAC (1990). Nitrogen was determined<br />

with a Kjeltec system (Tecator, Sweden). Protein was calculated<br />

from total nitrogen using a factor <strong>of</strong> 5.85 (Barba de la Rosa et al.,<br />

1992). Fat amount was obtained from 4 h hexane extraction. Ash<br />

was calculated from the weight remaining after heating the sample<br />

at 550 C for 2 h. Moisture measurement was determined from<br />

sample weight loss after oven drying at 110 C for 4 h. All samples<br />

were analyzed in triplicate.<br />

2.4. Essential amino acid analysis<br />

Amino acid analysis <strong>of</strong> amaranth flours was performed by<br />

reversed-phase high performance liquid chromatography (RP-HPLC)<br />

using a HPLC Waters 600 with a fluorescence detector 2475 (Waters)<br />

and performed by AccQ-Tag amino acid analysis kit and following<br />

the manual’s instructions (Waters). Samples <strong>of</strong> 5 mg <strong>of</strong> amaranth<br />

flours were hydrolyzed with 200 mL <strong>of</strong> 6 N HCl for 24 h at 110 C in<br />

vacuo system; one or two phenol crystals were added as an oxygen<br />

scavenger (Barba de la Rosa et al., 1992). The samples were<br />

neutralized with 1.2 N NaOH and dehydrated. Subsequently, samples<br />

were derivatized with 6-aminoquinolyl-N-hydroxysuccinimidyl<br />

carbamate. A 5 mL sample was applied to the column AccQ-Tag C18<br />

(Waters). All samples were analyzed in triplicate.<br />

2.5. Phytochemical analysis<br />

The freeze-dried amaranth flour samples were crushed and<br />

homogenized before phytochemicals were extracted using an<br />

A.P. Barba de la Rosa et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 117–121<br />

Accelerated Solvent Extraction 200 system (Dionex) (ASE),<br />

following the protocol reported by Carlsen et al. (2008). Briefly,<br />

0.1 g <strong>of</strong> the freeze-dried and homogenized sample was transferred<br />

to the 33 mL extraction cells followed by 5-g <strong>of</strong> chemically inert<br />

Ottawa sand (20–30 mesh particle size, Fisher Chemicals) after<br />

heating at 400 C. A filter was placed on top <strong>of</strong> the sample, and the<br />

extraction cell was filled with glass balls that had been heated at<br />

400 C. The eluent was 70% methanol, 30% water (v/v). The protocol<br />

for the ASE extraction was the following: preheat for 5 min, heat for<br />

5 min, static for 3 min, flush 80%, purge for 60 s, 4 cycles, pressure<br />

1500 psi, and temperature 80 C. Phytochemical extracts were<br />

collected in vials, which were filled with eluent to maintain 46.1 g<br />

( ¼ 50 mL) weight for all extracts, and stored at 20 C until<br />

chemical analysis. The extracts were filtered on a Sartorius SRP 15<br />

0.45-m filter (PTFE membrane) and diluted with water in a ratio<br />

<strong>of</strong> 1:1.<br />

Phytochemical analysis was carried out using an Applied<br />

Biosystems MDS Sciex API3200 Liquid chromatography-Ion Trap<br />

Quadrupole Mass Spectrometer (LC/MS/MS) with turbo electrospray<br />

ionization in a negative multiple reaction monitoring (MRM)<br />

mode. Pure reference compounds were used for identification <strong>of</strong><br />

the phenolic acids and polyphenols based on a comparison <strong>of</strong><br />

fragmentation pattern and retention times. Standard curves for<br />

commercially obtained pure standards were prepared in 50%<br />

methanol and 50% water (v/v) at the following concentrations: 3.13,<br />

6.25, 12.5, 25.0, 50.0, 100, 200, 400, and 800-mgL 1 . The standard<br />

curve was used for the quantitative determination. The analyzed<br />

compounds were three polyphenols, rutin (CAS no. 153-18-4,<br />

Extrasynthese, France), isoquercitrin (CAS no. 482-35-9, Fluka,<br />

Germany), and nicotiflorin (CAS no. 17650-84-9, Carl Roth,<br />

Germany), and three phenolic acids, vanillic acid (CAS no. 121-34-6,<br />

Fluka, Germany), 4-hydroxybenzoic acid (CAS no. 99-96-7, Fluka,<br />

Germany), and syringic acid (CAS no. 530-57-4, Sigma, Germany).<br />

2.6. Electrophoretic patterns<br />

Protein fractions were obtained by following the procedure<br />

described elsewhere with some modifications (Barba de la Rosa<br />

et al., 1992). Briefly, for albumin fraction extraction, a suspension <strong>of</strong><br />

flour/distilled water (1:10 w/v) was prepared and stirred for 1 h at<br />

4 C and centrifuged at 9000 g for 20 min at 4 C. The resulting<br />

pellet was re-suspended in 10 mM Na2HPO4, pH 7.5, 1 mM EDTA,<br />

0.1 M NaCl, stirred and centrifuged as before, the supernatant was<br />

named as globulins 7S. The pellet was again re-suspended in<br />

phosphate buffer but containing 0.8 M NaCl; the soluble fraction<br />

was named globulins 11S. The glutelin fraction was extracted with<br />

0.1 M NaOH. The protein patterns were analyzed by SDS-PAGE at<br />

15% w/v polyacrylamide in a Mini-Protean III system (Bio-Rad).<br />

Samples <strong>of</strong> 1 mg mL 1 were dissolved in Laemmli sample buffer<br />

(Laemmli, 1970). Electrophoresis was conducted at a constant<br />

current <strong>of</strong> 20 mA per gel for 2–3 h. After electrophoresis, the gel<br />

was stained with Coomassie Brilliant Blue G250 at a final concentration<br />

<strong>of</strong> 0.25%. Destaining was achieved by washing the gel for<br />

2–4 h with acetic acid/methanol/water (4.5:4.5:1 v/v/v).<br />

2.7. RAPD analysis<br />

DNA <strong>of</strong> the four cultivars was extracted from fresh young leaves<br />

by a standard procedure. DNA quantification was by measurement<br />

<strong>of</strong> absorbance at 260/280 nm. The extracted DNA was dissolved in<br />

double distilled water (Dellaporta et al., 1983). For amplification,<br />

the pair <strong>of</strong> oligonucleotides denominated as 23: 5 0 -CCC GCC TCC<br />

C-3 0 and 43: 5 0 -AAA ACC GGG C-3 0 were used as reported by Chan<br />

and Sun (1997) and following the protocol described by those<br />

authors: 3 mM MgCl2, 0.1% Triton X-100, 75 ng primer, and 30 ng<br />

genomic DNA was used in 25 mL reaction mix. PCR amplification


Table 1<br />

Physical properties <strong>of</strong> amaranth seeds<br />

Variety Plant height<br />

(cm)<br />

Tulyehualco 159.0 a<br />

DGETA 154.2 b<br />

Gabriela 149.1 b<br />

Nutrisol 161.3 a<br />

was performed in a iCycler (Bio-Rad) using the following cycle<br />

pr<strong>of</strong>ile: 1 cycle at 94 C for 2 min followed by 45 cycles at 94 C for<br />

1 min, 35 C for 2 min, and 72 C for 2 min; and the final cycle at<br />

72 C for 7 min. The amplification products were electrophoresed<br />

on 1.4% agarose gels and visualized using ethidium bromide<br />

staining. Negative controls were routinely used to check for<br />

possible contamination. For RAPD analysis, the banding patterns<br />

were recorded in a GelDoc photodocumentator (Bio-Rad) and gels<br />

analyzed with Quantity-one s<strong>of</strong>tware (Bio-Rad). The bands with the<br />

same molecular weight and mobility were treated as identical<br />

fragments; the presence <strong>of</strong> a band was coded as 1, and absence<br />

coded as 0. The data matrices were analyzed with the Squared<br />

Euclidean nearest neighbor method (Statgraphics Plus v5.0<br />

s<strong>of</strong>tware, Statistical Graphics Corp). Dendrograms were produced<br />

from the results.<br />

2.8. Statistical analysis<br />

Spike length<br />

(cm)<br />

Statistical analysis was performed and Tukey’s test was used to<br />

determine significance <strong>of</strong> differences among means. Significance<br />

was when compared means differed at P 0.05 (Reyes-Castañeda,<br />

1983).<br />

3. Results and discussion<br />

Seed yield<br />

(kg ha 1 )<br />

68.0 a<br />

1475 a<br />

54.5 a<br />

1422 a<br />

28.2 b<br />

1204 b<br />

nd 1121 b<br />

3.1. Physical properties <strong>of</strong> amaranth seeds<br />

Weight <strong>of</strong> Flour yield<br />

1000 seeds (%)<br />

(g)<br />

0.6 a<br />

87.3 a<br />

0.7 a<br />

89.1 a<br />

0.7 a<br />

87.6 a<br />

0.6 a<br />

88.2 a<br />

Means <strong>of</strong> three replicates within in the same column (physical property) with<br />

different letter are significantly different at P 0.05.<br />

The physical properties <strong>of</strong> amaranth plant and seeds are shown<br />

in Table 1. Tulyehualco and Nutrisol plants were the highest (161.3<br />

and 159.0 cm, respectively), but Tulyehualco had the longest spike<br />

(68 cm). Tulyehualco and DGETA had the highest seed yields <strong>of</strong><br />

1475 and 1422 kg ha 1 , respectively. When grown in drier environments,<br />

amaranth seeds yields <strong>of</strong> 1050 and 410 kg ha 1 were<br />

reported (Henderson et al., 2000). The amaranth seed yields<br />

obtained in our work compare favorably to commercially important<br />

crops like soybean (1532 kg ha 1 ) and maize (2315 kg ha 1 )<br />

(SAGARPA, 2002). There were no differences in the weight <strong>of</strong> 1000<br />

seeds among varieties. Flour yield is an important quality parameter<br />

<strong>of</strong> wheat, and since flour is an alternative use <strong>of</strong> amaranth<br />

seeds, this parameter was also measured. There were no differences<br />

among the varieties with a mean value <strong>of</strong> 88% (Table 1), values that<br />

are similar to wheat flour yields (Paredes-López et al., 1990). In<br />

Central Mexico, amaranth grain is paid at double the price <strong>of</strong> maize<br />

Table 2<br />

Proximal composition <strong>of</strong> amaranth seeds (% dry basis)<br />

Variety Protein* Lipids Ash Fiber<br />

Tulyehualco 15.0 b<br />

8.1 a<br />

3.3 a<br />

2.2 a<br />

DGETA 14.8 b/c<br />

7.9 a/b<br />

3.3 a<br />

1.9 b<br />

Gabriela 17.3 a<br />

8.9 a<br />

3.9 a<br />

2.5 a<br />

Nutrisol 15.3 b<br />

8.6 a<br />

3.5 a<br />

2.0 a/b<br />

*N 5.85. Means <strong>of</strong> triplicates in the same column with different letter are significantly<br />

different at P 0.05.<br />

A.P. Barba de la Rosa et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 117–121 119<br />

Table 3<br />

Essential amino acid composition <strong>of</strong> different amaranth variety flours (g <strong>of</strong> amino<br />

acid 100 g 1 <strong>of</strong> crude protein)<br />

Amino acid Tulyehualco DGETA Gabriela Nutrisol FAO/WHO/UNU (1986)<br />

Ile 3.2 2.2 3.1 3.5 1.3<br />

Leu 5.8 5.0 6.5 6.6 1.9<br />

Lys 6.9 5.2 7.5 6.7 1.6<br />

Met þ Cys a<br />

0.3 1.7 6.5 7.2 1.7<br />

Phe þ Tyr b<br />

7.9 3.7 8.3 7.3 1.9<br />

Thr 3.8 3.3 4.8 4.6 0.9<br />

Val 3.4 2.6 3.6 3.9 1.3<br />

a<br />

Requirements for methionine þ cysteine.<br />

b<br />

Requirements for phenylalanine þ tyrosine.<br />

grain. In this sense, farmers could have double benefits, seeds <strong>of</strong><br />

high nutritional quality and higher incomes for the rural families<br />

(Islas-Gutiérrez and Islas-Gutiérrez, 2001).<br />

3.2. Proximal composition <strong>of</strong> amaranth<br />

Amaranth protein content (Table 2) was higher than values<br />

reported for cereals (Paredes-López et al., 1990). The variety<br />

Gabriela had the highest seed protein content (17.3%) followed by<br />

Tulyehualco, Nutrisol and DGETA (15.0%, 15.3%, and 14.8%, respectively).<br />

The fat contents were similar among the varieties; however,<br />

at higher proportions (7–9%), than for cereals such as wheat (2.1%)<br />

and maize (4.5%). There were no differences in ash contents<br />

between varieties, but Gabriela had the highest fiber content, 2.5%<br />

(Table 2). Amaranth has been considered as a new source <strong>of</strong> fiber<br />

and methods for fractionation to get high-fiber amaranth products<br />

have been reported (Tosi et al., 2001). The essential amino acids in<br />

amaranth seeds (Table 3) are ideal according to FAO requirements<br />

for adults (FAO/WHO/UNU, 1986).<br />

3.3. Phytochemicals in amaranth seed flours<br />

It has been reported that amaranth seed flour contains polyphenols<br />

(flavonoids) with relatively high antioxidant status. For<br />

this reason, amaranth has been recommended for use in balanced<br />

diets (Gorinstein et al., 2007). However, to the best <strong>of</strong> our knowledge,<br />

no results have ever been published before on the types <strong>of</strong><br />

polyphenols present in amaranth seed flours. In this study, three<br />

polyphenols (rutin, isoquercitrin, and nicotiflorin), were identified<br />

and quantified (Table 4). Rutin was present at higher concentrations<br />

(10.1 mg g 1 flour) in Tulyehualco seed flour, while the<br />

highest amount <strong>of</strong> nicotiflorin (7.2 mg g 1 flour) was found in the<br />

Gabriela variety (Table 4). Rutin has been reported to be present in<br />

amaranth leaves (Suryavanshi et al., 2007), but this is the first time<br />

it has been reported in seed flours.<br />

Polyphenols in which sugar groups are beta-linked, as in the<br />

case <strong>of</strong> the three polyphenols found in this study, are easily<br />

degraded in the intestine <strong>of</strong> humans and animals due to the<br />

abundance <strong>of</strong> beta-glucosidase enzyme that liberates the aglyconic<br />

Table 4<br />

Phenolic acids and flavonoids present in different amaranth variety flour<br />

(mg metabolite g 1 flour) a<br />

Metabolite Tulyehualco DGETA Gabriela Nutrisol<br />

Isoquercitrin 0.5 0.5 0.3 nd<br />

Nicotiflorin 5.5 5.6 7.2 4.8<br />

Rutin 10.1 5.8 4.0 4.7<br />

4-Hydroxybenzoic acid 1.7 2.0 2.2 1.9<br />

Syringic acid 0.8 0.7 nd nd<br />

Vanillic acid 1.8 1.7 1.8 1.5<br />

a Mean <strong>of</strong> two replicates, nd ¼ not detected.


120<br />

moiety <strong>of</strong> the molecules. Several health effects have been published,<br />

which generally concern the uptake <strong>of</strong> the aglyconic groups,<br />

quercetin and kaempferol (Donovan et al., 2007). Rutin and its<br />

metabolites may effectively modulate advanced glycation end<br />

product (AGE) formation, which is associated with numerous<br />

pathologies. Clinical diseases possibly accelerated by AGEs include<br />

neuropathy, nephropathy, retinopathy, joint stiffness, senile cataracts,<br />

Alzheimer’s disease, and cardiovascular disease (Cervantes-<br />

Laurean et al., 2006). Polyphenols such as quercetin have been<br />

shown to serve as a protective defense against oxidative damage in<br />

vivo (Meyers et al., 2008). Nicotiflorin has been claimed to have<br />

protective effects on reducing memory dysfunction (Huang et al.,<br />

2007); recent results have proved a strong pharmacological basis<br />

for its potential therapeutic role in cerebral ischemic illness (Li<br />

et al., 2006). Phenolic compounds were proved to have antioxidant<br />

activity, where rutin, vanillic acid, ferulic acid, and quercetin have<br />

been reported in rice wine samples and correlated with antioxidant<br />

activities (Que et al., 2006).<br />

In relation to phenolic acids, one paper was published in which<br />

several phenolic acids were identified and quantified in amaranth<br />

seeds using HPLC-UV analysis (Klimczak et al., 2002). In our study,<br />

using LC/MS/MS, three phenolic acids; vanillic acid; 4-hydrozybenzoic<br />

acid, and syringic acid, were identified and quantified (Table 4). The<br />

levels <strong>of</strong> syringic acid and vanillic acid are similar to those reported in<br />

rye where syringic acid has been related with their bitterness (Heiniö<br />

et al., 2008).<br />

3.4. Electrophoretic patterns <strong>of</strong> seed storage amaranth proteins<br />

Electrophoretic patterns <strong>of</strong> the different protein fractions in<br />

amaranth seeds were obtained under denaturing conditions. The<br />

albumin fraction was the main fraction in amaranth proteins; all<br />

varieties had a characteristic band at 34 kDa (Fig. 1A) as reported by<br />

Barba de la Rosa et al. (1992). The albumin fraction was characterized<br />

by a group <strong>of</strong> molecular weight around 18 kDa; this group <strong>of</strong><br />

proteins is known as the MRPs (methionine rich proteins), proteins<br />

that contain 16–18% <strong>of</strong> methionine (Segura-Nieto et al., 1994).<br />

Another group <strong>of</strong> proteins in the albumins fraction was located at<br />

40–80 kDa. Albumins are comparable with egg-white proteins, and<br />

this protein fraction has been used as an egg substitute to prepare<br />

bread (Silva-Sánchez et al., 2004).<br />

There are two groups <strong>of</strong> globulin seed storage proteins, the 7S<br />

globulins were extracted with 0.1 M NaCl (Fig. 1B) and had a main<br />

band around 38 kDa. Globulin fractions extracted with 0.8 M NaCl<br />

(Fig. 1C) showed the characteristic pattern <strong>of</strong> 11S-like globulins, the<br />

group <strong>of</strong> acid polypeptides (35–38 kDa) and the group <strong>of</strong> basic<br />

polypeptides (22–25 kDa). The band <strong>of</strong> high molecular weight<br />

(55 kDa) was previously reported as a globulin precursor (Barba de<br />

la Rosa et al., 1992). Recently we reported that the 11S globulin<br />

A.P. Barba de la Rosa et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 117–121<br />

Fig. 1. Electrophoretic patterns <strong>of</strong> amaranth proteins fractions. A) albumins, B) 7S globulins, C) 11S globulins, D) glutelins. Lane S ¼ molecular weight standard; lane 1 ¼ Tulyehualco;<br />

lane 2 ¼ DGETA; lane 3 ¼ Gabriela; lane 4 ¼ Nutrisol. The arrow shows the differential band found in glutelin fraction.<br />

fraction is rich in peptides <strong>of</strong> angiotensin converting enzyme<br />

inhibitor (Silva-Sánchez et al., 2008).<br />

The amaranth glutelin pattern (Fig. 1D) had high similarity with<br />

11S globulins. There are three main polypeptides groups, one at<br />

22–25 kDa, the second at 35–38 kDa, and the third around 55 kDa.<br />

In this fraction, there was an additional band at 65 kDa in the<br />

Tulyehualco and DGETA varieties; these varieties were also those<br />

with the higher seed yields. Abugouch et al. (2003) suggested that<br />

glutelins are globulin-like proteins with hexameric structure <strong>of</strong><br />

approximately 300 kDa, and it is known that the amaranth globulin<br />

sequence has high homology with the Cupin domain that has been<br />

associated with resistance to extremes <strong>of</strong> environment, as heat,<br />

sulfur nutrition, water stress (Higashi et al., 2006; Khuri et al.,<br />

2001). Former studies by LC/MS/MS analysis showed that the<br />

glutelin fraction is rich in antihypertensive, and also contains the<br />

anticarcinogenic lunasin-like peptide (Silva-Sánchez et al., 2008).<br />

3.5. RADP analysis<br />

Morphological variations cause confusion or misidentification <strong>of</strong><br />

varieties <strong>of</strong> the same species so that genetic analyses could be<br />

necessary for correct identification. In comparative studies using<br />

RAPD, RFLP and/or allozyme markers, the most valuable tool for<br />

correct identification <strong>of</strong> genetic variations was RAPD (Chan and<br />

Sun, 1997). Four different varieties <strong>of</strong> amaranth were studied in this<br />

work; the amplified fragments were analyzed and were grouped,<br />

generating a dendogram (Fig. 2). DGETA and Gabriela are the closer<br />

varieties and probably both derived from Nutrisol. The data indicate<br />

Fig. 2. Dendrograms <strong>of</strong> Amaranth hypochondriacus varieties based on RAPD’s analysis<br />

using the Squared Euclidean nearest neighbor method from Statgraphics s<strong>of</strong>tware.<br />

Lane 1 ¼ Tulyehualco; lane 2 ¼ DGETA; lane 3 ¼ Gabriela; lane 4 ¼ Nutrisol; lane<br />

5 ¼ control.


that Tulyehualco variety is the one most distantly related to the<br />

other three.<br />

4. Conclusions<br />

As interest in amaranth cultivation has increased and breeders<br />

have produced a large number <strong>of</strong> new varieties adapted to different<br />

zones (Gimplinger et al., 2008; Henderson et al., 2000). However,<br />

some <strong>of</strong> these new varieties are only new names for old varieties.<br />

RAPD analysis showed that the old variety, Tulyehualco, could be<br />

the progenitor <strong>of</strong> the other three varieties used in this work; this<br />

variety also had the highest seed yield and has a good balance <strong>of</strong><br />

amino acids. Here, for the first time, some secondary metabolites<br />

are reported to be present in amaranth seed flours. These metabolites,<br />

namely rutin, isoquercitrin, and nicotiflorin are compounds<br />

that could have implications in the prevention <strong>of</strong> several illnesses.<br />

The highest rutin content was found in Tulyehualco, while nicotiflorin<br />

was found in highest concentration in the Gabriela variety.<br />

Tulyehualco, because <strong>of</strong> the high seed yield, high protein and rutin<br />

content, seems to be an excellent amaranth variety to grow in arid<br />

areas worldwide. The knowledge <strong>of</strong> amaranth as a source <strong>of</strong><br />

phytochemicals will increase their importance as a potential source<br />

<strong>of</strong> those compounds in the human diet. Amaranth therefore has<br />

great potential as a sustainable crop that could improve household<br />

food security and farm incomes. In arid areas where commercial<br />

crops such as beans, maize or rice cannot be grown, amaranth<br />

cultivation could contribute to the first Millennium Development<br />

Goal <strong>of</strong> reducing hunger and poverty.<br />

Acknowledgments<br />

Thanks are due AG Alpuche Solis for his help in plant harvesting,<br />

and to Fundación Produce San Luis Potosi for financial support. This<br />

work was partially supported financially by the European<br />

Commission 6th Framework Programme, AMARANTH:FUTURE-<br />

FOOD, Contract No. 032263. Thanks to Dagmar Janovská for<br />

reviewing the manuscript.<br />

References<br />

Abugouch, L.E., Martinez, E.N., Añon, M.C., 2003. Influence <strong>of</strong> the extracting solvent<br />

upon the structural properties <strong>of</strong> amaranth (Amaranthus hypochondriacus)<br />

glutelin. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 51, 4060–4065.<br />

AOAC, 1990. Official Methods <strong>of</strong> Analysis, 15th ed. Association <strong>of</strong> Official Analytical<br />

Chemists, Washington, DC.<br />

Barba de la Rosa, A.P., Gueguen, J., Paredes-López, O., Viroben, G., 1992. Fractionation<br />

procedures, electrophoretic characterization and amino acid composition<br />

<strong>of</strong> amaranth seed protein. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 40,<br />

931–936.<br />

Carlsen, S.C.K., Mortensen, A.G., Oleszek, W., Piacente, S., Stochmal, A.,<br />

Fomsgaard, I.S., 2008. Variation in Flavonoids in Leaves, Stems and Flowers <strong>of</strong><br />

White Clover Cultivars. Natural Products Communications 3, 1299–1306.<br />

Cervantes-Laurean, D., Schramm, D.D., Jacobson, E.L., Halaweish, I., Bruckner, G.G.,<br />

Biossonneault, G.A., 2006. Inhibition <strong>of</strong> advanced glycation end product<br />

formation on collagen by rutin and its metabolites. The <strong>Journal</strong> <strong>of</strong> Nutritional<br />

Biochemistry 17, 531–540.<br />

Chan, K.F., Sun, M., 1997. Genetic diversity and relationships detected by isozyme<br />

and RAPD analysis <strong>of</strong> crop and wild species <strong>of</strong> Amaranthus. Theoretical and<br />

Applied Genetics 95, 865–873.<br />

Dellaporta, S.L., Wood, J., Hicks, J.B., 1983. A plant DNA minipreparation. Plant<br />

Molecular Biology Reporter 1, 19–22.<br />

Dixon, J., Nalley, L., Kosina, P., La Rovere, R., Hellin, J., Aquino, P., 2006. Adoption and<br />

economic impact <strong>of</strong> improved wheat varieties in the developing world. <strong>Journal</strong><br />

<strong>of</strong> Agricultural <strong>Science</strong> 144, 489–502.<br />

Donovan, J.L., Manach, C., Faulks, R.M., Kroon, P.A., 2007. Absorption and metabolism<br />

<strong>of</strong> dietary plant secondary metabolites. In: Crozier, A., Clifford, M.N.,<br />

Ashihara, H. (Eds.), Plant Secondary Metabolites: Occurrence, Structure and<br />

Role in the Human Diet. Blackwell Publishing, London, pp. 303–351.<br />

A.P. Barba de la Rosa et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 117–121 121<br />

FAO/WHO/UNU, 1986. Special report. Energy and protein requirements. <strong>Cereal</strong><br />

Foods World 3, 694–695.<br />

Gimplinger, D.M., Schulte auf ´ m Erley, G., Dobos, G., Kaul, H.P., 2008. Optimum crop<br />

densities for potential yield and harvestable yield <strong>of</strong> grain amaranth are conflicting.<br />

European <strong>Journal</strong> <strong>of</strong> Agronomy 28, 119–125.<br />

Gorinstein, S., Medina-Vargas, O.J., Jaramillo, N.O., Arnao-Salas, I., Martínez-<br />

Ayala, A.L., Arancibia-Avila, P., Toledo, F., Katrich, E., Trakhtenberg, S., 2007. The<br />

total polyphenols and the antioxidant potentials <strong>of</strong> some selected cereals and<br />

pseudocereals. European Food Research and Technology 225, 321–328.<br />

He, H.P., Cai, Y., Sun, M., Corke, H., 2002. Extraction and purification <strong>of</strong> squelene<br />

from amaranthus grain. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 50,<br />

368–372.<br />

Heiniö, R.-L., Liukkonen, K.-H., Myllymäki, O., Pihlava, J.-M., Adlercreutz, H.,<br />

Heinonen, S.-M., Poutanen, K., 2008. Quantities <strong>of</strong> phenolic compounds and<br />

their impacts on the perceived flavor attributes <strong>of</strong> rye grain. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong><br />

<strong>Science</strong> 47, 566–575.<br />

Henderson, T.L., Joson, B.L., Schneiter, A.A., 2000. Row spacing, plant population,<br />

and cultivar effects on grain amaranth in the northern great plains. Agronomy<br />

<strong>Journal</strong> 92, 329–336.<br />

Higashi, Y., Hirai, M.Y., Fujiwara, T., Naito, S., Noji, M., Saito, K., 2006. Proteomic and<br />

transcriptomic analysis <strong>of</strong> Arabidopsis seeds: molecular evidence for successive<br />

processing <strong>of</strong> seed proteins and its implication in the stress response to sulfur<br />

nutrition. Plant <strong>Journal</strong> 48, 557–571.<br />

Hobbs, P.R., 2007. Conservation agriculture: what is it and why is it important for<br />

future sustainable food production. <strong>Journal</strong> <strong>of</strong> Agricultural <strong>Science</strong> 145,<br />

127–137.<br />

Huang, J.L., Fu, S.T., Jiang, Y.Y., Cao, Y.B., Guo, M.L., Wang, Y., Xu, Z., 2007. Protective<br />

effects <strong>of</strong> Nicotiflorin on reducing memory dysfunction, energy metabolism<br />

failure and oxidative stress in multi-infarct dementia model rats. Pharmacology,<br />

Biochemistry and Behavior 86, 741–748.<br />

Islas-Gutiérrez, J., Islas-Gutiérrez, F., 2001. Rentabilidad de los cultivos de amaranto<br />

y maíz para la zona central de México. Agricultura Técnica de México 27,<br />

143–151.<br />

Kauffman, C., 1992. The status <strong>of</strong> grain amaranth for the 1990s. Food Reviews<br />

International 8, 165–192.<br />

Khuri, S., Bakker, F.T., Dunwell, J.M., 2001. Phylogeny, function and evolution <strong>of</strong> the<br />

Cupins, a structurally conserved, functionally diverse superfamily <strong>of</strong> proteins.<br />

Molecular Biology and Evolution 18, 593–605.<br />

Klimczak, I., Malecka, M., Pacholek, B., 2002. Antioxidant activity <strong>of</strong> ethanolic<br />

extracts <strong>of</strong> amaranth seeds. Nahrung-Food 46, 184–186.<br />

Kozak, M., Bocianowski, J., Rybinski, W., 2008. Selection <strong>of</strong> promising genotypes<br />

based on path and cluster analyses. <strong>Journal</strong> <strong>of</strong> Agricultural <strong>Science</strong> 146, 85–92.<br />

Laemmli, U.K., 1970. Cleavage <strong>of</strong> structural protein during the assembly <strong>of</strong> head <strong>of</strong><br />

bacteriophage T4. Nature 227, 680–685.<br />

Li, R.P., Guo, M.L., Zhang, G., Xu, X.F., Li, Q., 2006. Neuroprotection <strong>of</strong> nicotiflorin in<br />

permanent focal cerebral ischemia and in neuronal cultures. Biological and<br />

Pharmaceutical Bulletin 29, 1868–1872.<br />

Meyers, K.J., Rudolf, J.L., Mitchell, A.E., 2008. Influence <strong>of</strong> dietary quercetin on<br />

glutathione redox status in mice. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 56,<br />

830–836.<br />

Omami, E.N., Hammes, P.S., Robbertse, P.J., 2006. Differences in salinity tolerance for<br />

growth and water-use efficiency in some amaranth (Amaranthus spp.) genotypes.<br />

New Zealand <strong>Journal</strong> <strong>of</strong> Crop and Horticultural <strong>Science</strong> 34, 11–22.<br />

Paredes-López, O., Barba de la Rosa, A.P., Hernández-López, D., Cárabez-Trejo, A.,<br />

1990. Amaranto: características y aprovechamiento agroindustrial. Programa<br />

Regional de Desarrollo Científico y Tecnológico, Washington, DC.<br />

Que, F., Mao, L., Pan, X., 2006. Antioxidant activities <strong>of</strong> five chinese rice wines and<br />

the involvement <strong>of</strong> phenolic compounds. Food Research International 39,<br />

581–587.<br />

Reyes-Castañeda, P. (Ed.), 1983. Bioestadística aplicada, Agronomía, Biología,<br />

Química. Trillas, México, DF.<br />

SAGARPA, 2002. Sistema de información agropecuaria de consulta. SIACON,<br />

México, DF.<br />

Segura-Nieto, M., Barba de la Rosa, A.P., Paredes-López, O., 1994. Biochemistry <strong>of</strong><br />

amaranth proteins. In: Paredes-López, O. (Ed.), Amaranth: Biology, Chemistry,<br />

and Technology. CRC Press, Boca Raton, FL, pp. 75–106.<br />

Silva-Sánchez, C., Barba de la Rosa, A.P., León-GalvánF, de Lumen, B.O., De León-<br />

Rodríguez, A., González de Mejía, E., 2008. Bioactive peptides in amaranth<br />

(Amaranthus hypochondriacus) seed storage proteins. <strong>Journal</strong> <strong>of</strong> Agricultural and<br />

Food Chemistry 56, 1233–1240.<br />

Silva-Sánchez, C., González-Castañeda, J., De León Rodríguez, A., Barba de la Rosa,<br />

A.P., 2004. Functional and rheological properties <strong>of</strong> amaranth albumins<br />

extracted from two mexican varieties. Plant Foods for Human Nutrition 59,<br />

169–174.<br />

Suryavanshi, V.L., Sathe, P.A., Baing, M.M., Singh, G.R., Lakshmi, S.N., 2007. Determination<br />

<strong>of</strong> Rutin in Amaranthus spinosus Linn. Whole Plant Powder by HPTLC.<br />

Chromatography 65, 767–769.<br />

Tosi, E.A., Ré, E., Lucero, H., Masciarelli, R., 2001. Dietary fiber obtained from<br />

amaranth (Amaranthus cruentus) grain by differential milling. Food Chemistry<br />

73, 441–443.


Relationship <strong>of</strong> milled grain percentages and flowering-related traits in rice<br />

Rodante E. Tabien a, *, Stanley Omar P.B. Samonte a , Emmanuel R. Tiongco b<br />

a Texas A&M University System, AgriLife Research and Extension Center, 1509 Aggie Drive, Beaumont, TX 77713, USA<br />

b Philippine Rice Research Institute, <strong>Science</strong> City <strong>of</strong> Muñoz, Nueva Ecija 3119, Philippines<br />

article info<br />

Article history:<br />

Received 29 December 2007<br />

Received in revised form 12 July 2008<br />

Accepted 15 July 2008<br />

Keywords:<br />

Rice<br />

Flowering duration<br />

Milling qualities<br />

Flowering traits<br />

1. Introduction<br />

abstract<br />

The economic value <strong>of</strong> harvested rice is determined by the grain<br />

yield and the percentages <strong>of</strong> head rice (at least 3/4 the length <strong>of</strong><br />

a head or kernel) and total milled rice. The high value <strong>of</strong> head rice<br />

relative to broken grains influences the cultivar choice and the preand<br />

post-harvest handling practices to maximize farm income.<br />

These milling traits had shown significant positive direct effects on<br />

gross income, thus these should be considered in selecting cultivars<br />

to plant and in developing a high income-grossing genotype<br />

(Samonte et al., 2006).<br />

Milling traits are influenced by the environment, cultural practices,<br />

milling processes, drying, and genotype. Environmental<br />

factors reported to affect head rice percentages include meteorological<br />

conditions such as relative humidity, air temperature, and<br />

rainfall (Banaszek and Siebenmorgen, 1990; Jodari and Linscombe,<br />

1996; Thompson and Mutters, 2006). Cultural practices include<br />

time <strong>of</strong> draining (Counce et al., 1990; McCauley and Way, 2002),<br />

harvesting (McCauley and Way, 2002) and nitrogen fertilization<br />

(Jongkaewwattana et al., 1993; Leesawatwong et al., 2005; Seetanum<br />

and De Datta, 1973). Milling processes that influence grain<br />

milling traits include the type <strong>of</strong> mill (Bautista and Siebenmorgen,<br />

2002), milling cylinder speeds (Dilday, 1989), degree <strong>of</strong> milling<br />

(Reid et al., 1998), temperature and and duration (Cnossen et al.,<br />

* Corresponding author. Tel.: þ1 409 752 2741.<br />

E-mail address: retabien@ag.tamu.edu (R.E. Tabien).<br />

0733-5210/$ – see front matter Published by Elsevier Ltd.<br />

doi:10.1016/j.jcs.2008.07.015<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 122–127<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

The economic value <strong>of</strong> harvested rice is determined by the grain yield and the percentages <strong>of</strong> head rice<br />

(at least 3/4 the length <strong>of</strong> a head or kernel) and total milled rice. This study was conducted to determine<br />

the effects <strong>of</strong> flowering-related traits such as duration <strong>of</strong> flowering, rate <strong>of</strong> flowering, heading, and<br />

duration from heading to maturity on head rice and total milling percentages <strong>of</strong> rice. Flowering data,<br />

gathered for two years from 105 long grain rice genotypes grown in Beaumont, Texas were analyzed for<br />

their effects on and relationship with milling traits. A positive linear relationship was obtained for rate <strong>of</strong><br />

flowering and the duration from heading to maturity but negative for duration <strong>of</strong> flowering and days to<br />

heading. Genotypes with early heading had relatively shorter flowering durations, and genotypes with<br />

shorter flowering duration had higher head rice and total milled rice. A faster rate in attaining 100%<br />

flowering and more days from heading to maturity were favorable in increasing head rice and total<br />

milled grains. The duration from the start <strong>of</strong> flowering to heading or to 100% flowering can be used in the<br />

evaluation and selection for high head rice and total milled rice percentages in rice.<br />

Published by Elsevier Ltd.<br />

2003; Schluterman and Siebenmorgen, 2007), cooling methods and<br />

milling procedures (Pan et al., 2005), and head rice separation<br />

methods (Lloyd et al., 2001). Drying, as it relates to grain moisture<br />

content, is critical in getting high percentage <strong>of</strong> head rice (Banaszek<br />

and Siebenmorgen, 1990; Dilday, 1989; Siebenmorgen and Jindal,<br />

1986; Siebenmorgen et al., 1992). Grain moisture content (MC) <strong>of</strong><br />

15–18% has been found ideal for milling grain rice (Jodari and<br />

Linscombe, 1996), but the most recent study indicated that the<br />

optimal harvest moisture content (HMC) at which milling traits<br />

peaked for long grain cultivar was 18–22% while medium grains<br />

needed 22–24% HMC (Siebenmorgen et al., 2007).<br />

Cultivars differed significantly in their milled rice percentages<br />

(Dilday, 1989). Genotypic traits reported to affect head rice include<br />

tillering and kernel weights at lower seeding rates (Gravois and<br />

Helms, 1996), kernel thickness (Lu and Siebenmorgen, 1995),<br />

panicle type and grain weight (Wang et al., 2007), and panicle<br />

length and maturity (Jongkaewwattana and Geng, 2002). Nitrogen<br />

fertilization has been shown to increase head rice percentage<br />

(Jongkaewwattana et al., 1993), but this increase was dependent on<br />

genotype (Perez et al., 1996). The optimum N rates and harvest<br />

grain moisture to obtain the maximum head rice percentage <strong>of</strong><br />

different cultivars were also determined by Leesawatwong et al.<br />

(2005). Siebenmorgen et al. (2007) found that the magnitude <strong>of</strong><br />

head rice reduction at different moisture content was dependent<br />

on cultivar.<br />

In estimating the number <strong>of</strong> days from emergence to heading<br />

(50% <strong>of</strong> the panicles are flowering) for each breeding line, rice<br />

breeders have observed that flowering duration from the onset to


100% flowering and maturity vary across lines. Chau and Kunze<br />

(1982) reported large variation in grain maturity and consequently<br />

MC within a field (13–43%). There was a large MC variation in<br />

individual kernels in panicle and in plant, and among plants<br />

(Bautista and Siebenmorgen, 2005; Holloway et al., 1995; Kocher<br />

et al., 1990). Differences in grain weights and quality within panicle<br />

were found to be dependent on variety (Cheng et al., 2007; Liu<br />

et al., 2005).<br />

Rice flowers asynchronously. The flowers at the top <strong>of</strong> the<br />

panicles open first and the bottom florets open last, and the flowering<br />

duration may last up to 15 days (Holloway et al., 1995; Luh<br />

and Luh, 1991). These variations have resulted in multi-modal MC<br />

distribution among rice kernels and to a variation in amylose<br />

content (Cheng et al., 2007).<br />

Grain structure was also found to be related to grain filling<br />

which had an impact on head rice (Jongkaewwattana and Geng,<br />

2001). Decreasing kernel dimensions, volume and density, and<br />

amylose content from top to lower part <strong>of</strong> the panicle was observed<br />

in cultivars with different maturities (Cheng et al., 2007; Jongkaewwattana<br />

and Geng, 2002) and this non-uniformity influenced<br />

head rice recovery. It is hypothesized that flowering-related traits<br />

such as duration <strong>of</strong> flowering, rate <strong>of</strong> flowering, heading, and<br />

duration from heading to harvest can affect head rice and total<br />

milling percentages. Hence, this study determined these<br />

relationships.<br />

2. Experiment<br />

The Uniform Regional Rice Nursery (URRN) located at the Texas<br />

A&M University System, AgriLife Research and Extension Center,<br />

Beaumont, TX (29 57 0 N, 94 30 0 W), is part <strong>of</strong> a multi-state yield<br />

trials. The nursery is composed <strong>of</strong> elite genotypes from various rice<br />

breeding programs in the USA, and it has 200 entries per year. The<br />

105 long grain genotypes that were common across the 2005 and<br />

2006 URRN were used in this study. The soil in the research center<br />

is Beaumont, an Entic Pelludert (fine, montmorillonitic, and<br />

thermic). The planting dates <strong>of</strong> the breeding materials in both years<br />

were mid-April. Each cropping year, the field experiments were<br />

fertilized with 225 kg N ha 1 equally split into three and applied at<br />

planting, one month after planting, and at panicle differentiation.<br />

The entries were laid out in randomized complete block design<br />

(RCBD) with two to four replications for entries at the preliminary<br />

and the advanced trials, respectively, but only two replications<br />

were used for data gathering in this study. Each plot had six 3-m<br />

long rows spaced at 25 cm.<br />

The field data obtained on a plot basis for each line from two<br />

replications included: days from emergence to onset (start), 25, 50<br />

(heading), 75, and 100% flowering, and days to harvest. Milling<br />

samples were obtained at maturity from rice grown in a 1-sq m<br />

area in each plot. These were threshed and dried right after harvest.<br />

The head rice and total milled rice percentages were obtained<br />

following the standard procedures (Fan et al., 2000).<br />

To produce brown rice, the 125-g rough rice samples were<br />

hulled by a single pass through a rubber roll huller (Satake Engineering<br />

Co., LTD., Tokyo, Japan). The brown rice was then milled for<br />

54 s using a McGill No. 2 mill (Rapsilver Supply Co Inc., Brookshire,<br />

TX). The generated milled rice was separated into head rice and<br />

broken fractions with a shaker/separator (#12 screen ¼ 4.76 mm;<br />

Seedburo Equipment Co., Chicago, IL); The weight <strong>of</strong> the grain<br />

recovered was used to calculate the percentage <strong>of</strong> total milled rice<br />

and head rice based on 125 g <strong>of</strong> rough rice.<br />

From field data, the following flowering durations from seedling<br />

emergence were estimated on a plot basis: start <strong>of</strong> flowering<br />

(onset) to 50% flowering (FD0–50), 50% flowering to 100% flowering<br />

(FD50–100), start <strong>of</strong> flowering to 100% flowering (FD0–100), and rate <strong>of</strong><br />

flowering (slope <strong>of</strong> the number <strong>of</strong> days to 25, 50, 75, and 100%).<br />

Date <strong>of</strong> harvest was used in estimating the number <strong>of</strong> days from<br />

50% flowering to maturity.<br />

Analysis <strong>of</strong> variance was used to evaluate the effect <strong>of</strong> genotype,<br />

year and its interaction to days from emergence to 50% flowering<br />

(heading), duration <strong>of</strong> flowering (FD0–50, FD50–100, FD0–100), rate <strong>of</strong><br />

flowering, days from heading to maturity, and milling traits<br />

(percent head rice and percent total milled rice). Regression and<br />

correlation analyses were conducted to determine the relationships<br />

between flowering-related traits (flowering duration, rate <strong>of</strong><br />

flowering, heading, days from heading to maturity) and the<br />

percentages <strong>of</strong> head rice and total milled rice, and the days to<br />

heading with flowering duration. Using the relationships derived,<br />

estimated values for the flowering-related traits at 70 and 55% total<br />

and head rice percentages, respectively, were generated.<br />

3. Results and discussion<br />

3.1. Year, genotype, and genotype year effects on flowering and<br />

milling traits<br />

In breeding trials across years or locations, the number <strong>of</strong> days<br />

from emergence to heading was significantly affected by genotype<br />

and genotype year (Table 1). The number <strong>of</strong> days to heading<br />

ranged from 82.0 days (recorded in line RU0401084, cvs. ‘Spring’<br />

and ‘Trenasse’ (known as very early rice cultivars), to 96.5 days (line<br />

RU0503126) in 2005 and 72.0 days (cv. Trenasse) to 90.0 days (line<br />

RU0501139) in 2006. Mean number <strong>of</strong> days to heading was 88.1<br />

days in 2005 and 79.9 days in 2006. Although it was much shorter<br />

for the genotypes to have 50% flowering in 2006, the variations<br />

were not enough to indicate significance between the two years.<br />

Earlier genotypes or later genotypes in 2005 showed the same<br />

flowering response in 2006 indicating a strong genetic control for<br />

this trait. A QTL in chromosome 8 that explains more than 94.9% <strong>of</strong><br />

variation in heading date was reported recently (Zhang et al., 2006).<br />

The same locus was identified in several different populations (Lin<br />

et al., 2003; Xiong et al., 1999; Zhuang et al., 1997); thus this QTL<br />

could be present in some <strong>of</strong> the 105 genotypes evaluated in this<br />

study.<br />

Table 1<br />

Means squares from the analysis <strong>of</strong> variance for flowering-related traits <strong>of</strong> 105 long grain genotypes grown at Beaumont, Texas in 2005 and 2006<br />

Source <strong>of</strong><br />

variation<br />

DF Days to<br />

heading<br />

Duration<br />

0–50%<br />

flowering<br />

Duration<br />

50–100%<br />

flowering<br />

Duration<br />

0–100%<br />

flowering<br />

Flowering<br />

rate<br />

Duration<br />

from<br />

heading to<br />

maturity<br />

Whole<br />

milled rice<br />

percentage<br />

Total milled<br />

rice<br />

percentage<br />

Replication 1 21.94 0.09 0.68 0.38 13.30 21.94 0.92 0.19<br />

Year (Y) 1 7109.49 85.95 0.68 103.09 889.76 237.75 5637.87* 1323.77*<br />

Error 1 60.95 9.75 3.62 24.38 341.23 60.95 11.55 0.97<br />

Genotype (G) 104 36.11** 1.84** 0.68** 3.24** 32.24** 47.92** 72.48** 17.65**<br />

Y G interaction 104 4.72** 1.04 0.52* 1.66 18.17 13.51** 36.58** 6.31**<br />

Error b 208 1.60 1.16 0.37 1.50 16.86 1.60 8.48 1.20<br />

*, ** Significant at the 5 and 1% level <strong>of</strong> probability, respectively.<br />

R.E. Tabien et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 122–127 123


124<br />

Genotype significantly affected FD0–50, FD50–100, and FD0–100,<br />

while genotype year significantly affected FD50–100 (Table 1).<br />

These indicate that there was enough variation in the duration <strong>of</strong><br />

flowering among 105 genotypes. Genotype FD0–50 (averaged across<br />

years) ranged from 2.0 days (line RU0401145 and cv. Spring) to 5.5<br />

days (line RU0503098), FD0–100 ranged from 3.5 days (cv. Spring) to<br />

9.5 days (line RU0503098), while FD50–100 ranged from 1.0 days (cv.<br />

Trenasse) to 4.0 days (line RU0503098). Mean FD0–50, FD0–100, and<br />

FD50–100 were 3.3, 5.5, and 2.2 days, respectively.<br />

Shorter duration <strong>of</strong> flowering (0–50%, 50–100%, or 0–100%) was<br />

mostly obtained on very early season varieties like cv. Spring and<br />

Trenasse. Duration <strong>of</strong> flowering, however, was not taken in yield<br />

trials, unlike heading date or the number <strong>of</strong> days from emergence<br />

to 50% flowering. The highly significant variation obtained and<br />

interaction noted suggests that these traits may be important in<br />

obtaining high quantity and quality harvests. Long flowering<br />

duration may increase variation in kernels at harvest. Since the first<br />

floret will be at a different stage relative to the late flowering<br />

florets, there is likely to be high variability in size, degree <strong>of</strong> grain<br />

filling, and moisture content. Reports indicated significant variation<br />

in kernel qualities within the panicle, plant or hill, and fields<br />

(Bautista et al., 2007; Jongkaewwattana and Geng, 2002), and these<br />

were related to head rice recovery.<br />

The rate <strong>of</strong> flowering (% days 1 ) was significantly affected by<br />

genotype but not by year, and by the genotype year interaction<br />

(Table 1). The rate <strong>of</strong> flowering across years ranged from 9.9%<br />

days 1 (line RU0503098) to 25.08% days 1 (line RU0502177) in<br />

2005 and 10.81% days 1 (line RU0503098) to 29.33% days 1 (line<br />

RU0203032) in 2006. The mean across genotypes for the rate <strong>of</strong><br />

flowering was 17.65% days 1 in 2005 and 20.56% days 1 in 2006.<br />

The average rate <strong>of</strong> flowering in two years was slowest in<br />

RU0503098 (10.36 % days 1 ) and fastest in cultivar Spring (25.66 %<br />

days 1 ). Among the released cultivars, early maturing cultivars,<br />

such as Spring and Trenasse, the hybrid XP 723, and the newly<br />

released Presidio were consistently fast to reach 100% flowering in<br />

two years. The wide variation observed in two years among<br />

genotypes suggests that some genotypes had much faster or slower<br />

rate to attain 100% flowering than the rest <strong>of</strong> the entries. Similar to<br />

duration <strong>of</strong> flowering, the rate <strong>of</strong> flowering may affect the uniformity<br />

<strong>of</strong> the grains. The grains will be more uniform for fast flowering<br />

genotypes than for slow flowering genotypes.<br />

The number <strong>of</strong> days from heading to maturity was significantly<br />

affected by genotype, and by genotype year interaction, but not<br />

by year (Table 1). In 2005, the shortest duration from heading to<br />

harvest was 29 days (line RU0503126) and the longest was 47 days<br />

(line RU0503181). In 2006, the shortest duration was 26 days (line<br />

RU0401179) and the longest was 49 days (line RU0501102). Average<br />

duration to harvest per year was 38 days in 2005 and 40 days in<br />

2006. Among the released cultivars, Cocodrie, Trenasse, and hybrid<br />

XP 723 matured longest at 45, 44 and 42 days, respectively. The<br />

shortest duration among cultivars was from Banks at 34 days followed<br />

by Saber and Spring at 35 days. The variation among genotypes<br />

again reflects potential differences in the duration <strong>of</strong> grain<br />

filling that can affect milling traits.<br />

Genotype, genotype year, and year had highly significant<br />

effect on both head rice and total milled rice percentages (Table 1).<br />

Head rice percentages (averaged across years) ranged from 37.1%<br />

(line RU0501096) to 64.3% (line RU0504198) in 2005 and from<br />

42.2% (cv. ‘Priscilla’) to 67.5% (cv. ‘Cybonnet’) in 2006. Mean (averaged<br />

across genotypes) head rice percentage was 51.5% in 2005 and<br />

58.9% in 2006. Total milled rice percentages (averaged across years)<br />

ranged from 63.0% (line RU0301188) to 72.9% (cv. ‘XP723’) in 2005,<br />

and from 63.0% (line RU0501139) to 76.9% (line RU0502068) in<br />

2006. Mean (averaged across genotypes) total milled rice<br />

percentage was 68.0% in 2005 and 71.5% in 2006. The results clearly<br />

indicate that better milling traits were obtained in 2006 compared<br />

R.E. Tabien et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 122–127<br />

to 2005, and this can be due to differences in environmental factors<br />

such as temperature, rainfall, and relative humidity (Banaszek and<br />

Siebenmorgen, 1990; Cooper et al., 2006; Jodari and Linscombe,<br />

1996; Thompson and Mutters, 2006). The reported significant<br />

interaction <strong>of</strong> year and genotype (Gravois et al., 1991) was considered<br />

an important source <strong>of</strong> variation for head rice.<br />

3.2. Days to heading and milling percentages<br />

Regression analysis shown in Fig. 1 revealed a linear relationship<br />

<strong>of</strong> days to heading with head rice and total milled rice. Although the<br />

amount <strong>of</strong> variation that can be explained by this trait was low in<br />

2005 and much higher at 30–38% in 2006, the same negative and<br />

highly significant relationships were obtained in both years. The<br />

head rice and total milled rice percentages decreased with longer<br />

duration to heading. A reduction <strong>of</strong> 0.74–0.80% in head rice and<br />

0.35–0.44% for total milled rice is possible per day increase in days<br />

to heading.<br />

Based on the 2006 linear relationship, a desirable total milled<br />

rice percentage <strong>of</strong> 70% and head rice percentage <strong>of</strong> 55% should have<br />

83.44 days and 85.23 days to heading, respectively. Longer periods<br />

than these estimates will mean lower milling percentages. Very<br />

early heading genotypes are usually not selected in most rice<br />

breeding programs. In most cases, plants that head too early had<br />

fewer tillers and lighter biomass, thereby reducing grain yield (Xiao<br />

et al., 1998); these also had prolonged vegetative stage resulting in<br />

poor grain filling and lower grain yield.<br />

Grain yield and days to heading were positively correlated (data<br />

not shown); thus there should be a compromise to obtain both high<br />

yield and high total and head rice percentages. Using the two year<br />

data, correlation <strong>of</strong> heading to total and head rice percentages was<br />

very high, with r-values <strong>of</strong> 0.70 and 0.75, respectively. This indicates<br />

that heading may be useful for the indirect selection <strong>of</strong> high head<br />

rice and total milled rice percentages. However, one must consider<br />

its interaction with year and the potential impact <strong>of</strong> the environment.<br />

Actual use <strong>of</strong> this trait in selection will verify its effectiveness<br />

in improving milling traits.<br />

The overlapping <strong>of</strong> values for total and head rice as shown in<br />

Fig. 1, for instance, indicates variability if this trait will be used as<br />

selection index. These variations can be due to timing at which the<br />

50% heading has occurred. Using historical data <strong>of</strong> two cultivars,<br />

average daily low temperature (high night time temperature) at<br />

50% heading was shown to significantly affect head rice (Cooper<br />

et al., 2006). High night time temperature was reported to decrease<br />

Fig. 1. Relationship <strong>of</strong> number <strong>of</strong> days from emergence to heading, and head and total<br />

milled rice percentages in 105 long grain genotypes grown at Beaumont, Texas in 2005<br />

and 2006.


grain dimensions, increase number <strong>of</strong> chalky grains, lower amylose<br />

content, and increase length <strong>of</strong> amylopectin chains, thereby<br />

affecting milling qualities (Cooper et al., 2008; Counce et al., 2005).<br />

3.3. Flowering duration and milling percentages<br />

Flowering duration to 50% and 100%, and 50 to 100% flowering<br />

were negatively correlated with milling traits (Table 2). The highest<br />

correlation was found between FD0–100 and percentage head grain<br />

(r ¼ 0.464) followed by FD0–50 and head rice percentage<br />

(r ¼ 0.433). The FD50–100 was also significantly and negatively<br />

correlated with head rice and total milled rice percentages, but<br />

these correlation coefficients were lower than both FD0–50 and<br />

FD0–100. These relationships indicate that the duration to reach 50%<br />

or 100% flowering is important in obtaining high total and head rice<br />

percentages than the duration from 50–100%. Moreover, these<br />

indicate the need for shorter duration <strong>of</strong> flowering to insure better<br />

total and head grain percentages. The negative relationship<br />

between flowering duration and milled rice percentages is shown<br />

in Fig. 2. There was a 2.4% and 1.1% decrease in head rice and total<br />

milled, respectively for every one day increase in flower duration.<br />

Grains from the rice panicle or from the entire plant, develop<br />

and mature asynchronously. At harvest, the distribution <strong>of</strong> MC <strong>of</strong><br />

grains is multi-modal (Bautista and Siebenmorgen, 2005; Kocher<br />

et al., 1990). The range <strong>of</strong> grain MC decreases as the MC at harvest<br />

decreases suggesting that the ranges <strong>of</strong> MC are directly related to<br />

the maturity <strong>of</strong> the grains (Bautista and Siebenmorgen, 2005).<br />

These ranges <strong>of</strong> variation will be more at prolonged flowering<br />

duration since the maturity <strong>of</strong> the grains will be different. The<br />

differences in flowering result in wider variation in grain development,<br />

and the final weight and quality <strong>of</strong> the grains (Wang et al.,<br />

2006). Any non-uniformity expressed in variable grain size and<br />

shape, grain filling, and maturity has negative effects on rice milling<br />

traits (Jongkaewwattana and Geng, 2001).<br />

Rice breeders commonly estimate the number <strong>of</strong> days from<br />

emergence to heading <strong>of</strong> their breeding lines. Adding days to start<br />

<strong>of</strong> flowering to the list <strong>of</strong> evaluation criteria will enable the estimation<br />

<strong>of</strong> FD0–50 for indirect selection criterion for relatively higher<br />

head rice and total milled rice percentage. The mean <strong>of</strong> 105 genotypes<br />

for FD0–50 and FD0–100, were 3.3 and 5.5 days, respectively,<br />

and the mean head rice and total milled rice percentages (averaged<br />

across years and genotypes) were 55.2 and 69.8%, respectively.<br />

Based on the regression equations obtained, the FD0–50 and FD0–100<br />

should be at less than 5.5 days in order to produce higher than<br />

average head rice and total milled rice percentages.<br />

Unaffected by year and the genotype year interaction, the trait<br />

<strong>of</strong> flowering duration flowering can be used in indirect selection<br />

relative to the number <strong>of</strong> days to heading. Although the correlation<br />

was lower compared to heading date, this trait can be a more<br />

reliable indicator <strong>of</strong> milling potentials. Actual use in the breeding<br />

program will prove its efficiency in improving milling traits.<br />

Table 2<br />

Correlation between flowering-related traits and milling percentages <strong>of</strong> 105 long<br />

grain rice genotypes grown at Beaumont, Texas in 2005 and 2006<br />

Flowering duration Correlation coefficient (r)<br />

Whole milled rice Total milled rice<br />

Start <strong>of</strong> flowering to 50% flowering 0.43** 0.43**<br />

50–100% flowering 0.27** 0.190**<br />

Start to 100% flowering 0.46** 0.42**<br />

Heading date 0.70** 0.75**<br />

Flowering rate þ0.42** þ0.40**<br />

Duration for heading to maturity<br />

** Significant at the 1% probability level.<br />

þ0.18** þ0.43**<br />

R.E. Tabien et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 122–127 125<br />

Fig. 2. Relationship <strong>of</strong> flowering duration and whole and total milled rice percentages<br />

in 105 long grain genotypes grown at Beaumont, Texas in 2005 and 2006.<br />

3.4. Rate <strong>of</strong> flowering and milling percentages<br />

The rate <strong>of</strong> flowering, similar to the duration <strong>of</strong> flowering, was<br />

linearly and positively related to both head rice and total milled rice<br />

percentages as shown in Fig. 3. Increasing rate <strong>of</strong> flowering was<br />

favorable in improving milling traits. This result indicates that<br />

a genotype should reach 100% flowering at a faster rate in order to<br />

produce high quality grain for milling. The relationship <strong>of</strong> rate <strong>of</strong><br />

flowering and milling can be explained by the uniformity <strong>of</strong> grains<br />

at harvest. If the genotype has a fast rate <strong>of</strong> flowering, the grains<br />

will be filled nearly at the same time and will be in a near-uniform<br />

stage during harvest. Synchronous grain development related to<br />

grain filling and maturity has a positive impact on milling qualities<br />

(Jongkaewwattana and Geng, 2001). Correlation <strong>of</strong> flowering rate<br />

to milling percentages was significant with r-values comparable to<br />

that <strong>of</strong> duration <strong>of</strong> flowering, thus indicating the former’s potential<br />

for indirect selection. Estimation <strong>of</strong> flowering rate, however, needs<br />

several data points compared to the simpler gathering <strong>of</strong> data for<br />

duration <strong>of</strong> flowering.<br />

3.5. Duration <strong>of</strong> heading to maturity and milling percentages<br />

The number <strong>of</strong> days from heading to maturity did not have<br />

a significant relationship with head rice percentages in 2005 and<br />

2006. However, there were significant linear relationships between<br />

Fig. 3. Relationship <strong>of</strong> flowering rate and whole and total milled rice percentages in<br />

105 long grain genotypes grown at Beaumont, Texas in 2005.


126<br />

the number <strong>of</strong> days from heading to maturity and total milled rice<br />

percentages in 2005 and 2006 (Fig. 4). For each day that the<br />

number <strong>of</strong> days from heading to maturity increased, total milled<br />

rice percentage increased by 0.24% in 2005 and 0.29% in 2006. The<br />

increase in percent total grain at longer duration from flowering to<br />

harvest can be attributed to better grain filling. It was observed that<br />

grain filling affected grain traits such as weight and density, shape<br />

and size, which, in turn, affected milling traits (Siebenmorgen et al.,<br />

1992). Moreover, at 40 days after flowering, most <strong>of</strong> the grains had<br />

20% moisture content, and maximum total and head rice grain was<br />

obtained at 12–22%. An opposite relationship was noted in Californian<br />

cultivars with different maturities. There was a quadratic<br />

relationship between maturity and head rice grain, and the<br />

percentage head rice increased from early and intermediate to late<br />

and very late maturing cultivars (Jongkaewwattana and Geng,<br />

2002). This difference can be attributed to grain shape and size.<br />

Most <strong>of</strong> the Californian cultivars were medium grain, but this<br />

study focused on long grains that are popular in the Gulf Coast.<br />

Variation in grain size and shape contributed to the variation in rate<br />

and duration <strong>of</strong> grain filling (Jongkaewwattana and Geng, 2001)<br />

and ultimately milling quality. It was reported that late-heading<br />

entries had larger panicle, and the date <strong>of</strong> heading was found to be<br />

correlated with the number <strong>of</strong> grains per panicle (0.52) and<br />

spikelets per panicle (0.42) (Zhang et al., 2006). These results<br />

further support increased grain variations in late heading<br />

genotypes.<br />

Number <strong>of</strong> days to heading and maturity are important for rice<br />

adaptation to certain cultivation areas and to cropping seasons in<br />

most rice growing countries. Shorter maturity or early flowering is<br />

needed in rainfed and upland areas and during the rainy season.<br />

However, a relatively longer maturity is needed to have higher<br />

grain yield in irrigated areas during the dry season planting.<br />

Therefore, it is important to consider the number <strong>of</strong> days to heading<br />

and maturity in selecting for higher total and head rice percentages,<br />

and high grain yield.<br />

3.6. Duration <strong>of</strong> flowering and the number <strong>of</strong> days from emergence<br />

to 50% flowering<br />

The number <strong>of</strong> days from emergence to heading had a significant<br />

linear relationship with flowering duration (Fig. 5). The<br />

number <strong>of</strong> days from the start <strong>of</strong> flowering to 50% was longer at<br />

prolonged duration to heading (50% flowering). Across 2005 and<br />

2006, the daily flowering duration increased by 0.12 days as the<br />

number <strong>of</strong> days from emergence to heading increased. This<br />

Fig. 4. Relationship <strong>of</strong> number <strong>of</strong> days from heading and maturity on total milled rice<br />

percentages in 105 long grain genotypes grown at Beaumont, Texas in 2005 and 2006.<br />

R.E. Tabien et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 122–127<br />

Fig. 5. Relationship <strong>of</strong> number <strong>of</strong> days from emergence to start <strong>of</strong> flowering and the<br />

flowering duration in 105 long grain genotypes grown at Beaumont, Texas in 2005 and<br />

2006.<br />

relationship suggests that late flowering genotypes should be<br />

avoided to have genotypes with shorter duration <strong>of</strong> flowering. Early<br />

flowering genotypes should give better total and head grain<br />

percentages considering the positive relationship <strong>of</strong> heading date<br />

and the duration <strong>of</strong> flowering, and the negative relationship <strong>of</strong><br />

milling traits and the duration <strong>of</strong> flowering. Breeders are not<br />

focusing on very early genotypes because <strong>of</strong> lower grain yields but<br />

since earliness is related to duration <strong>of</strong> flowering and the duration<br />

<strong>of</strong> flowering is correlated with milling traits, a trade-<strong>of</strong>f is necessary.<br />

Based on the linear relationships obtained, a line should<br />

flower between 80 and 85 days. The US-released varieties included<br />

in the trials had 72–86 days to 50% flowering and the majority <strong>of</strong><br />

cultivars were within the estimated range (data not shown).<br />

Based on these results, flowering-related traits such as days to<br />

heading, rate <strong>of</strong> flowering, duration <strong>of</strong> flowering, and duration from<br />

heading to maturity can impact head rice and total milled rice<br />

percentages. These milling traits were positively related with rate<br />

<strong>of</strong> flowering, and duration from heading to maturity but negatively<br />

related with duration <strong>of</strong> flowering (0–50%, 50–100% and 0–100%)<br />

and days to heading. Rice genotypes with early heading had relatively<br />

shorter flowering durations, and genotypes with shorter<br />

flowering duration had higher head rice and total milled rice<br />

percentages. Faster duration to attain 100% flowering and more<br />

days from heading to maturity were favorable in increasing<br />

percentage head rice and total grain. Estimating the duration from<br />

start <strong>of</strong> flowering to heading or to 100% flowering can be potential<br />

selection criteria in the indirect evaluation and selection <strong>of</strong><br />

breeding lines for high head rice and total milled rice percentages.<br />

The number <strong>of</strong> days from heading to maturity can be considered in<br />

indirectly selecting for high total milled rice percentage.<br />

Acknowledgements<br />

The authors would like to acknowledge the financial support<br />

from Texas Rice Research Foundation and the field and laboratory<br />

assistance, especially the harvesting and milling <strong>of</strong> samples,<br />

provided by Chersty Harper, Patrick Frank, Joel Pace, Richard Boyd,<br />

and Pat Carre.<br />

References<br />

Banaszek, M.M., Siebenmorgen, T.J., 1990. Head rice yield reduction rates caused by<br />

moisture adsorption. Transactions <strong>of</strong> the ASAE 33, 1263–1269.<br />

Bautista, R.C., Siebenmorgen, T.J., 2005. Individual kernel rice moisture content<br />

variability trends. Applied Engineering in Agriculture 21 (4), 637–643.


Bautista, R.C., Siebenmorgen, T.J., 2002. Evaluation <strong>of</strong> laboratory mills for milling<br />

small samples <strong>of</strong> rice. Applied Engineering in Agriculture 18 (5), 577–583.<br />

Bautista, R.C., Siebenmorgen, T.J., Counce, P.A., 2007. Rice kernel dimensional variability<br />

trends. Applied Engineering in Agriculture 23 (2), 207–217.<br />

Chau, N.N., Kunze, O.R., 1982. Moisture content variation among harvested rice<br />

grains. Transactions <strong>of</strong> the ASAE 25, 1037–1040.<br />

Cheng, F.M., Zhang, Q.F., Zhu, H.J., Zhao, N.C., Wang, F., Chen, K.M., Zhang, G.P., 2007.<br />

The difference in amylose content within a panicle as affected by the panicle<br />

morphology <strong>of</strong> rice cultivars. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 46, 49–57.<br />

Cnossen, A.G., Jiménez, M.J., Siebenmorgen, T.J., 2003. Rice fissuring response to high<br />

drying and tempering temperatures. <strong>Journal</strong> <strong>of</strong> Food Engineering 59, 61–69.<br />

Cooper, N.T.W., Siebenmorgen, T.J., Counce, P.A., Meullenet, J.-F., 2006. Explaining<br />

rice milling quality variation using a historical weather data analysis. <strong>Cereal</strong><br />

Chemistry 83 (4), 447–450.<br />

Cooper, N.T.W., Siebenmorgen, T.J., Counce, P.A., 2008. Effects <strong>of</strong> night time<br />

temperature during kernel development on rice physicochemical properties.<br />

<strong>Cereal</strong> Chemistry 83 (4), 447–450.<br />

Counce, P.A., Siebenmorgen, T.J.,Vories, E.D.,Pitts, D.J.,1990. Time <strong>of</strong>draining and harvest<br />

effects on rice grain yield and quality. <strong>Journal</strong> <strong>of</strong> Production Agriculture 3, 436–445.<br />

Counce, P.A., Bryant, R.J., Bergman, C.J., Bautista, R.C., Wang, Y.-J., Siebenmorgen, T.J.,<br />

Moldenhauer, K.A.K., Meullenet, J.-F.C., 2005. Rice milling quality, grain<br />

dimensions, and starch branching as affected by high night temperatures.<br />

<strong>Cereal</strong> Chemistry 82 (6), 645–648.<br />

Dilday, R.H., 1989. Milling quality <strong>of</strong> rice: cylinder speed vs. grain-moisture content<br />

at harvest. Crop <strong>Science</strong> 29, 1532–1535.<br />

Fan, J., Siebenmorgen, T.J., Yang, W., 2000. A study <strong>of</strong> head rice yield reduction <strong>of</strong><br />

long and medium grain rice varieties in relation to various harvest and drying<br />

conditions. Transactions <strong>of</strong> the ASAE 43, 1709–1714.<br />

Gravois, K.A., Moldenhauer, K.A.K., Rohman, P.C., 1991. Genotypic and genotype<br />

environment effects for rough rice and head rice yields. Crop <strong>Science</strong> 31, 907–911.<br />

Gravois, K.A., Helms, R.S., 1996. Seeding rate effects on rough rice yield, head rice,<br />

and total milled rice. Agronomy <strong>Journal</strong> 88, 82–84.<br />

Holloway, G.E., Siebenmorgen, T.J., Counce, P.A., Lu, R., 1995. Causes <strong>of</strong> multi-modal<br />

moisture content frequency distribution among rice kernels. Applied Engineering<br />

in Agriculture 11 (4), 561–565.<br />

Jodari, F., Linscombe, S.D., 1996. Grain fissuring and milling yields <strong>of</strong> rice cultivars as<br />

influenced by environmental conditions. Crop <strong>Science</strong> 36, 1496–1502.<br />

Jongkaewwattana, S., Geng, S., 2001. Inter-relationships amongst grain characteristics,<br />

grain-filling parameters and rice (Oryza sativa L.) milling quality. <strong>Journal</strong><br />

<strong>of</strong> Agronomy and Crop <strong>Science</strong> 187, 223–229.<br />

Jongkaewwattana, S., Geng, S., 2002. Non-uniformity <strong>of</strong> grain characteristics and<br />

milling quality <strong>of</strong> California rice (Oryza sativa L.) <strong>of</strong> different maturities. <strong>Journal</strong><br />

<strong>of</strong> Agronomy and Crop <strong>Science</strong> 188, 161–167.<br />

Jongkaewwattana, S., Geng, S., Brandon, D.M., Hill, J.E., 1993. Effect <strong>of</strong> nitrogen and<br />

harvest grain moisture on head rice yield. Agronomy <strong>Journal</strong> 85, 1143–1146.<br />

Kocher, M.F., Siebenmorgen, T.J., Norman, R.J., Wells, B.R., 1990. Rice kernel moisture<br />

content variation at harvest. Transactions <strong>of</strong> the ASAE 33 (2), 541–548.<br />

Leesawatwong, M., Jamjod, S., Kuo, J., Dell, B., Rerkasem, B., 2005. Nitrogen fertilizer<br />

increases seed protein and milling quality <strong>of</strong> rice. <strong>Cereal</strong> Chemistry 82, 588–593.<br />

Lin, H.X., Liang, Z.W., Sasaki, T., Yano, M., 2003. Fine mapping and characterization<br />

<strong>of</strong> quantitative trait loci Hd4 and Hd5 controlling heading date in rice. Breeding<br />

<strong>Science</strong> 53, 51–59.<br />

Liu, Z.H., Cheng, F.M., Zhang, G.P., 2005. Positional variations in phytic acid and<br />

protein content within a panicle <strong>of</strong> japonica rice. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 41,<br />

297–303.<br />

R.E. Tabien et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 122–127 127<br />

Lloyd, B.B., Cnossen, A.G., Siebenmorgen, T.J., 2001. Evaluation <strong>of</strong> two methods for<br />

separating head rice from brokens for head rice yield determination. Applied<br />

Engineering in Agriculture 17 (5), 643–648.<br />

Lu, R., Siebenmorgen, T.J., 1995. Correlation <strong>of</strong> head rice yield to selected physical<br />

and mechanical properties <strong>of</strong> rice kernels. Transaction <strong>of</strong> the ASAE 38 (2),<br />

889–894.<br />

Luh, S., Luh, B., 1991. Properties <strong>of</strong> rice caryopsis. In: Rice Production, vol. 2. Van<br />

Nostrand Reinhold, New York. 394.<br />

McCauley, G.N., Way, M.O., 2002. Drain and harvest timing affects on rice grain<br />

drying and whole-milled grain. Field Crops Research 74, 163–172.<br />

Pan, Z., Thompson, J.F., Amarantunga, K.S.P., Anderson, T., Zheng, X., 2005. Effect <strong>of</strong><br />

cooling methods and milling procedures on the appraisal <strong>of</strong> rice milling quality.<br />

Transactions <strong>of</strong> the ASAE 48 (5), 1865–1871.<br />

Perez, C.M., Juliano, B.O., Liboon, S.P., Alcantara, J.M., Cassman, K.G., 1996. Effects <strong>of</strong><br />

late nitrogen fertilizer application on head rice yield, protein content, and grain<br />

quality <strong>of</strong> rice. <strong>Cereal</strong> Chemistry 73, 556–560.<br />

Reid, J.D., Siebenmorgen, T.J., Mauromoustakos, A., 1998. Factors affecting the slope<br />

<strong>of</strong> head rice yield vs. degree <strong>of</strong> milling. <strong>Cereal</strong> Chemistry 75, 738–741.<br />

Samonte, S.O.P.B., Wilson, L.T., McClung, A.M., Tabien, R.E., 2006. Selection for stable<br />

high income grossing rice genotypes. In: Proceedings, 31st Rice Technical<br />

Working Group, p. 60. The Woodlands, Texas.<br />

Schluterman, D.A., Siebenmorgen, T.J., 2007. Relating rough rice moisture content<br />

reduction and tempering duration to head rice yield reduction. Transactions <strong>of</strong><br />

the ASABE 50 (1), 137–142.<br />

Seetanum, W., De Datta, S.K., 1973. Grain yield, milling quality and seed viability <strong>of</strong><br />

rice as influenced by time <strong>of</strong> nitrogen application and time <strong>of</strong> harvest.<br />

Agronomy <strong>Journal</strong> 65, 390–394.<br />

Siebenmorgen, T.J., Jindal, V.K., 1986. Effects <strong>of</strong> moisture adsorption on the head rice<br />

yields <strong>of</strong> long-grain rough rice. Transactions <strong>of</strong> the ASAE 29, 1767–1771.<br />

Siebenmorgen, T.J., Counce, P.A., Lu, R., Kocher, M.R., 1992. Correlation <strong>of</strong> head rice<br />

yield with individual kernel moisture content distribution at harvest. Transactions<br />

<strong>of</strong> the ASAE 35, 1879–1884.<br />

Siebenmorgen, T.J., Bautista, R.C., Counce, P.A., 2007. Optimal harvest moisture<br />

contents for maximizing milling quality <strong>of</strong> long- and medium-grain rice cultivars.<br />

Applied Engineering in Agriculture 23 (4), 517–527.<br />

Thompson, J.F., Mutters, R.G., 2006. Effect <strong>of</strong> weather and rice moisture at harvest<br />

on milling quality <strong>of</strong> California medium grain rice. Transactions <strong>of</strong> the ASABE<br />

49 (2), 435–440.<br />

Wang, F., Chen, S., Cheng, F., Liu, Y., Zhang, G., 2007. Differences in grain weight and<br />

quality within a rice (Oryza sativa L.) panicle as affected by panicle type and<br />

source-sink relation. <strong>Journal</strong> <strong>of</strong> Agronomy and Crop <strong>Science</strong> 193, 63–73.<br />

Xiao, J., Li, L., Grandillo, S., Ahn, S.N., Yuan, L.P., Tanksley, S.D., McCouch, S.R., 1998.<br />

Identification <strong>of</strong> trait improving quantitative loci alleles from wild rice relatives,<br />

Oryza rufipogon. Genetics 150, 899–909.<br />

Xiong, L., Lui, K., Dai, X., Xu, C., Zhang, Q., 1999. Identification <strong>of</strong> genetics factors<br />

controlling domestication-related traits in rice using F2 population <strong>of</strong><br />

a cross between Oryza sativa and O. rufipogon. Theoretical and Applied<br />

Genetics 98, 243–251.<br />

Zhang, Y., Luo, L., Xu, C., Zhang, Q., Xing, Y., 2006. Quantitative trait loci for panicle<br />

size, heading date and plant height co-segregating in trait performance<br />

derived near isogenic lines <strong>of</strong> rice (Oryza sativa). Theoretical and Applied<br />

Genetics 113, 361–368.<br />

Zhuang, J.Y., Lin, H.X., Lu, J., Qian, H.R., Hittalmani, S., Huang, N., Zheng, K.L., 1997.<br />

Analysis <strong>of</strong> QTL x environment interaction for yield components and plant<br />

height. Theoretical and Applied Genetics 95, 799–808.


Texture, processing and organoleptic properties <strong>of</strong> chickpea-fortified spaghetti<br />

with insights to the underlying mechanisms <strong>of</strong> traditional durum pasta quality<br />

Jennifer Ann Wood *<br />

NSW Department <strong>of</strong> Primary Industries, Tamworth Agricultural Institute, 4 Marsden Park Road, Calala, NSW 2340, Australia<br />

article info<br />

Article history:<br />

Received 10 March 2008<br />

Received in revised form 9 July 2008<br />

Accepted 22 July 2008<br />

Keywords:<br />

Cicer arietinum L<br />

Pasta quality<br />

Rheology<br />

Sensory<br />

1. Introduction<br />

abstract<br />

Nutritionists recommend pulses (grain legumes) such as<br />

chickpea (Cicer arietinum) in the diet as they have many nutritional<br />

benefits. Chickpea seed has a high protein digestibility, contains<br />

high levels <strong>of</strong> complex carbohydrates (low glycaemic index), is rich<br />

in vitamins and minerals and is relatively free from anti-nutritional<br />

factors (Muzquiz and Wood, 2007; Wood and Grusak, 2007).<br />

Consumers are becoming increasingly health conscious and<br />

while many admit to knowing pulses are good for them, they are<br />

not sure how to use them in their diet. There is also a perception<br />

that cooking pulses is difficult and/or time consuming.<br />

Australian production <strong>of</strong> desi chickpea is currently around<br />

200,000 tonnes per year (Knights et al., 2007), yet less than 0.5% <strong>of</strong><br />

this is consumed by Australians. In comparison, pasta, produced<br />

from durum wheat (Triticum turgidum) flour, is consumed worldwide.<br />

It is relatively non-perishable, inexpensive, easy to prepare,<br />

and readily accepted by all age groups.<br />

The nutritional benefits, deep yellow–orange cotyledon colour<br />

and lack <strong>of</strong> <strong>of</strong>fensive ‘beany’ flavours compared to other pulses<br />

make chickpea a suitable candidate for incorporation with durum<br />

* Tel.: þ61 2 67631157; fax: þ61 2 67631222.<br />

E-mail address: jenny.wood@dpi.nsw.gov.au<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.07.016<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 128–133<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

Nutritionally enhanced spaghetti was prepared from durum semolina fortified with 0–30% desi chickpea<br />

‘besan’ flour. This study examined the dough rheology, processing ease and quality attributes <strong>of</strong> the<br />

fortified spaghetti including protein, starch, texture (firmness, resilience and stickiness), colour, cooking<br />

loss, and organoleptic acceptability. Chickpea-fortified spaghetti was acceptable to consumers, had<br />

reasonable pasta quality, including lower cooking loss and less stickiness than the control spaghetti and<br />

retained firmness better than durum after refrigeration. This study suggests that chickpea-fortified<br />

spaghetti may be suited to uses such as fresh pasta, in soups, canning, and microwave re-heating. In<br />

addition, this study has added to the understanding <strong>of</strong> the underlying mechanisms <strong>of</strong> pasta quality. The<br />

main findings were: (1) gluten content/composition appears to be more important than protein content<br />

for pasta firmness; (2) the protein–polysaccharide matrix appears to be more important than the starch<br />

composition for cooking loss; (3) increased protein and amylose contents were associated with<br />

decreased pasta stickiness; (4) cooking loss and stickiness were not necessarily as strongly related as<br />

commonly believed. Further research into these theories is necessary to fully understand the underlying<br />

mechanisms <strong>of</strong> pasta quality.<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

into pasta. In this way consumers could increase their consumption<br />

<strong>of</strong> pulses with little extra effort or thought.<br />

The inclusion <strong>of</strong> pulses in cereal based foods is known to<br />

increase the nutritive value by improving protein content and<br />

lysine availability (Bahnassey et al., 1986; Hernandez and Sotelo,<br />

1984; Kurien et al., 1971; Reyes-Moreno et al., 2004; Siddique et al.,<br />

1996; Wood and Grusak, 2007). Several studies have examined<br />

various aspects <strong>of</strong> chickpea incorporation into pasta (Goni and<br />

Valentin-Gamazo, 2003; Sabanis et al., 2006) however the endproduct<br />

qualities <strong>of</strong> pasta produced from ‘besan’ (dehulled desi<br />

chickpea flour) has not been thoroughly investigated.<br />

Goni and Valentin-Gamazo (2003) showed that spaghetti containing<br />

25% chickpea flour had a significantly lower glycaemic<br />

index (GI) than traditional durum spaghetti. Chickpea inclusion<br />

also increased the mineral and fat content without affecting the<br />

total starch content. Zhao et al. (2005) incorporated 5-30% <strong>of</strong><br />

different pulse flours into spaghetti and found that firmness and<br />

colour intensity increased, while overall quality decreased.<br />

However, the chickpeas used in this study were kabuli types and<br />

the seed coats were not removed prior to grinding. Sabanis et al.<br />

(2006) investigated 5-50% inclusion <strong>of</strong> chickpea flour in durum<br />

lasagne and found that the physical properties <strong>of</strong> the dough were<br />

improved; however, processing, handling and cooking characteristics<br />

deteriorated with the higher substitution levels.


This study aimed to more comprehensively identify whether,<br />

besan, could be successfully incorporated into durum spaghetti to<br />

obtain a convenient health product with organoleptic properties<br />

acceptable to consumers. Furthermore, this research provides<br />

additional information to the existing literature on the effects <strong>of</strong><br />

starch and protein on the quality <strong>of</strong> traditional durum pasta.<br />

2. Experimental<br />

2.1. Samples<br />

Unconditioned seed (50.0 g) <strong>of</strong> desi chickpea (cv. Amethyst)<br />

was dehulled in the ‘Sheller’ component (attrition-style) <strong>of</strong> an SK<br />

Engineering mill (SK Engineering and Allied Works, Bahraich,<br />

India) to produce dhal (split cotyledons) according to Wood et al.<br />

(2008). The resulting dhal was subsequently milled through a<br />

1.0 mm sieve in a Newport Scientific mill (Newport Scientific<br />

Pty Ltd, Narabeen, Australia) to obtain coarse, besan flour.<br />

Commercial durum semolina was obtained from Goodman Fielders<br />

(Tamworth, Australia) and blended with 0, 10, 15, 20, 25 and 30% <strong>of</strong><br />

the chickpea flour by weight, in duplicate. No significant difference<br />

was found between the moisture contents <strong>of</strong> the blended flours<br />

as determined by the AACC approved method 44-15A (AACC,<br />

1995b).<br />

2.2. Dough quality and pasta colour evaluation<br />

Dough properties were determined with a farinograph and<br />

constant dough weight using the AACC approved method 54-21<br />

(AACC, 1995d). Pasta was produced under vacuum with standard<br />

33.3% water addition and dried in a pasta oven at low temperature<br />

according to Wood et al. (2001). Pasta colour (L*, a*, b*) <strong>of</strong> multiple<br />

layers <strong>of</strong> parallel spaghetti strands were measured (mean <strong>of</strong> six<br />

readings) using a Minolta chroma meter CR-310 (Minolta Camera<br />

Co., Osaka, Japan) on both dry and cooked spaghetti. The change in<br />

colour due to cooking was determined by calculating<br />

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

DE* ¼ ðDL Þ 2 þðDa Þ 2 þðDb Þ 2<br />

q<br />

2.3. Cooked pasta quality evaluation<br />

The quality <strong>of</strong> the cooked spaghetti was evaluated according to<br />

Wood et al. (2001). Spaghetti strands (10 g, 7 cm lengths) were<br />

placed in 250 ml <strong>of</strong> vigorously boiling distilled water containing<br />

5 ml <strong>of</strong> stock solution (0.125 g NaHCO3, 175 g NaCl in 500 ml<br />

distilled H2O). Optimum cooking time <strong>of</strong> the spaghetti was recorded<br />

when the core in the middle <strong>of</strong> the spaghetti strand was no<br />

longer visible when squashed between two perspex sheets. Texture<br />

analysis (firmness, resilience and stickiness) and cooking loss were<br />

performed on spaghetti cooked to optimum cooking time according<br />

to Wood et al. (2001). Cooking loss was calculated as described<br />

by Matsuo et al. (1992).<br />

2.4. Rapid visco analyser (RVA)<br />

Spaghetti was milled into flour through a 0.5 mm sieve in<br />

a Newport Scientific mill (Newport Scientific Pty, Ltd, Narabeen,<br />

Australia). Moisture content <strong>of</strong> the flours and spaghetti blends were<br />

determined by the AACC approved method 44-15A (AACC, 1995b).<br />

Pasting properties were evaluated in duplicate using 29 g <strong>of</strong> sample<br />

(13.8% solids, moisture corrected) in the Rapid Visco Analyser<br />

(Newport Scientific Pty, Ltd, Narabeen, Australia). The temperature<br />

was held at 25 C for 2 min, increased to 95 C over 5 min, held<br />

at 95 C for 3 min and decreased to 25 C over 5 min. Peak<br />

J.A. Wood / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 128–133 129<br />

time, peak viscosity, breakdown, final viscosity and setback were<br />

recorded.<br />

2.5. Protein content and amino acid analysis<br />

Total protein content <strong>of</strong> the fortified products and controls were<br />

determined in duplicate with a LECO nitrogen analyser using AACC<br />

approved method 46-30 (AACC, 1995c). Protein determinations<br />

were based on nitrogen factors <strong>of</strong> 5.7 for wheat flour and 6.25<br />

for chickpea flour. Nitrogen factors for spaghetti blends were<br />

adjusted on the basis <strong>of</strong> the relative proportion <strong>of</strong> proteins in the<br />

ingredients. Following extraction by acid hydrolysis, amino acid<br />

analysis <strong>of</strong> the 0%, 15% and 30% fortified spaghetti were determined<br />

by HPLC using AACC approved method 07-01 (AACC, 1995a).<br />

Tryptophan content was not determined.<br />

2.6. Amylose content<br />

The amylose content <strong>of</strong> the flours and spaghetti were determined<br />

using the Megazyme amylose and amylopectin assay kit<br />

(Megazyme International, Ireland).<br />

2.7. Sensory evaluation<br />

Sensory evaluation was performed on cooked spaghetti cooled<br />

in distilled water (5 min), drained, covered with plastic, and<br />

refrigerated overnight. Spaghetti fortified with %0, %15 and 30%<br />

chickpea flour was randomly numbered and scored by 27 untrained<br />

volunteers the following day. Texture analysis was performed on<br />

this spaghetti as previously described. A hedonic scale was chosen<br />

for the sensory evaluation questionnaire. The volunteers were<br />

asked to give a ranking <strong>of</strong> the attributes (colour, firmness, flavour,<br />

overall acceptability) for the three samples on a linear scale. The<br />

distance from the undesirable end <strong>of</strong> the scale was measured for<br />

each sample and subjected to ANOVA.<br />

2.8. Statistical analyses<br />

All the parameters were analysed in duplicate. The data was<br />

statistically analysed by analysis <strong>of</strong> variance (ANOVA). Each<br />

parameter was tested for significance (P < 0.05) between the<br />

fortified and control spaghetti samples. When significant differences<br />

were found, the least significant difference (LSD) test was<br />

used to determine the differences among means. A correlation<br />

coefficient between cooking loss and stickiness was determined<br />

using Micros<strong>of</strong>t Excel.<br />

3. Results and discussion<br />

3.1. Dough and processing properties<br />

Increasing chickpea fortification significantly (P < 0.05) decreased<br />

water absorption and resulted in longer development times and less<br />

stable doughs (Table 1). Similar farinograph trends have been reported<br />

previously for wheat incorporating legume flours or their protein<br />

concentrates (Lorenz et al., 1979; Rasmay et al., 2000; Yanez-Farias<br />

et al., 1999). Many <strong>of</strong> these effects can be attributed to weakening <strong>of</strong><br />

the gluten matrix due to the incorporation <strong>of</strong> chickpea flour which<br />

contains no gluten. Chickpea proteins are comprised mainly <strong>of</strong> globulins<br />

(53–60%) with lesser concentrations <strong>of</strong> albumins, prolamins and<br />

glutelins (Dhawan et al., 1991).<br />

Fortification made the dough particles increasingly sticky,<br />

causing them to aggregate during mixing. This made extrusion to<br />

produce the spaghetti increasingly difficult. For this reason,<br />

spaghetti production above 30% fortification was not undertaken.<br />

Chickpea flour contains significant levels <strong>of</strong> soluble non-starch


130<br />

Table 1<br />

Effect <strong>of</strong> chickpea flour fortification on spaghetti quality parameters<br />

Quality parameters Fortification (chickpea %) P<br />

0% 10% 15% 20% 25% 30%<br />

Water absorption (%) 57.6 a<br />

57.6 a<br />

57.2 b<br />

56.3 c<br />

55.8 d<br />

54.2 e<br />

0.00<br />

DDT (min) 3.75 d<br />

4.50 c<br />

4.50 c<br />

4.50 c<br />

5.00 b<br />

5.75 a<br />

0.00<br />

B10 (BU) 50 d<br />

60 c<br />

70 b<br />

70 b<br />

75 ab<br />

80 a<br />

0.00<br />

Protein (%) 12.43 e<br />

14.33 d<br />

13.82 d<br />

15.37 c<br />

16.99 b<br />

17.42 a<br />

0.00<br />

Glutamine/glutamicacid (g/100 g) 4.81 a<br />

NA 4.48 b<br />

NA NA 4.22 c<br />

0.00<br />

Proline (g/100 g) 1.67 a<br />

NA 1.57 b<br />

NA NA 1.51 c<br />

0.00<br />

Lysine(g/100 g) 0.22 c<br />

NA 0.36 b<br />

NA NA 0.62 a<br />

0.00<br />

Cysteine (g/100 g) 0.33 a<br />

NA 0.34 a<br />

NA NA 0.33 a<br />

0.12<br />

Methionine (g/100 g) 0.21 a<br />

NA 0.21 a<br />

NA NA 0.20 a<br />

0.08<br />

Colour, L* (dry) 57.73 a<br />

57.25 ab<br />

55.98 abc<br />

55.20 bc<br />

54.61 c<br />

53.96 c<br />

0.00<br />

Colour, a* (dry) 3.56 c<br />

6.25 b<br />

7.13 ab<br />

7.61 ab<br />

8.00 a<br />

8.28 a<br />

0.00<br />

Colour, b* (dry) 40.74 a<br />

33.76 b<br />

36.33 ab<br />

36.72 ab<br />

37.86 ab<br />

38.74 ab<br />

0.04<br />

Colour, L* (cooked) 78.37 a<br />

75.85 b<br />

74.01 c<br />

73.10 cd<br />

72.33 d<br />

71.26 e<br />

0.00<br />

Colour, a* (cooked) 0.33 e<br />

3.42 d<br />

4.66 c<br />

5.21 bc<br />

6.10 ab<br />

6.52 a<br />

0.00<br />

Colour, b* (cooked) 32.52 a<br />

27.03 c<br />

28.65 bc<br />

29.95 abc<br />

30.27 ab<br />

30.61 ab<br />

0.01<br />

Colour change, DE 32.74 a<br />

28.16 b<br />

28.18 b<br />

27.07 b<br />

27.21 b<br />

27.19 b<br />

0.01<br />

Amylose (%) 23.01 b<br />

22.52 b<br />

23.63 b<br />

23.54 b<br />

23.45 b<br />

26.65 a<br />

0.01<br />

Peak time (min) 8.50 a<br />

8.30 a<br />

8.30 a<br />

7.73 b<br />

7.70 b<br />

7.70 b<br />

0.02<br />

Peak viscosity (cP) 2191 c<br />

2303 b<br />

2231 c<br />

2246 bc<br />

2518 a<br />

2529 a<br />

0.00<br />

Breakdown (cP) 922 c<br />

957 c<br />

959 c<br />

965 c<br />

1089 b<br />

1180 a<br />

0.00<br />

Final viscosity (cP) 5976 a<br />

5802 c<br />

5891 b<br />

4532 e<br />

5345 d<br />

5407 d<br />

0.00<br />

Setback (cP) 4707 a<br />

4456 b<br />

4619 a<br />

3250 d<br />

3916 c<br />

4058 c<br />

0.03<br />

Firmness (g) 624.6 a<br />

641.1 ab<br />

671.6 a<br />

574.8 bc<br />

550.9 cd<br />

510.1 d<br />

0.01<br />

Resilience (g) 62.28 a<br />

46.02 ab<br />

51.81 a<br />

33.07 bc<br />

15.45 c<br />

13.93 c<br />

0.00<br />

Stickiness (gs) 5.91 a<br />

4.72 bc<br />

4.74 bc<br />

4.67 bc<br />

4.81 b<br />

3.81 c<br />

0.00<br />

Cooking loss (%) 5.15 a<br />

4.84 ab<br />

4.64 ab<br />

4.66 ab<br />

4.53 b<br />

4.50 b<br />

0.03<br />

NA, not analysed. For each quality parameter, means within rows followed by different letters are significantly different (P < 0.05) by least significant difference (LSD) test.<br />

polysaccharides (NSP), about ten times higher than bread wheat<br />

flour (Naivikul and D’Appolonia, 1979), and this may have<br />

contributed to the increased stickiness. Based on these observations<br />

and the lower farinograph water absorptions <strong>of</strong> the chickpeafortified<br />

doughs, it is possible that reducing the amount <strong>of</strong> water<br />

added may improve mixing and extrudability to some extent.<br />

3.2. Dry pasta properties<br />

Protein content and amino acid composition generally increased<br />

(P < 0.05) with fortification. Protein content <strong>of</strong> control spaghetti was<br />

12.4% compared to 17.4% for the 30% fortified spaghetti (Table 1). A<br />

similar trend was observed by Zhao et al. (2005). Chickpea blends<br />

showed increased levels <strong>of</strong> all amino acids except for cysteine and<br />

methionine (no significant change) and glutamine/glutamic acid<br />

and proline which significantly (P < 0.05) decreased with fortification<br />

(Table 1). Chickpea flour is higher in protein but limiting in<br />

cysteine and methionine, hence no significant change in their<br />

contents were observed in the fortified spaghetti. Glutamine/glutamic<br />

acid and proline are indicative <strong>of</strong> gluten-type proteins which<br />

explains their lower levels since chickpea flour is gluten free. On the<br />

other hand, wheat is limiting in lysine. Lysine content significantly<br />

(P < 0.05) increased by 64 and 182% in the 15 and 30% fortified<br />

spaghetti respectively (Fig. 1).<br />

The colour <strong>of</strong> all the dried spaghetti generally became significantly<br />

(P < 0.05) less bright (lower L*), more red (higher a*) and less<br />

yellow (lower b*) as the percentage <strong>of</strong> chickpea flour increased<br />

(Fig. 2). The 10% and 15% chickpea-fortified spaghetti were rated by<br />

the sensory panel as visually similar to the control, however the<br />

spaghetti fortified with 20% or more tended to display an undesirable<br />

brownish tint (supported by L*a*b* measurements, Table 1).<br />

This is similar to the findings <strong>of</strong> Zhao et al. (2005).<br />

3.3. Rapid visco analysis<br />

The chickpea besan flour slurry had a much lower and flatter RVA<br />

viscosity pr<strong>of</strong>ile than the durum semolina slurry (Fig. 3). Wood<br />

J.A. Wood / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 128–133<br />

(2004) found similar comparisons in RVA pr<strong>of</strong>iles between chickpea<br />

besan flour (cv. Howzat) and a hard bread wheat flour (cv. Drysdale).<br />

Chickpea has a lower starch content with a higher proportion <strong>of</strong><br />

amylose and a greater non-starch polysaccharide (NSP) content and<br />

lipid content than durum and bread wheats (Fabriani and Lintas,<br />

1988; Wood and Grusak, 2007). This probably explains much <strong>of</strong> the<br />

differences in the RVA pr<strong>of</strong>iles <strong>of</strong> the besan and semolina slurries.<br />

The 10% and 15% fortified spaghetti displayed similar pasting<br />

pr<strong>of</strong>iles to the control spaghetti. They generally had higher peak<br />

viscosities and lower final viscosities than the control spaghetti<br />

(Table 1). The higher peak viscosities indicate larger water binding<br />

capacities <strong>of</strong> the starches in these blends. The 25% and 30% fortified<br />

spaghetti peaked earlier and higher than the control spaghetti with<br />

slightly lower final viscosities and much lower setbacks. The slurries<br />

<strong>of</strong> these blends swelled faster and bound more water; however,<br />

less retrogradation occurred, resulting in a less viscous gel after<br />

cooking and cooling. Most <strong>of</strong> these observations can be explained<br />

by the higher NSP content <strong>of</strong> the fortified spaghetti. NSP has a high<br />

Lysine Content (g / 100 g)<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

Lysine Content<br />

Protein Content<br />

0 15 30<br />

% Chickpea in pasta<br />

Fig. 1. Protein and lysine content <strong>of</strong> spaghetti increases with desi chickpea flour<br />

fortification.<br />

17<br />

16<br />

15<br />

14<br />

13<br />

12<br />

11<br />

10<br />

Protein Content (%)


L* (brightness)<br />

60<br />

59<br />

58<br />

57<br />

56<br />

55<br />

54<br />

53<br />

52<br />

51<br />

50<br />

L<br />

b<br />

0 10 15 20 25 30<br />

% Chickpea in pasta (dry)<br />

Fig. 2. Colour changes in dry spaghetti with increasing desi chickpea flour fortification.<br />

water absorbing capacity which has been shown to increase the<br />

viscosity <strong>of</strong> bread wheat slurries (Sasaki et al., 2000). On cooling,<br />

the NSP can impede the formation <strong>of</strong> a three dimensional starch<br />

network, preventing hydrogen bonds between amylose and<br />

amylopectin from forming during retrogradation, resulting in a less<br />

viscous gel (Kim and D’Appolonia, 1977; Yoshimura et al., 1996).<br />

The 20% fortified spaghetti pasting curve did not track between<br />

the 10–15% and 25–30% spaghetti as expected (Fig. 3). Instead it<br />

peaked at a similar viscosity to the control and had a much lower<br />

final viscosity. This indicates that the starch granules in the blend<br />

swelled similarly to the control, but again with less retrogradation<br />

and reduced viscosity after cooking and cooling. The sample was reanalysed<br />

twice more but all generated the same pasting curve. This<br />

unexpected result is difficult to explain. There may be some sort <strong>of</strong><br />

synergistic process between the besan and semolina constituents<br />

occurring at 20% fortification.<br />

Viscosity cP<br />

8000<br />

6000<br />

4000<br />

2000<br />

0<br />

0 4 8 12 16 20<br />

Time mins<br />

Semolina<br />

0% Pasta<br />

10% Pasta<br />

30% Pasta<br />

20% Pasta<br />

Chickpea<br />

Flour<br />

Fig. 3. RVA pr<strong>of</strong>ile <strong>of</strong> raw flours (durum semolina and desi chickpea ‘course besan’)<br />

and spaghetti with different levels <strong>of</strong> chickpea fortification.<br />

J.A. Wood / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 128–133 131<br />

120<br />

80<br />

40<br />

0<br />

41<br />

39<br />

37<br />

35<br />

33<br />

31<br />

29<br />

27<br />

25<br />

b* (yellowness)<br />

Temp 'C<br />

3.4. Cooked pasta properties<br />

The optimum cooking time was determined to be 10.5 minutes<br />

for all the spaghetti samples. After cooking, all the spaghetti<br />

samples were more bright (higher L*) and less yellow (lower b*)<br />

due to leaching and/or degradation <strong>of</strong> colour pigments such as<br />

carotenoids and xanthophyll. Bleaching was most pronounced in<br />

the control spaghetti, as calculated by DE* (Table 1). Despite this,<br />

the cooked chickpea-fortified spaghetti samples were generally less<br />

bright (L*), more red (a*) and less yellow (b*) than the control<br />

spaghetti (P > 0.05). However, there was no statistical difference<br />

between the 0%, 15% and 30% fortified spaghetti in panellist<br />

perceptions <strong>of</strong> cooked spaghetti colour.<br />

The 10% and 15% substituted spaghetti had similar firmness and<br />

resilience to the durum control spaghetti. Similarly, Zhao et al.<br />

(2005) found no significant difference in the firmness <strong>of</strong> most <strong>of</strong><br />

their chickpea–spaghetti blends compared to the durum control.<br />

However, our results showed that firmness and resilience<br />

decreased with fortification <strong>of</strong> 20–30% (Table 1; Fig. 4).<br />

Increasing the protein content <strong>of</strong> durum spaghetti has been<br />

shown to increase firmness (Nobile et al., 2005; Sissons et al.,<br />

2005). In addition, lowering the amylose content has been shown<br />

to decrease firmness (Gianibelli et al., 2005). This suggests that the<br />

fortified spaghetti firmness should have increased as protein<br />

content and amylose content increased due to fortification. This<br />

appears to be the case up to a threshold <strong>of</strong> 15% fortification.<br />

However, chickpea flour contains no gluten. As more chickpea flour<br />

was added, the semolina gluten was effectively diluted (observed as<br />

a decrease in glutamine/glutamic acid contents; P < 0.05) leading<br />

to weakening <strong>of</strong> the gluten matrix and a decrease in spaghetti<br />

firmness. This result suggests that it may not be the protein content<br />

per se (nor the high amylose content), rather the gluten content and<br />

possibly gluten composition that may be more important in<br />

determining spaghetti firmness. This finding is consistent with the<br />

work <strong>of</strong> Sissons et al. (2005) who showed that gluten content<br />

increased spaghetti firmness with no consistent trend relating to<br />

glutenin/gliadin composition. The larger NSP content <strong>of</strong> the<br />

chickpea flour may similarly contribute to weakening <strong>of</strong> the protein<br />

matrix.<br />

Legume starches generally contain more amylose than cereal<br />

starches. Zhao et al. (2005) found an increase in cooking loss<br />

(measured by solids loss) with fortification but this method is not<br />

comparable to amylose loss. The cooking loss test used in this study<br />

mainly measures the amount <strong>of</strong> amylose leached into the water<br />

during cooking (Matsuo et al., 1992). Sharma et al. (2002) and<br />

Gianibelli et al. (2005) found cooking loss to decrease in spaghetti<br />

with a low amylose content. Hence, it was initially presumed that<br />

the chickpea-fortified spaghetti would have increased cooking loss<br />

Firmness (g)<br />

700<br />

650<br />

600<br />

550<br />

500<br />

450<br />

400<br />

Firmness (g)<br />

Resilience (g)<br />

0 10 15 20 25 30<br />

% Chickpea in pasta<br />

Fig. 4. General decrease in firmness and resilience <strong>of</strong> spaghetti with increasing desi<br />

chickpea flour fortification.<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Resilience


132<br />

due to the higher amylose contents. However, cooking loss was<br />

shown to decrease with fortification (Table 1; Fig. 5). This result<br />

suggests that amylose content is not the major factor in cooking<br />

loss.<br />

Sissons et al. (2005) found no relationship <strong>of</strong> cooking loss with<br />

protein content in durum spaghetti. However, our results showed<br />

decreased cooking loss with increased protein and NSP contents <strong>of</strong><br />

fortified spaghetti. This suggests that the protein–polysaccharide<br />

matrix (involving both starch and NSP) may be responsible for the<br />

retention <strong>of</strong> amylose during spaghetti cooking and not necessarily<br />

the starch composition per se, as suggested by Sissons et al. (2005).<br />

The chickpea-fortified spaghetti was also less sticky (P < 0.05)<br />

than the control (Fig. 5). Pasta surface stickiness is believed to be<br />

influenced by both the surface structure <strong>of</strong> the spaghetti strand and<br />

starch exuded onto the strand surface during cooking (Cunin et al.,<br />

1995; Dexter et al., 1985; Perovic, 2000). Lowering the amylose<br />

content <strong>of</strong> spaghetti has been shown to increase stickiness (Gianibelli<br />

et al., 2005; Sharma et al., 2002). Furthermore, increasing<br />

the protein content <strong>of</strong> pasta has been associated with decreased<br />

stickiness (Nobile et al., 2005; Sissons et al., 2005). Hence, the<br />

reduced stickiness <strong>of</strong> the chickpea -fortified spaghetti is probably<br />

a result <strong>of</strong> both higher protein and higher amylose contents.<br />

However, whilst both the cooking loss and stickiness generally<br />

improved with chickpea fortification, they were only weakly<br />

correlated with one another (r ¼ 0.51).<br />

Since cooking loss, stickiness and firmness all decreased with<br />

the addition <strong>of</strong> besan chickpea flour, it appears that gluten, per se,<br />

may have little effect on the retention <strong>of</strong> amylose and other<br />

carbohydrates during cooking. The protein–polysaccharide matrix<br />

as a whole may be more likely to influence spaghetti quality in<br />

terms <strong>of</strong> firmness, stickiness and cooking loss. Additional components<br />

<strong>of</strong> chickpea flour that may aid the protein–polysaccharide<br />

matrix in carbohydrate retention include monomeric proteins and<br />

starch-bound phospholipids. This clearly requires further investigation<br />

and the knowledge may be useful for genotype selection in<br />

durum wheat breeding programs.<br />

3.5. Cooked and refrigerated pasta properties<br />

The sensory evaluation study gave an indication <strong>of</strong> consumer<br />

preference. Most panellists scored the 15 and 30% cooked fortified<br />

spaghetti to be equally acceptable as the durum control. No<br />

statistically significant (P < 0.05) difference could be found for<br />

colour, flavour or overall acceptability. However, the 30% blend was<br />

scored significantly (P < 0.05) firmer than the 0% and 15% blends.<br />

This was unexpected, as the texture analyser results <strong>of</strong> freshly<br />

cooked spaghetti showed a decreasing firmness with higher<br />

chickpea fortification.<br />

Stickiness (gs)<br />

6.0<br />

5.5<br />

5.0<br />

4.5<br />

4.0<br />

3.5<br />

Stickiness (gs)<br />

Cooking Loss (%)<br />

0 10 15 20 25 30<br />

% Chickpea in pasta<br />

Fig. 5. General decrease in cooking loss and stickiness <strong>of</strong> spaghetti with increasing<br />

desi chickpea flour fortification.<br />

J.A. Wood / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 128–133<br />

5.2<br />

5.1<br />

5.0<br />

4.9<br />

4.8<br />

4.7<br />

4.6<br />

4.5<br />

4.4<br />

% Cooking Loss<br />

Firmness (g)<br />

800.0<br />

700.0<br />

600.0<br />

500.0<br />

400.0<br />

300.0<br />

200.0<br />

100.0<br />

0.0<br />

When re-assessed by the texture analyser, the panel’s findings<br />

were confirmed. The fortified spaghetti were slightly firmer than<br />

the control after refrigeration (P < 0.05). Whilst all the spaghetti<br />

s<strong>of</strong>tened substantially after refrigeration, the effect was greater in<br />

the control durum spaghetti: a firmness decrease <strong>of</strong> 69% for the<br />

durum spaghetti compared to 66% for the fortified spaghetti<br />

(Fig. 6). Refrigeration differentially influenced the chickpea-fortified<br />

spaghetti and durum control spaghetti to alter the firmness<br />

trend.<br />

Riva et al. (2000) found that the rate and degree <strong>of</strong> starch<br />

retrogradation was exaggerated at cold temperatures. This too<br />

could cause firmer spaghetti, if the chickpea-fortified spaghetti<br />

were more predisposed to retrogradation under cold temperatures<br />

then the durum control. However, the RVA results showed the<br />

fortified spaghetti to have less retrogradation than the durum<br />

control.<br />

No difference in spaghetti diameter between the durum and<br />

fortified spaghetti was detected. However, water absorption may<br />

have occurred by diffusion through the gelatinised system to cause<br />

an increase in weight with negligible volume change (Riva et al.,<br />

2000). If the chickpea blends were more susceptible to water<br />

absorption, spaghetti density would increase which may increase<br />

firmness. This is a possibility, as the RVA results showed the fortified<br />

spaghetti to have a larger water binding capacity during<br />

cooking and this may remain after cooling. This finding requires<br />

further investigation.<br />

4. Conclusion<br />

Normal Fridge<br />

100% Durum pasta<br />

20% Chickpea pasta<br />

Fig. 6. Effect <strong>of</strong> refrigeration on the firmness <strong>of</strong> 100% durum spaghetti and spaghetti<br />

fortified with 20% desi chickpea flour.<br />

Chickpea-fortified spaghetti was acceptable to consumers and<br />

provided an enhanced nutritional status via the amino acid pr<strong>of</strong>ile.<br />

Lysine content increased by 64 and 182% in the 15 and 30% blends<br />

respectively whilst the total protein content and the content <strong>of</strong><br />

most amino acids increased with fortification. Spaghetti processing<br />

and handling characteristics deteriorated as the level <strong>of</strong> fortification<br />

increased. Functional dough properties and spaghetti firmness<br />

were generally hindered by increasing amounts <strong>of</strong> chickpea flour.<br />

However, spaghetti stickiness improved with increasing fortification<br />

and cooking loss was reduced. Chickpea-fortified spaghetti<br />

retained firmness much better than durum after refrigeration. A<br />

marketing advantage may exist if this desirable firmness is retained<br />

when blended pasta is subjected to canning, microwave re-heating<br />

and inclusion in soups. This firmness retention property may also<br />

be <strong>of</strong> interest to fresh pasta manufacturers. Chickpea-fortified pasta<br />

would make a cheap, attractive and convenient health food which<br />

is acceptable to consumers in western society. In addition, this


study has added to the understanding <strong>of</strong> the underlying mechanisms<br />

<strong>of</strong> pasta quality. The findings were: (1) gluten content/<br />

composition appears to be more important than protein content for<br />

pasta firmness; (2) the protein–polysaccharide matrix appears to<br />

be more important than the starch composition for cooking loss;<br />

(3) supportive <strong>of</strong> previous findings that increased protein and<br />

amylose contents are associated with decreased pasta stickiness;<br />

(4) cooking loss and stickiness are not necessarily as strongly<br />

related as commonly believed. Further research into these theories<br />

is necessary to fully understand the underlying mechanisms <strong>of</strong><br />

pasta quality.<br />

Acknowledgements<br />

The author wishes to thank the Grains Research and Development<br />

Corporation (GRDC) for funding (DAN402), Goodman<br />

Fielders, Tamworth for providing semolina, L. Ayre (NSW Department<br />

<strong>of</strong> Primary Industries, Tamworth) for statistical analyses, and<br />

M. Sissons (NSW Department <strong>of</strong> Primary Industries, Tamworth) for<br />

use <strong>of</strong> the pasta extruder/dryer.<br />

References<br />

American Association <strong>of</strong> <strong>Cereal</strong> Chemists (AACC), 1995a. Approved Methods <strong>of</strong> the<br />

AACC. Method 07-01, ninth ed. AACC, St Paul, MN.<br />

American Association <strong>of</strong> <strong>Cereal</strong> Chemists (AACC), 1995b. Approved Methods <strong>of</strong> the<br />

AACC. Method 44-15A, ninth ed. AACC, St Paul, MN.<br />

American Association <strong>of</strong> <strong>Cereal</strong> Chemists (AACC), 1995c. Approved Methods <strong>of</strong><br />

the AACC. Method 46-30, ninth ed. AACC, St Paul, MN.<br />

American Association <strong>of</strong> <strong>Cereal</strong> Chemists (AACC), 1995d. Approved Methods <strong>of</strong> the<br />

AACC. Method 54-21, ninth ed. AACC, St Paul, MN.<br />

Bahnassey, Y., Khan, K., Harrold, R., 1986. Fortification <strong>of</strong> spaghetti with edible<br />

legumes. 1. Physicochemical, antinutritional, amino acid, and mineral composition.<br />

<strong>Cereal</strong> Chem 63, 210–215.<br />

Cunin, C., Handschin, S., Walther, P., Escher, F., 1995. Structural changes <strong>of</strong> starch<br />

during cooking <strong>of</strong> durum wheat pasta. Lebensm. Wiss. Technol 28, 323–328.<br />

Dexter, J.E., Matsuo, R.R., MacGregor, A.W., 1985. Relationship <strong>of</strong> instrumental<br />

assessment <strong>of</strong> spaghetti cooking quality to the type and the amount <strong>of</strong> material<br />

rinsed from cooked spaghetti. J. <strong>Cereal</strong> Sci 3, 39–53.<br />

Dhawan, K., Malhotra, S., Dahiya, B., Singh, D., 1991. Seed protein fractions and amino<br />

acid composition in gram (Cicer arietinum). Plant Food Hum. Nutr 41, 225–232.<br />

Fabriani, G., Lintas, C. (Eds.), 1988. Durum Wheat: Chemistry and Technology. AACC,<br />

St. Paul, MN.<br />

Gianibelli, M.C., Sissons, M.J., Batey, I.L., 2005. Effect <strong>of</strong> source and proportion <strong>of</strong><br />

waxy starches on pasta cooking quality. <strong>Cereal</strong> Chem 82, 321–327.<br />

Goni, I., Valentin-Gamazo, C., 2003. Chickpea flour ingredient slows glycemic<br />

response to pasta in healthy volunteers. Food Chem 81, 511–515.<br />

Hernandez, M., Sotelo, A., 1984. Nutritional evaluation <strong>of</strong> wheat flour cookies<br />

supplemented with chickpea, cheese whey and amino acids. Nutr. Rep. Int 29,<br />

845–858.<br />

Kim, S.K., D’Appolonia, B.L., 1977. Effect <strong>of</strong> pentosans on the retrogradation <strong>of</strong> wheat<br />

starch gels. <strong>Cereal</strong> Chem 54, 150–160.<br />

J.A. Wood / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 128–133 133<br />

Knights, E.J., Açikgöz, N., Warkentin, T., Bejiga, G., Yadav, S.S., Sandhu, J.S., 2007.<br />

Area, production and distribution. In: Yadav, S.S., Redden, B., Chen, W.,<br />

Sharma, B. (Eds.), Chickpea Breeding and Management. CAB International,<br />

Wallingford, UK, pp. 167–178.<br />

Kurien, S., Narayanaswamy, D., Daniel, V.A., Swaminathan, M., Parpia, H.A., 1971.<br />

Supplementary value <strong>of</strong> pigeonpea (Cajanus cajan) and chickpea to poor diets<br />

based on kaffir corn and wheat. Nutr. Rep. Int 4, 229–236.<br />

Lorenz, K., Dilsaver, W., Wolt, M., 1979. Fababean flour and protein concentrate in<br />

baked goods and in pasta products. Bakers’ Dig 53, 39–45.<br />

Matsuo, R.R., Malcolmson, L.J., Edwards, N.M., Dexter, J.E., 1992. A colorimetric<br />

method for estimating spaghetti cooking losses. <strong>Cereal</strong> Chem 69, 27–29.<br />

Muzquiz, M., Wood, J.A., 2007. Antinutritional factors. In: Yadav, S.S., Redden, B.,<br />

Chen, W., Sharma, B. (Eds.), Chickpea Breeding and Management. CAB International,<br />

Wallingford, UK, pp. 143–166.<br />

Naivikul, O., D’Appolonia, B.L., 1979. Carbohydrates <strong>of</strong> legume flours compared with<br />

wheat flour. 3. Nonstarchy polysaccharides. <strong>Cereal</strong> Chem 56, 45–49.<br />

Nobile, M.A., Baiano, A., Conte, A., Mocci, G., 2005. Influence <strong>of</strong> protein content on<br />

spaghetti cooking quality. J. <strong>Cereal</strong> Sci 41, 347–356.<br />

Perovic, B., 2000. Role <strong>of</strong> starches in pasta cooking. Tecnica Molitoria 51, 726–731.<br />

Rasmay, N.M.H., El-Shatanovi, G.A., Hassan, K.E.W., 2000. High-protein macaroni<br />

from legume flours and their protein concentrates. Ann. Agric. Sci 45, 555–570.<br />

Reyes-Moreno, C., Cuevas-Rodriguez, E.O., Milan-Carrillo, J., Cardenas-<br />

Valenzuela, O.G., Barron-Hoyos, J., 2004. Solid state fermentation process for<br />

producing chickpea (Cicer arietinum L.) tempeh flour: physicochemical and<br />

nutritional characteristics <strong>of</strong> the product. J. Sci. Food Agric 84, 271–278.<br />

Riva, M., Fessas, D., Schiraldi, A., 2000. Starch retrogradation in cooked pasta and<br />

rice. <strong>Cereal</strong> Chem 77, 433–438.<br />

Sabanis, D., Makri, E., Doxastakis, G., 2006. Effect <strong>of</strong> durum flour enrichment with<br />

chickpea flour on the characteristics <strong>of</strong> dough and lasagne. J. Sci. Food Agric 86,<br />

1938–1944.<br />

Sasaki, T., Yasui, T., Matsuki, J., 2000. Influence <strong>of</strong> non-starch polysaccharides isolated<br />

from wheat flour on the gelatinization and gelation <strong>of</strong> wheat starches.<br />

Food Hydrocolloids 14, 295–303.<br />

Sharma, R., Sissons, M.J., Rathjen, A.J., Jenner, C.F., 2002. The null-4A allele at the<br />

waxy locus in durum wheat affects pasta cooking quality. J. <strong>Cereal</strong> Sci 35,<br />

287–297.<br />

Siddique, I.M., Awan, J.A., Salim-ur-RehmanGilani, A.H., Ali, M., 1996. Evaluation <strong>of</strong><br />

composite flour based on wheat (Triticum aestivum) and chickpea (Cicer arietinum)<br />

for chapati making. Sci. Int 8, 165–168.<br />

Sissons, M.J., Egan, N.E., Gianibelli, M.C., 2005. New insights into the role <strong>of</strong> gluten<br />

on durum pasta quality using reconstitution method. <strong>Cereal</strong> Chem 82, 601–608.<br />

Wood, J.A., 2004. RVAÔ Viscosity <strong>of</strong> Chickpea Flours. Tech. J. Newport Sci 1, 1.<br />

Wood, J.A., Batey, I.L., Hare, R.A., Sissons, M.J., 2001. A comparison <strong>of</strong> Australian and<br />

imported spaghetti. Food Aust 53, 349–354.<br />

Wood, J.A., Grusak, M.A., 2007. Nutritional Value <strong>of</strong> Chickpea. In: Yadav, S.S.,<br />

Redden, B., Chen, W., Sharma, B. (Eds.), Chickpea Breeding and Management.<br />

CAB International, Wallingford, UK, pp. 101–142.<br />

Wood, J.A., Knights, E.J., Harden, S., 2008. Milling performance in desi type chickpea<br />

(Cicer arietinum L.): effects <strong>of</strong> genotype, environment and seed size. J. Sci. Food<br />

Agric 88, 108–115.<br />

Yanez-Farias, G.A., Bernal-Aguilar, V., Ramirez-Rodriguez, L., Barron-Hoyos, J.M.,<br />

1999. Fortification <strong>of</strong> some cereal foods with a chickpea protein concentrate.<br />

Food Sci. Tech. Int 5, 89–93.<br />

Yoshimura, M., Takaya, T., Nishinari, K., 1996. Effects <strong>of</strong> konjac-glucomannan on the<br />

gelatinization and retrogradation <strong>of</strong> corn starch as determined by rheology and<br />

differential scanning calorimetry. J. Agric. Food Chem 44, 2970–2976.<br />

Zhao, Y.H., Manthey, F.A., Chang, S.K.C., Hou, H.J., Yuan, S.H., 2005. Quality characteristics<br />

<strong>of</strong> spaghetti as affected by green and yellow pea, lentil, and chickpea<br />

flours. J. Food Sci 70, S371–S376.


Impact <strong>of</strong> re-grinding on hydration properties and surface composition<br />

<strong>of</strong> wheat flour<br />

M. Mohamad Saad a , C. Gaiani a, *, J. Scher a , B. Cuq b , J.J. Ehrhardt c , S. Desobry a<br />

a LSGA, Laboratoire de <strong>Science</strong>s et Génie Alimentaires, Nancy University, 2 avenue de la Forêt de Haye, B.P. 172, 54505 Vandoeuvre Les Nancy Cedex, France<br />

b Montpellier SupAgro, INRA, UMR 1208 Unit for Emerging Technology and Polymer Engineering, 2, place Viala, 34060 Montpellier Cedex 1, France<br />

c LCPME, Laboratoire de Chimie Physique et de Microbiologie pour l’Environnement, Nancy Université, CNRS, 405, rue de Vandoeuvre, 54600 Villers Lès Nancy, France<br />

article info<br />

Article history:<br />

Received 30 November 2007<br />

Received in revised form 28 July 2008<br />

Accepted 5 August 2008<br />

Keywords:<br />

DVS<br />

Sorption isotherm<br />

Wheat flour<br />

XPS<br />

1. Introduction<br />

abstract<br />

One <strong>of</strong> the major control variables in food preservation technology<br />

is water activity (aw). This term indicating the ‘‘quality’’ <strong>of</strong><br />

water content in food, describes the degree <strong>of</strong> ‘‘boundness’’ <strong>of</strong><br />

water and thus its availability to participate in physical, chemical,<br />

and microbiological reactions (Brunauer et al., 1938; Van den Berg<br />

and Bruin, 1981). The moisture sorption isotherm is defined as the<br />

relationship between the total moisture content and water activity<br />

<strong>of</strong> the food at a constant temperature and under equilibrium<br />

conditions. Moisture sorption isotherms are useful, not only in<br />

showing at which moisture content certain desirable or undesirable<br />

levels <strong>of</strong> aw is achieved, but also in indicating what significance<br />

small changes in moisture content will have in terms <strong>of</strong> changes in<br />

aw. It is therefore a useful guide to the storage life <strong>of</strong> foods held at<br />

moderate temperatures and preserved only by reduced aw<br />

(Abdullah et al., 2000).<br />

The mathematical description <strong>of</strong> the sorption phenomena in<br />

foods is <strong>of</strong> interest. More than 200 equilibrium moisture content–<br />

equilibrium relative humidity relationships have been reported in<br />

Abbreviations: BET, Brunauer–Emmett–Teller; DVS, dynamic vapor sorption;<br />

GAB, Guggenheim–Andersen–de Boer; RH, relative humidity; TSS, third stage<br />

sorption; XPS, X-ray photoelectron spectroscopy.<br />

* Corresponding author. Tel.: þ33 3 83 59 58 78; fax: þ33 3 83 59 58 04.<br />

E-mail address: claire.gaiani@ensaia.inpl-nancy.fr (C. Gaiani).<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.08.001<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 134–140<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

The combination <strong>of</strong> two analytical methodologies (water vapor sorption isotherm by using the DVS and<br />

chemical surface composition by using the XPS) has been used to enhance the understanding <strong>of</strong> the<br />

impact <strong>of</strong> re-grinding on the wheat flour hydration mechanism. A controlled atmosphere microbalance<br />

was used to construct water sorption isotherms at 25 C <strong>of</strong> different samples <strong>of</strong> wheat flours obtained by<br />

successive re-grinding <strong>of</strong> native wheat flour.<br />

Experimental water adsorption isotherms were modeled using different complementary models, based<br />

on two-parameter (BET), three-parameter (GAB), and four-parameter (TSS) models. A slight increase in<br />

water sorption capacity <strong>of</strong> wheat flour due to the re-grinding process was observed. The most affected<br />

parameters <strong>of</strong> the sorption isotherm models were C (the energy constant) and Xm (the monolayer water<br />

content capacity). The X-ray photoelectron spectroscopy (XPS) analysis showed changes in chemical<br />

bonds on wheat particle surfaces due to re-grinding process and particularly a significant increase in<br />

hydrophilic and decrease in hydrophobic bonds.<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

the literature (McMinn et al., 2004). Both the International Union <strong>of</strong><br />

Pure and Applied Chemistry (IUPAC) and the Commission on<br />

Colloid and Surface Chemistry (report <strong>of</strong> 1985) recommend the<br />

usefulness <strong>of</strong> the two-parameter model BET (Brunauer, Emmett,<br />

and Teller) because <strong>of</strong> its simplicity <strong>of</strong> application. This model is<br />

based on a multilayer sorption and is used in the low water activity<br />

range (0.05 < aw < 0.40). A three-parameter model (GAB model)<br />

was recommended by the European Project Group COST 90 (Wolf<br />

et al., 1984) and has been successfully applied to various foods (Van<br />

den Berg, 1985). The GAB (Guggenheim, Andersen, and de Boer)<br />

equation represents a model based on multi-layers and condensation<br />

and covers a wider range <strong>of</strong> water activity (0.05 < aw < 0.8)<br />

(Chirife et al., 1992; Timmermann and Chirife, 1991; Van den Berg<br />

and Bruin, 1981). Timmermann (1989) and Timmermann and<br />

Chirife (1991) developed a four-parameter equation, named the<br />

third stage sorption isotherm (TSS) to extend the GAB isotherm<br />

model to water activity ranges approaching unity. It is based on the<br />

premise that after a certain number <strong>of</strong> moisture layers exist, the<br />

moisture behaves as liquid water which has a dilution effect. The<br />

TSS model has thus the ability to predict infinite moisture<br />

adsorption (Bronlund and Paterson, 2004) and to give experimental<br />

data for many food systems a better fit when compared to the GAB<br />

equation. At the experimental level, the water adsorption capacity<br />

<strong>of</strong> food products has recently been evaluated using controlled<br />

atmosphere microbalances and dynamic automated sorption<br />

methods (Bell and Labuza, 2000; Johnson and Brennan, 2000). The


small size <strong>of</strong> samples and the dynamic airflow around the samples<br />

enable generation <strong>of</strong> a complete isotherm in less than a week (Bell<br />

and Labuza, 2000). Several studies using controlled atmosphere<br />

microbalances have already been reported (Ketal et al., 2004;<br />

Levoguer and Willians, 1997)<br />

The surface composition and particle properties can be considered<br />

as critical parameters <strong>of</strong> powders as they are supposed to play<br />

an important role within the powder’s end use processes and<br />

should certainly not be neglected (Fäldt, 1995; Kim et al., 2002).<br />

Understanding the mechanism <strong>of</strong> the powder surface formation in<br />

terms <strong>of</strong> the compositional aspect will be highly useful in powder<br />

quality improvement and new product development (Kim et al.,<br />

2002). A classical way to characterise the powder’s surface<br />

composition is the use <strong>of</strong> X-ray photoelectron spectroscopy, also<br />

referred in the literature to electron spectroscopy for chemical<br />

analysis (ESCA) (Mistry et al., 1992). XPS is a well established<br />

technique for identifying elements and determining differences in<br />

surface chemistry <strong>of</strong> particle surfaces (Sionkowska et al., 2006). XPS<br />

has been commonly applied to spray-dried dairy powders (skim<br />

milk powder, whole milk powder, cream powder and whey protein<br />

concentrate) in order to investigate the relationship between milk<br />

powder processing, particle surface structure, and wetting properties<br />

to make advances in the understanding <strong>of</strong> the rehydration<br />

process or storage effects (Gaiani et al., 2006, 2007; Kim et al.,<br />

2002).<br />

The main objective <strong>of</strong> the current work was to study the impact<br />

<strong>of</strong> the re-grinding process on hydration properties <strong>of</strong> wheat flour.<br />

This study is primarily concerned with determining the experimental<br />

sorption data <strong>of</strong> different re-ground wheat flour samples<br />

coming from the same batch <strong>of</strong> wheat flour and describing them by<br />

using the two-, three-, and four-parameter isotherm models. Water<br />

adsorption measurements by using the dynamic vapor sorption<br />

have been conducted in combination with measurement <strong>of</strong> changes<br />

in surface bonds on the surface <strong>of</strong> wheat flour particles using the Xray<br />

photoelectron spectroscopy.<br />

2. Materials and methods<br />

2.1. Wheat flour sample preparation and conservation<br />

The different samples were prepared from the selected<br />

commercial wheat flour (Grands Moulins de Paris, Ivry Sur Seine,<br />

France) intended for bread making (extraction rate 75%). Initial<br />

water content <strong>of</strong> the native wheat flour (FNative) was 13.7%. The<br />

different flour samples (F1, F2, and F3) were obtained by successive<br />

re-grinding (R1,R2, and R3) <strong>of</strong> the native flour to reduce particle size<br />

using a laboratory grinder, under ambient relative humidity<br />

conditions (ZM 200, Retsch, France). For the first re-grinding<br />

process R1, a 50 g sample <strong>of</strong> native wheat flour was placed inside<br />

the laboratory grinder. The resulting wheat flour was named F1. The<br />

flour F1 was twice re-ground under the same conditions R2 and R3,<br />

and then the re-ground flours F2 and F3 were produced. During our<br />

study, all samples were stored at 18 C in hermetically sealed<br />

cans, in order to preserve their physicochemical characteristics (Da<br />

Costa-Correia, 1997; Kusunose et al., 2002). Before use, samples<br />

were defrosted for 12 h at 20 C.<br />

2.2. Wheat flour characterization<br />

2.2.1. Chemical analyses<br />

Water and damaged starch contents <strong>of</strong> flour samples were<br />

estimated according to American Association <strong>of</strong> <strong>Cereal</strong> Chemists<br />

(AACC) methods 44-15A and 76-30A, respectively. The total<br />

nitrogen content (TN) was determined by the Kjeldahl method and<br />

protein content was calculated according to TN 5.7 AACC (2000)<br />

Method (No: 46-10). Fat content was determined by acid hydrolysis<br />

M.M. Saad et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 134–140 135<br />

with HCl followed by extraction <strong>of</strong> hydrolysed lipids with ethers<br />

according to AACC Method (No: 30-10). All results represent the<br />

average <strong>of</strong> three tests.<br />

2.2.2. Physical properties<br />

The particle size distributions were measured by static light<br />

scattering (Mastersizer S, Malvern Instruments Ltd, Malvern, UK)<br />

with a 5 mW He–Ne laser operating at a wavelength <strong>of</strong> 632.8 nm<br />

with a 300F lens. Five grams <strong>of</strong> wheat sample were taken and<br />

introduced in 100 mL ethanol to reach a correct obscuration (Berton<br />

et al., 2002). The criterion selected was d50 which means that<br />

50% <strong>of</strong> the particles have a diameter lower than this criterion (i.e.<br />

midpoint <strong>of</strong> volume cumulative distribution). Results are the<br />

average <strong>of</strong> triplicate experiments carried out on different days.<br />

2.3. Sorption isotherms<br />

Water sorption isotherms were determined gravimetrically<br />

using DVS technique (Surface Measurement Systems, London, UK).<br />

The DVS apparatus monitors the moisture sorption capacities <strong>of</strong><br />

wheat flour as function <strong>of</strong> relative humidity (RH). The changes in<br />

sample weight over time at 25 C and at any desired RH (between<br />

0% and 98%) were recorded. About 15–20 mg <strong>of</strong> sample were loaded<br />

onto the quartz sample pan. The program was initially set to control<br />

the humidity at 0% for 12 h (drying phase). This step allowed the<br />

sample water activity to decrease to zero and internally equilibrate.<br />

The sample was then subjected to successive steps <strong>of</strong> 10% RH<br />

increase, up to 98%. For each step, mass changes (m) and the rate <strong>of</strong><br />

mass changes (dm/dt) were plotted against time. The equilibrium<br />

was considered to be reached when changes in mass with time<br />

(dm/dt) were lower than 0.002%/min (i.e. 2 g water/100 g db/day).<br />

All experiments were run at 25 C and 2–4 tests were carried out<br />

for each sample. The accuracy <strong>of</strong> the system was 1.0% RH<br />

and 0.2 C, respectively.<br />

2.4. XPS analysis<br />

The XPS analyses were carried out with a Kratos Axis Ultra<br />

(Kratos Analytical, Manchester, UK) spectrometer using a monochromatic<br />

Al Ka source. The delay-line detector allows a high count<br />

rate and the power applied to the X-ray anode was reduced to 90 W<br />

in order to avoid the X-ray induced degradation <strong>of</strong> the sample. The<br />

instrument work function was calibrated to give a binding energy<br />

(BE) <strong>of</strong> 83.96 eV for the Au 4f7/2 line for metallic gold and the<br />

spectrometer dispersion was adjusted to give a binding energy <strong>of</strong><br />

932.62 eV for Cu 2p3/2 line for metallic copper. The wheat flour<br />

samples were attached to the sample holder using a double sided<br />

conductive tape and then evacuated overnight prior to analyses. All<br />

spectra were recorded at a 90 take-<strong>of</strong>f angle, the analysed area<br />

being currently about 700 300 mm. Survey spectra were recorded<br />

with 1.0 eV steps and 160 eV analyser pass energy and the high<br />

resolution regions with 0.05 eV steps and 20 eV pass energy. In<br />

both cases the hybrid lens mode was used. During the data acquisition<br />

the Kratos charge neutralizer system was used on all specimens<br />

with the following settings: filament current 1.6 A, charge<br />

balance 2.4 V, filament bias 1.0 V and magnetic lens trim coil<br />

0.375 A. As overcompensation is always observed, the C1s line for<br />

adventitious carbon and C–C carbon was set to 284.60 eV and<br />

therefore used as an internal energy reference. With these<br />

parameters we could obtain a C1s signal with sharp, symmetric<br />

components with a FWHM <strong>of</strong> 1.2 eV. Spectra were analysed using<br />

the Vision s<strong>of</strong>tware from Kratos (Vision 2.2.2). A Shirley baseline<br />

was selected to subtract the background and Gaussian–Lorentzian<br />

(70–30%) shapes were used for spectral decomposition. Quantification<br />

was performed using the photoemission cross-sections and<br />

the transmission coefficients given in the Vision package.


136<br />

Table 1<br />

Wheat flour samples chemical composition<br />

Wheat flour Composition (g/100 g flour)<br />

Water Protein Damaged starch Lipids<br />

FNative 13.7 0.1 12.5 0.1 4.7 0.4 1.5 0.2<br />

F1 13.6 0.2 12.4 0.1 8.5 0.0 1.4 0.1<br />

F2 13.4 0.2 12.4 0.2 9.1 0.4 1.6 0.1<br />

F3 13.4 0.3 12.4 0.3 12.6 0.8 1.2 0.3<br />

2.5. Data analysis<br />

The water vapor adsorption isotherms were described by using<br />

three models: the BET (Brunauer–Emmett–Teller) model (Eq. (1)),<br />

GAB (Guggenheim–Andersen–de Boer) model (Eq. (2)), and TSS<br />

(third stage sorption) model (Eq. (3a–c)).<br />

C BETaw<br />

X ¼ Xm<br />

ð1 awÞð1 aw þ CBETawÞ C GABK GABaw<br />

X ¼ Xm<br />

ð1 KGABawÞð1 KGABaw þ CGABKGABawÞ HðawÞH 0 ðawÞC TSSK TSSaw<br />

X ¼ Xm<br />

ð1 KTSSawÞ½1 þðCTSSHðawÞ 1ÞKTSSawŠ HðawÞh1 þ 1 K TSS<br />

K TSS<br />

ðK TSSawÞ hTSS<br />

ð1 awÞ<br />

M.M. Saad et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 134–140<br />

(1)<br />

(2)<br />

(3a)<br />

(3b)<br />

Table 2<br />

Wheat flour samples particle size distribution<br />

Wheat flour d10 (mm) d50 (mm) d90 (mm)<br />

FNative 15.1 0.8 67.4 0.6 147.8 1.4<br />

F1 11.4 2.6 49.7 4.6 122.2 3.7<br />

F2 10.0 2.6 40.4 4.6 108.4 3.7<br />

F3 4.3 0.9 17.2 3.8 44.5 3.8<br />

H 0 ðawÞh1 þ HðawÞ 1<br />

HðawÞ<br />

ð1 KTSSawÞ ½hTSS þ ð1 h<br />

ð1 awÞ<br />

TSSÞawŠ (3c)<br />

where Xm is the monolayer moisture content (% db), X is the<br />

equilibrium water content (% db), CBET, CGAB, and CTSS are characteristic<br />

energy constants, KBET, KGAB, KTSS are the characteristic<br />

constants correcting the properties <strong>of</strong> the multilayer molecules<br />

with respect to the bulk liquid, and hTSS is the third stage sorption<br />

isotherm constant.<br />

The model parameters were directly determined from experimental<br />

water vapor sorption isotherms. The parameters <strong>of</strong> the BET<br />

model (Xm, C), GAB model (Xm,C, and K), and TSS model (Xm, C, K,<br />

and hTSS) were identified, respectively, with Eqs. (1)–(3). The<br />

parameters were calculated by an optimization procedure according<br />

to the Gauss–Newton algorithm using the s<strong>of</strong>tware Excel 2007<br />

(Micros<strong>of</strong>t). The minimized objective function was the sum <strong>of</strong> the<br />

absolute difference between experimental and predicted points.<br />

2.6. Statistical analyses<br />

All statistical analysis was carried out by using Micros<strong>of</strong>t Excel<br />

2007 (Micros<strong>of</strong>t Corporation, USA).<br />

Fig. 1. Sorption isotherm pr<strong>of</strong>ile obtained for the FNative, F1, F2 and F3 wheat flour samples estimated at 25 C from 0% to 98% RH and modeled up with the TSS model.


Table 3<br />

Parameters obtained from the fitted curves with BET, GAB, and TSS models for FNative, F1, F2 and F3 wheat flour samples<br />

Model Flour Xm C K hTSS R 2<br />

Abs. difference<br />

BET FNative 6.41 0.69 18.9 10.1 – – 0.992 0.266<br />

F1 6.28 0.46 18.2 1.2 – – 0.990 0.296<br />

F2 6.36 0.03 17.4 1.0 – – 0.994 0.249<br />

F3 6.47 0.08 21.1 2.2 – – 0.994 0.263<br />

3. Results and discussion<br />

3.1. Wheat flour characterization<br />

The chemical composition <strong>of</strong> the selected samples <strong>of</strong> wheat<br />

flour is presented in Table 1. As reported by Berton et al. (2002) and<br />

Wang and Flores (2000), the re-grinding process produced<br />

a significant increase in damaged starch content (from 5% to 13%),<br />

while the protein and lipid contents remain stable (about 12.5% and<br />

1.5%, respectively). The water content slightly decreases from 13.7%<br />

to 13.2% with re-grinding and this could be due to the slight heat<br />

effects during the re-grinding process (þ10 C). As expected, the<br />

size distribution <strong>of</strong> flour samples decreases as re-grinding progresses<br />

(Table 2). The d50 <strong>of</strong> the samples is around 67 mm for native<br />

flour (FNative) and decreases to 50, 40, and 17 mm, respectively, for F1,<br />

F2 and F3 flours. The decrease in particle size with re-grinding is due<br />

to the successive breaking <strong>of</strong> flour particles inside the grinder. The<br />

breaks in particles occurring during re-grinding have been<br />

considered as responsible for starch granule rupture and thus<br />

induce increase in damaged starch content (Hoseney, 1994).<br />

According to Dubois (1949), starch granule exhibits elastic properties<br />

that lead to different types <strong>of</strong> damage such as cracks and<br />

breaks during grinding, which play an essential role in increasing<br />

the surface subjected to water vapor.<br />

3.2. Water vapor adsorption isotherms<br />

The adsorption isotherm pr<strong>of</strong>iles at 25 C, from 0% to 98% RH are<br />

presented in Fig. 1 for the selected wheat flour samples. As<br />

expected, these isotherms demonstrate an increase in equilibrium<br />

moisture content with increasing water activity. As has been shown<br />

by Roman-Gutierrez et al. (2002), this behavior is obviously<br />

a sigmoidal shaped curve reflecting a Type II isotherm according to<br />

the Brunauer classification (Brunauer et al., 1938). This form can be<br />

described in terms <strong>of</strong> a three-step moisture adsorption process. As<br />

previously observed by other researchers (Quirijns et al., 2005),<br />

sorption isotherms involve, during the first step, polar groups <strong>of</strong><br />

high binding energy to hydrophilic components (starches, proteins,<br />

and pentosans) being saturated with water molecules (i.e. the<br />

monolayer coverage). During the second stage, additional water<br />

molecules are bound onto the monolayer (i.e. multilayer coverage)<br />

and water clusters begin to form. During the third step, accumulation<br />

<strong>of</strong> water in intermolecular free spaces occurs and results in<br />

partial swelling that in turn may expose additional hydrophilic<br />

binding sites. It can be noticed that very slight differences seem to<br />

be observed between the experimental adsorption isotherms for<br />

the native wheat flours and the three re-ground flour samples.<br />

M.M. Saad et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 134–140 137<br />

GAB FNative 8.50 0.75 12.9 4.9 0.686 0.015 – 0.997 0.159<br />

F1 8.36 0.94 13.4 1.5 0.701 0.046 – 0.997 0.184<br />

F2 8.54 0.02 12.8 0.5 0.678 0.014 – 0.998 0.138<br />

F3 8.64 0.04 13.6 0.3 0.609 0.008 – 0.997 0.190<br />

TSS FNative 8.84 1.14 11.4 3.0 0.665 0.037 14.3 1.9 0.998 0.234<br />

F1 8.22 1.21 14.5 3.0 0.710 0.067 17.1 5.8 0.988 0.178<br />

F2 9.32 3.19 10.9 0.1 0.634 0.036 17.5 1.1 0.997 0.290<br />

F3 9.09 0.64 11.5 1.6 0.678 0.042 14.4 0.5 0.998 0.224<br />

Xm: monolayer moisture content g <strong>of</strong> water/100 g <strong>of</strong> dry base; C: constant; K: constant; hTSS: TSS model constant; Abs. difference: the average value <strong>of</strong> absolute difference<br />

between experimental data and calculated value <strong>of</strong> water content.<br />

3.3. Water vapor sorption mathematical modeling<br />

3.3.1. Brunauer–Emmett–Teller Model<br />

The experimental sorption data <strong>of</strong> wheat flour samples were<br />

first fitted with the two parameter BET model (Eq. (1)), between 0%<br />

and 40% RH (Table 3). There is a good agreement between experimental<br />

data and predicted values (average R 2 ¼ 0.992 and average<br />

absolute difference between experimental and calculated<br />

data ¼ 0.27%). The calculated Xm (monolayer moisture content) and<br />

CBET values are presented in Table 3 for native and re-ground flour<br />

samples.<br />

No differences have been found between the calculated values<br />

<strong>of</strong> BET model parameters for the four different products. We can<br />

report almost the same values <strong>of</strong> Xm (6.28–6.47 g/100 g dry bases)<br />

for the native wheat flour and the re-ground flour samples. Almost<br />

Fig. 2. XPS spectra obtained for the FNative wheat flour (survey scan).


138<br />

the same values for monolayer water contents (Xm) have already<br />

been reported in the literature for wheat flours (Roman-Gutierrez<br />

et al., 2002).<br />

We observe a slight increase in the CBET values with re-grinding,<br />

from CBET ¼ 18.9 for the native flour to CBET ¼ 21.1 for the sample F3.<br />

That is to say, when using the BET model, the energy constant (CBET)<br />

is found to be slightly increased as re-grinding rate increases. This<br />

behavior may be manifested by the liberation <strong>of</strong> hydrophilic sites<br />

during the re-grinding process, whereas different types <strong>of</strong> starch<br />

damage as cracks and breaks take place.<br />

3.3.2. Guggenheim–Andersen–de Boer Model<br />

The experimental sorption data <strong>of</strong> wheat flour samples have<br />

also been fitted with the three-parameter GAB model (Eq. (2)),<br />

between 0% and 80% RH (Table 3). As expected, there is good<br />

agreement between experimental data and predicted values<br />

(average R 2 ¼ 0.997 and average absolute difference between<br />

experimental and calculated data ¼ 0.17%). The calculated GAB<br />

model parameter Xm (monolayer moisture content), KGAB, and CGAB<br />

values are presented in Table 3 for the native wheat flour and the<br />

three re-ground flour samples. No significant differences have been<br />

found between the calculated values <strong>of</strong> GAB model parameters for<br />

the four products. We can report almost the same values <strong>of</strong> Xm<br />

(8.36–8.64 g/100 g dry bases) for the native wheat flour and the<br />

re-ground samples, nonetheless the highest values were observed<br />

for the most re-ground flour (sample F3). It can be noticed that the<br />

M.M. Saad et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 134–140<br />

Xm values determined using the GAB model (8.36–8.64 g/100 g dry<br />

bases) are slightly higher than those calculated using the BET model<br />

(6.28–6.47 g/100 g dry bases). Timmermann (2003) also stated that<br />

the value Xm given by the BET model is always smaller than the<br />

monolayer value corresponding to the GAB model.<br />

No differences were observed among CGAB values (12.8–13.7 g/<br />

100 g dry bases) for the native wheat flour and the re-ground flour<br />

samples. Still, the highest values were observed for the most<br />

re-ground flour (sample F3).<br />

The KGAB values remain constant (between 0.678 and 0.701) for<br />

all flour samples. The estimated value confirms observations,<br />

carried out on many starchy foods (Chirife et al., 1992), which<br />

reveals that KGAB values are about 0.74. Probably, this stability in<br />

KGAB values during the re-grinding process may be attributable to<br />

the steady chemical composition <strong>of</strong> all studied flour samples.<br />

Similar values <strong>of</strong> Xm, CGAB, and KGAB have already been reported in<br />

the literature for wheat flour (Chirife et al., 1992; Roman-Gutierrez<br />

et al., 2002).<br />

3.3.3. Third stage sorption model<br />

The experimental sorption data <strong>of</strong> wheat flour samples have<br />

then been fitted with the four-parameter TSS model (Eq. (3a–c))<br />

between 0% and 98% RH as illustrated in Fig. 1 and Table 3. There is<br />

good agreement between experimental data and predicted values<br />

(average R 2 ¼ 0.995 and average absolute difference between<br />

experimental and calculated data ¼ 0.23%). The calculated TSS<br />

Fig. 3. Example <strong>of</strong> XPS narrow spectra <strong>of</strong> O1s obtained for FNative, F1, F2 and F3 wheat flour samples, respectively.


model parameters Xm (monolayer moisture content), CTSS, KTSS, and<br />

hTSS values are presented in Table 3 for the native wheat flour and<br />

the three re-ground flour samples.<br />

The calculated values <strong>of</strong> the TSS model (Xm, CTSS, KTSS, and hTSS)<br />

appear to stay constant for all the selected samples, with Xm values<br />

ranging between 8.84 and 9.32 g/100 g dry bases, CTSS values<br />

ranging between 10.9 and 14.5, KTSS ranging between 0.634 and<br />

0.710, and with hTSS values ranging between 14.3 and 17.5. Therefore,<br />

it may be stated that no significant change was observed<br />

between the native wheat flour and the three re-ground flour<br />

samples, when describing the whole adsorption isotherms<br />

between 0% and 98% relative humidity with the TSS model. The<br />

calculated values <strong>of</strong> the hTSS parameter appear to be low in<br />

comparison with ‘‘usual’’ values that have been found (Timmermann,<br />

1989). Bronlund and Paterson (2004) have found that hTSS<br />

value is around 30. Because <strong>of</strong> wheat flour insolubility in water<br />

(finite sorption at RH ¼ 100%), the TSS model takes into consideration<br />

the relative insolubility behavior <strong>of</strong> wheat components by<br />

lowering the hTSS value. This could explain why low values <strong>of</strong> h<br />

parameter have been obtained.<br />

3.4. Wheat flour chemical surface analyses using X-ray<br />

photoelectron spectroscopy<br />

The survey scan <strong>of</strong> the native wheat flour sample is represented<br />

as an example with the identification <strong>of</strong> the O1s,N1s,C1s,S2p and P2p<br />

peaks as illustrated in Fig. 2. The detection <strong>of</strong> S2p and P2p elements<br />

was possible thanks to the high sensitivity <strong>of</strong> the XPS equipment<br />

(Gaiani et al., 2006). Nevertheless, the concentration <strong>of</strong> these two<br />

elements was found below 1% for all the samples. Almost the same<br />

XPS curve shapes have been previously observed for milk powders<br />

(Gaiani et al., 2006). According to a model for biochemical<br />

compounds, the carbon (C), oxygen (O) and azotes (N) peaks were<br />

decomposed (Gerin et al., 1995). The C1s peak was decomposed into<br />

four peaks corresponding to the C–(C, H), C–(O, N), C]O and O–<br />

C]O functions. The O1s peak was decomposed in three peaks<br />

attributed to the O]C, O–C and H2O functions. The N1s peak was<br />

decomposed into C–NH and C–NH 3þ functions.<br />

As a typical example, decomposition <strong>of</strong> the oxygen peak for the<br />

native wheat flour and the three re-ground flour samples are<br />

presented in Fig. 3. Similar decomposition curves have been<br />

constructed for carbon and nitrogen (data not shown). From these<br />

decompositions, the data were analysed in terms <strong>of</strong> rate <strong>of</strong> surface<br />

bonds (Table 4). From the C1s peak, it appears that the re-grinding<br />

process (from FNative, F1, F2, toF3 flour) induces a slight decrease <strong>of</strong><br />

Table 4<br />

Elemental surface composition (bold) and rate <strong>of</strong> surface bond for FNative,F1,F2 and F3<br />

wheat flour samples<br />

Element Bond Wheat flour sample<br />

FNative F1 F2 F3<br />

% C 77.0 77.4 76.7 74.7<br />

% C–H, C–C 61.0 60.1 59.1 57.2<br />

% C–O, C–N 28.2 27.9 29.3 31.0<br />

%C]O 6.2 8.6 8.0 8.7<br />

% O–C]O 4.6 3.4 3.6 3.1<br />

% O 18.6 18.0 18.9 19.3<br />

% O–C 82.0 77.5 76.9 75.2<br />

%O]C 13.7 14.1 15.2 18.9<br />

%H2O 4.3 8.4 7.9 5.9<br />

% N 3.9 4.1 4.0 5.5<br />

% C–NH 93.2 92.7 96.5 93.6<br />

% C–NH 3þ<br />

6.8 7.3 3.5 6.4<br />

% S 0.3 0.3 0.2 0.3<br />

% P 0.2 0.2 0.2 0.2<br />

M.M. Saad et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 134–140 139<br />

Fig. 4. Changes in surface chemical bonds percentage as re-grinding rate increases for<br />

FNative, F1, F2 and F3 wheat flour samples, where white columns represent C–(H,C) and<br />

black columns represent C–(O–N), C]O, O–C]O.<br />

the C–(C, H) bond (from 61, 60, 59 to 57%, respectively, for FNative,F1,<br />

F2,toF3 samples). Concurrently the total <strong>of</strong> C–O functions (C–(O, N),<br />

C]O and O–C]O) increased. From the O1s peaks, the O]C functions<br />

increased significantly whereas the O–C functions decreased<br />

(from 82.0, 77.5, 76.9 to 75.2, respectively, for FNative, F1, F2, toF3<br />

samples). These results clearly reflect a decrease in the number <strong>of</strong><br />

hydrophobic bonds in contrast to hydrophilic bonds which increase<br />

in its turn (Fig. 4). This behavior may be due to physical changes<br />

induced by re-grinding process resulting in starch granule rupture.<br />

4. Conclusion<br />

Whatever the model applied (BET, GAB, or TSS), wheat flour<br />

shows a slight trend to adsorb more water vapor after the regrinding<br />

process. This bonded water was adsorbed more strongly.<br />

One could assume that starch damages occurring during the regrinding<br />

process could affect the surface and modify interactions <strong>of</strong><br />

particles with water vapor, whereas the hydrophilic bonds increase.<br />

Hence, the TSS (third stage sorption) model can be used coupled<br />

with the DVS technique to describe sorption isotherms <strong>of</strong> wheat<br />

flour throughout the entire range <strong>of</strong> water activity. Techniques such<br />

as DVS can be used coupled with XPS to obtain valuable information<br />

related to total and partial contribution <strong>of</strong> chemical components<br />

<strong>of</strong> wheat flours in their hydration properties. Therefore,<br />

further studies including surface composition evolution during<br />

milling coupled with sorption isotherms <strong>of</strong> pure wheat flour<br />

components (starch, protein,.) would improve understanding<br />

contributions <strong>of</strong> wheat flour components with water vapor<br />

adsorption.<br />

Acknowledgements<br />

We are grateful to J. Lambert, LCPME engineer CNRS (Nancy), for<br />

providing XPS analysis and technical assistance. The authors would<br />

like also to acknowledge the support <strong>of</strong> Syrian Ministry <strong>of</strong> Higher<br />

Education for financial support.<br />

References<br />

American Association <strong>of</strong> <strong>Cereal</strong> Chemists, 2000. Official Methods <strong>of</strong> the AACC.<br />

Method 44-15A approved October 1975, revised October 1981; revised<br />

November 1999; Method 76-30A, approved May 1969, revised October 1982,<br />

and October 1984; revised November 1999; Method 46-10, approved April<br />

1961, revised November 1999; Method 30-10, approved April 1961, revised<br />

November 1999, 10th ed. The Association, St. Paul, MN, USA.<br />

Abdullah, N., Nawawi, A., Othman, I., 2000. Fungal spoilage <strong>of</strong> starch-based foods in<br />

relation to its water activity (aw). <strong>Journal</strong> <strong>of</strong> Stored Products Research 36, 47–54.


140<br />

Bell, L.N., Labuza, T.P., 2000. Moisture Sorption: Practical Aspects <strong>of</strong> Isotherm<br />

Measurement and Use. American Association <strong>of</strong> <strong>Cereal</strong> Chemists, Inc., St Paul, MN.<br />

Berton, B., Scher, J., Villieras, F., Hardy, J., 2002. Measurement <strong>of</strong> hydration capacity<br />

<strong>of</strong> wheat flour: influence <strong>of</strong> composition and physical characteristics. Powder<br />

Technology 128, 326–331.<br />

Bronlund, J., Paterson, T., 2004. Moisture sorption isotherms for crystalline, amorphous<br />

and predominantly crystalline lactose powders. International Dairy<br />

<strong>Journal</strong> 14, 247–254.<br />

Brunauer, S., Emmett, P.H., Teller, E., 1938. Adsorption <strong>of</strong> gases in multimolecular<br />

layers. <strong>Journal</strong> <strong>of</strong> American Chemical Society 60, 309–319.<br />

Chirife, J., Timmermann, E.O., Iglesias, H.A., Boquet, R., 1992. Some features <strong>of</strong> the<br />

parameter k <strong>of</strong> the GAB equation as applied to sorption isotherms <strong>of</strong> selected<br />

food materials. <strong>Journal</strong> <strong>of</strong> Food Engineering 15, 75–82.<br />

Da Costa-Correia, J.M., 1997. Contribution à l’étude des propriétés rhéologiques des<br />

farines de blé: Influence de la composition, de l’hydratation de la distribution<br />

granulométrique des particules. Ph.D. thèsis, Institue Nationale Polytechnique<br />

de Lorraine.<br />

Dubois, M., 1949. Incidences du réglage de la mouture sur les propriétés des farines.<br />

Bulletin ENSMIC 113, 170–187.<br />

Fäldt, P., 1995. Surface Composition <strong>of</strong> Spray-dried Emulsions. Ph.D. thesis, Lund<br />

University.<br />

Gaiani, C., Ehrhardt, J.J., Scher, J., Hardy, J., Desobry, S., Banon, S., 2006. Surface<br />

composition <strong>of</strong> dairy powders observed by X-ray photoelectron spectroscopy<br />

and effects on their rehydration properties. Colloids and Surfaces B: Biointerfaces<br />

49, 71–78.<br />

Gaiani, C., Scher, J., Ehrhardt, J.J., Linder, M., Schuck, P., Desobry, S., Banon, S., 2007.<br />

Relationships between dairy powder surface composition and wetting properties<br />

during storage: importance <strong>of</strong> residual lipids. <strong>Journal</strong> <strong>of</strong> Agricultural and<br />

Food Chemistry 55, 6561–6567.<br />

Gerin, P.A., Dengis, P.B., Rouxhet, P.G., 1995. Performance <strong>of</strong> XPS analysis <strong>of</strong> model<br />

biochemical compounds. <strong>Journal</strong> <strong>of</strong> Chemical Physics 92, 1043–1065.<br />

Hoseney, R.C., 1994. Dry milling <strong>of</strong> cereals. In: Principles <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

and Technology. American Association <strong>of</strong> <strong>Cereal</strong> Chemists, St. Paul, MN, USA,<br />

pp. 125–145.<br />

Johnson, P.N.T., Brennan, J.G., 2000. Kinetics <strong>of</strong> moisture absorption by plantain<br />

flour. <strong>Journal</strong> <strong>of</strong> Food Engineering 45, 33–36.<br />

Ketal, L.V., Graham, B., David, C., 2004. The use <strong>of</strong> dynamic vapour sorption and near<br />

infra-red spectroscopy (DVS-NIR) to study the crystal transitions <strong>of</strong> theophylline<br />

and the report <strong>of</strong> a new solid-state transition. European <strong>Journal</strong> <strong>of</strong><br />

Pharmaceutical <strong>Science</strong>s 22, 97–105.<br />

Kim, E.H.J., Chen, X.D., Pearce, D., 2002. Surface characterization <strong>of</strong> four industrial<br />

spray dried dairy powders in relation to chemical composition,<br />

M.M. Saad et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 134–140<br />

structure and wetting properties. Colloids and Surfaces B: Biointerfaces 26<br />

(3), 197–212.<br />

Kusunose, C., Noguchi, S., Yamagishi, T., Seguichi, M., 2002. Binding starch to tailing<br />

fraction by proteins in stored wheat flour. Food Hydrocolloids 1, 73–77.<br />

Levoguer, C.L., Willians, D.R., 1997. Moisture sorption properties <strong>of</strong> food products<br />

and packaging materials studied by dynamic vapour sorption. Food Technology<br />

Europe 6, 28–30.<br />

McMinn, A.M., Al-Muhtaseb, A., Ronald, A., Magi, T., 2004. Assessment <strong>of</strong> two- and<br />

three-parameter Lewicki models for description <strong>of</strong> sorption phenomena <strong>of</strong><br />

starch materials. <strong>Journal</strong> <strong>of</strong> the <strong>Science</strong> <strong>of</strong> Food and Agriculture 84, 1695–1700.<br />

Mistry, V.V., Hassan, H.N., Robinson, D.J., 1992. Effect <strong>of</strong> lactose and protein on the<br />

microstructure <strong>of</strong> dried milk. Food Structure 11, 73–82.<br />

Quirijns, E.J., Boxtel, A.J.B.V., Straten, W., 2005. Sorption isotherms, GAB parameters<br />

and isosteric heat <strong>of</strong> sorption. <strong>Journal</strong> <strong>of</strong> the <strong>Science</strong> <strong>of</strong> Food and Agriculture 85,<br />

1805–1814.<br />

Roman-Gutierrez, A., Guilbert, S., Cuq, B., 2002. Distribution <strong>of</strong> water between<br />

wheat flour components: a dynamic water vapor adsorption study. <strong>Journal</strong> <strong>of</strong><br />

<strong>Cereal</strong> <strong>Science</strong> 36, 347–355.<br />

Sionkowska, A., Wisniewski, M., Kaczmarek, H., Skopinska, J., Chevallier, P.,<br />

Mantovani, D., Lazare, S., Tokarev, V., 2006. The influence <strong>of</strong> UV irradiation on<br />

surface composition <strong>of</strong> collagen/PVP blended films. Applied Surface <strong>Science</strong><br />

253, 1970–1977.<br />

Timmermann, E.O., 1989. A BET-like three sorption stage isotherm. <strong>Journal</strong> <strong>of</strong><br />

Chemical Society 85, 1631–1645.<br />

Timmermann, E.O., Chirife, J., 1991. The physical state <strong>of</strong> water sorbed at high<br />

activities in starch in terms <strong>of</strong> the GAB sorption equation. <strong>Journal</strong> <strong>of</strong> Food<br />

Engineering 13, 171–179.<br />

Timmermann, E.O., 2003. Multilayer sorption parameters: BET or GAB values?<br />

Colloids and Surfaces A: Physicochemical and Engineering Aspects 220,<br />

235–260.<br />

Van den Berg, C., Bruin, S., 1981. Water activity and its estimation in food systems.<br />

In: Rockland, L.B., Stewart, F. (Eds.), Water Activity: Influence on Food Quality.<br />

Academic Press, New York, pp. 147–177.<br />

Van den Berg, C., 1985. Development <strong>of</strong> B.E.T.-like models for sorption <strong>of</strong> water on<br />

foods, theory and relevance. In: Simatos, D., Multon, J.L. (Eds.), Properties <strong>of</strong><br />

Water in Foods. Martinus Nijh<strong>of</strong>f, Dordrecht, p. 119.<br />

Wang, L., Flores, R.A., 2000. Effects <strong>of</strong> flour particle size on the textural properties <strong>of</strong><br />

flour tortillas. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 31, 263–272.<br />

Wolf, W., Spiess, W.E.L., Jung, G., Weisser, H., Bizot, H., Duckworth, R.B., 1984. The<br />

water vapor sorption isotherms <strong>of</strong> microcrystalline cellulose (MCC) and <strong>of</strong><br />

purified potato starch: results <strong>of</strong> a collaborative study. <strong>Journal</strong> <strong>of</strong> Food Engineering<br />

3, 51–72.


Identification <strong>of</strong> novel haze-active beer proteins by proteome analysis<br />

Takashi Iimure a, *, Nami Nankaku b , Megumi Watanabe-Sugimoto c , Naohiko Hirota a , Zhou Tiansu a ,<br />

Makoto Kihara a , Katsuhiro Hayashi a , Kazutoshi Ito d , Kazuhiro Sato b<br />

a Bioresources Research and Development Department, Sapporo Breweries Ltd., 37-1, Nittakizaki, Ota, Gunma 370-0393, Japan<br />

b Barley Germplasm Center, Research Institute for Bioresources, Okayama University, 2-20-1, Chuo, Kurashiki, Okayama 710-0046, Japan<br />

c Graduate school <strong>of</strong> Natural <strong>Science</strong> and Technology, Okayama University, 1-1-1, Tsushimanaka, Okayama 700-8530, Japan<br />

d Frontier Laboratories <strong>of</strong> Value Creation, Sapporo Breweries Ltd., 10 Okatohme, Yaizu, Shizuoka 425-0013, Japan<br />

article info<br />

Article history:<br />

Received 30 June 2008<br />

Received in revised form 31 July 2008<br />

Accepted 15 August 2008<br />

Keywords:<br />

Beer colloidal haze<br />

Barley<br />

Proteome analysis<br />

1. Introduction<br />

abstract<br />

In clear beers, colloidal storage haze is one <strong>of</strong> the principal<br />

indicators <strong>of</strong> beer quality. Consumers rely greatly on visual<br />

impressions and will generally judge a beer as stale or not fit to<br />

drink if it displays haze when it is supposed to be clear. The cause <strong>of</strong><br />

storage haze has been identified as due to interactions between<br />

haze-active proteins and certain polyphenols (Asano et al., 1982;<br />

Siebert, 1999; Siebert et al., 1996). The addition <strong>of</strong> silica gel during<br />

the beer filtration process is effective in removing haze-active<br />

proteins (Leiper et al., 2003; Siebert and Lynn, 1997). Haze-active<br />

polyphenols can be removed by the addition <strong>of</strong> polyvinylpolypyrolidone<br />

(PVPP) in beer filtration (McMurrough et al.,<br />

Abbreviations: BDAI-1, barley dimeric alpha-amylase inhibitor; CMb, CMb<br />

component <strong>of</strong> tetrameric alpha-amylase inhibitor; CMe, trypsin inhibitor CMe<br />

precursor; 2DE, two-dimensional gel electrophoresis; LC-MS/MS, liquid chromatography<br />

mass spectrometry/mass spectrometry; LTP, lipid transfer protein; MALDI<br />

TOF-MS, matrix-assisted laser desorption/ionization time-<strong>of</strong>-flight mass spectrometry;<br />

NCBI-nr, the non-redundant amino acid database <strong>of</strong> the National Center<br />

for Biotechnology Information; PAS, proteins adsorbed onto the silica gel; TAI,<br />

trypsin/amylase inhibitor pUP13; Trx-2p, thioredoxin from yeast.<br />

* Corresponding author. Tel.: þ81 276 56 1454; fax: þ81 276 56 1605.<br />

E-mail address: takashi.iimure@sapporobeer.co.jp (T. Iimure).<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.08.004<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 141–147<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

Colloidal haze reduces beer quality considerably. Four haze samples were analyzed by two-dimensional<br />

gel electrophoresis (2DE) in order to identify haze-active proteins. Several protein spots were observed in<br />

all <strong>of</strong> the four haze samples. Using mass spectrometry analysis followed by a database search identified<br />

these spots as barley dimeric alpha-amylase inhibitor (BDAI-1), CMb component <strong>of</strong> tetrameric alphaamylase<br />

inhibitor (CMb) and trypsin inhibitor CMe precursor (CMe). These proteins were considered to<br />

be haze-active. Since haze-active proteins are adsorbed by silica gel in the beer filtration process, we<br />

eluted proteins adsorbed onto silica gel (PAS) and identified their species. These major PAS were identified<br />

as protein Z4, protein Z7 and trypsin/amylase inhibitor pUP13 (TAI), rather than BDAI-1, CMb and<br />

CMe. Furthermore, we analyzed proline compositions in the beer proteins, PAS and the haze proteins.<br />

Consequently, we found that the proline compositions <strong>of</strong> PAS were higher (ca. 20 mol%) than those in the<br />

beer proteins (ca. 10 mol%), although those <strong>of</strong> the haze-active proteins such as BDAI-1, CMb and CMe<br />

were 6.6–8.7 mol%. Our results suggest that BDAI-1, CMb and CMe are not predominant haze-active<br />

proteins, but growth factors <strong>of</strong> beer colloidal haze.<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

1995; Mikysˇka et al., 2002). Using barley cultivars that lack<br />

proanthocyanidins is also effective to produce beer without hazeactive<br />

polyphenols (Fukuda et al., 1999; von Wettstein et al., 1980,<br />

1985). Although research has been conducted on haze-active<br />

proteins and polyphenols (Apperson et al., 2002; Evans et al., 2003;<br />

Leiper et al., 2003; Robinson et al., 2004), the factors controlling<br />

colloidal haze formation are still not clear.<br />

A series <strong>of</strong> beer proteins such as hordeins (Asano et al., 1982),<br />

lipid transfer protein 1 (Evans and Hejgaard, 1999; van Nierop et al.,<br />

2004) and protein Z (Evans and Hejgaard, 1999) are derived from<br />

barley malt. Hordeins are the predominant barley storage proteins<br />

and contain high levels <strong>of</strong> proline and glutamine. They are known<br />

to be involved in haze formation (Asano et al., 1982). Silica gel<br />

adsorbs proline rich proteins such as hordeins (Leiper et al., 2003).<br />

Robinson et al. (2007a,b) suggested that the barley trypsin inhibitor<br />

CMe precursor (CMe) was a haze-active protein based on the<br />

analysis <strong>of</strong> eluate proteins from silica gel. They distinguished the<br />

cultivars by the presence <strong>of</strong> CMe bands on immuno blot analysis<br />

with antibodies raised against silica eluate proteins, and revealed<br />

that beer brewed from cultivars without the CMe band formed less<br />

haze than the one with the CMe band (Evans et al., 2003; Robinson<br />

et al., 2004). Although several non-hordein proteins are regarded as<br />

haze-active, functions <strong>of</strong> these proteins in colloidal haze formation<br />

have not been well characterized. To reveal the mechanism <strong>of</strong> haze


142<br />

formation, beer proteins should be more comprehensively<br />

investigated.<br />

Proteome analysis, such as two-dimensional gel electrophoresis<br />

(2DE), followed by protein mass spectrometry, is a powerful tool to<br />

identify potential proteins <strong>of</strong> interest. Using this technique, barley<br />

grain, barley malt (Bak-Jensen et al., 2004; Østergaard et al., 2002)<br />

and beer (Perrocheau et al., 2005) have been analyzed. We also<br />

analyzed the beer proteins using 2DE and identified a possible<br />

foam-promoting protein (Okada et al., 2008).<br />

In this study, we prepared four haze-positive beer samples,<br />

analyzed proteins in haze by 2DE and identified haze-active<br />

proteins. We assumed that the protein spots observed on all four<br />

haze samples were from haze-active proteins, and that any proteins<br />

not appearing on the gel were haze-inactive proteins. In addition, in<br />

order to investigate the composition <strong>of</strong> proteins adsorbed onto the<br />

silica gel (PAS), we analyzed the PAS by 2DE and compared the<br />

proline compositions <strong>of</strong> the beer proteins, PAS and the haze<br />

proteins.<br />

2. Experimental<br />

2.1. Beer and haze samples<br />

To prepare haze-positive beer samples, commercial beers were<br />

stored for over one year at 4 C. These haze-positive beer samples<br />

were used for the analyses. Beer brewed from the malt <strong>of</strong> North<br />

American cultivar A, as described in our previous paper (Okada<br />

et al., 2008), was used to prepare a 25% salt-precipitated protein<br />

fraction. Beer samples filtered with and without silica gel<br />

(SiO 2 þbeer and SiO 2 beer) were brewed using only malt and hop,<br />

according to the standard method <strong>of</strong> Sapporo Breweries Ltd., as<br />

described by Okada et al. (2008). To prepare SiO 2 þ beer, the beer<br />

sample was filtered with 200 ppm <strong>of</strong> silica gel.<br />

2.2. Beer quality analysis<br />

Beer characteristics were analyzed according to the standard<br />

methods <strong>of</strong> the European Brewery Convention (EBC Analytica,<br />

1987). Colloidal stability was scored by a forcing test (FT-3). The<br />

bottled beer samples were stored at 60 C for 3 days, and then 0 C<br />

for 1 day. Subsequently, the beer clarity, FT-3, was assessed in terms<br />

<strong>of</strong> haze formation.<br />

2.3. Protein sample preparation<br />

To separate the haze and the beer fractions, the haze-positive<br />

beer samples were centrifuged at 2000 g, for 30 min at 4 C. The<br />

precipitate was washed twice in 5% ethanol solution, before being<br />

dissolved in 8 M urea (Wako, Japan) þ 2% 3-[(3-cholamidopropyl)<br />

dimethylammonio] propansulfonic acid (CHAPS) (Dojindo Laboratories,<br />

Japan) solution containing 0.28% dithiothreitol (Wako).<br />

Three milliliter <strong>of</strong> this solution and supernatant were applied to<br />

a PD-10 column (GE Healthcare Biosciences, Japan), and the<br />

desalted proteins were eluted by 4 ml <strong>of</strong> distilled water. The protein<br />

concentration was determined by the Bradford’s (1976) method<br />

using bovine serum albumin as a standard. These solutions were<br />

lyophilized and the lyophilized protein samples were subsequently<br />

used in 2DE and amino acid composition analyses. The 25% saltprecipitated<br />

protein fraction <strong>of</strong> the beer brewed from the malt <strong>of</strong><br />

cultivar A was prepared according to Okada et al. (2008). The<br />

proteins adsorbed onto the silica gel (PAS) were prepared as<br />

described below by modification <strong>of</strong> the method described by Evans<br />

et al. (2003). Silica gel was added to the degassed beer sample upto<br />

200 ppm, whereupon the solution was stirred for 1 h at room<br />

temperature and then centrifuged at 2000 g for 30 min at 4 C. 2%<br />

(v/v) <strong>of</strong> ammonium solution was added to the precipitate.<br />

T. Iimure et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 141–147<br />

Subsequently, to separate the silica gel and proteins, the solution<br />

was stirred for 1 h at room temperature and centrifuged at<br />

2000 g for 30 min at 4 C. By adding HCl, the pH <strong>of</strong> the solution<br />

was adjusted to 8.0. Subsequently, the solution was desalted using<br />

a PD-10 column (GE Healthcare Biosciences) and then lyophilized.<br />

The lyophilized protein was defined as PAS.<br />

2.4. Two-dimensional gel electrophoresis (2DE)<br />

The 2DE <strong>of</strong> both the first and second dimensions using the<br />

Multiphor II system (GE Healthcare Biosciences), and silver staining<br />

were carried out according to Okada et al. (2008).<br />

2.5. Mass spectrometry analysis and protein identification<br />

Selected protein spots were excised from the 2DE gel and<br />

digested with trypsin as described in a previous paper (Okada et al.,<br />

2008). Eluted proteins were applied to matrix-assisted laser<br />

desorption/ionization time-<strong>of</strong>-flight mass spectrometry (MALDI<br />

TOF-MS) as described in Okada et al. (2008). Proteins were identified<br />

by peptide mass finger printing on the non-redundant amino<br />

acid database <strong>of</strong> the National Center for Biotechnology Information<br />

(NCBI-nr) using MASCOT s<strong>of</strong>tware (Perkins et al., 1999). When<br />

a protein was not identified, it was digested using trypsin and/or<br />

lysil endopeptidase and the sample was analyzed using liquid<br />

chromatography mass spectrometry/mass spectrometry (LC-MS/<br />

MS). A Tris–HCl buffer (pH 8.0) containing lysil endopeptidase was<br />

added to decolorized sample gel prepared according to Okada et al.<br />

(2008). Then the gel was incubated for 3 h at 35 C. Subsequently,<br />

trypsin was added to this sample and incubated for 20 h at 35 C.<br />

Each sample solution was applied to LC-MS/MS. The condition <strong>of</strong><br />

the LC-MS/MS analysis is described below. Equipment: MAGIC<br />

2002 (Michrom BioResources, Inc., USA), column: Magic C18<br />

(0.1 150 mm, Michrom BioResources, Inc.), mobile phase: 2%<br />

acetonitrile þ 0.1% formic acid, and 90% acetonitrile þ 0.1% formic<br />

acid, flow rate: 250–300 nL min 1 , mass spectrometer: Q-T<strong>of</strong>2<br />

(Micromass, U.K.), ionization method: Nan<strong>of</strong>low-LC ESI, ionization<br />

mode: positive mode, electric potential <strong>of</strong> capillary: 1.8 kV, collision<br />

energy: 20–56 eV. To identify the protein species, NCBI-nr was<br />

searched for using the resultant values <strong>of</strong> product ion from all<br />

precursor ions using the MASCOT search engine (Perkins et al.,<br />

1999).<br />

2.6. Analysis <strong>of</strong> the proline composition <strong>of</strong> protein fraction<br />

The proline compositions <strong>of</strong> beer proteins, haze proteins and<br />

PAS were analyzed using the Waters Pico$Tag system (Waters,<br />

Japan). A 20 mL aliquot <strong>of</strong> 6 M HCl (1% phenol) was added to the<br />

lyophilized protein sample, and then the proteins were completely<br />

hydrolyzed at 110 C for 20 h. Subsequently, the amino acid<br />

concentration was measured. Amino acid concentrations were<br />

determined as described below. The sample was ultra-filtered by<br />

Microcon Ultracel YM-10, MW 10,000 (Nihon Millipore Ltd., Japan).<br />

Once the filtrate had been diluted to a suitable concentration,<br />

fluorescence derivatization was carried out. A 20 mL aliquot <strong>of</strong><br />

sample solution, 60 mL <strong>of</strong> AccQ$Fluorborate buffer (Waters) and<br />

20 mL <strong>of</strong> AccQ Fluor reagent (Waters) were mixed, and the mixture<br />

was subsequently incubated for 10 min at 55 C. The conditions for<br />

high performance liquid chromatography are described as follows:<br />

equipment: 2695 separation module, detector: 2475 multi l fluorescent<br />

detector (Waters) with Ex. 250 nm and Em. 395 nm,<br />

column: an AccQ Tag column (Waters), mobile phase: 100 mM<br />

sodium acetate þ 5.6 mM triethylamine (pH 5.7), 100 mM sodium<br />

acetate þ 5.6 mM triethylamine (pH 6.8), acetonitrile and distilled<br />

water. The chromatography data produced was analyzed using<br />

Empower personal s<strong>of</strong>tware (Waters).


3. Results<br />

3.1. 2DE analysis <strong>of</strong> the proteins in haze<br />

After we prepared the haze-positive beer samples, the haze and<br />

beer protein fractions were separated by centrifugation. Then the<br />

protein fractions were analyzed by 2DE (Fig. 1). On the 2DE gel <strong>of</strong><br />

the beer proteins, a large intensely staining spot (spot b0) at about<br />

pI 4.5–5.5, and a Mr <strong>of</strong> 35–45 kDa was observed. However, the<br />

intensity <strong>of</strong> spot b0 from the haze protein was lower than that <strong>of</strong><br />

the beer (Fig. 1A, Table 1). In the haze proteins, intense spots were<br />

observed in region I-I, b8–10, b13, and b14 (Fig. 1). Spots in region<br />

I-I and b8 were intense both in the beer and the haze proteins. On<br />

the other hand, spots b9, b10, b13 and b14 were only intense in the<br />

haze proteins, and spots b11 and b12 were only intense in the beer<br />

proteins (Fig. 1, Table 1). The low molecular weight region <strong>of</strong> the<br />

2DE image <strong>of</strong> the 25% salt-precipitated protein fraction was similar<br />

to those <strong>of</strong> the haze proteins, but different from those <strong>of</strong> the beer<br />

proteins (Fig. 1, Table 1). The spots <strong>of</strong> the 25% salt-precipitated<br />

fraction were distinct in region I-I, b0 and b8–b10, but faint or<br />

invisible in b5 and b11–b14. Fig. 2 shows the 2DE images in region I<br />

<strong>of</strong> the four haze protein samples. The protein spots were classified<br />

into three groups i.e. (1) haze-active, because the protein spot was<br />

observed in all four haze samples, (2) moderately haze-active<br />

because the protein spot was observed in few samples as a relatively<br />

faint spot or observed in all samples but as a faint spot, and<br />

(3) haze-inactive because the protein spot was faint or invisible in<br />

all four samples. Protein spots in region I-I, b10, b13 and b14 were<br />

observed in all four samples. Conversely, protein spots <strong>of</strong> b11 and<br />

b12 were invisible, and b0 (data not shown) was faint in all four<br />

samples. In addition, spot b5 was only observed in sample i faintly,<br />

and spots b8 and b9 were observed in 2 and 3 samples respectively<br />

(Table 1, Fig. 2). Overall, the protein spots are: (1) haze-active in<br />

A<br />

Mr<br />

(kDa)<br />

97.4<br />

66.2<br />

45.0<br />

31.0<br />

21.5<br />

14.4<br />

97.4<br />

66.2<br />

45.0<br />

31.0<br />

21.5<br />

14.4<br />

97.4<br />

66.2<br />

45.0<br />

31.0<br />

21.5<br />

14.4<br />

T. Iimure et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 141–147 143<br />

pI3 pI 10<br />

Beer (supernatant)<br />

Mw<br />

b0<br />

(kDa)<br />

14.4<br />

Haze (precipitate)<br />

25% salt precipitate<br />

14.4<br />

14.4<br />

B<br />

region I-I, b10, b13 and b14, (2) moderately haze-active in spots b0,<br />

b8 and b9, and (3) haze-inactive in spots b5, b11 and b12, respectively<br />

(see also Table 1).<br />

3.2. Mass spectrometry analysis and protein identification<br />

To identify protein species, we analyzed the protein spots in<br />

region I-I, b0, b5, and b8–b14 by MALDI TOF-MS or LC-MS/MS<br />

followed by a database search. Table 2 shows the identified protein<br />

spots by the analyses. Considering the classification <strong>of</strong> regions by<br />

haze-activity, barley dimeric alpha-amylase inhibitor (BDAI-1),<br />

CMb component <strong>of</strong> tetrameric alpha-amylase inhibitor (CMb), and<br />

trypsin inhibitor CMe precursor (CMe) were found in haze-active<br />

regions. Protein Z-type serpin (protein Z4), serpin (protein Z7), and<br />

chain A, non-specific lipid transfer protein 1 (LTP) were found in<br />

moderately haze-active regions. Thioredoxin from yeast (Trx-2p)<br />

and trypsin/amylase inhibitor pUP13 (TAI) were found in hazeinactive<br />

regions.<br />

3.3. 2DE analysis <strong>of</strong> the proteins adsorbed onto the silica gel (PAS)<br />

Silica gel is known to adsorb haze-active proteins (Leiper et al.,<br />

2003; Siebert and Lynn, 1997) therefore, an addition <strong>of</strong> silica gel at<br />

beer filtration is effective to prevent beer colloidal haze. We<br />

prepared two beer samples with and without silica gel in filtration,<br />

described as SiO 2 þ beer and SiO 2 beer, respectively. The quality<br />

pr<strong>of</strong>iles <strong>of</strong> these beer samples were similar except for FT-3 i.e. the<br />

beer clarity after the forcing test (Table 3). The SiO 2 þ beer showed<br />

resistance to beer colloidal haze. On the other hand, the SiO 2 beer<br />

was shown to be prone to haze formation in response to force testing<br />

and contained haze-active protein, which was partly adsorbed onto<br />

the silica gel. To prepare the proteins adsorbed onto the silica gel<br />

(PAS), we added silica gel to the SiO 2 beer, and then eluted proteins<br />

b11 b12<br />

Region I Region I-I<br />

Fig. 1. The two-dimensional gel electrophoresis (2DE) images <strong>of</strong> the beer proteins <strong>of</strong> sample i, the haze proteins <strong>of</strong> sample i and the 25% salt-precipitated proteins <strong>of</strong> the beer<br />

brewed from the malt <strong>of</strong> cultivar A. A: The whole images <strong>of</strong> 2DE, pI 3–10, B: the enlarged images <strong>of</strong> region I in A. Arrows indicate the spot numbers.<br />

b5<br />

b9<br />

b2<br />

b10<br />

b8<br />

b8<br />

b13<br />

b14


144<br />

Table 1<br />

Summary <strong>of</strong> the spot intensities <strong>of</strong> region I-I, b0, b5 and b8–b14 in the beers, the hazes, the 25% salt-precipitated proteins and the proteins adsorbed onto the silica gel (PAS).<br />

The beer proteins The haze proteins The 25%<br />

Haze-activity PAS<br />

salt-precipitated<br />

proteins<br />

a<br />

Sample i Sample ii Sample iii Sample iv<br />

b0 B b 6 6 6 6 B 6c B<br />

Region I-I B B B B B B B 6<br />

b5 6 6<br />

b8 B B 6 B 6 6<br />

b9 B 6 B B 6 6<br />

b10 B 6 6 6 B B 6<br />

b11 B B<br />

b12 B B<br />

b13 B 6 6 B B<br />

b14 B 6 B B B<br />

a PAS, proteins adsorbed onto the silica gel.<br />

b ‘B’ indicates that the spot was distinct, ‘6’ indicates that the spot was faint, and ‘ ’ indicates that the spot was invisible.<br />

c ‘B’ indicates that the protein was haze-active because the protein spot was observed in all four haze samples, ‘6’ indicates that the protein was moderately haze-active<br />

because the protein spot was observed in few samples as relatively faint spot or observed in all samples but as faint spot, and ‘ ’ indicates that the protein was haze-inactive<br />

because the protein spot was faint or invisible in all four samples.<br />

from the silica gel. Subsequently, we analyzed the proteins <strong>of</strong> two<br />

beer samples and PAS. The 2DE images <strong>of</strong> the two beer proteins were<br />

quite similar (Fig. 3). The 2DE images in PAS and the haze proteins<br />

were different both at the intensity <strong>of</strong> the spot b0 (protein Z) (Figs.1A<br />

and 4A) and the region I in the low molecular region (Figs. 2 and 4).<br />

The intensity <strong>of</strong> spot b0 in PAS was higher than those in the haze<br />

T. Iimure et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 141–147<br />

sample (Figs.1A and 4A). In the PAS at region I (Fig. 4B), intensities <strong>of</strong><br />

spots b8 (CMb), b9 (LTP), b10 (CMb), b13 (CMe) and b14 (CMe) were<br />

faint or invisible, those in region I-I (BDAI-1) were low except for the<br />

spot b2, and those in b11 and b12 (TAI) were distinct, respectively.<br />

These results suggest that the major PAS are protein Z4, protein Z7,<br />

and TAI, rather than BDAI-1, CMb and CMe.<br />

Fig. 2. The two-dimensional gel electrophoresis (2DE) images <strong>of</strong> the low molecular region, region I (Fig. 1A) <strong>of</strong> the haze samples i, ii, iii and iv. Arrows indicate the spot numbers.


Table 2<br />

Summary <strong>of</strong> protein identification by mass spectrometry analysis followed by<br />

a database search.<br />

Protein spot<br />

number<br />

Protein name Accession number Organism<br />

b0 protein Z-type serpin CAA66232 barley<br />

serpin CAA64599 barley<br />

Region I-I a<br />

barley dimeric<br />

alpha-amylase<br />

inhibitor; BDAI-1<br />

CAA08836 barley<br />

b5 thioredoxin; Trx-2p NP_011725 yeast<br />

b8 CMb component <strong>of</strong><br />

tetrameric alpha-amylase<br />

inhibitor<br />

CAA49556 barley<br />

b9 chain A, non-specific<br />

lipid transfer protein 1<br />

1MID-A barley<br />

b10 CMb component <strong>of</strong><br />

tetrameric alpha-amylase<br />

inhibitor<br />

CAA49556 barley<br />

b11 trypsin/amylase<br />

inhibitor pUP13<br />

1208404A barley<br />

b12 trypsin/amylase<br />

inhibitor pUP13<br />

1208404A barley<br />

b13 trypsin inhibitor<br />

CMe precursor<br />

P01086 barley<br />

b14 trypsin inhibitor<br />

CMe precursor<br />

P01086 barley<br />

a All spots in region I-I.<br />

Table 3<br />

The characteristics <strong>of</strong> the beer samples filtrated with silica gel (SiO 2 þ beer) and filtrated<br />

without silica gel (SiO 2 beer).<br />

SiO2 beer SiO2þ beer<br />

original gravity (%) 11.67 11.49<br />

final extract (%) 3.95 3.88<br />

apparent extract (%) 2.11 2.07<br />

alcohol (vol.%) 5.07 4.99<br />

pH 4.49 4.50<br />

color (EBC) 7.9 7.6<br />

bitter unit (mg/L) 28.2 28.1<br />

FT-3 a<br />

>10.0 1.68<br />

a FT-3 means the beer clarity after the forcing test (see Section 2).<br />

3.4. Proline compositions <strong>of</strong> the beer, the haze and the PAS<br />

It is well known that hordeins, one <strong>of</strong> the haze-active proteins,<br />

are proline rich (Asano et al., 1982). We examined the proline<br />

compositions <strong>of</strong> the beer proteins, the haze proteins and PAS (Table<br />

4). The proline compositions in the beer proteins were ca. 10 mol%,<br />

while those in PAS were ca. 20 mol%. These results suggest that<br />

proline rich proteins are adsorbed onto the silica gel. On the other<br />

Mr<br />

(kDa)<br />

97.4<br />

66.2<br />

45.0<br />

31.0<br />

21.5<br />

14.4<br />

T. Iimure et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 141–147 145<br />

hand, the proline compositions <strong>of</strong> the two haze proteins (samples i<br />

and iii) differed considerably (Table 4).<br />

4. Discussion<br />

To identify haze-active proteins, we prepared four haze-positive<br />

beer samples and analyzed the haze proteins using 2DE and mass<br />

spectrometry followed by a database search. Consequently, we<br />

assume that BDAI-1, CMb and CMe are haze-active, protein Z4,<br />

protein Z7 and LTP are moderately haze-active, and Trx-2p and TAI<br />

are haze-inactive proteins, respectively. Among the haze-active<br />

proteins the spots b10 (CMb), b13 (CMe) and b14 (CMe) were not<br />

observed in the beer protein sample i (Fig. 1) but spots in region I-I<br />

(BDAI-1) were intense (Fig. 1B). It is suggested, therefore, that CMb<br />

and CMe are concentrated in the haze fraction. In salting-out <strong>of</strong> the<br />

beer proteins, proteins did not precipitate at up to 20% saturation,<br />

but precipitated at 25% saturation <strong>of</strong> ammonium sulfate (data not<br />

shown). Therefore, the 25% salt-precipitated protein fraction may<br />

contain relatively hydrophobic proteins. The intense protein spots<br />

in region I-I (BDAI-1), b8 and b10 (CMb) were observed in a 25%<br />

salt-precipitated fraction (Fig. 1B), suggesting that BDAI-1 and CMb<br />

are potentially haze-active due to their high hydrophobicity. CMe is<br />

a possible haze-active protein although the mechanism is unknown<br />

at present.<br />

The spot intensities <strong>of</strong> regions I-I (BDAI-1), b10 (CMb), b13 and<br />

b14 (CMe) were different among the four haze samples (Fig. 2). In<br />

haze sample ii in Fig. 2 the protein spots b10 (CMb), b13 and b14<br />

(CMe) were less prominent despite the high intensity <strong>of</strong> the protein<br />

spots in region I-I (BDAI-1). This could have been due to the<br />

difference in the level <strong>of</strong> haze formability and suggested that BDAI-<br />

1 was more haze-active than both CMb and CMe.<br />

From the analyses <strong>of</strong> our four haze-positive beer samples, we<br />

consider that BDAI-1, CMb and CMe contribute to haze formation.<br />

The association <strong>of</strong> CMe with haze formation has already been<br />

suggested by Evans et al. (2003) and Robinson et al. (2004). They<br />

indicated that beer brewed from the malt <strong>of</strong> several barley varieties<br />

without the CMe band in western blot analysis using the antibody<br />

<strong>of</strong> silica eluate proteins showed improved haze stability compared<br />

to those with the CMe band. In order to further confirm the function<br />

<strong>of</strong> other haze-active proteins, i.e. BDAI-1 and CMb, we need to<br />

perform further experiments; for example force tests where these<br />

proteins are added to beer, and/or the development <strong>of</strong> novel<br />

cultivars without BDAI-1 or CMb which are then subject to the<br />

same trials.<br />

It has been shown that the addition <strong>of</strong> silica gel during beer<br />

filtration contributes to haze reduction due to adsorption <strong>of</strong> hazeactive<br />

proteins onto silica gel (Leiper et al., 2003; Siebert and Lynn,<br />

1997). We prepared two beer samples that were filtered with and<br />

without silica gel, i.e. SiO 2 þ beer and SiO 2 beer, respectively. The<br />

pI 3 pI 10 pI 3 pI 10<br />

SiO2 + beer SiO2-beer Fig. 3. The two-dimensional gel electrophoresis images <strong>of</strong> the beer filtered with silica gel (SiO 2 þ beer) and <strong>of</strong> that filtered without silica gel (SiO 2 beer).


146<br />

A<br />

Mr<br />

(kDa)<br />

B<br />

97.4<br />

66.2<br />

45.0<br />

31.0<br />

21.5<br />

14.4<br />

pI 3 pI 10<br />

Region I<br />

b11<br />

Region I-I<br />

b9<br />

b12<br />

PAS a<br />

PAS<br />

Fig. 4. The two-dimensional gel electrophoresis (2DE) images <strong>of</strong> the proteins adsorbed<br />

onto the silica gel (PAS). A: The whole image <strong>of</strong> the 2DE, pI 3–10, B: the enlarged image<br />

<strong>of</strong> region I. Arrows indicate the spot numbers.<br />

haze stability <strong>of</strong> SiO 2 þ beer was improved (Table 3). We then<br />

analyzed PAS using 2DE and identified protein species using mass<br />

spectrometry followed by a database search. As a result, the major<br />

proteins adsorbed onto the silica gel were identified as protein Z4,<br />

protein Z7 and TAI, rather than BDAI-1, CMb and CMe. Therefore,<br />

the majority <strong>of</strong> BDAI-1, CMb and CMe were transferred to beer after<br />

filtration. If BDAI-1, CMb and CMe were the predominant hazeactive<br />

proteins, the level <strong>of</strong> haze stability in SiO 2 þ beer might be<br />

comparable with that <strong>of</strong> SiO 2 beer. However, the haze stability in<br />

SiO 2 þ beer was improved compared to SiO 2 beer (Table 3). This<br />

evidence demonstrates that current silica treatment practices are<br />

only partially effective in improving haze stability because the<br />

majority <strong>of</strong> the BDAI-1, CMb and CMe remain in the treated beer. It<br />

follows that if these proteins could be removed, either by selection<br />

<strong>of</strong> novel barley varieties that do not contain these proteins, or<br />

improved haze stability treatments, the colloidal stability <strong>of</strong> beer<br />

could be further improved.<br />

Table 4<br />

The proline compositions <strong>of</strong> the beer proteins, the proteins adsorbed onto the silica<br />

gel (PAS) and the haze proteins.<br />

Sample name Proline (mol%)<br />

Beer SiO2þ beer 11.4<br />

SiO2 beer 10.7<br />

PAS a<br />

b2<br />

b10<br />

b8<br />

b13<br />

SiO 2 þ beer 16.2<br />

SiO 2 beer 23.4<br />

Haze sample i 5.6<br />

sample iii 27.9<br />

a PAS, proteins adsorbed onto the silica gel.<br />

T. Iimure et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 141–147<br />

b14<br />

It is conventionally accepted that haze proteins have elevated<br />

levels <strong>of</strong> proline (Siebert and Lynn, 1997). The proline composition<br />

<strong>of</strong> PAS was ca. 20 mol% (Table 4), while those <strong>of</strong> BDAI-1, CMb and<br />

CMe were as low as 6.6, 8.7 and 6.8 mol%, respectively. On the other<br />

hand, the proline composition <strong>of</strong> the haze proteins varied considerably<br />

between samples i (5.6 mol%) and iii (27.9 mol%), respectively<br />

(Table 4). If the beer colloidal haze consists <strong>of</strong> only proline<br />

rich proteins, such as hordein, the proline composition in all haze<br />

samples would be expected to be high (ca. 20 mol%). In addition, if<br />

the beer colloidal haze consists <strong>of</strong> only relatively proline poor<br />

proteins, such as BDAI-1, CMb and CMe, the proline compositions in<br />

all haze samples would be low (6–9 mol%). We estimate, therefore,<br />

that the proline composition <strong>of</strong> colloidal haze proteins might be<br />

determined by the degree <strong>of</strong> interaction between proline poor<br />

proteins, such as BDAI-1, CMb and CMe, and a core colloidal haze<br />

consisting <strong>of</strong> proline rich proteins such as hordeins. It is suggested<br />

that BDAI-1, CMb and CMe are not predominant haze-active<br />

proteins, but in fact initiation or growth factors in the formation <strong>of</strong><br />

colloidal haze.<br />

Proline rich protein such as hordein was not detected in the 2DE<br />

analysis <strong>of</strong> haze proteins, because hordein may not have been<br />

detectable by the 2DE analysis applied in this study due to being<br />

poorly silver stained. Therefore, our results do not exclude the<br />

contribution <strong>of</strong> hordein-derived polypeptides to haze formation. To<br />

further improve the beer colloidal haze, we need to develop novel<br />

silica gel having adsorbing affinity to BDAI-1, CMb and CMe or<br />

develop cultivars with low haze-active proteins. Okada et al. (2008)<br />

suggested that BDAI-1 is a possible foam-positive protein. Therefore<br />

the amount <strong>of</strong> BDAI-1 may well need to be optimized to<br />

produce beer with high qualities such as foam retention and haze<br />

stability.<br />

Acknowledgements<br />

We are grateful to T. Yazawa, K. Ito and H. Kato <strong>of</strong> the Bioresources<br />

Research and Development Department, Sapporo Breweries<br />

Ltd. for their technical assistance. This study was supported by<br />

the Program for Promotion <strong>of</strong> Basic Research Activities for Innovative<br />

Biosciences, Japan (PROBRAIN).<br />

References<br />

Apperson, K., Leiper, K.A., McKeown, I.P., Birch, D.J.S., 2002. Beer fluorescence and<br />

the isolation, characterization and silica adsorption <strong>of</strong> haze-active beer<br />

proteins. <strong>Journal</strong> <strong>of</strong> the Institute <strong>of</strong> Brewing 108, 193–199.<br />

Asano, K., Shinagawa, K., Hashimoto, N., 1982. Characterization <strong>of</strong> haze-forming<br />

proteins <strong>of</strong> beer and their roles in chill haze formation. <strong>Journal</strong> <strong>of</strong> the American<br />

Society <strong>of</strong> Brewing Chemists 40, 147–154.<br />

Bak-Jensen, K.S., Laugesen, S., Roepstorff, P., Svensson, B., 2004. Two-dimensional<br />

gel electrophoresis pattern (pH 6–11) and identification <strong>of</strong> water-soluble barley<br />

seed and malt proteins by mass spectrometry. Proteomics 4, 728–742.<br />

Bradford, M.M., 1976. A rapid and sensitive method for the quantization <strong>of</strong> microgram<br />

quantities <strong>of</strong> protein utilizing the principle <strong>of</strong> protein–dye binding.<br />

Analytical Biochemistry 72, 248–254.<br />

European Brewery Convention, Analytica, 1987. European Brewery Convention,<br />

fourth ed. Brauerei und Getraenke Rundschau, Zurich.<br />

Evans, D.E., Hejgaard, J., 1999. The impact <strong>of</strong> malt derived proteins on beer foam<br />

quality part I: the effect <strong>of</strong> germination and kilning on the level <strong>of</strong> proteins Z4,<br />

protein Z7 and LTP1. <strong>Journal</strong> <strong>of</strong> the Institute <strong>of</strong> Brewing 105, 159–169.<br />

Evans, D.E., Robinson, L.H., Sheehan, M.C., Tolhurst, R.L., Hill, A., Skerritt, J.S.,<br />

Barr, A.R., 2003. Application <strong>of</strong> immunological methods to differentiate<br />

between foam-positive and haze-active proteins originating from malt. <strong>Journal</strong><br />

<strong>of</strong> the American Society <strong>of</strong> Brewing Chemists 61, 55–62.<br />

Fukuda, K., Saito, W., Arai, S., Aida, Y., 1999. Production <strong>of</strong> a novel proanthocyanidin-free<br />

barley line with high quality. <strong>Journal</strong> <strong>of</strong> the Institute <strong>of</strong> Brewing 105,<br />

179–183.<br />

Leiper, K.A., Stewart, G.G., McKeown, I.P., 2003. Beer polypeptides and silica gel part<br />

I. Polypeptides involved in haze formation. <strong>Journal</strong> <strong>of</strong> the Institute <strong>of</strong> Brewing<br />

109, 57–72.<br />

McMurrough, I., Madigan, D., Smyth, M.R., 1995. Adsorption by polyvinylpolypyrrolidone<br />

<strong>of</strong> catechins and proanthocyanidins from barley. <strong>Journal</strong><br />

<strong>of</strong> Agricultural and Food Chemistry 43, 2687–2691.


Mikysˇka, A., Hrabák, M., Hasˇková, D., Sˇrogl, J., 2002. The role <strong>of</strong> malt and hop<br />

polyphenols in beer quality, flavour and haze stability. <strong>Journal</strong> <strong>of</strong> the Institute <strong>of</strong><br />

Brewing 108, 78–85.<br />

Okada, Y., Iimure, T., Takoi, K., Kaneko, T., Kihara, M., Hayashi, K., Ito, K., Sato, K.,<br />

Takeda, K., 2008. The influence <strong>of</strong> barley malt protein modification on beer<br />

foam stability and their relationship to the barley dimeric a-amylase inhibitor-I<br />

(BDAI-I) as a possible foam-promoting protein. <strong>Journal</strong> <strong>of</strong> Agricultural and Food<br />

Chemistry 56, 1458–1464.<br />

Østergaard, O., Melchior, S., Roepstorff, P., Svensson, B., 2002. Initial proteome<br />

analysis <strong>of</strong> mature barley seeds and malt. Proteomics 2, 733–739.<br />

Perkins, D.N., Pappin, D.J., Creasy, D.M., Cottrell, J.S., 1999. Probability-based protein<br />

identification by searching sequence databases using mass spectrometry data.<br />

Electrophoresis 20, 3551–3567.<br />

Perrocheau, L., Rogniaux, H., Boivin, P., Marion, D., 2005. Probing heat-stable watersoluble<br />

proteins from barley to malt and beer. Proteomics 5, 2849–2858.<br />

Robinson, L.H., Evans, D.E., Kaukovirta-Norja, A., Vilpola, A., Aldred, P., Home, S.,<br />

2004. The interaction between malt protein quality and brewing conditions and<br />

their impact on beer colloidal stability. Technical Quarterly – Master Brewers<br />

Association <strong>of</strong> the Americas 41, 353–362.<br />

Robinson, L.H., Healy, P., Stewart, D.C., Eglinton, J.K., Ford, C.M., Evans, D.E., 2007a.<br />

The identification <strong>of</strong> a barley haze active protein that influences beer haze<br />

stability: the genetic basis <strong>of</strong> a barley malt haze active protein. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong><br />

<strong>Science</strong> 45, 335–342.<br />

T. Iimure et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 141–147 147<br />

Robinson, L.H., Juttner, J., Milligan, A., Lahnstein, J., Eglinton, J.K., Evans, D.E., 2007b.<br />

The identification <strong>of</strong> a barley haze active protein that influences beer haze<br />

stability: cloning and characterization <strong>of</strong> the barley SE protein as a barley<br />

trypsin inhibitor <strong>of</strong> the chlor<strong>of</strong>orm/methanol type. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 45,<br />

343–352.<br />

Siebert, K.J., Carrasco, A., Lynn, P.Y., 1996. Formation <strong>of</strong> protein–polyphenol haze in<br />

beverages. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 44, 1997–2005.<br />

Siebert, K.J., Lynn, P.Y., 1997. Mechanisms <strong>of</strong> beer colloidal stabilization. <strong>Journal</strong> <strong>of</strong><br />

the American Society <strong>of</strong> Brewing Chemists 55, 73–78.<br />

Siebert, K.J., 1999. Effects <strong>of</strong> protein–polyphenol interactions on beverage haze, stabilization,<br />

and analysis. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 47, 353–362.<br />

van Nierop, S.N.E., Evans, D.E., Axcell, B.C., Cantrell, I.C., Rautenbach, M., 2004.<br />

Impact <strong>of</strong> different wort boiling temperatures on the beer foam stabilizing<br />

properties <strong>of</strong> lipid transfer protein 1. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry<br />

52, 3120–3129.<br />

von Wettstein, D., Jende-Strid, B., Ahrenst-Larsen, B., Erdal, K., 1980. Proanthocyanidin-free<br />

barley prevents the formation <strong>of</strong> beer haze. Technical Quarterly –<br />

Master Brewers Association <strong>of</strong> the Americas 17, 16–23.<br />

von Wettstein, D., Nilan, R.A., Ahrenst-Larsen, B., Erdal, K., Ingversen, J., Jende-<br />

Strid, B., Kristiansen, K.N., Larsen, J., Outtrup, H., Ullrich, S.E., 1985. Proanthocyanidin-free<br />

barley for brewing: progress in breeding for high yield and<br />

research tool in polyphenol chemistry. Technical Quarterly – Master Brewers<br />

Association <strong>of</strong> the Americas 22, 41–52.


Mixing properties and dough functionality <strong>of</strong> transgenic lines <strong>of</strong> a commercial<br />

wheat cultivar expressing the 1Ax1, 1Dx5 and 1Dy10 HMW glutenin subunit<br />

genes<br />

Elena León a , Santiago Marín a , María J. Giménez a , Fernando Piston a , Marta Rodríguez-Quijano b ,<br />

Peter R. Shewry c , Francisco Barro a, *<br />

a<br />

Instituto de Agricultura Sostenible, CSIC, Alameda del Obispo s/n, 14080 Córdoba, Spain<br />

b<br />

Universidad Politécnica de Madrid, 28040 Madrid, Spain<br />

c<br />

Rothamsted Research, Harpenden, Hertfordshire AL5 2JQ, UK<br />

article info<br />

Article history:<br />

Received 18 April 2008<br />

Received in revised form 13 August 2008<br />

Accepted 18 August 2008<br />

Keywords:<br />

1Ax1<br />

1Dx5<br />

1Dy10<br />

Anza<br />

GM wheat<br />

1. Introduction<br />

abstract<br />

The viscoelastic properties <strong>of</strong> wheat flour dough allow wheat to<br />

be used for making bread and many other food products such as<br />

cake, biscuits, pasta and noodles. The unique properties <strong>of</strong> wheat<br />

result from the unusual biomechanical properties <strong>of</strong> the gluten<br />

proteins, which form a network conferring elasticity and<br />

* Corresponding author. Tel.: þ34 957 499240; fax: þ34 957 499252.<br />

E-mail address: ge1bal<strong>of</strong>@uco.es (F. Barro).<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.08.002<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 148–156<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

In this work we report the effects <strong>of</strong> the HMW-GS 1Ax1, 1Dx5 and 1Dy10 on the breadmaking quality <strong>of</strong><br />

the bread wheat cultivar Anza that contains the HMW-GS pairs 1Dx2 þ 1Dy12 and 1Bx7* þ 1By8, and is<br />

null for the Glu-A1 locus. This allows the characterization <strong>of</strong> individual subunits 1Dx5 and 1Dy10 in the<br />

absence <strong>of</strong> subunit 1Dx5, and the interactions between these subunits and subunits 1Dx2 and 1Dy12 to<br />

be determined. Three transgenic lines termed T580, T581 and T590, containing, respectively, the HMW-<br />

GS 1Ax1, 1Dx5 and 1Dy10 were characterized over 3 years using a range <strong>of</strong> widely-used grain and dough<br />

testing methods. The transgenic subunits 1Ax1, 1Dx5 and 1Dy10 accounted for 25.2%, 20.3% and 17.9%,<br />

respectively, <strong>of</strong> the total HMW-GS in the three transgenic lines. Although lines T581 and T590 expressed<br />

similar levels <strong>of</strong> subunits 1Dx5 and 1Dy10 they had different effects on other aspects <strong>of</strong> protein<br />

composition, including changes in the ratios <strong>of</strong> glutenin/gliadin, <strong>of</strong> HMW/LMW-GS, the 1Dx2/1Dy12, the<br />

x-type/y-type HMW-GS and the proportions <strong>of</strong> high molecular mass glutenin polymers. In contrast, lines<br />

transformed to express subunits 1Ax1 and 1Dx5 showed similar changes in protein composition, with<br />

higher protein contents and decreased ratios <strong>of</strong> glutenin/gliadin and 1Dx2/1Dy12. In addition, both<br />

transgenic lines showed similar increases in the ratio <strong>of</strong> x-type/y-type subunits compared to the control<br />

line. The transgenic lines were analysed using Farinograph, Mixograph and Alveograph. This confirmed<br />

that the expression <strong>of</strong> all three subunits resulted in increased dough strength (and hence breadmaking<br />

quality) <strong>of</strong> the cultivar Anza. A beneficial effect <strong>of</strong> subunit 1Dx5 has not been reported previously,<br />

transgenic wheat lines expressing this subunit giving overstrong dough unsuitable for breadmaking.<br />

However, the expression <strong>of</strong> subunit 1Dy10 had a greater effect on breadmaking quality than subunits<br />

1Ax1 and 1Dx5. The Farinograph parameters such as dough stability and peak time were increased by<br />

9.2-fold and 2.4-fold, respectively, in line T590 (expressing 1Dy10) with respect to the control line.<br />

Similarly, the Mixograph mixing time was increased by four-fold and the resistance breakdown<br />

decreased by two-fold in line T590 compared with the control line. The Alveograph W value was also<br />

increased by 2.7-fold in line T590 compared to the control line. These transgenic lines are <strong>of</strong> value for<br />

studying the contribution <strong>of</strong> specific HMW-GS to wheat flour functional properties.<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

extensibility to the dough. One group <strong>of</strong> gluten proteins, the high<br />

molecular weight glutenin subunits (HMW-GS), play an important<br />

role in determining the functional properties <strong>of</strong> wheat dough. The<br />

HMW-GS <strong>of</strong> wheat have been studied in detail, demonstrating that<br />

allelic differences in their composition result in effects on glutenin<br />

polymers and hence breadmaking quality (Payne, 1987; Shewry<br />

et al., 2003). Bread wheat contains six HMW-GS genes, with tightly<br />

linked pairs <strong>of</strong> genes encoding x- and y- type subunits being<br />

present at each <strong>of</strong> the Glu-A1, Glu-B1, and Glu-D1 loci on the long<br />

arms <strong>of</strong> chromosomes 1A, 1B, and 1D, respectively. Wheat cultivars


containing HMW-GS 1Dx5 and 1Dy10 have stronger doughs than<br />

those containing the allelic subunits 1Dx2 and 1Dy12 (Shewry<br />

et al., 2003). However, because the x- and y-type genes are so<br />

tightly linked in the D genome, it has not been possible to separate<br />

them by recombination and hence to determine their individual<br />

contributions to the breadmaking quality <strong>of</strong> wheat.<br />

Genetic engineering provides an opportunity to develop cultivars<br />

with new HMW-GS combinations in order to explore the<br />

molecular and biochemical bases for the breadmaking quality <strong>of</strong><br />

wheat and to generate lines with improved or novel functional<br />

properties. This approach has also been used to investigate the<br />

effects <strong>of</strong> HMW-GS genes when expressed in other cereals. Thus,<br />

genes encoding HMW-GS have been expressed in wheat (Altpeter<br />

et al., 1996; Barro et al., 1997; Blechl and Anderson, 1996), tritordeum<br />

(Rooke et al., 1999a), maize (Sangtong et al., 2002), and rye<br />

(Altpeter et al., 2004).<br />

Expression <strong>of</strong> additional HMW-GS genes in wheat has resulted<br />

in increased amounts <strong>of</strong> HMW-GS protein and changes in the<br />

grain functional properties, the effects varying depending on the<br />

specific genes and genetic backgrounds. Thus, expression <strong>of</strong><br />

subunit 1Ax1 in transgenic wheat resulted in lines with improved<br />

rheological properties while lines expressing subunit 1Dx5<br />

resulted in unsuitable properties for breadmaking (Barro et al.,<br />

2003; Darlington et al., 2003). In particular, high levels <strong>of</strong><br />

expression <strong>of</strong> subunit 1Dx5 resulted in greatly increased dough<br />

strength, being too strong for conventional breadmaking (Alvarez<br />

et al., 2001; Darlington et al., 2003; Rooke et al., 1999b). The<br />

differences between the results obtained with subunits 1Ax1 and<br />

1Dx5 may be due to the fact that subunit 1Dx5 has an additional<br />

cysteine residue which promotes a more highly cross-linked<br />

glutenin network (Darlington et al., 2003; Popineau et al., 2001).<br />

Also, because subunit 1Dx5 only occurs with 1Dy10 it has been<br />

suggested that the effects might be mitigated by increasing<br />

subunit 1Dy10 at the same time (Barro et al., 2003; Darlington<br />

et al., 2003).<br />

More recently, the effects <strong>of</strong> expression <strong>of</strong> transgenic subunits<br />

1Dx5 and/or 1Dy10 on flour properties in a wheat line which<br />

already expresses the endogenous forms <strong>of</strong> these subunits have<br />

been reported (Blechl et al., 2007). The results confirmed the<br />

effects <strong>of</strong> the increased levels <strong>of</strong> subunit 1Dx5 on dough strength<br />

but showed that increases in subunit 1Dy10 alone also resulted in<br />

curves typical <strong>of</strong> strong doughs. Furthermore, the expression <strong>of</strong><br />

subunits 1Dx5 and 1Dy10 in the same lines failed to alleviate the<br />

overstrong characteristics associated with subunit 1Dx5 (Blechl<br />

et al., 2007). However, the effects <strong>of</strong> subunits 1Dx5 and 1Dy10<br />

have not been tested in wheat backgrounds that do not contain<br />

these subunits. We report here the transformation <strong>of</strong> wheat with<br />

subunits 1Ax1, 1Dx5 and 1Dy10 using a genetic background<br />

which does not express endogenous forms <strong>of</strong> subunits 1Dx5 and<br />

1Dy10 but the allelic subunits 1Dx2 and 1Dy12. This allows the<br />

effects <strong>of</strong> individual subunits 1Dx5 and 1Dy10, and the interactions<br />

between these and subunits 1Dx2 and 1Dy12 to be<br />

determined.<br />

2. Material and methods<br />

2.1. Plasmids<br />

Three plasmids, p1Dx5, p1Ax1 and pK-Dy10A containing,<br />

respectively, the genomic fragment for the HMW glutenin subunits<br />

1Dx5 (Anderson et al., 1989), 1Ax1 (Halford et al., 1992), and 1Dy10<br />

(Anderson et al., 1989), were used in combination with plasmid<br />

pACH25 (Christensen and Quail, 1996) which contains the bar gene<br />

that confers tolerance to phosphinothricin (PPT) under the control<br />

<strong>of</strong> maize ubiquitin promoter.<br />

E. León et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 148–156 149<br />

2.2. Plant transformation and in vitro culture<br />

Immature scutella (0.5–1.5 mm in length) <strong>of</strong> wheat were used as<br />

target tissue for transformation by particle bombardment. Caryopses<br />

were harvested 16 days after anthesis and explants isolated<br />

as described by Barcelo and Lazzeri (1995). For each bombardment,<br />

30 scutella were excised and placed in the centre <strong>of</strong> a 9 cm Petri<br />

dish containing MP4 medium which consisted <strong>of</strong> MS (Murashige<br />

and Skoog, 1962) induction medium supplemented with 30 g l 1<br />

sucrose and 4 mg l 1 picloram. Explants were cultured in the dark<br />

at 25 C for 5 days prior to bombardment. Explants were subjected<br />

to a 4 h osmotic treatment, both before and after bombardment, in<br />

the same induction medium described above, but supplemented<br />

with 0.4 M mannitol.<br />

For bombardment, plasmids were precipitated onto 0.6 mm gold<br />

particles at 0.5 pmol mg 1 gold for pAHC25 and 0.75 pmol mg 1<br />

gold for the plasmids containing the HMW glutenin subunit genes.<br />

For each shot, 58 mg <strong>of</strong> coated particles were delivered at a pressure<br />

<strong>of</strong> 1100 psi using a PDS 1000/He particle gun (BioRad). Following<br />

bombardment, plates containing explants were cultured in MP4<br />

medium (see above) in the dark at 25 C for 3 weeks and then<br />

transferred to a RZ regeneration medium (Barro et al., 1998) supplemented<br />

with 2 mg l 1 <strong>of</strong> PPT and cultured for a further 3 weeks<br />

at 25 C with a photoperiod <strong>of</strong> 12 h. Surviving explants were<br />

transferred to R regeneration medium (Barro et al., 1998) supplemented<br />

with 2 mg l 1 PPT and cultured for a further two rounds <strong>of</strong><br />

3 weeks each. Putative transgenic plants were then transferred to<br />

soil and grown to maturity in the greenhouse. Homozygous<br />

progeny <strong>of</strong> plants containing HMW-GS transgenes were identified<br />

by SDS–PAGE <strong>of</strong> endosperm proteins by single half-seed descent<br />

(see below). Homozygous lines were self-pollinated for three<br />

generations and assayed during 3 years, using a randomized<br />

complete block design with two replicates, as described (Barro<br />

et al., 2002).<br />

2.3. Genomic Southern blot analysis<br />

Total genomic DNA was extracted from leaf tissue using a CTAB<br />

method (Stacey and Isaac, 1994). Genomic DNA was digested with<br />

either EcoRV (which cuts once within the p1Ax1 and pK-Dy10A<br />

constructs) or ScaI (which cuts once within the p1Dx5 construct).<br />

Digested DNA was separated by electrophoresis using 0.8% (w/v)<br />

agarose gel, blotted onto Hybond-N þ membrane (Amersham<br />

Biosciences) and baked for 2 h at 80 C. Hybridization and signal<br />

detection were performed according to Roche Diagnostic protocols.<br />

Briefly, the membrane was prehybridized for 1 h with DIG Easy Hyb<br />

Granules (Roche Diagnostics) at 48 C. Southern filters were<br />

hybridized overnight at 48 C with PCR-generated digoxigeninlabelled<br />

probes produced using primers for the 1Ax1 (Rooke et al.,<br />

1999a), 1Dx5 (D’Ovidio and Anderson, 1994) and 1Dy10 genes (5 0 -<br />

CCA CAA CAC CGA GCA CCA CAA-3 0 ,5 0 -GGG CGG CAC CAC AGT TTG<br />

CTC-3 0 ), with p1Ax1, p1Dx5 and pK-Dy10A as the DNA template,<br />

respectively. After hybridization the membrane was washed twice<br />

with a solution <strong>of</strong> 0.1% (w/v) SDS in 2 SSC for 15 min at 65 C. The<br />

DIG Luminescent Detection Kit (Roche Diagnostics) was used for<br />

signal detection. Finally, the membrane was exposed to AGFA<br />

medical X-ray (Agfa-Gevaert NV) film for 1–6 h.<br />

2.4. Protein and SDS–PAGE analysis<br />

The protein content <strong>of</strong> flour was calculated from the Kjeldahl<br />

nitrogen content (%N 5.7) and expressed on a dry matter basis.<br />

For protein extraction and quantification <strong>of</strong> the different protein<br />

fractions, all subplots in a year from the same line were bulked.<br />

Therefore, 30 individual seeds per line (ten seeds from each year)<br />

were crushed into a fine powder and used to extract the endosperm


150<br />

storage proteins. Gliadins were extracted in 60% (v/v) <strong>of</strong> aqueous<br />

ethanol using a rotary shaker for 40 min. Samples were centrifuged<br />

at 13,000 g for 5 min and the supernatant collected. Glutenins were<br />

extracted in 625 mM Tris–HCl pH 6.8, 5% (v/v) 2-mercaptoethanol,<br />

10% (v/v) glycerol, 0.02% (w/v) bromophenol blue and 2% (w/v)<br />

dithiothreitol in a 5:1 ratio (ml:mg) to wholemeal and separated<br />

using a Tris–borate buffer system and 10% (w/v) acrylamide gels<br />

(Shewry et al., 1995). Gliadin and glutenin contents were determined<br />

in the above extracts using a Bradford assay system (Bradford,<br />

1976). For densitometry analysis, glutenins were separated by<br />

SDS–PAGE gels and analysed using a Kodak Image Station 440CF<br />

and Kodak 1D Image Analysis S<strong>of</strong>tware using the SDS–PAGE<br />

Molecular Weight Standards from Bio-Rad as reference.<br />

Falling Number was determined according to the standard ICC<br />

method no. 107 (ICC, 1995). Sedimentation volume was determined<br />

according to Zeleny using the standard ICC method no. 116 (ICC,<br />

1994).<br />

SE-HPLC was carried out in duplicate at the Campden and<br />

Chorleywood Food Research Association (Chipping Campden, UK)<br />

as described by Morel et al (2000) and Millar (2003).<br />

2.5. Alveograph<br />

Approximately 0.8 kg samples <strong>of</strong> seeds per line and year were<br />

milled with a Chopin CD1 mill to obtain white flour. Alveograph<br />

tests were performed using an Alveograph MA 82 (Tripette et<br />

Renaud, France) and the dough properties determined in five<br />

replicates according to standard ICC method no. 121 (ICC, 1992a).<br />

The alveograms were evaluated to determine overpressure<br />

(tenacity) (P), the average length <strong>of</strong> the curve (extensibility) (L), the<br />

P/L ratio, the deformation work (W), the swelling index (G) and the<br />

elasticity index.<br />

2.6. Farinograph<br />

Approximately 0.2 kg samples <strong>of</strong> seeds per line and year were<br />

milled and mixed using a 50 g mixer for standard Farinograph<br />

analysis (Brabender GmbH & Co. KG, Germany) test. The Farinograph<br />

characteristics were determined in two different replicates<br />

per year according to standard ICC method no. 115/1 (ICC, 1992b).<br />

Farinograph parameters evaluated included water absorption,<br />

arrival time, departure time, stability, peak time, mixing tolerance<br />

time and degree <strong>of</strong> s<strong>of</strong>tening.<br />

2.7. Mixograph<br />

Dough mixing properties were determined with a 10 g Mixograph<br />

(National Manufacturing Co., Lincoln NE). Samples were<br />

mixed to optimum water absorption following 54-40A method<br />

(AACC, 2000). The mixing parameters determined were mixing<br />

time, peak resistance and resistance breakdown.<br />

2.8. Statistics<br />

Data were analysed using the SPSS version 11.0 statistical s<strong>of</strong>tware<br />

package (SPSS Inc., Chicago, IL, USA). The general analysis <strong>of</strong><br />

variance and the least significant difference pairwise comparisons<br />

<strong>of</strong> means were used to determine significant differences.<br />

3. Results<br />

Plasmids p1Ax1, p1DX5 and pKS-Dy10A, containing, respectively,<br />

the HMW-GS 1Ax1, 1Dx5 and 1Dy10, were used for transformation<br />

<strong>of</strong> the bread wheat cultivar Anza. Transformation<br />

efficiency (transformants per scutellum bombarded) for this<br />

cultivar was 0.4%. This cultivar contains four endogenous HMW-GS:<br />

E. León et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 148–156<br />

1Bx7*, 1By8, 1Dx2 and 1Dy12 (Fig. 1A). Single seed descent was<br />

used to obtain homozygotes for lines containing HMW glutenin<br />

subunits transgenes by screening 30 half-seeds in each generation<br />

by SDS–PAGE. Homozygous transgenic lines expressing the HMW-<br />

GS 1Ax1, 1Dx5 and 1Dy10 transgenes (Fig. 1A) were termed T580,<br />

T581 and T590, respectively. These lines were multiplied by selfpollinating<br />

for three generations and assayed for 3 years as<br />

described (Barro et al., 2002).<br />

Southern blot analysis was complicated by the presence <strong>of</strong><br />

cross-hybridizing bands in the control line, which was probably<br />

due to the sequence similarity between the endogenous and<br />

transgenic HMW-GS. Nevertheless, the banding pattern after<br />

hybridization showed that transgenic lines contained four, two and<br />

three insertion events for HMW-GS 1Ax1, 1Dx5 and 1Dy10,<br />

respectively (Fig. 1B).<br />

3.1. Protein characterization<br />

The total protein contents <strong>of</strong> the flours ranged from 11.7% to<br />

13.7% for the control line and line T581 transformed with HMW-GS<br />

1Dx5, respectively (Table 1). All three transgenic lines had higher<br />

protein contents than the control line but only the differences<br />

between lines T580 and T581, containing HMW-GS 1Ax1 and 1Dx5,<br />

respectively, were statistically significant in comparison to the<br />

control line (Table 1). As shown in Table 1, the glutenin fraction<br />

decreased from 4.21 mg/mg flour in the control line to 3.25 and 3.62<br />

in lines T580 and T581, but increased to 5.28 in line T590. In<br />

contrast, the gliadin contents increased significantly in all three<br />

transgenic lines with respect to the control line, with no significant<br />

differences among the three transgenic lines. These changes are<br />

also reflected in the glutenin/gliadin ratio. Thus, lines T580 and<br />

T581 had lower glutenin/gliadin ratios than the control line, but<br />

line T590 had a similar ratio to that <strong>of</strong> control line (Table 1).<br />

The expression <strong>of</strong> HMW-GS in the transgenic lines resulted in<br />

changes in amounts and proportions <strong>of</strong> HMW-GS in the endosperm.<br />

Thus, the transgenic subunits 1Ax1, 1Dx5 and 1Dy10<br />

accounted for 25.2%, 20.3% and 17.9%, respectively, <strong>of</strong> the total<br />

HMW in the three transgenic lines (Table 1). These increases in<br />

HMW subunit content were associated with decreases in LMW<br />

subunit content in all transgenic lines, but the differences were only<br />

significant for line T581 expressing the 1Ax1 subunit (Table 1).<br />

Expression <strong>of</strong> the transgenic HMW-GS was also associated with<br />

a significant decrease in the amounts <strong>of</strong> endogenous HMW-GS<br />

(Table 1). Furthermore, although all endogenous subunits were<br />

decreased, the relative effects varied. The expression <strong>of</strong> subunits<br />

1Ax1 and 1Dx5 in lines T580 and T581, respectively, led to greater<br />

decreases in the contents <strong>of</strong> the endogenous subunit 1Dx2 (Table<br />

1), with the Dx2/Dy12 ratio decreasing significantly from 1.11 in the<br />

control line to 0.83 and to 0.88 in lines T580 and T581, respectively.<br />

Furthermore, the ratio <strong>of</strong> x-type/y-type HMW-GS also increased in<br />

lines T580 and T581, respectively. In contrast, the contents <strong>of</strong> the<br />

endogenous subunits 1Dx2 and 1Dy12 were decreased in line T590<br />

expressing subunit 1Dy10 but to a similar extent (Table 1). Therefore,<br />

the ratio <strong>of</strong> Dx2/Dy12 in this line was comparable to that in the<br />

control line (Table 1), although the ratio <strong>of</strong> the x-type/y-type was<br />

decreased from 1.25 in the control line to 0.91 as a consequence <strong>of</strong><br />

the addition <strong>of</strong> a new y-type subunit.<br />

There is a well-established correlation between the breadmaking<br />

performance <strong>of</strong> flour and the proportion <strong>of</strong> high molecular mass<br />

glutenin polymers (Gupta and MacRitchie, 1994; Popineau et al.,<br />

1994). The proportions <strong>of</strong> monomeric, oligomeric and polymeric<br />

gluten proteins in the flours were therefore determined by SE-HPLC<br />

essentially as described by Morel et al. (2000). This method uses<br />

sonication in 1% (w/v) SDS in 0.1 M phosphate buffer, pH 6.9, to<br />

extract the total grain proteins, which are then separated into five<br />

fractions corresponding broadly to high molecular mass glutenin


A<br />

B<br />

1Ax1<br />

polymers (F1), lower molecular mass glutenin polymers (F2), ugliadins<br />

(F3), a- and g-type gliadins (F4) and albumins and globulins<br />

(F5). The sonication procedure results in shearing <strong>of</strong> glutenin<br />

polymers and hence the sizes <strong>of</strong> the polymers separated by SE-<br />

HPLC do not accurately reflect the sizes <strong>of</strong> those present in vivo.<br />

However, the relative amounts <strong>of</strong> the peaks do relate to dough<br />

strength with %F1/%F2 and (%F3 þ %F4)/%F1 showing particularly<br />

strong correlations (Millar, 2003).<br />

All three transgenic lines had higher values for %F1 and %F1/%F2,<br />

indicating that they had higher proportions <strong>of</strong> high molecular mass<br />

glutenin polymers. This effect was also greater in line T590<br />

expressing the 1Dy10 subunit than in lines T580 and T581.<br />

However, values for (%F3 þ %F4)/%F1 decreased in all the transgenic<br />

lines relative to the untransformed parent (Table 1).<br />

3.2. Technological properties <strong>of</strong> flour<br />

T580<br />

1Dx2<br />

1Bx7*<br />

1By8<br />

1Dy12<br />

The potential breadmaking quality <strong>of</strong> the flours was determined<br />

over 3 years using a range <strong>of</strong> widely-used grain and dough testing<br />

methods (Table 2). The Falling numbers <strong>of</strong> all <strong>of</strong> the samples were<br />

higher than the recommended limit <strong>of</strong> 180 s with no differences<br />

among all four lines. The grain test weight was also similar for all<br />

four lines. The thousand seed weight <strong>of</strong> line T580 was higher than<br />

that <strong>of</strong> the other lines. The Zeleny sedimentation was increased in<br />

all three transgenic lines and this was greatest in the line<br />

expressing subunit 1Dy10.<br />

The rheological properties and water absorption <strong>of</strong> the dough<br />

were determined using a Brabender Farinograph (Fig. 2) and10g<br />

Mixograph (Fig. 3) while the dough viscoelastic properties were<br />

determined using a Chopin Alveograph (Fig. 4). The Farinograph<br />

E. León et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 148–156 151<br />

WT<br />

WT<br />

T581<br />

1Dx5<br />

T590<br />

T580 WT WT T581 WT T590<br />

Fig. 1. (A) SDS–PAGE <strong>of</strong> seed protein extracts from transgenic wheat lines and their non-transformed parent (WT). The location <strong>of</strong> the HMW-GS native to Anza (1Dx2, 1Bx7*, 1By9<br />

and 1Dy12) are indicated. (B) Southern blot analysis <strong>of</strong> the transgenic lines. Genomic DNA from lines T580, T581 and T590 was digested with EcoRV, ScaI and EcoRV, respectively,<br />

and hybridized with probes amplified from plasmids p1Ax1, p1Dx5 and pK-Dy10A. Four, three and two extra hybridized bands, compare to non-transformed parent (WT), are<br />

present in DNA <strong>of</strong> lines T580, T581 and T590, respectively.<br />

WT<br />

1Dy10<br />

characteristics <strong>of</strong> transgenic lines are shown in Table 2. No<br />

significant differences were detected for water absorption but the<br />

rheological properties differed with the arrival and peak times,<br />

increasing significantly in all three transgenic lines in comparison<br />

to the control line. The transgenic lines expressing subunits 1Ax1<br />

and 1Dx5 gave similar values for arrival time and peak time while<br />

the line expressing subunit 1Dy10 gave the highest values for<br />

these two parameters (Table 2). The dough stability and departure<br />

time were also significantly higher for the line transformed with<br />

subunit 1Dy10. Dough stability is an indicator <strong>of</strong> the overall<br />

quality <strong>of</strong> the protein in the flour and this increased from 1.3 in the<br />

control line to 11.9 in the line expressing subunit 1Dy10 (Table 2).<br />

The mixing tolerance index (MTI) is an indicator <strong>of</strong> how well the<br />

dough will perform during the critical final stages <strong>of</strong> mixing and is<br />

related to strength (Shuey, 1984). Thus, a high MTI means that the<br />

dough will tend to break down quickly while a low MTI may<br />

indicate that the dough will require longer mixing to fully<br />

develop. The MTI was decreased significantly in all three transgenic<br />

lines with respect to that <strong>of</strong> the control line. However, this<br />

decrease was greater for transgenic line T590 expressing subunit<br />

1Dy10, for which the MTI value fell from 132.8 (BU) in the control<br />

line to 35.2 (BU) (Table 2). The degree <strong>of</strong> s<strong>of</strong>tening varied from<br />

132.5 (BU) in the control line to 56.1 (BU) in the line transformed<br />

with subunit 1Dy10. Higher values <strong>of</strong> degree <strong>of</strong> s<strong>of</strong>tening,<br />

measured in Brabender units, are characteristic <strong>of</strong> weak flours and<br />

lower values <strong>of</strong> strong flours.<br />

The rheological properties <strong>of</strong> the transgenic wheats were also<br />

determined using a 10 g Mixograph (Fig. 3). The Mixograph is<br />

widely used in cereal research as it measures a range <strong>of</strong> rheological<br />

parameters that relate to the behaviour <strong>of</strong> the dough in bread


152<br />

Table 1<br />

Characterization <strong>of</strong> the storage proteins in transgenic and control lines <strong>of</strong> wheat cultivar Anza<br />

Parameter Line<br />

Anza T580 T581 T590<br />

Endogenous HMW 7* þ 8, 2 þ 12 7* þ 8, 2 þ 12 7* þ 8, 2 þ 12 7* þ 8, 2 þ 12<br />

Transgenic HMW NA 1Ax1 1Dx5 1Dy10<br />

Number <strong>of</strong> insertions NA 4 2 3<br />

Crude protein (%) a<br />

11.7 0.3 13.2 0.4 13.7 0.6 12.2 0.3<br />

Glutenin (mgmg 1 flour) b<br />

4.2 0.2 3.3 0.2 3.6 0.4 5.3 0.2<br />

Gliadin (mgmg 1 flour) b<br />

2.7 0.1 3.4 0.1 3.7 0.4 3.5 0.1<br />

Glutenin/gliadin<br />

HMW composition<br />

1.53 0.04 0.96 0.05 0.98 0.06 1.51 0.04<br />

c<br />

1Ax1 NA 25.2 0.6 NA NA<br />

1Dx2 27.8 0.6 16.8 0.5 19.1 0.4 21.5 0.7<br />

1Dx5 NA NA 20.3 0.8 NA<br />

1Bx7* 27.7 0.9 22.2 0.3 23.4 0.5 26.2 0.2<br />

1By8 19.4 0.9 15.5 0.3 15.6 0.9 15.9 0.9<br />

1Dy10 NA NA NA 17.9 0.6<br />

1Dy12 25.1 0.4 20.3 0.8 21.6 0.9 18.5 0.7<br />

making and other food processing systems. The machine measures<br />

the resistance as the dough is mixed, the most important parameters<br />

being the mixing time (time to maximum resistance), the<br />

peak resistance (maximum resistance obtained during mixing) and<br />

the resistance breakdown (the loss <strong>of</strong> resistance on over-mixing). In<br />

E. León et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 148–156<br />

HMW (% glutenins) 31.7 0.79 32.4 1.21 36.8 0.52 32.9 0.45<br />

LMW (% glutenins) 68.3 0.61 67.6 1.13 63.2 0.49 67.1 0.51<br />

HMW/LMW ratio 0.47 0.01 0.49 0.02 0.58 0.01 0.49 0.01<br />

1Dx2/1Dy12 ratio 1.11 0.03 0.83 0.05 0.88 0.06 1.16 0.07<br />

1Bx7/1By8 ratio 1.43 0.11 1.43 0.04 1.50 0.12 1.64 0.09<br />

x-type/y-type ratio 1.25 0.06 1.80 0.07 1.69 0.06 0.91 0.03<br />

SE-HPLC d<br />

%F1 12.7 0.35 13.6 0.35 13.0 0.05 14.1 0.30<br />

%F1/%F2 0.60 0.02 0.61 0.01 0.62 0.00 0.68 0.01<br />

(%F3 þ %F4)/%F1<br />

NA, not applicable.<br />

4.08 0.12 3.63 0.12 3.88 0.03 3.54 0.10<br />

a<br />

Kjeldahl method. Average <strong>of</strong> six replications standard error.<br />

b<br />

Glutenin and gliadins were determined using a Bradford assay system. Average <strong>of</strong> 30 replications standard error.<br />

c<br />

HMW composition was determined by densitometry analysis. Average <strong>of</strong> 30 replications standard error.<br />

d<br />

SE-HPLC was carried out in duplicate by bulking samples from the third year. Average standard error.<br />

general, strong doughs have long mixing times, high peak resistances<br />

and low resistance breakdown. Transformation with the<br />

three HMW subunit genes led in all cases to increased dough<br />

strength, with quantitative differences depending on the subunit<br />

expressed (Fig. 3, Table 2). Thus, the mixing time increased from<br />

Table 2<br />

Technological properties <strong>of</strong> the transgenic and control lines <strong>of</strong> wheat cultivar Anza<br />

Parameter a<br />

Line<br />

Anza T580 T581 T590<br />

Falling number (s) 350 9.5 350 9.8 350 3.0 346 1.2<br />

Grain test weight (g l 1 ) 79.9 3.0 79.8 5.0 79.4 4.3 80.2 4.1<br />

Thousand seed weight (g) 38.4 0.9 42.1 1.6 37.5 0.7 39.7 0.9<br />

Zeleny sedimentation (ml) 20.2 0.4 28.2 2.0 31.3 3.8 41.1 3.0<br />

Farinograph<br />

Water absorption 72.2 1.5 72.8 0.9 73.8 2.8 73.7 1.2<br />

Arrival time (min) 2.1 0.1 3.4 0.6 3.7 0.2 4.6 0.3<br />

Peak time (min) 2.7 0.3 4.4 0.5 4.8 0.4 6.6 0.8<br />

Dough stability (min) 1.3 0.3 2.7 0.2 2.7 0.3 11.9 2.2<br />

Departure time (min) 3.5 0.4 6.0 0.6 6.3 0.6 16.5 2.1<br />

Mixing Tolerance Index (BU) 132.8 14.4 77.5 9.2 80.8 7.6 35.2 3.8<br />

Degree <strong>of</strong> s<strong>of</strong>tening (BU) 132.5 18.9 76.7 8.5 71.7 9.2 56.1 6.2<br />

Mixograph<br />

Mixing time (s) 30.0 72.0 87.0 120.0<br />

Peak resistance (AU) 58.0 75.0 60.0 61.5<br />

Resistance breakdown (%) 37.9 26.7 18.3 18.7<br />

Alveograph<br />

Tenacity, P (mm) 42.2 5.1 65.8 3.1 71.7 3.7 79.0 5.4<br />

Extensibility, L (mm) 94.2 4.5 91.8 3.0 61.0 0.6 82.3 5.9<br />

P/L ratio 0.4 0.03 0.7 0.04 1.2 0.07 1.0 0.12<br />

W-value ( 10 4 J) 97 8.2 170 13.2 180 12.1 258 4.9<br />

a<br />

Values are the average <strong>of</strong> six replications standard error, except for Mixograph parameters for which bulked samples from the third year were used.


30 s in the control line to 72 in that with subunit 1Ax1, to 87 in that<br />

with subunit 1Dx5, and to 120 in that with subunit 1Dy10 (Table 2).<br />

However, the peak resistance values were similar for all lines except<br />

for the line T580 expressing subunit 1Ax1 where it was higher<br />

(Table 2). The resistance breakdown decreased from 37.9% in the<br />

control line to 26.7, to 18.3, and to 18.7 in the lines with subunits<br />

1Ax1, 1Dx5 and 1Dy10, respectively (Table 2).<br />

The Chopin Alveograph is a dough-testing instrument that<br />

simulates the expansion <strong>of</strong> bubbles in fermenting dough and as<br />

such the parameters derived from this test should give some<br />

indication as to how the dough will act during fermentation. The<br />

Resistance (AU)<br />

Resistance (AU)<br />

100<br />

50<br />

0<br />

100<br />

50<br />

0<br />

E. León et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 148–156 153<br />

Fig. 2. Farinograms <strong>of</strong> the doughs prepared from non-transformed control (Anza) and the flour from transgenic lines T580, T581 and T590, expressing the HMW-GS 1Ax1, 1Dx5 and<br />

1Dy10, respectively.<br />

most important parameters obtained from an alveogram are the<br />

tenacity (P) <strong>of</strong> the dough (expressing the resistance <strong>of</strong> the dough to<br />

deformation), the dough extensibility (L), the P/L ratio (which gives<br />

a general indication <strong>of</strong> the viscoeslastic properties <strong>of</strong> the dough)<br />

and the deformation energy (W), which is proportional to the<br />

energy required for deformation. This W-value is used to measure<br />

the dough strength by bakers. In general, low W values are related<br />

to weak flours and high values to strong flours. Transformation<br />

with the HMW subunit genes led to increased tenacity (P) in all<br />

three transgenic lines with respect to the control line (Table 2).<br />

However, the dough extensibility (L) decreased significantly from<br />

Anza T580 (1Ax1)<br />

0<br />

0 90 180 270 360<br />

0 90 180 270 360<br />

Time (s) Time (s)<br />

T581 (1Dx5) T590 (1Dy10)<br />

0<br />

0 90 180 270 360<br />

0 90 180 270 360<br />

Time (s) Time (s)<br />

Fig. 3. Mixograph curves <strong>of</strong> the doughs prepared from non-transformed control (Anza) and the flour from transgenic lines T580, T581 and T590, expressing the HMW-GS 1Ax1,<br />

1Dx5 and 1Dy10, respectively. Resistance in arbitrary units (AU) is plotted against time in minutes.<br />

Resistance (AU)<br />

Resistance (AU)<br />

100<br />

50<br />

100<br />

50


154<br />

Fig. 4. Alveograph curves <strong>of</strong> the doughs prepared from non-transformed control (Anza) and the flour from transgenic lines T580, T581 and T590, expressing the HMW-GS 1Ax1,<br />

1Dx5 and 1Dy10, respectively.<br />

94.2 in the control line to 61.0 in the line transformed with subunit<br />

1Dx5. The P/L ratio increased significantly from 0.4 in the control<br />

line to 0.7, 1.2, and 1.0 in that for lines transformed, respectively,<br />

with subunits 1Ax1, 1Dx5 and 1Dy10 (Table 2). Finally, the W value<br />

increased from 96.7 in the control line to 170.0, 180.3, and 258.3 in<br />

the lines with subunits 1Ax1, 1Dx5 and 1Dy10, respectively.<br />

4. Discussion<br />

Dough mixing and development are critical processes in bread<br />

production. Much research has therefore been conducted to<br />

determine the parameters that influence them and to facilitate the<br />

production <strong>of</strong> high quality bread that satisfies not only processing<br />

requirements but also customer expectations. The HMW-GS <strong>of</strong><br />

wheat are major determinants <strong>of</strong> the breadmaking quality <strong>of</strong> wheat<br />

and, in particular, the subunit pair 1Dx5 þ 1Dy10 and subunit 1Ax1<br />

have high quality scores (Payne, 1987; Shewry et al., 2003). The<br />

production <strong>of</strong> transgenic lines <strong>of</strong> wheat differing in their HMW-GS<br />

composition makes it possible to determine the effects <strong>of</strong> individual<br />

subunits on the rheological and technological properties <strong>of</strong><br />

wheat gluten. However, the behaviour <strong>of</strong> individual subunits may<br />

also be affected by the HMW-GS composition in which they are<br />

expressed as subunit interactions are crucial in forming the glutenin<br />

polymers. In this work we report the effects <strong>of</strong> the HMW-GS<br />

1Ax1, 1Dx5 and 1Dy10 on the breadmaking quality <strong>of</strong> the bread<br />

wheat cultivar Anza that contain the HMW-GS pairs 1Dx2 þ 1Dy12<br />

and 1Bx7* þ 1By8 and is null for the Glu-A1 locus. Wheat cultivars<br />

containing subunits 1Dx2 þ 1Dy12 generally have lower dough<br />

resistance than those containing subunits 1Dx5 þ 1Dy10 (Shewry<br />

et al., 2003). Therefore, this genotype is an excellent background to<br />

determine separately the effects <strong>of</strong> subunits 1Dx5 and 1Dy10 and<br />

subunit 1Ax1 on breadmaking quality. Although subunits 1Dx5 and<br />

1Dy10 have been separately introduced into the bread wheat<br />

Bobwhite (Blechl et al., 2007), this cultivar already contain both<br />

subunits which could influence the behaviour <strong>of</strong> the additional<br />

subunit protein encoded by the transgenes.<br />

This is therefore the first report on the characterization <strong>of</strong> wheat<br />

expressing subunit 1Dy10 in the absence <strong>of</strong> subunit 1Dx5. The<br />

E. León et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 148–156<br />

results show that subunit 1Dy10 had a greater effect on breadmaking<br />

quality than subunits 1Ax1 and 1Dx5. However, the<br />

expression <strong>of</strong> subunits 1Ax1 and 1Dx5 also led to increased<br />

breadmaking quality. This is not surprising for subunit 1Ax1 as this<br />

protein has been reported to have positive effects in a range <strong>of</strong><br />

wheat backgrounds (Barro et al., 2003; Darlington et al., 2003).<br />

However, the beneficial effect <strong>of</strong> subunit 1Dx5 was unexpected as<br />

previous transgenic wheat lines expressing this subunit had overstrong<br />

doughs unsuitable for breadmaking (Darlington et al., 2003;<br />

Rooke et al., 1999b).<br />

The expression levels <strong>of</strong> the transgenic 1Ax1, 1Dx5 and 1Dy10<br />

HMW-GS were comparable to those <strong>of</strong> the endogenous subunits.<br />

However, although lines T581 and T590 expressed similar levels <strong>of</strong><br />

subunits 1Dx5 and 1Dy10 (Table 1) they had different effects on<br />

other aspects <strong>of</strong> protein composition, including changes in the<br />

ratios <strong>of</strong> glutenin/gliadin, the HMW/LMW, the 1Dx2/1Dy12, x-type/<br />

y-type HMW-GS and the proportions <strong>of</strong> high molecular mass glutenin<br />

polymers (Table 1). In contrast, the lines transformed to<br />

express subunits 1Ax1 and 1Dx5 showed more similar changes in<br />

protein composition, with higher protein contents and decreased<br />

ratios <strong>of</strong> glutenin/gliadin and <strong>of</strong> 1Dx2/1Dy12. In addition, both<br />

transgenic lines showed similar increases in the ratio <strong>of</strong> x-type/ytype<br />

subunits compared to the control line (Table 1) while the line<br />

expressing subunit 1Dx5 also showed a significant increase in the<br />

HMW to LMW ratio. The increases in the HMW/LMW ratio and<br />

the changes in the glutenin/gliadin ratio as a consequence <strong>of</strong> the<br />

expression <strong>of</strong> transgenic HMW-GS have been reported previously<br />

(Alvarez et al., 2001; Blechl et al., 2007; Rakszegi et al., 2005). These<br />

changes also indicate that the new HMW-GS affect the structure <strong>of</strong><br />

the glutenin network with the x-type subunits (i.e. 1Ax1 and 1Dx5)<br />

having similar effects which are different to those <strong>of</strong> subunit 1Dy10.<br />

In the case <strong>of</strong> T590 (expressing 1Dy10) the composition is more<br />

comparable to that <strong>of</strong> the control line. For example, both the control<br />

line and line T590 had similar protein contents and ratios <strong>of</strong> glutenin/gliadin,<br />

HMW/LMW and 1Dx2/1Dy12 (Table 1). However, the<br />

x-type/y-type ratio was lower in line T590 than in the control line<br />

while the proportion <strong>of</strong> high molecular mass polymers was<br />

significantly increased.


The transgenic lines were analysed over 3 years using standard<br />

dough testing methods (Farinograph, Mixograph and Alveograph).<br />

This confirmed that the expression <strong>of</strong> all three subunits increased<br />

the dough strength (and hence breadmaking quality) <strong>of</strong> the cultivar<br />

Anza. Furthermore, a stepwise increase was observed depending on<br />

the HMW-GS expressed. Subunits 1Ax1 and 1Dx5 resulted in<br />

similar changes in the protein composition and structure, as<br />

described above, and this is reflected in their similar effects on<br />

breadmaking quality. Thus, the Farinograph showed increased<br />

dough resistance and stability and increased dough tolerance to<br />

mixing, the values being characteristic <strong>of</strong> medium grade flours<br />

(Table 2). Improvements in the rheological properties <strong>of</strong> the<br />

transgenic lines expressing subunits 1Ax1 and 1Dx5 were<br />

confirmed using the 10 g Mixograph (Fig. 3). Both transgenic lines<br />

had longer mixing times and lower values for resistance breakdown<br />

than the control line (Table 2). Subunit 1Ax1 has previously<br />

been reported to either increase dough resistance (Barro et al.,<br />

2003; Darlington et al., 2003) or to have little or no effect on dough<br />

resistance (Alvarez et al., 2001). In the former case, the genotype<br />

was derived from an Olympic Gabo cross and contained no 1Ax<br />

subunit and in the latter report the genotypes (BW and Federal)<br />

both expressed subunit 1Ax2*. However, the expression <strong>of</strong> subunit<br />

1Ax2* was suppressed in the transgenic line <strong>of</strong> Federal and the<br />

breadmaking quality was comparable to those <strong>of</strong> control lines<br />

expressing subunit 1Ax2*. It was therefore suggested that subunit<br />

1Ax1 had a similar effect to subunit 1Ax2*.<br />

Most previous studies <strong>of</strong> lines expressing the 1Dx5 transgene<br />

reported overstrong dough properties which were unsuitable for<br />

breadmaking. In contrast, in the present study the expression <strong>of</strong><br />

subunit 1Dx5 in the breadmaking variety Anza resulted in<br />

improved dough properties for breadmaking. Rooke et al. (1999b)<br />

suggested that although the transgenic subunits are incorporated<br />

into glutenin polymers, their organization within the polymers may<br />

differ from that in the corresponding control lines. An imbalance in<br />

the ratio <strong>of</strong> x-type/y-type subunits may therefore be responsible for<br />

the effects <strong>of</strong> expression <strong>of</strong> the 1Dx5 transgene without expression<br />

<strong>of</strong> subunit 1Dy10 (Alvarez et al., 2001). Our results indicate that<br />

subunits 1Dx5 and 1 Ax1 are similar in this respect, having similar<br />

effects on protein composition and dough properties.<br />

In contrast, transformation with subunit 1Dy10 led to stronger<br />

flours than those made from the control line and lines transformed<br />

with subunits 1Ax1 and 1Dx5. Farinograph parameters such as<br />

dough stability and peak time were increased by 9.2-fold and 2.4fold,<br />

respectively, in line T590 (expressing 1Dy10) with respect to<br />

the control line, while the mixing tolerance index and the degree <strong>of</strong><br />

s<strong>of</strong>tening decreased by 3.8-fold and 2.4-fold, respectively. The<br />

Mixograph and Alveograph parameters similarly showed that<br />

subunit 1Dy10 had a greater effect on breadmaking quality. Thus,<br />

the mixing time was increased by four-fold and resistance breakdown<br />

decreased by two-fold in line T590 compared with the<br />

control line. The W value was also increased by 2.7-fold in line T590<br />

compared to the control line.<br />

In this study, subunits 1Dx5 and 1Dy10 were compared in<br />

a cultivar <strong>of</strong> low breadmaking quality which provides an appropriate<br />

background to explore their effects on dough properties.<br />

Small scale mixing experiments incorporating purified 1Dx5 and<br />

1Dy10 subunits into dough (Uthayakumaran et al., 2000) showed<br />

that subunit 1Dx5 resulted in longer mixing times and lower<br />

resistance breakdown than subunit 1Dy10. In contrast, our results<br />

showed that expression <strong>of</strong> subunits 1Dx5 and 1Dy10 in transgenic<br />

wheat resulted in doughs with similar low resistance breakdown<br />

but that dough made from flour expressing the subunit 1Dy10<br />

transgene had a longer mixing time. This effect <strong>of</strong> the subunit<br />

1Dy10 transgene on dough strength is in agreement with the<br />

results reported by Blechl et al. (2007) for the expression <strong>of</strong> subunit<br />

1Dy10 in Bobwhite. This expression <strong>of</strong> subunit 1Dy10 was also<br />

E. León et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 148–156 155<br />

associated with a greater effect on the proportion <strong>of</strong> high molecular<br />

mass glutenin polymers than those observed with subunits 1Dx5.<br />

Acknowledgements<br />

The authors acknowledge funding by the Spanish CICYT (project<br />

AGL2007-65685-C02-01). Rothamsted Research receives grantaided<br />

support from the Biotechnology and Biological <strong>Science</strong>s<br />

Research Council (BBSRC) <strong>of</strong> the UK. Plasmid pK-Dy10A was<br />

provided by Dr. Ann E. Blechl. The technical assistance <strong>of</strong> Azahara<br />

Vida is also acknowledged.<br />

References<br />

AACC, 2000. Physical Dough Tests. Mixograph Method tenth ed Method 54-40A.<br />

Altpeter, F., Vasil, V., Srivastava, V., Vasil, I.K., 1996. Integration and expression <strong>of</strong> the<br />

high molecular weight glutenin subunit 1Ax1 gene into wheat. Nature<br />

Biotechnology 14, 1155–1159.<br />

Altpeter, F., Popelka, J.C., Wieser, H., 2004. Stable expression <strong>of</strong> 1Dx5 and 1Dy10<br />

high-molecular-weight glutenin subunit genes in transgenic rye drastically<br />

increases the polymeric glutelin fraction in rye flour. Plant Molecular Biology<br />

54, 783–792.<br />

Alvarez, M.L., Gómez, M., Carrillo, J.M., Vallejos, R.H., 2001. Analysis <strong>of</strong> dough<br />

functionality <strong>of</strong> flours from transgenic wheat. Molecular Breeding 8, 103–108.<br />

Anderson, O.D., Greene, F.C., Yip, R.E., Halford, N.G., Shewry, P.R., Malpica-<br />

Romero, J.M., 1989. Nucleotide sequences <strong>of</strong> the two high-molecular-weight<br />

glutenin genes from the D-genome <strong>of</strong> a hexaploid bread wheat, Triticum aestivum<br />

L. cv Cheyenne. Nucleic Acids Research 17, 461–462.<br />

Barcelo, P., Lazzeri, P.A., 1995. Transformation <strong>of</strong> cereals by microprojectile<br />

bombardment <strong>of</strong> immature inflorescence and scutellum tissues. In: Jones, H.<br />

(Ed.), Methods in Molecular Biology: Plant Gene Transfer and Expression<br />

Protocols. Humana Press, Totowa, NJ, pp. 113–123.<br />

Barro, F., Rooke, L., Bekes, F., Gras, P., Tatham, A.S., Fido, R., Lazzeri, P.A., Shewry, P.R.,<br />

Barcelo, P., 1997. Transformation <strong>of</strong> wheat with high molecular weight subunit<br />

genes results in improved functional properties. Nature Biotechnology 15,<br />

1295–1299.<br />

Barro, F., Cannell, M.E., Lazzeri, P.A., Barcelo, P., 1998. The influence <strong>of</strong> auxins on<br />

transformation <strong>of</strong> wheat and tritordeum and analysis <strong>of</strong> transgene integration<br />

patterns in transformants. Theoretical and Applied Genetics 97, 684–695.<br />

Barro, F., Barcelo, P., Lazzeri, P.A., Shewry, P.R., Martín, A., Ballesteros, J., 2002. Field<br />

evaluation and agronomic performance <strong>of</strong> transgenic wheat. Theoretical and<br />

Applied Genetics 105, 980–984.<br />

Barro, F., Barcelo, P., Lazzeri, P.A., Shewry, P.R., Ballesteros, J., Martin, A., 2003.<br />

Functional properties <strong>of</strong> flours from field grown transgenic wheat lines<br />

expressing the HMW subunit 1Ax1 and 1Dx5 genes. Molecular Breeding 12,<br />

223–229.<br />

Blechl, A.E., Anderson, O.D., 1996. Expression <strong>of</strong> a novel high-molecular-weight<br />

glutenin subunit gene in transgenic wheat. Nature Biotechnology 14, 875–879.<br />

Blechl, A., Lin, J., Nguyen, S., Chan, R., Anderson, O.D., Dupont, F.M., 2007. Transgenic<br />

wheats with elevated levels <strong>of</strong> Dx5 and/or Dy10 high-molecular-weight glutenin<br />

subunits yield doughs with increased mixing strength and tolerance.<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 45, 172–183.<br />

Bradford, M.M., 1976. A rapid and sensitive method for quantitation <strong>of</strong> microgram<br />

quantities <strong>of</strong> protein utilizing the principle <strong>of</strong> protein-dye-binding. Analytical<br />

Biochemistry 72, 248–254.<br />

Christensen, A.H., Quail, P.H., 1996. Ubiquitin promoter-based vectors for high-level<br />

expression <strong>of</strong> selectable and/or screenable marker genes in monocotyledonous<br />

plants. Transgenic Research 5, 213–218.<br />

Darlington, H., Fido, R., Tatham, A.S., Jones, H., Salmon, S.E., Shewry, P.R., 2003.<br />

Milling and baking properties <strong>of</strong> field grown wheat expressing HMW subunit<br />

transgenes. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 38, 301–306.<br />

D’Ovidio, R., Anderson, O.D., 1994. PCR analysis to distinguish between alleles <strong>of</strong><br />

a member <strong>of</strong> a multigene family correlated with wheat bread-making quality.<br />

Theoretical and Applied Genetics 88, 759–763.<br />

Gupta, R.B., MacRitchie, F., 1994. Allelic variation at glutenin subunit and gliadin<br />

loci, Glu-1, Glu-3 and Gli-1 <strong>of</strong> bread wheats: biochemical basis <strong>of</strong> the allelic<br />

effects on dough properties. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 19, 19–28.<br />

Halford, N.G., Field, J.M., Blair, H., Urwin, P., Moore, K., Robert, L., Thompson, R.,<br />

Flavell, R.B., Tatham, A.S., Shewry, P.R., 1992. Analysis <strong>of</strong> HMW glutenin subunits<br />

encoded by chromosome 1A <strong>of</strong> bread wheat (Triticum aestivum L.) indicates<br />

quantitative effects on grain quality. Theoretical and Applied Genetics 83,<br />

373–378.<br />

ICC, 1992a. Method for using <strong>of</strong> the Chopin-Alveograph. International Association<br />

for <strong>Cereal</strong> <strong>Science</strong> and Technology Method No. 121.<br />

ICC, 1992b. Method for using the Brabender Farinograph. International Association<br />

for <strong>Cereal</strong> <strong>Science</strong> and Technology Method No. 115/1.<br />

ICC, 1994. Determination <strong>of</strong> the sedimentation value (according to Zeleny) as an<br />

approximate measure <strong>of</strong> baking quality. International Association for <strong>Cereal</strong><br />

<strong>Science</strong> and Technology Method No. 116/1.<br />

ICC, 1995. Determination <strong>of</strong> the ‘‘Falling Number’’ according to Hagberg-Perten as<br />

a mesure <strong>of</strong> the degree <strong>of</strong> Alpha-Amylase activity in grain and flour. International<br />

Association for <strong>Cereal</strong> <strong>Science</strong> and Technology Method No. 107/1.


156<br />

Millar, S.J., 2003. The development <strong>of</strong> near infrared (NIR) spectroscopy calibrations<br />

for the prediction <strong>of</strong> wheat and flour quality. Project Report No. 310. HGCA, UK.<br />

101.<br />

Morel, M.-H., Dehlon, P., Autran, J.C., Leygue, J.P., Bar-L’Helgouac’h, C., 2000. Effects<br />

<strong>of</strong> temperature, sonication time and power settings on size distribution and<br />

extractability <strong>of</strong> total wheat flour proteins as determined by size-exclusion<br />

high-performance liquid chromatography. <strong>Cereal</strong> Chemistry 77, 685–691.<br />

Murashige, T., Skoog, F., 1962. A revised medium for rapid growth and bio assays<br />

with tobacco tissue cultures. Physiologia Plantarum 15, 473–497.<br />

Payne, P.I., 1987. Genetics <strong>of</strong> wheat storage proteins and the effect <strong>of</strong> allelic variation<br />

on bread-making quality. Annual Review <strong>of</strong> Plant Physiology 38, 141–153.<br />

Popineau, Y., Cornec, M., Lefebvre, J., Marchylo, B., 1994. Influence <strong>of</strong> high Mr glutenin<br />

subunits on glutenin polymers and rheological properties <strong>of</strong> glutens and<br />

gluten subfractions <strong>of</strong> near-isogenic lines <strong>of</strong> wheat Sicco. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong><br />

<strong>Science</strong> 19, 231–241.<br />

Popineau, Y., Deshayes, G., Lefebvre, J., Fido, R., Tatham, A.S., Shewry, P.R., 2001.<br />

Prolamin aggregation, gluten viscoelasticity, and mixing properties <strong>of</strong> transgenic<br />

wheat lines expressing 1Ax and 1Dx high molecular weight glutenin<br />

subunit transgenes. <strong>Journal</strong> <strong>of</strong> Agricultural and Food Chemistry 49, 395–401.<br />

Rakszegi, M., Bekes, F., Lang, L., Tamas, L., Shewry, P.R., Bedo, Z., 2005. Technological<br />

quality <strong>of</strong> transgenic wheat expressing an increased amount <strong>of</strong> a HMW glutenin<br />

subunit. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 42, 15–23.<br />

Rooke, L., Barro, F., Tatham, A.S., Fido, R., Steele, S., Bekes, F., Gras, P., Martin, A.,<br />

Lazzeri, P.A., Shewry, P.R., Barcelo, P., 1999a. Altered functional properties <strong>of</strong><br />

E. León et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 148–156<br />

tritordeum by transformation with HMW glutenin subunit genes. Theoretical<br />

and Applied Genetics 99, 851–858.<br />

Rooke, L., Bekes, F., Fido, R., Barro, F., Gras, P., Tatham, A.S., Barcelo, P., Lazzeri, P.,<br />

Shewry, P.R., 1999b. Overexpression <strong>of</strong> a gluten protein in transgenic wheat<br />

results in greatly increased dough strength. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 30,115–120.<br />

Sangtong, V., Moran, D., Chikwamba, R., Wang, K., Woodman-Clikeman, W.,<br />

Long, M., Lee, M., Scott, M., 2002. Expression and inheritance <strong>of</strong> the wheat Glu-<br />

1DX5 gene in transgenic maize. Theoretical and Applied Genetics 105, 937–945.<br />

Shewry, P.R., Tatham, A.S., Fido, R.J., 1995. Separation <strong>of</strong> plant proteins by electrophoresis.<br />

In: Jones, H. (Ed.), Methods in Molecular Biology: Plant Gene Transfer<br />

and Expression Protocols. Humana Press, Totowa, NJ, pp. 399–422.<br />

Shewry, P.R., Halford, N.G., Tatham, A.S., Popineau, Y., Lafiandra, D., Belton, P.S.,<br />

2003. The high molecular weight subunits <strong>of</strong> wheat glutenin and their role in<br />

determining wheat processing properties. Advances in Food and Nutrition<br />

Research 45, 219–302.<br />

Shuey, W.C., 1984. Interpretation <strong>of</strong> the Farinogram. In: D’Appolonia, B.L.,<br />

Kunerth, W.H. (Eds.), The Farinograph Handbook. American Association <strong>of</strong><br />

<strong>Cereal</strong> Chemistry, St. Paul, MN, pp. 31–32.<br />

Stacey, J., Isaac, P.G., 1994. Isolation <strong>of</strong> DNA from plants. In: Isaac, P.G. (Ed.), Methods<br />

in Molecular Biology: Protocols for nucleic acid analysis by nonradiactive<br />

probes. Humana Press, Totowa, NJ, pp. 9–15.<br />

Uthayakumaran, S., Stoddard, F.L., Gras, P.W., Bekes, F., 2000. Effects <strong>of</strong> incorporated<br />

glutenins on functional properties <strong>of</strong> wheat dough. <strong>Cereal</strong> Chemistry 77,<br />

737–743.


Wheat glutenin proteins assemble into a nanostructure with unusual<br />

structural features<br />

Sarah H. Mackintosh a , Susie J. Meade a , Jackie P. Healy b , Kevin H. Sutton a , Nigel G. Larsen a ,<br />

Adam M. Squires c,1 , Juliet A. Gerrard b, *<br />

a<br />

Food and Biomaterials Innovation, New Zealand Institute for Crop & Food Research Ltd., Private Bag 4704, Christchurch, New Zealand<br />

b<br />

School <strong>of</strong> Biological <strong>Science</strong>s, University <strong>of</strong> Canterbury, 20 Kirkwood Ave, Private Bag 4800, Christchurch 8020, New Zealand<br />

c<br />

Cavendish Laboratory, Cambridge University, Cambridge, UK<br />

article info<br />

Article history:<br />

Received 11 December 2007<br />

Received in revised form 19 August 2008<br />

Accepted 22 August 2008<br />

Keywords:<br />

Glutenins<br />

X-ray fibre diffraction<br />

Wheat protein<br />

Protein nanostructure<br />

1. Introduction<br />

abstract<br />

Proteins have an innate ability to adopt specific three dimensional<br />

forms with a plethora <strong>of</strong> functions. This affords the potential<br />

for design <strong>of</strong> nanoscale materials via the manufacture <strong>of</strong> selfassembled<br />

protein nanostructures, with properties that are not<br />

easily duplicated with traditional organic molecules and other<br />

polymers (Sanford and Kumar, 2005). Significantly, protein nanomaterials<br />

incorporate surface chemical functionality via the amino<br />

acid side chains, therefore providing additional possibilities for the<br />

construction <strong>of</strong> advanced nanomaterials, with a combination <strong>of</strong><br />

structural and (bio)chemical function. The enormous promise for<br />

the use <strong>of</strong> peptides and protein as smart materials has attracted<br />

much recent attention (Shen et al., 2006; Waterhouse and Gerrard,<br />

2004).<br />

Amyloid fibrils are highly ordered nan<strong>of</strong>ibres <strong>of</strong> insoluble<br />

protein, rich in b-strand content (Dobson, 1999). The mature fibril<br />

comprises a number <strong>of</strong> prot<strong>of</strong>ilaments (sub-structures) that typically<br />

twist together to form a nanostructure with a diameter <strong>of</strong><br />

* Corresponding author. Tel.: þ64 3 3667001; fax: þ64 3 3642590.<br />

E-mail address: juliet.gerrard@canterbury.ac.nz (J.A. Gerrard).<br />

1<br />

Present address: School <strong>of</strong> Chemistry, University <strong>of</strong> Reading, Whiteknights,<br />

Reading RG6 6AD, UK.<br />

0733-5210/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.jcs.2008.08.003<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 157–162<br />

Contents lists available at <strong>Science</strong>Direct<br />

<strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong><br />

journal homepage: www.elsevier.com/locate/jcs<br />

The proteins <strong>of</strong> wheat have a known propensity to aggregate into a variety <strong>of</strong> forms. We report here<br />

a novel nanostructure from wheat proteins, derived from a crude extract <strong>of</strong> high molecular weight<br />

glutenins. The structure was characterised by a significant thi<strong>of</strong>lavin T (ThT) fluorescence and a fibrillar<br />

morphology by transmission electron microscopy (TEM). The ThT fluorescence and TEM data are<br />

suggestive <strong>of</strong> an amyloid structure, but the X-ray fibre diffraction data show a reflection pattern (4.02,<br />

4.2–4.3, 4.6, 12.9, 19.3 and 38.7 Å) inconsistent with both the classic amyloid form and the previously<br />

described b-helix structure. The 4.6 Å reflection is consistent with that predicted for the amyloid inter-bstrand,<br />

and the absence <strong>of</strong> the inter-b-sheet distance at z10–11 Å is not unprecedented in amyloid-like<br />

structures. However, our observed X-ray reflection pattern has not been previously reported and suggests<br />

a novel wheat glutenin nanostructure.<br />

Ó 2008 Elsevier Ltd. All rights reserved.<br />

approximately 5–15 nm, and a rope-like appearance when imaged<br />

at high magnifications (Serpell, 2000). Whilst there are a number <strong>of</strong><br />

competing structural models (Makin and Serpell, 2005), amyloid<br />

fibrils are defined by observations <strong>of</strong> characteristic electron<br />

microscopy, specific chemical staining and X-ray fibre diffraction.<br />

The specific arrangement <strong>of</strong> the b-strands within the mature fibril<br />

is defined as a cross-b pattern and leads to characteristic reflections<br />

in the X-ray fibre diffraction pattern <strong>of</strong> these fibrils, which is<br />

increasingly accepted as a diagnostic indicator <strong>of</strong> the presence <strong>of</strong><br />

amyloid fibrils (Makin and Serpell, 2004). As well as their well<br />

recognised importance in the diseases <strong>of</strong> protein misfolding<br />

(Dobson, 2001), the use <strong>of</strong> amyloid fibrils as nanomaterials is<br />

attracting increasing attention (Gras, 2007; Waterhouse and Gerrard,<br />

2004). Other amyloid-like nanostructure models have been<br />

proposed, including b-helix and spiral structures (DeMarco et al.,<br />

2006; Makin and Serpell, 2005).<br />

Since polyglutamine proteins are associated with amyloid<br />

formation in the body (Chen et al., 2002), our attention was drawn<br />

to the proteins <strong>of</strong> wheat as a potential raw material for amyloid<br />

fibril manufacture. Wheat proteins are known to have a high<br />

glutamine content (Sugiyama et al., 1985), including the high<br />

molecular weight glutenin subunits (HMW-GS) associated with<br />

flour quality which have a (glutamic acid þ glutamine content) in<br />

the order <strong>of</strong> 30% (Anjum et al., 2008; Kipp et al., 1996). A BLAST<br />

search (Altschul et al., 1997) on the wheat (Triticum aestivum)


158<br />

genome returned matches for polyglutamine sequences up to<br />

eleven amino acids in length. Additionally, gluten proteins are well<br />

known to form aggregated structures; indeed it is this ability that<br />

facilitates their use in the production <strong>of</strong> a variety <strong>of</strong> baked goods<br />

and pasta products (Shewry et al., 2002). They have also been<br />

explored for biomaterial use (Guilbert et al., 2001).<br />

The gluten network itself is extremely complex, with the<br />

molecular weight <strong>of</strong> these HMW-GS components ranging from<br />

approx. 67–88 kDa (Shewry et al., 2002). The units themselves<br />

appear to be comprised <strong>of</strong> three distinct domains: a central<br />

repetitive domain <strong>of</strong> variable length (440–680 residues), and much<br />

shorter flanking non-repetitive terminal domains (N – 81–104 and<br />

C – 42 residues) (Tatham, 1995). Structural information to date has<br />

been based on predictive modeling, microscope imaging, indirect<br />

methods such as FTIR and CD, and limited SAXS analysis (Lindsay<br />

and Skerritt, 1999). Indications are that the secondary structure <strong>of</strong><br />

the proteins in the matrix includes a high proportion <strong>of</strong> b-turn<br />

structure for the central domain, while the terminal domains<br />

comprise a-helices (Tatham, 1995 and references therein). The low<br />

molecular weight group <strong>of</strong> glutenin proteins (LMW-GS) is less well<br />

characterised. It is known they have a glutamine rich N-terminal<br />

domain (again predicted to form b-reverse turns), and are rich in<br />

cysteine residues located primarily in the C-terminal end, but also<br />

in the central domain (predicted to be a-helical). This group <strong>of</strong><br />

proteins predominantly exists in monomeric form, and contains<br />

numerous internal disulfide bonds (Lew et al., 1992). Subgroups<br />

within the LMW-GS are also known to possess similar secondary<br />

structure to the gliadin protein group, which has been documented<br />

to aggregate under appropriate conditions (Kasarda et al., 1967).<br />

The gliadin fibril aggregates were observed to be 80 Å thick, up to<br />

several thousand Ångstroms long and dissociated to globular<br />

protein subunits at very low ionic strength and low pH.<br />

Despite considerable research into their aggregation properties,<br />

the tendency for gluten proteins to form amyloid-like structures<br />

has not been previously explored. We report herein an exploration<br />

<strong>of</strong> whether or not gluten proteins can form amyloid structures and<br />

the characterisation <strong>of</strong> a novel nanostructure derived from wheat<br />

glutenins, characterised by spectroscopic binding assays, transmission<br />

electron microscopy and X-ray fibre diffraction.<br />

2. Experimental<br />

2.1. Materials<br />

Unless otherwise stated, all chemicals and reagents were<br />

purchased from Sigma Chemical Company Ltd. (St Louis, U.S.A.),<br />

Aldrich Chemicals (Milwaukee, U.S.A.) or BDH Laboratory Supplies<br />

(Poole. U.K.), and were <strong>of</strong> analytical grade. The wheat flour derived<br />

protein was extracted from a Domino cultivar, sourced by Crop and<br />

Food Research Ltd. and milled in-house.<br />

2.2. Extraction protocol<br />

Wheat proteins were crudely fractionated using an extraction<br />

method based on the methods <strong>of</strong> Gerrard et al. (2001), which were<br />

derived from the protocols <strong>of</strong> Osborne (1907). The albumins and<br />

globulins were extracted in dilute saline (2% w/v). The mixture was<br />

mixed by pulse vortex every 5 min for 30 min to allow extraction <strong>of</strong><br />

the appropriate protein group, then centrifuged at 10 000 g for<br />

5 min. The resulting pellet was resuspended in a concentrated<br />

ethanol solution (70% v/v) and mixed by pulse vortex every 5 min<br />

for 30 min to resuspend the pellet and to separate the soluble and<br />

insoluble components. After centrifugation (10 000 g, 5 min) the<br />

gliadin extract was decanted. The pellet was resuspended in a SDSphosphate<br />

buffer (0.05% w/v SDS, 0.05 M phosphate, pH 6.9) with<br />

mixing by pulse vortex (every 5 min for 30 min), to extract the<br />

S.H. Mackintosh et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 157–162<br />

LMW-GS and HMW-GS. After centrifugation (10 000 g, 5 min) the<br />

pellet was again resuspended in SDS-phosphate buffer (0.05% w/v<br />

SDS, 0.05 M phosphate, pH 6.9) with mixing by pulse vortex (every<br />

5 min for 30 min). This final solution <strong>of</strong> glutenins was sonicated at<br />

30 W for 15 s using a Branston model 250 sonic disruptor, to<br />

solubilise any protein components still held within the pellet.<br />

2.3. Protein concentration<br />

Protein concentrations were determined using a modified<br />

version <strong>of</strong> the Bradford’s (1976) method. A standard curve was<br />

determined using bovine serum albumin.<br />

2.4. PAGE<br />

Polyacrylamide gel electrophoresis was routinely run using an<br />

Invitrogen XCell SureLock Mini-Cell powered by a Bio-Rad 300<br />

power pack, and precast 4–12% BisTris gels from Invitrogen. The gel<br />

system was run in a 1 3-(N-morpholino) propanesulfonic acid<br />

(MOPS) SDS running buffer, over approx. 90 min at 150 V, and<br />

stained with Coomassie blue. The marker was wide molecular<br />

range SigmaMarker (product no. S8445 from Sigma).<br />

2.5. Screening protocols<br />

In preliminary screens, all four fractions <strong>of</strong> wheat protein were<br />

incubated in an array <strong>of</strong> solution conditions and periodically tested<br />

using thi<strong>of</strong>lavin T (ThT, see below) for the presence <strong>of</strong> amyloid<br />

fibrils. In particular, additions <strong>of</strong> the following compounds to the<br />

extraction buffers were tested: acid (HCl or H2SO4); sodium chloride<br />

(0–0.2 M); denaturing compounds (2–6 M urea, 0–30% (v/v)<br />

TFE, 1 M mercaptoethanol or 0.5% (w/v) SDS); and insulin fibrils (2%<br />

v/v 5.8 mg/ml solution <strong>of</strong> preformed insulin fibrils/wheat protein<br />

solution). The extracted glutenin protein fractions proved to be the<br />

most interesting and were explored further. No further experiments<br />

were carried out on the albumin and globulin or gliadin<br />

fractions. Glutenin fractions were then extracted, and the proteins<br />

lyophilised and resuspended at a concentration <strong>of</strong> 10 mg/ml (w/v)<br />

in a variety <strong>of</strong> modified buffers. A screen <strong>of</strong> conditions was trialled<br />

in order to optimise conditions for fibril formation.<br />

The solution conditions <strong>of</strong> the glutenin extracts were varied in<br />

order to optimise the conditions for fibril formation, as judged by<br />

the ThT assay. A range <strong>of</strong> conditions and variables were investigated<br />

both individually and in combination and they are summarized in<br />

Table 1. The solutions were then incubated at temperatures <strong>of</strong><br />

either 25 Cor37 C and monitored for periods <strong>of</strong> up to 105 days<br />

and compared to control samples kept frozen at 20 C. As<br />

required, trypsin solution was prepared with 1 mM HCl and mercaptoethanol<br />

(10 ml/ml) and the glutenin extracts were treated at<br />

a ratio <strong>of</strong> 1:20 (v/v) prior to incubation. Insulin fibrils were formed<br />

by incubation <strong>of</strong> insulin from bovine pancreas (Sigma product no.<br />

Table 1<br />

Range <strong>of</strong> conditions for glutenin incubations.<br />

Variable Range<br />

Glutenin concentration 2–10 mg/ml<br />

Temperature 20 to 37 C<br />

Time 0–105 days<br />

pH 2–7<br />

Urea 0–6 M<br />

TFE 0–30%<br />

Sodium chloride 0–0.2 M<br />

Sonication þ/<br />

Trypsin pre-treatment þ/<br />

Insulin seeding 2% v/v (5.8 mg/ml solution <strong>of</strong> preformed<br />

insulin fibrils/wheat protein solution)


I5500) at pH 2 (HCl in distilled water) at 50 C for 48 h according to<br />

the method <strong>of</strong> Nilsson and Dobson (2003). Fibril formation was<br />

confirmed by TEM and ThT assay. These seed fibrils were added to<br />

selected glutenin samples prior to incubation. In addition to<br />

changing the solution conditions, the impact <strong>of</strong> shaking and/or<br />

periodic sonication during the incubation was explored. Aliquots<br />

removed from the incubating reactions were stored at 20 C until<br />

required for analysis by ThT and TEM (below).<br />

2.6. ThT assay<br />

The ThT method used was based on the protocols <strong>of</strong> LeVine<br />

(1999). The protein samples were diluted to a concentration <strong>of</strong> 10–<br />

20 mg/ml in Tris buffer (50 mM Tris, 100 mM NaCl, pH 7.5). To this<br />

solution, the ThT dye was added to a final concentration <strong>of</strong> 5 mM.<br />

The solution was mixed and left to stand at room temperature for<br />

3 min to allow binding between the dye and protein to equilibrate,<br />

before the emission spectra were recorded. The spectra <strong>of</strong> buffer<br />

only, buffer with ThT and buffer with protein were used as the<br />

experimental controls. Fluorescence spectroscopy was performed<br />

using a Cary Eclipse Varian fluorescence spectrophotometer interfaced<br />

with Cary Eclipse operating s<strong>of</strong>tware (V 2.0), using 1 cm path<br />

length quartz cuvettes and the excitation was set to 450 10 nm<br />

and emission set to 470–520 nm. The relative fluorescence is<br />

graphed. This reading is the incubated sample fluorescence with<br />

the control (non-incubated protein with ThT) fluorescence subtracted.<br />

The error bars are standard errors for a triplicate<br />

measurement.<br />

2.7. Transmission electron microscopy<br />

Samples were analysed by TEM, based on the protocols<br />

described by Brenner and Horne (1959), to confirm the presence <strong>of</strong><br />

fibrils. The preparation involved placing small aliquots (5 ml) <strong>of</strong><br />

protein fibril samples on formvar-coated copper grids. After 90–<br />

120 s, the excess solution was drained with filter paper, the grids<br />

washed with double distilled water and the samples negatively<br />

stained with 1% (w/v) uranyl acetate for 45–60 s, before the excess<br />

solution was again drained with filter paper and the grids left to<br />

dry. For each sample, four images were recorded and a representative<br />

image was selected from amongst these.<br />

2.8. X-ray diffraction<br />

The glutenin samples were prepared for X-ray fibre diffraction<br />

by placing a 10 ml droplet <strong>of</strong> the incubated protein solution between<br />

two wax-filled glass capillary ends (Serpell et al., 1999). The capillaries<br />

were manually separated as the sample dried, to promote<br />

fibril alignment. The resulting fibre was then analysed using an Xray<br />

beam produced by a Rigaku Cu-K rotating-anode source with<br />

the data collected using an R-AXIS IV image-plate X-ray detector.<br />

The data were analysed using Fit2D (V12.043) s<strong>of</strong>tware. Images<br />

were recorded in the Biochemistry Department, University <strong>of</strong><br />

Cambridge.<br />

3. Results and discussion<br />

3.1. Screening for fibril-forming conditions<br />

All four wheat protein fractions – the albumins and globulins,<br />

the gliadins, the SDS-soluble glutenins and the SDS-insoluble<br />

glutenins – were put through an initial screen including a range <strong>of</strong><br />

solution conditions that have been shown to induce fibril<br />

formation in other systems (Kim et al., 2004). The initial screen<br />

showed promising results for the SDS-soluble glutenins and the<br />

SDS-insoluble glutenins only (data not shown) and no further<br />

S.H. Mackintosh et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 157–162 159<br />

work was carried out on the albumin and globulin and gliadin<br />

fractions.<br />

The solution conditions <strong>of</strong> the two glutenin extracts were then<br />

varied in a systematic, combinatorial fashion, as described in<br />

Section 2. These included acidic conditions and addition <strong>of</strong><br />

a denaturant. Also explored was the effect <strong>of</strong> shaking the solution<br />

during the incubation, sonicating the solution periodicially during<br />

the incubation, and temperature. At various time points throughout<br />

the incubation, aliquots were extracted from each <strong>of</strong> the incubated<br />

protein mixture samples and analysed by ThT assay. Evidence for<br />

increased b-sheet, consistent with fibril-like structures, was found<br />

under some conditions, particularly in acidic conditions in the<br />

presence <strong>of</strong> urea, consistent with the requirement for partial<br />

unfolding <strong>of</strong> the proteins prior to fibril formation (Chiti and Dobson,<br />

2006). The most successful conditions involved a pH <strong>of</strong> 6–7 in the<br />

presence <strong>of</strong> either 2 M urea or 30% (v/v) trifluoroethanol. The SDSinsoluble<br />

glutenins gave more encouraging results under these<br />

conditions than the SDS-soluble glutenins. The SDS-soluble glutenin<br />

conditions that gave significant increases in ThT fluorescence<br />

are shown in Table 2.<br />

An example <strong>of</strong> the results is shown in Fig. 1. A significant lag<br />

period was observed for all treatments in which positive results<br />

were obtained, characteristic <strong>of</strong> a period <strong>of</strong> equilibrium where<br />

fibril growth is at a minimum and interactions between the glutenin<br />

subunits are limited and reversible, prior to the growth<br />

phase (Nilsson and Dobson, 2003). The lag phase observed is<br />

significantly longer than any lag phase reported for amyloid<br />

formation in vitro in the literature to date. This increase in ThT<br />

fluorescence is consistent with a structure with increased b-sheet<br />

content.<br />

3.2. Characterising the aggregated structures: SDS-PAGE<br />

In an attempt to establish which <strong>of</strong> the many proteins in the<br />

glutenin fractions were forming the aggregated structure, the<br />

incubated solutions were analysed by SDS-PAGE. The solution was<br />

spun down to remove any aggregated protein and the supernatant<br />

solution analysed as a function <strong>of</strong> time, for many <strong>of</strong> the conditions<br />

tested. The electrophoretic pr<strong>of</strong>ile <strong>of</strong> this fraction is well studied<br />

(Gerrard et al., 2001) but typically, no specific proteins were found<br />

to be selectively lost from solution; rather, all proteins were lost at<br />

a similar rate, although larger proteins were removed from the<br />

supernatant somewhat faster than smaller ones. A typical result is<br />

found in Fig. 2. These results suggest that a mixture <strong>of</strong> proteins is<br />

incorporated into the aggregate, or that different proteins form<br />

different structures that are incorporated into the assembly, and/or<br />

that those proteins not incorporated were subject to hydrolysis<br />

and/or proteolysis during the long timeframe <strong>of</strong> the incubation.<br />

Extensive further work would be required to distinguish between<br />

these possibilities, and due to the long timeframe and low yield <strong>of</strong><br />

the process, these experiments were not continued further.<br />

Table 2<br />

Summary <strong>of</strong> SDS-insoluble glutenin incubation conditions that resulted in significant<br />

increase ThT fluorescence after 105 days. Frozen controls showed no increase in<br />

ThT fluorescence.<br />

Treatment SDS-insoluble<br />

Temp. ( C) Conc. (mg/ml) Conditions<br />

glutenin<br />

25 10 pH 6–7 þThT<br />

25 10 pH 6–7, 0.5 M H2SO4 þThT<br />

37 10 pH 6–7 þThT<br />

37 10 pH 6–7, 0.5 M HCl þThT<br />

37 10 pH 6–7, 2 M urea þThT<br />

37 10 pH 6–7, 30% (v/v) TFE þThT


160<br />

Relative Fluorescence<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 10 20 30 40 50 60 70<br />

Time (Days)<br />

Fig. 1. Formation <strong>of</strong> amyloid-like structures in the SDS-soluble glutenins <strong>of</strong> wheat<br />

incubated at 25 C in the presence <strong>of</strong> 2 M urea.<br />

3.3. Characterising the aggregated structures: TEM and X-ray fibre<br />

diffraction<br />

ThT is commonly used to monitor amyloid fibril formation, as it<br />

forms a fluorescent species when bound to protein material with<br />

a high content <strong>of</strong> b-structure (LeVine, 1999). A positive ThT result is<br />

thus strongly suggestive <strong>of</strong> the presence <strong>of</strong> b-sheet structure such<br />

as an amyloid fibril, but other methods are required to corroborate<br />

this observation.<br />

Samples which showed a positive ThT test were observed under<br />

TEM. A variety <strong>of</strong> morphologies were apparent in the sample, as<br />

shown in Fig. 3. Fibril structures were typically unbranched and <strong>of</strong><br />

indeterminate length (at least several microns), with widths <strong>of</strong><br />

between 5 and 15 nm, <strong>of</strong> similar dimensions to amyloid fibrils<br />

described in the literature (Serpell, 2000). Some displayed the<br />

Fig. 2. SDS-PAGE gel <strong>of</strong> glutenins, in 2 M urea, incubated for various times up to 70<br />

days. 1. Marker (205, 116, 97.4, 84, 66, 55, 45, 36, 29, 24, 20.1, 14.2, and 6.5 kDa); 2. 7<br />

days; 3. 14 days; 4. 21 days; 5. 28 days; 6. 42 days; 7. 56 days; and 8. 70 days.<br />

S.H. Mackintosh et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 157–162<br />

classical twisted rope appearance (Fig. 3A, B). Consistent with the<br />

low yield <strong>of</strong> ThT positive material (Fig. 1), the nan<strong>of</strong>ibrillar structures<br />

were found to be contaminated with amorphously aggregated<br />

protein. The sticky nature <strong>of</strong> the amorphous aggregates made it<br />

impracticable to separate and purify the nanostructures. We estimate<br />

that the yield <strong>of</strong> these nanostructures was less than 1%.<br />

X-ray fibre diffraction is increasingly used as a diagnostic indicator<br />

for amyloid fibril formation (Makin et al., 2005). The diffraction<br />

pattern produced when an aligned fibril sample is exposed to<br />

an X-ray beam can be used to detect the cross-b pattern indicative<br />

<strong>of</strong> b-sheet structure, <strong>of</strong> indeterminate length, with the b-strands<br />

aligned perpendicular to the fibril axis and the b-sheets running<br />

parallel to it. For amyloid fibrils, the interstrand separation classically<br />

appears as a 4.7 Å meridonal reflection (i.e. in a parallel<br />

direction to the mounted fibril fibre) and the intersheet separation<br />

as an equatorial reflection at approx. 10–11 Å. However, polyglutamine<br />

peptide fibre patterns have been reported with a reflection<br />

at 4.7 Å, but the 10 Å reflection is <strong>of</strong>ten faint or not observed (Perutz<br />

et al., 2002).<br />

The diffraction pattern obtained from the wheat glutenin<br />

protein fibrillar structure is shown in Fig. 4. It is clear that the<br />

sample is anisotropic, but the reflections show distinct differences<br />

from the classic amyloid form. Equatorially aligned bands at 12.9,<br />

19.3 and 38.7 Å, and meridionally aligned bands at 4.6, 4.2–4.3, and<br />

4.02 Å were observed. A number <strong>of</strong> fainter bands are also visible.<br />

The faint band at 4.6 Å is suggestive <strong>of</strong> an aligned structure with an<br />

intra-strand distance corresponding to that <strong>of</strong> a b-sheet. The<br />

absence <strong>of</strong> a diffraction ring at approximately 10–11 Å is also<br />

consistent with amyloid fibrils formed from high glutamine<br />

protein. In classic amyloid fibrils (i.e. those linked with disease) this<br />

distance corresponds to the interstrand spacing, and forms an<br />

integral part <strong>of</strong> definition for cross-b-structure.<br />

Thus, whilst it is tempting to speculate that the observed<br />

nanostructures are amyloid-like, this model does not fit all the data.<br />

The other wide-angle meridional reflections and the inner equatorial<br />

reflections (12.9, 19.3, and 38.7 Å) have not, to our knowledge,<br />

been reported in any fibrillar system; the equatorial reflections are<br />

more consistent with a lateral assembly (i.e. perpendicular to the<br />

fibril axis), with a spacing <strong>of</strong> 38.7 Å. In light <strong>of</strong> this, the 4.6 Å<br />

reflection may be due to another feature within the fibrils <strong>of</strong> a nonamyloid<br />

structure, or perhaps indicates that the sample may<br />

contain a small amount <strong>of</strong> amyloid-like fibrils mixed with nonamyloid<br />

fibrils. The glutenin nanostructure thus appears to contain<br />

novel features distinct from both the previously reported classical<br />

amyloid fibril structure and the hypothesised b-sheet helix structure<br />

(DeMarco et al., 2006).<br />

3.4. Attempts to improve the rate and yield <strong>of</strong> fibril formation<br />

The novel wheat protein structure was formed at a low rate and<br />

the overall yield was too small to be considered as a viable process<br />

for commercial manufacturing. All treatments consistently resulted<br />

in extremely low yields. In a subsequent screen, therefore, two<br />

further methods were explored in order to increase the rate <strong>of</strong><br />

formation <strong>of</strong> fibrillar structures, and boost the yield: seeding the<br />

solution with preformed fibrils, and pre-treatment <strong>of</strong> the protein<br />

with trypsin.<br />

An analysis <strong>of</strong> the sequences <strong>of</strong> glutenin proteins using the<br />

aggregating predicting programme Zygregattor (Pawar et al., 2005)<br />

revealed that the regions <strong>of</strong> local sequence most likely to hinder<br />

aggregation <strong>of</strong> the glutenin proteins were those containing lysine<br />

or arginine residues. Since trypsin cuts proteins adjacent to both<br />

these residues in a selective manner (Wang and Carpenter, 1967), it<br />

was hoped that pre-digestion with trypsin might improve the rate<br />

and yield <strong>of</strong> formation <strong>of</strong> fibrils from the wheat protein incubations.<br />

Unfortunately, despite confirmation by SDS-PAGE that the trypsin


Fig. 3. Wheat glutenin protein structure images derived from transmission electron microscopy analysis. Images were taken from SDS-insoluble glutenin proteins incubated at<br />

25 C (2 mg/ml, pH 7.4) for between 54 and 105 days. Scale bars represent 200 nm.<br />

had indeed cut the glutenins into the desired smaller fragments, no<br />

subsequent increase in rate <strong>of</strong> formation <strong>of</strong> ThT positive species, or<br />

yield was observed.<br />

Since amyloid formation is a seeded process (Ban et al., 2006),<br />

attempts were made to enhance the rate <strong>of</strong> production by<br />

Fig. 4. Diffraction patterns obtained from dried SDS-insoluble glutenin proteins at<br />

2 mg/ml, pH 7.4 for 105 days.<br />

S.H. Mackintosh et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 157–162 161<br />

sonication and addition <strong>of</strong> preformed fibril seeds. These processes<br />

have been reported to accelerate fibril formation (Stathopulos et al.,<br />

2004). The sonication provides strong agitation that can fragment<br />

the fibrils, thus potentially providing more ‘sticky ends’ for fibril<br />

elongation. Samples were sonicated periodically during the incubation<br />

<strong>of</strong> wheat protein in a range <strong>of</strong> different buffer conditions.<br />

Insulin seeds were formed by sonication <strong>of</strong> insulin fibrils and these<br />

insulin seeds were added to the glutenin protein extracts prior to<br />

incubation. Neither method was successful. The lack <strong>of</strong> sequence<br />

identity between the insulin seeds and the glutenin structure may<br />

account for the failure <strong>of</strong> the latter method.<br />

3.5. Concluding comments<br />

Although our original aim <strong>of</strong> manufacturing large quantities <strong>of</strong><br />

protein nanostructures from a readily available heterogeneous<br />

protein substrate was not met, a novel aggregated structure has<br />

been assembled and characterised from the wheat glutenin<br />

proteins, without prior purification. The structure displays some<br />

characteristic features <strong>of</strong> amyloid fibrils, including a detectable<br />

increase in ThT fluorescence assays and a fibrillar morphology<br />

when viewed under an electron microscope. While the aggregated<br />

material does not display the classic cross-b pattern commonly<br />

observed for amyloid material, bands do exist within the pattern<br />

at z4.8 Å (corresponding to inter-b-strand distance). The absence<br />

<strong>of</strong> an inter-b-sheet distance at z10–11 Å is not unprecedented<br />

and a number <strong>of</strong> recent articles describing amyloid-like<br />

structures, that may represent a distinct but related form <strong>of</strong><br />

amyloid-like nanostructures (Kwan et al., 2006). However, our<br />

observed X-ray reflection pattern has not been previously reported<br />

and appears to represent a new nanostructured form <strong>of</strong> wheat<br />

proteins.<br />

Rauscher et al. (2006) have discussed the role <strong>of</strong> proline and<br />

glycine in control <strong>of</strong> protein self-organization into elastomeric or


162<br />

amyloid-like fibrils. Given that HMW-GS contain repeating<br />

sequences <strong>of</strong> proline, glycine and glutamine (Tatham and Shewry,<br />

2002), it may well be that this novel structure is dependent on the<br />

interaction <strong>of</strong> these three amino acids and could shed light on the<br />

relationship between sequence, structure and elastomeric qualities<br />

<strong>of</strong> these intriguing proteins.<br />

Acknowledgements<br />

The authors thank Harald Dobberstein and Manfred Ingerfeld<br />

for assistance with TEM, Grant Pearce and Amol Pawar for assistance<br />

with the Zygreggator algorithm, and Drs Cait MacPhee, Kate<br />

McGrath, Sarah Meehan, and Margie Sunde for useful discussions.<br />

Funding was provided by the New Zealand Foundation for<br />

Research, <strong>Science</strong> and Technology via the New Economy Research<br />

Fund contracts C02X0204 (Advanced Biological Materials) and<br />

C02X0404 (Amyloid Fibrils).<br />

References<br />

Altschul, S.F., Madden, T.L., Schäffer, A.A., Zhang, J., Zhang, Z., Miller, W., Lipman, D.J.,<br />

1997. Gapped BLAST and PSI-BLAST: a new generation <strong>of</strong> protein database<br />

search programs. Nucleic Acids Research 25, 3389–3402.<br />

Anjum, F.M., Khan, M.R., Din, A., Saaed, M., Pasha, I., Arshad, M.U., 2008. Wheat<br />

gluten: high molecular weight glutenin subunits – structure, genetics, and<br />

relation to dough elasticity. <strong>Journal</strong> <strong>of</strong> Food <strong>Science</strong> 72, R56–R63.<br />

Ban, T., Yamaguchi, K., Goto, Y., 2006. Direct observation <strong>of</strong> amyloid fibril growth,<br />

propagation, and adaptation. Accounts <strong>of</strong> Chemical Research 39, 663–670.<br />

Bradford, M.A., 1976. A rapid and sensitive method for the quantitation <strong>of</strong> microgram<br />

quantities <strong>of</strong> protein utilising the principle <strong>of</strong> protein–dye binding.<br />

Analytical Biochemistry 72, 248–254.<br />

Brenner, S., Horne, R.W., 1959. A negative staining method for high resolution<br />

electron microscopy <strong>of</strong> viruses. Biochimica et Biophysica Acta 34, 103–110.<br />

Chen, S., Berthelier, V., Hamilton, J.B., O’Nuallain, B., Wetzel, R., 2002. Amyloid-like<br />

features <strong>of</strong> polyglutamine aggregates and their assembly kinetics. Biochemistry<br />

41, 7391–7399.<br />

Chiti, F., Dobson, C.M., 2006. Protein misfolding, functional amyloid, and human<br />

disease. Annual Review <strong>of</strong> Biochemistry 75, 333–366.<br />

DeMarco, M.L., Silveira, J., Caughey, B., Daggett, V., 2006. Structural properties <strong>of</strong><br />

prion protein prot<strong>of</strong>ibrils and fibrils: an experimental assessment <strong>of</strong> atomic<br />

models. Biochemistry 45, 15573–15582.<br />

Dobson, C.M., 1999. Protein misfolding, evolution and disease. Trends in<br />

Biochemical <strong>Science</strong>s 24, 329–332.<br />

Dobson, C.M., 2001. The structural basis <strong>of</strong> protein folding and its links with human<br />

disease. Philosophical Transactions <strong>of</strong> the Royal Society <strong>of</strong> London. Series B,<br />

Biological <strong>Science</strong>s 356, 133–145.<br />

Gerrard, J.A., Fayle, S.E., Brown, P.K., Sutton, K.H., Simmons, L., Rasiah, I., 2001.<br />

Effects <strong>of</strong> microbial transglutaminase on the wheat proteins <strong>of</strong> bread and<br />

croissant dough. <strong>Journal</strong> <strong>of</strong> Food <strong>Science</strong> 66, 782–786.<br />

Guilbert, S., Gontard, N., Morel, M.H., Chalier, P., Micard, V., Redl, A., 2001. Formation<br />

and properties <strong>of</strong> wheat gluten films and coatings. In: Gennadios, A. (Ed.),<br />

Protein-based Films and Coatings. CRC Press, Boca Raton, pp. 69–122.<br />

Gras, S., 2007. Amyloid fibrils: from disease to design. New biomaterial applications<br />

for self-assembling cross-b fibrils. Australian <strong>Journal</strong> <strong>of</strong> Chemistry 60, 333–342.<br />

Kasarda, D.D., Bernardin, J.E., Thomas, R.S., 1967. Reversible aggregation <strong>of</strong> a-gliadin<br />

to fibrils. <strong>Science</strong> 155, 203–205.<br />

Kim, J.R., Muresan, A., Lee, K.Y., Murphy, R.M., 2004. Urea modulation <strong>of</strong> b-amyloid<br />

fibril growth: experimental studies and kinetic models. Protein <strong>Science</strong> 13,<br />

2888–2898.<br />

S.H. Mackintosh et al. / <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 49 (2009) 157–162<br />

Kipp, B., Belitz, H.-D., Seilmeier, W., Wieser, H., 1996. Comparative studies <strong>of</strong> high Mr<br />

subunits <strong>of</strong> rye and wheat. I. Isolation and biochemical characterisation and<br />

effects on gluten extensibility. <strong>Journal</strong> <strong>of</strong> <strong>Cereal</strong> <strong>Science</strong> 23, 227–234.<br />

Kwan, A.H., Winefield, R.D., Sunde, M., Matthews, J.M., Haverkamp, R.G.,<br />

Templeton, M.D., Mackay, J.P., 2006. Structural basis for rodlet assembly in<br />

fungal hydrophobins. Proceedings <strong>of</strong> the National Academy <strong>of</strong> <strong>Science</strong>s 103,<br />

3621–3626.<br />

LeVine III, H., 1999. Quantification <strong>of</strong> beta-sheet amyloid fibril structures with thi<strong>of</strong>lavin<br />

T. In: Wetzel, R. (Ed.), Methods in Enzymology 309: Amyloid, Prions, and<br />

Other Protein Aggregates. Academic Press, US, pp. 274–284.<br />

Lew, E.J.-L., Kuzmicky, D.D., Kasarda, D.D., 1992. Characterization <strong>of</strong> low molecular<br />

weight glutenin subunits by reversed-phase high-performance liquid chromatography,<br />

sodium dodecyl sulfate-polyacrylamide gel electrophoresis, and Nterminal<br />

amino acid sequencing. <strong>Cereal</strong> Chemistry 69, 505–515.<br />

Lindsay, M.P., Skerritt, J.H., 1999. The glutenin macropolymer <strong>of</strong> wheat flour doughs<br />

structure–function perspectives. Trends in Food <strong>Science</strong> & Technology 10,<br />

247–253.<br />

Makin, O.S., Serpell, L.C., 2004. Structural characterisation <strong>of</strong> islet amyloid polypeptide<br />

fibrils. <strong>Journal</strong> <strong>of</strong> Molecular Biology 335, 1279–1288.<br />

Makin, O.S., Serpell, L.C., 2005. Structures for amyloid fibrils. FEBS <strong>Journal</strong> 272,<br />

5950–5961.<br />

Makin, O.S., Atkins, E., Sikorski, P., Johansson, J., Serpell, L.C., 2005. Molecular basis<br />

for amyloid fibril formation and stability. Proceedings <strong>of</strong> the National Academy<br />

<strong>of</strong> <strong>Science</strong>s <strong>of</strong> the United States <strong>of</strong> America 102, 315–320.<br />

Nilsson, M.R., Dobson, C.M., 2003. Chemical modification <strong>of</strong> insulin in amyloid<br />

fibrils. Protein <strong>Science</strong> 12, 2637–2641.<br />

Osborne, T.B., 1907. The proteins <strong>of</strong> the wheat kernel. In: Judd, Detweiller (Eds.),<br />

Publications <strong>of</strong> the Carnegie Institution Washington No. 84, pp. 1–119. Washington,<br />

DC.<br />

Pawar, A.P., DuBay, K.F., Zurdo, J., Chiti, F., Vendruscolo, M., Dobson, C.M., 2005.<br />

Prediction <strong>of</strong> ‘‘aggregation-prone’’ and ‘‘aggregation-susceptible’’ regions in<br />

proteins associated with neurodegenerative diseases. <strong>Journal</strong> <strong>of</strong> Molecular<br />

Biology 350, 379–392.<br />

Perutz, M.F., Finch, J.T., Berriman, J., Lesk, A., 2002. Amyloid fibers are water-filled<br />

nanotubes. Proceedings <strong>of</strong> the National Academy <strong>of</strong> <strong>Science</strong>s 99, 5591–5595.<br />

Rauscher, S., Baud, S., Miao, M., Keeley, F.W., Pomes, R., 2006. Proline and glycine<br />

control protein self-organization into elastomeric or amyloid-like fibrils.<br />

Structure 14, 1667–1676.<br />

Sanford, K., Kumar, M., 2005. New proteins in a materials world. Current Opinion in<br />

Biotechnology 16, 416–421.<br />

Serpell, L.C., Fraser, P.E., Sunde, M., 1999. In: Wetzel, R. (Ed.), Methods in Enzymology<br />

309: Amyloid, Prions, and Other Protein Aggregates. Academic Press,<br />

US, pp. 526–536.<br />

Serpell, L.C., 2000. Alzheimer’s amyloid fibrils: structure and assembly. Biochimica<br />

et Biophysica Acta 1502, 16–30.<br />

Shen, X., Mo, X., Moore, R., Frazier, S.J., Iwamoto, T., Tomich, J.M., Sun, X.S., 2006.<br />

Adhesion and structural properties <strong>of</strong> protein nanomaterials containing<br />

hydrophobic and charged amino acids. <strong>Journal</strong> <strong>of</strong> Nanoscience and Nanotechnology<br />

6, 837–844.<br />

Shewry, P.R., Halford, N.G., Belton, P.S., Tatham, A.S., 2002. The structure and<br />

properties <strong>of</strong> gluten: an elastic protein from wheat grain. Philosophical Transactions<br />

B 357, 133–142.<br />

Stathopulos, P.B., Scholz, G.A., Hwang, Y.M., Rumfeldt, J.A., Lepock, J.R.,<br />

Meiering, E.M., 2004. Sonication <strong>of</strong> proteins causes formation <strong>of</strong> aggregates that<br />

resemble amyloid. Protein <strong>Science</strong> 13, 3017–3027.<br />

Sugiyama, T., Rafalski, A., Peterson, D., Soll, D., 1985. A wheat HMW glutenin subunit<br />

gene reveals a highly repeated structure. Nucleic Acids Research 13, 8729–8737.<br />

Tatham, A.S., 1995. The structures <strong>of</strong> wheat proteins. In: Sch<strong>of</strong>ield, J.D. (Ed.), Wheat<br />

Structure, Biochemistry and Functionality. The Royal Society <strong>of</strong> Chemistry,<br />

Reading, U.K., pp. 53–62.<br />

Tatham, A.S., Shewry, P.R., 2002. Comparative structures and properties <strong>of</strong> elastic<br />

proteins. Philosophical Transactions B 357, 229–234.<br />

Wang, S.S., Carpenter, F.H., 1967. Kinetics <strong>of</strong> the tryptic hydrolysis <strong>of</strong> the oxidized B<br />

chain <strong>of</strong> bovine insulin. Biochemistry 6, 215–224.<br />

Waterhouse, S.H., Gerrard, J.A., 2004. Amyloid fibrils in bionanotechnology.<br />

Australian <strong>Journal</strong> <strong>of</strong> Chemistry 57, 519–523.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!