Anda di halaman 1dari 225

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/275517565

Assessing Quality of Novel Plant Proteins for


Salmonids

THESIS JANUARY 2012


DOI: 10.13140/RG.2.1.1688.0167

READS

23

1 AUTHOR:

M A Kabir Chowdhury
University of Guelph
47 PUBLICATIONS 207 CITATIONS

SEE PROFILE

Available from: M A Kabir Chowdhury


Retrieved on: 23 February 2016
Assessing Quality of Novel Plant Proteins for Salmonids
by

M A Kabir Chowdhury

A Thesis
Presented to
The University of Guelph

In partial fulfilment of requirements


for the degree of
Doctor of Philosophy
in
Animal and Poultry Science

Guelph, Ontario, Canada

M A Kabir Chowdhury, January, 2012


ABSTRACT

ASSESSING QUALITY OF NOVEL PLANT PROTEINS FOR SALMONIDS

M A Kabir Chowdhury Advisor:


University of Guelph, 2012 Professor D. P Bureau

Approaches for the evaluation of plant protein ingredients for salmonid feeds were

investigated in a series of four trials. The first trial compared the apparent digestibility

coefficients (ADC) of crude protein (CP) and amino acids (AAs) of two novel products -

Indian mustard protein concentrate (IMC, 62% CP) and Indian mustard protein meal (IMM,

42% CP), to a commercially available soy protein concentrate (SPC, 57% CP) for two

salmonid species, rainbow trout and Atlantic salmon. The second trial involved assessment

of relative bioavailability of arginine (Arg) from IMC, IMM and SPC compared to that of a

crystalline Arg (L-Arg) in rainbow trout using slope-ratio assay. In the third trial, the effects

of phytic acid (PA) and lignin on nutrient utilization and partitioning in rainbow trout were

assessed. Finally, a series of experiments was conducted in the final trial to establish the

evaluation criteria for pellet quality assessment. The ADC of CP and most AAs in IMC and

IMM were high (>90%). Differences in the ADCs of some AAs can be attributed to the high

PA intake by fish fed 30%-IMC diet. The significantly higher (P<0.05) bioavailability of

Arg from IMC (123 to 187%) and IMM (116 to 211%) relative to that of L-Arg, as

determined by various regression approaches, reaffirmed the findings of the first trial that

these ingredients are of excellent protein quality and can readily be used in compounded fish

feeds. It can be inferred from the lack of effects of PA, lignin or PA plus lignin on most

indices of physiology, performance, and nutrient utilization in the pair-fed fish, that like any
other animal, controls feed intake when in the presence of one or more dietary ANF. It was

also shown in the pellet quality assessment trial that minor changes in dietary composition

can significantly alter physical properties of aquaculture feed. This study highlighted the

importance of a comprehensive assessment for the effective evaluation of the nutritive value

of plant protein ingredients for use in aquaculture feeds.


Dedication
To my father
Dr. Badiuzzaman Chowdhury (1936-1981)

iv
ACKNOWLEDGMENTS

First and foremost, I would like to thank my advisor Dr. Dominique P. Bureau for his

guidance, caring help and support during this endeavour to shape my ideas and knowledge

around the basic principles of fish nutrition. His straightforward but no stones unturned

approach to science and nutrition in particular, has taught me some very valuable lessons. I have

nothing but utmost respect for you. Thank you, Dom.

Thanks to my advisory committee, Dr. Cornelis F. M. de Lange, Dr. Kendall Swanson, and

Dr. Vernon Osborne for all the help and guidance they have provided me along the way. Thank

you for all your patience and open-door approach to something relatively uncommon. Thank you

Kees for your valuable insights and ideas that pushing me to think differently. Thank you Vern

for the opportunities you have given me, whether it is teaching or discussing new ideas or other

forms of support.

I would also like to thank Professor Richard Moccia and Dr. Albert Tacon for their

contribution to the thesis. As examiners, their broad perspectives and positive attitudes have

made a stressful day more rewarding.

I would like to thank Dr. Margaret Quinton for her help in understanding some difficult

statistical analyses. I would also like to thank Kattia Peciado Iniguez and Thomas Martie for

their significant role in two of my experiments. Thanks to my past and present lab mates Dr.

Andre Dumas, Dr. Katheline Hua, Dr. Guillaume Salze, Jamie Hooft, Patricio Saez, and many

others who have been a true inspiration with their attitude towards building a good team and

helping each other out at difficult times.

v
The study was funded by the BIOEXX Specialty Proteins Ltd. (Toronto, ON, Canada),

Mathematics of Information Technology and Complex Systems (MITACS Inc., Vancouver, BC,

Canada), Martin Mills Inc. (Elmira, ON, Canada), and the Ontario Ministry of Natural

Resources. I would also to recognize DANISCO Animal Nutrition (MO, USA), Evonik

Industries AG (Hanau, Germany) and Pure Lignin Environmental Laboratories (BC, Canada) for

their contribution in analyzing phytic acid, amino acid and providing purified lignin,

respectively.

Finally, I would like to thank my family for their love, patience, support and help during

this study. Without their support, it would have been impossible to pursuit this dream.

vi
TABLE OF CONTENTS

Acknowledgment..................................................................................................................... v

Table of Contents.................................................................................................................... vii

List of Tables........................................................................................................................... xiii

List of Figures......................................................................................................................... xvii

List of Abbreviations.............................................................................................................. xx

Chapter 1 General Introduction.......................................................................................... 1

Chapter 2 Review of Literature........................................................................................... 7

2.1. Introduction................................................................................................................ 7

2.2. Review of Aquaculture Feed Production and Aquaculture Nutrition........................ 9

2.2.1. Evolution of salmonid feed formulation............................................................. 9

2.2.2. Evolution of fish meal usage.......................................................... 15

2.2.3. Desired characteristics of a potential feed ingredient......................................... 15

2.3. Overview of World Feed Ingredients......................................................................... 16

2.3.1. Production of oilseed meals............................................ 18

2.3.2. Production of protein concentrates and isolates............................................. 18

2.3.3. Further improvement of the protein quality............................................ 19

2.3.4. Production approaches................................................ 20

2.3.2. Rapeseed / canola / mustard (Brassicacae)........................................................ 22

2.3.2.1. Indian mustard (Brassica juncea)............................................................... 22

2.3.2.2. Nutritive value of mustard products: issues and challenges....................... 23

2.4. Nutritive Value of Protein Ingredients to Fish........................................................... 24

2.4.1. Overview of methodological approaches........................................................... 25

vii
2.4.2. Digestibility............................................................ 26

2.4.2.1.. Digestive process related to protein and amino acids... 26

2.4.2.2. Assessing digestibility in fish: techniques, issues and limitations......... 30

2.4.2.2.1. Fecal collection methods.................................................................. 30

2.4.2.2.2. Indigestible markers.......................................................................... 31

2.4.2.2.3. In-vitro assessment of digestibility................................................... 32

2.4.2.2.4. Limitations of digestibility of proteins in feed ingredients.............. 33

2.4.3. Bioavailability of amino acids............................................................................ 33

2.4.3.1. Methodological approaches........................................................................ 35

2.4.3.2. Limitations of slope-ratio approaches......................................................... 36

2.4.4. The presence and effects of ANF and other non-nutritive components............. 37

2.4.4.1. General overview........................................................................................ 37

2.4.4.2. Phytic acid................................................................................................... 41

2.4.4.2.1. Interactions with minerals................................................................... 41

2.4.4.2.2. Interactions with protein and amino acids.......................................... 45

2.4.4.2.3. Interactions with proteolytic enzymes................................................. 46

2.4.4.2.4. Effect of PA on nutrient utilization................................................. 46

2.4.4.3. Fibre components and their effects............................................................. 47

2.4.4.4. Lignin.......................................................................................................... 48

2.4.4.4.1. Lignin biosynthesis............................................................................. 48

2.4.4.4.2. Effects of purified lignin..................................................................... 50

2.4.4.4.3. Effects of lignin on minerals and amino acid availability................... 50

2.4.5. Physical characteristics of aquaculture feed...................................................... 51

viii
2.4.5.1. Overview of feed production technology.................................................... 51

2.4.5.2. Physical properties of pellets and their importance.................................... 54

2.4.5.3. Different factors affecting physical properties........................................... 56

2.4.5.4. Components of plant proteins influencing pellet characteristics................ 56

2.5. Conclusion........................................................................................................................ 57

Chapter 3 Digestibility of Amino Acids in Two Novel Ingredients, Indian Mustard

Protein Concentrate and Indian Mustard Meal Compared to That of A

Soy Protein Concentrate in Rainbow Trout and Atlantic Salmon............... 59

Abstract.............................................................................................................................. 59

3.1. Introduction................................................................................................................ 60

3.2. Methods...................................................................................................................... 62

3.2.1. Dietary design..................................................................................................... 63

3.2.2. Experimental conditions..................................................................................... 67

3.2.3. Chemical analyses.............................................................................................. 68

3.2.4. Calculations........................................................................................................ 69

3.2.5. Statistical analysis.............................................................................................. 69

3.3. Results........................................................................................................................ 70

3.3.1. Apparent digestibility of energy and nutrients................................................... 70

3.3.2. Apparent digestibility of amino acids................................................................ 71

3.4. Discussion.................................................................................................................. 75

Chapter 4 Relative Bioavailability of Arginine in Novel Indian Mustard Protein

Concentrate and Indian Mustard Meal Compared to That of A Soy

Protein Concentrate in Rainbow Trout (Oncorhynchys Mykiss) as

ix
Determined by Slope-Ratio Assay................................................................... 81

Abstract............................................................................................................................. 81

4.1. Introduction................................................................................................................ 82

4.2. Methods...................................................................................................................... 85

4.2.1. Experimental diets.............................................................................................. 87

4.2.2. Experimental conditions..................................................................................... 87

4.2.3. Chemical analyses.............................................................................................. 91

4.2.4. Calculations........................................................................................................ 92

4.2.5. Statistical analysis.............................................................................................. 92

4.3. Results........................................................................................................................ 93

4.3.1. Growth performance and nutrient utilization..................................................... 94

4.3.2. Relative bioavailability of arginine.................................................................... 94

4.4. Discussion.................................................................................................................. 103

Chapter 5 Effect of Dietary Phytic Acid and Purified Lignin and Their Interactions

on Physiological Indices, Growth Performance, Nutrient and Energy

Partitioning of Rainbow Trout, Oncorhynchus Mykiss................................. 107

Abstract............................................................................................................................ 107

5.1. Introduction................................................................................................................ 108

5.2. Methods...................................................................................................................... 111

5.2.1. Experimental diets.............................................................................................. 112

5.2.2. Experimental conditions and feeding management............................................ 112

5.2.3. Physiological indices.......................................................................................... 113

5.2.4. Chemical analysis............................................................................................... 113

x
5.2.5. Calculations........................................................................................................ 115

5.2.6. Statistical analysis.............................................................................................. 115

5.3. Results........................................................................................................................ 120

5.4. Discussion.................................................................................................................. 125

Chapter 6 Assessing The Effects of Lignin on Physical Properties of Steam-Pelleted

Experimental Trout Feed................................................................................. 130

Abstract............................................................................................................................. 130

6.1. Introduction................................................................................................................ 131

6.2. Methods...................................................................................................................... 132

6.2.1. Oil and water absorption by the mash................................................................ 135

6.2.2. Physical properties of the pellets........................................................................ 135

6.2.3. Statistical analysis.............................................................................................. 136

6.3. Results........................................................................................................................ 137

6.3.1. Oil and water absorption.................................................................................... 137

6.3.2. Physical properties............................................................................................. 137

6.4. Discussion.................................................................................................................. 142

6.4.1. Oil and water absorption.................................................................................... 142

6.4.2. Pellet quality....................................................................................................... 143

6.4.3. Conclusion.......................................................................................................... 145

Chapter 7 General Discussion............................................................................................ 146

7.1. Evaluation of A Novel Plant Protein Ingredient: Why is It Important? .................... 146

7.2. Digestibility of Amino Acids from Plant Protein Sources......................................... 147

7.3. Bio-availability of Amino Acid and Slope-ratio Assay............................................. 149

xi
7.4. Effect of Phytic Acid and Ligin on Nutrient Utilization............................................ 151

7.3. Beyond Tradition........................................................................................................ 153

Bibliography............................................................................................................................ 157

xii
LIST OF TABLES
Table 2.1 A chronological account of the evolution of salmonid feed type, feed

ingredients, crude protein and lipid contents since

1920................................................................................................................ 14

Table 2.2 Major co-products from grain, oilseed and animal production industries

available for compounded feed productidson (estimated global output, in

MMT)............................................................................................................. 17

Table 2.3 Major enzymes in fish for the digestion of protein, peptides, amino acids

and other non-nitrogenous compounds. 29

Table 2.4 Major anti-nutrients in soybean meals and their effects on salmonid fishes. 38

Table 2.5 Total phosphorus and phyate-P content in the major plant ingredients

currently utilized in compounded aquafeed.................................................. 44

Table 2.6 Indication of final pellets bulk density in relation to floating or sinking

properties........................................................................................................ 55

Table 2.7 Final fat contents (%) in the pellets of different aquaculture species in

relation to floating or sinking properties........................................................ 55

Table 3.1 Ingredient composition of the basal diet (%)................................................ 64

Table 3.2 Proximate chemical composition, gross energy, essential amino acid

profiles, total phosphorus and phytic acid content of the test ingredients

(as is basis).................................................................................................... 65

Table 3.3 Proximate analysis and, gross energy, essential amino acid profiles, total

phosphorus and phytic acid concentrations of the diets (as is basis)............ 66

Table 3.4 ADC of nutrients, phosphorus, gross energy and essential amino acids in

xiii
the reference diet (diet 1) and the three test diets for rainbow trout and

Atlantic salmon............................................................................................. 72

Table 3.5 ADC of nutrients and gross energy in Indian mustard protein concentrate

(IMC), Indian mustard meal (IMM) and soy protein concentrate (SPC) for

two rainbow trout and Atlantic salmon......................................................... 73

Table 3.6 ADC of eight essential amino acids in Indian mustard protein concentrate

(IMC), Indian mustard meal (IMM) and soy protein concentrate (SPC) for

rainbow trout and Atlantic salmon................................................................ 74

Table 4.1 Proximate chemical composition, gross energy, essential amino acid

profiles, total phosphorus and phytic acid content of the test ingredients... 86

Table 4.2 Ingredients composition of the diets (g kg-1)............................................... 89

Table 4.3 Nutrient, gross energy and essential amino acid (EAA) composition of the

test diets (DM basis)..................................................................................... 90

Table 4.4 Weight gain, TGC and feed intake and feed efficiency of rainbow trout fed

diets with increasing levels of arginine supplied by L-Arg or three protein-

bound sources (mean SD)............................................................................ 95

Table 4.5 Protein (PD) and lipid (LD) deposition, recovered energy (RE), N- (NRE)

and energy (ERE) retention efficiencies for rainbow trout fed diets with

increasing levels of arginine (mean SD)..................................................... 96

Table 4.6 Slopes (b) and confidence intervals (CI) of thermal-unit growth

coefficients (TGC) and protein deposition (PD) in fish fed diets from

different sources of arginine (L-arginine, Indian mustard protein

concentrate - IMC, Indian mustard protein meal - IMM, soy protein

xiv
concentrate - SPC) as determined by linear regressions using two and one

level of arginine (%) and arginine intake (g fish-1)...................................... 97

Table 4.7 Bioavailability of arginine from Indian mustard protein concentrate - IMC,

Indian mustard protein meal - IMM, soy protein concentrate - SPC in

rainbow trout in compare to L-arginine based on thermal-unit growth

coefficients (TGC) and protein deposition (PD) as determined by linear

regressions using two and one level of arginine (%).................................... 98

Table 5.1 Basal mix for the test diets............................................................................ 117

Table 5.2 Ingredient and proximate chemical composition of the test diets (DM

basis).............................................................................................................. 118

Table 5.3 Analyzed essential and non-essential amino acid composition of the test

diets (digestible AA, % DM)........................................................................ 119

Table 5.4 Performance and physiological indices of fish (initial weight = 40 g fish-1)

fed the experimental diets for 84d at 15C. Significance between the

diets............................................................................................................... 122

Table 5.5 Whole carcass nutrient deposition and retention efficiency of rainbow

trout fed test diets with various levels of phytic acid and lignin................... 123

Table 5.6 Nutrient deposition and retention efficiencies in dressed carcass and

viscera of rainbow trout fed the experimental diets...................................... 124

Table 6.1 Ingredient composition of the basal mix prepared for assessing oil and

water absorption capacity with the addition of lignin. 16.0% fish oil was

added before mixing with kaolin and lignin to prepare three diets of 0%,

1.5% and 3% lignin....................................................................................... 134

xv
Table 6.2 Oil (fish oil and vegetable oil) and water absorption capacity (in %) of the

diet mash with various levels of lignin (Mean SE).................................... 138

Table 6.3 Sinking velocity (cm s-1) and hardness (Newton) of the diet pellets with

various levels of lignin (mean SE)............................................................. 139

Table 6.4 Parameter estimates (b) of the effect of lignin levels (0%, 1.5% and 3%),

pellet size, duration in water (time) and their interactions on pellet water

stability (%), parameter estimates (b) of the effect of the three lignin levels

on sinking velocity and hardness of the pellets, and oil and water

absorption of the mash (lignin levels 0%, 1.5%, 3%, 6%, and 12%)........... 141

xvi
LIST OF FIGURES
Figure 2.1 The world production of fish and shrimp (*excluding filter feeding

fishes, dark grey), and the fraction produced fed compounded feed (in

%, light grey) since 1995.......................................................................... 10

Figure 2.2 Global output of farmed salmonids in weight (MMT) and value (billion

USD)........................................................................................................ 11

Figure 2.3 Production of farmed salmonids (in thousand metric tonne) in

continental America, Asia, and Europe in 2008 and 2009....................... 12

Figure 2.4 Flow-chart of soybean processing showing the techniques used to

produce various soy products for consumption........................................ 21

Figure 2.5 Digestion and absorption of dietary proteins in the intestine of fish....... 28

Figure 2.6 Dietary fibre components......................................................................... 40

Figure 2.7 The structure of phytic acid (myo-inositol hexaphosphate, IP6)............. 43

Figure 2.8 Structure of three monolignol precursors to lignin (adapted from

Hatfield and Vermerris, 2001).................................................................. 49

Figure 2.9 Steps in a typical aquaculture feed manufacturing process...................... 53

Figure 3.1 Apparent digestibility coefficients (%) of arginine - Arg (pKa 12.5),

lysine Lys (pKa 10.5) and histidine - His (pKa 6.1) in Indian

mustard protein concentrate (IMC), Indian mustard meal (IMM) and

soy protein concentrate (SPC) for rainbow trout and Atlantic salmon

and their relationships to the dietary phytic acid intake. Dietary phytic

acid intake in 30%-IMC, 30%-IMM and 30%SPC diets were 29

mg.fish-1.d-1, 15 mg.fish-1.d-1 and 14 mg.fish-1.d-1 for rainbow trout and

xvii
10 mg.fish-1.d-1, 6 mg.fish-1.d-1 and 5 mg.fish-1.d-1 for Atlantic salmon,

respectively............................................................................................. 79

Figure 3.2 Broken line regression of ADC (%) of phenylalanine in Indian mustard

protein concentrate (IMC), Indian mustard meal (IMM) and soy

protein concentrate (SPC) for rainbow trout and Atlantic salmon with

the dietary phytic acid intake (R2=0.83). Dietary phytic acid intake in

30%-IMC, 30%-IMM and 30%SPC diets were 29 mg.fish-1.d-1, 15

mg.fish-1.d-1 and 14 mg.fish-1.d-1 for rainbow trout and 10 mg.fish-1.d-1,

6 mg.fish-1.d-1 and 5 mg.fish-1.d-1 for Atlantic salmon,

respectively............................................................................................. 80

Figure 4.1 Weight gain (g fish-1) and thermal-unit growth coefficient (TGC) of

rainbow trout in response to dietary arginine intake.............................. 99

Figure 4.2 Protein (PD) and lipid (LD) deposition in rainbow trout in response to

dietary arginine intake (g fish-1)............................................................. 100

Figure 4.3 Thermal-unit growth coefficient (TGC) and protein deposition (PD) of

rainbow trout in response to arginine level (%) in diets from four

arginine sources (L-Arg, IMC - Indian mustard protein concentrate,

IMM -Indian mustard protein meal, SPC -soy protein concentrate).

Solid and dotted lines represent response based on two and one level of

arginine, respectively............................................................................ 101

Figure 4.4 Thermal-unit growth coefficient (TGC) and protein deposition (PD) of

rainbow trout in response to arginine intake (g fish-1) from four

arginine sources (L-Arg, IMC - Indian mustard protein concentrate,

xviii
IMM -Indian mustard protein meal, SPC -soy protein concentrate).

Solid and dotted lines represent response based on two and one level of

arginine, respectively............................................................................ 102

Figure 5.1 Total feed intake of fish fed experimental diets when fed to satiety

during the first two weeks of the trial. Diets 1, 2, 3 (0% lignin), diets 4,

5, 6 (1.5% lignin) and diets 7, 8, 9 (3% lignin)..................................... 121

Figure 6.1 Effect of lignin inclusion rate (0%, 1.5% and 3%) on dry matter loss

and nitrogen leaching from pellets of two sizes (2-mm and 3-mm)

following immersion in shaking water bath at 100 rpm min-1 for >1-

min, 10-min and 20-min of durations...................................................... 140

Figure 7.1 A flow-diagram for characterization of the compoenents of a plant

protein ingredient..................................................................................... 155

Figure 7.2 A frame-work for evaluation of a plant protein

ingredient.................................................................................................. 156

xix
LIST OF ABBREVIATIONS

AA Amino acid
ADC Apparent digestibility coefficient
ADF Acid detergent fibre
ADL Acid detergent lignin
ANF Anti-nutritional factors
ATP Adenosine tri-phosphate
BBM Brush border membrane
BBS Bobesin
CART Cocaine- and amphetamine-regulated transcript
CCK Cholecystokinin
CGM Corn gluten meal
CI Confidence interval
CP Crude protein
CRF Corticotrophin-releasing factor
DDGS Dried distillers grain solubles
DEI Digestible energy intake
DHA Docosahexaenoic acid
DM Dry matter
DNI Digestible nitrogen intake
EAA Essential amino acid
EFA Essential fatty acid
EPA Eicosapentaenoic acid
ERE Energy retention efficiency
EU European Union
FA Fatty acid
FAS Fatty acid synthetase
FBW Final body weight
FCI Fultons condition index
FCR Feed conversion ratio
xx
FE Feed efficiency
FO Fish oil
G Guaicyl
GDAS Gastric dilation air sacculitis
GE Gross energy
GIT Gastro-intestinal tract
GLP Glucagon like peptide
GRP Gastrin releasing peptideds
IBW Initial body weight
H Hydroxyphenyl
IMC Indian mustard protein concentrate
IMM Indian mustard protein meal
IP6 Myo-inositol hexaphosphate
IPF Intra-peritoneal fat
LD Lipid deposition
LPL Lupoprotein lipase
LRE Lipid retention efficiency
MBM Meat and bone meal
MBW Metabolic body weight
MUFA Mono-unsaturated fatty acid
MMT Million metric tons
NA North America
NDF Neutral detergent fibre
NEAA Non-essential amino acid
NPY Neuropeptide-Y
NRC National Research Council
NRE Nitrogen retention efficiency
PA Phytic acid
PBM Poultry by-product meal
PD Protein deposition
PDI Protein dispersibility index
xxi
PUFA Poly-unsaturated fatty acid
RE Recoevred energy
S Syringyl
SBM Soybean meal
SPC Soy protein concentrate
TAG Tri-acyl glyceraldehyde
TGC Thermal-unit growth coefficient
USDA United States Department of Agriculture
VO Vegetable oil
WRO World Renderers Organization

xxii
CHAPTER 1

GENERAL INTRODUCTION

Aquaculture, the fastest growing food-producing sector in the world, has been supplying

49% of the worlds food fish since 2010 (FAO, 2011). A majority (22.2 million metric tons

MMT) of this is comprised of various carp species mainly produced in China and Indian sub-

continent followed by 3.2 MMT of tilapia (FAO, 2011). The significant production increase of

these two groups of fish species in the past few years is mainly attributable to the culture

intensification resulting in higher growth in the worldwide demand for compounded aquaculture

feed than that of the production (Naylor et al., 2009; Tacon and Metian, 2008). Global

production of various salmonid species is third after carps and tilapias at 2.5 MMT in 2009, an

increase of 67% from its 2000 production of 1.5 MMT (FAO, 2011). Salmonids are generally fed

high protein and lipid diets (nutrient-dense), which can support high growth performance and

feed efficiency, and low waste outputs (Bureau and Hua, 2010). Fish meal has traditionally been

the main protein source in nutrient-dense feeds (Bakke et al., 2011; Bureau and Hua, 2010).

Salmonid feeds are increasingly formulated to contain high levels of economical plant

protein ingredients due to the rising demand, stagnating supply, and price volatility of fish

meal. Improved understanding of the nutritional requirements of salmonids has allowed

formulation of successful feeds based on plant protein ingredients. But the inclusion of many

plant protein ingredients in salmonid feed formulations is limited because of their low protein

and high fibre contents, and/or the presence of various anti-nutritional factors (ANF)

(Dabrowska and Wojno, 1977; Dabrowski et al., 1989; Ketola, 1975; Page and Andrews, 1973;

Refstie et al., 2000; Wee and Shu, 1989). Growth and fish health problems have frequently been

1
reported for salmonid fish fed diets formulated with very high levels of plant proteins (Gaylord

et al., 2010).

In recent years, application of various extraction, fractionation, and processing techniques

resulted in innovative high quality plant protein products (>45% CP) low in fibre and starch and

ANF from oilseed and grain industries. Processing addressed many of the limitations associated

with feeding diets with very high levels of plant ingredients. Protein concentrates of soybean and

canola, for example, are of very high nutritive value to salmonids and very valuable ingredients

for nutrient dense diets formulation (Drew et al., 2007; Higgs et al., 1994; Mambrini et al., 1999;

Mwachireya et al., 1999; Storebakken et al., 1998b; 2000; Thiessen et al., 2004). These

ingredients are more widely available and increasingly economical, and their use in commercial

salmonid feed formulations is growing steadily. However, the cost of many of these ingredients

is still often prohibitive and many studies have shown that growth and feed efficiency of

salmonids can suffer when high levels of these ingredients are used in diets.

Increasing use of a wide variety of plant protein ingredients in aquaculture feeds has

highlighted the need for better characterization and evaluation of the nutrients of these

ingredients (Bureau, 2008; Gatlin et al., 2007). Glencross et al. (2007b) reviewed strategies

employed in evaluating aquafeed ingredients, and described five key components including

ingredient characterization (proximate, macro- and micro-nutrients composition), digestibility of

nutrients, palatability, nutrient utilization, and functionality. Characterization of a plant protein

ingredient should include information on proximate composition, its amino acid profile, macro-

and micro-mineral contents, composition of various fibre components and anti-nutritional

factors.

2
The determination of nutrient digestibility is the first step to evaluate the potential of an

ingredient for use in an aquaculture diet. The digestibility of hundreds of different feed

ingredients to different aquaculture species has been estimated in the past (NRC, 2011).

However, most digestibility studies have focused on apparent digestibility of proximate analysis

components (crude protein, lipids, gross energy) and very limited effort has been invested in

estimating the digestibility or availability of individual nutrients (e.g. amino acids) in feed

ingredients (Bureau, 2008; De Silva et al., 2000; Gaylord et al., 2010; Kaushik and Seiliez, 2010;

NRC, 2011). Fish feeds are increasingly formulated on digestible amino acid basis and more

effort should be invested in determining the amino acid digestibility of different protein sources

(El-Haroun et al., 2009; Poppi et al., 2011; Yamamoto et al., 2005).

Digestibility is a measure of disappearance of nutrients from the gastrointestinal tract and

estimates of digestibility may not necessarily reflect the amount of nutrient useable

(bioavailable) to the animal. More direct assessment of proportions of ingested dietary nutrients

that is potentially available for metabolism or synthesis is important (Ammerman et al., 1995;

Batterham, 1992; Lewis and Baily, 1995; Stein et al., 2007). Among several nutrient bio-

availability protocols, slope-ratio assays have been shown to provide direct and practical

assessment of the level of of individual nutrients in feed ingredients, and have been used to

assess the bioavailability of individual amino acids in dfferent ingredients for livestock (Lewis

and Bailey, 1995). Application of slope-ratio assays to determine bioavailability of amino acids

in fish is relatively recent, and significant variations exist among the methods adopted (El-

Haroun and Bureau, 2007; Li et al., 2008, 2009; Poppi et al., 2011; Zarate and Lovell, 1997).

Another important step in evaluation of plant protein ingredients is to assess the effects of

ANF and fibre components on utilization of nutrients, and on health and well-being of the

3
species. Although fibre components and most ANF are either reduced or removed during

processing of ingredients, some are difficult to remove and their concentration may increase

during fractionation of protein from fibre components. Increasing inclusion of novel plant

proteins in compounded fish feed has accentuated the need for comprehensive species-specific

evaluation of the effect of ANF and fibre components of individual ingredients.

Phytic acid is probably one of the most studied ANF in fish nutrition because of its

effects on availability of phosphorus, and other divalent minerals such as Ca2+ and Mg2+. Proton

disassociation sites on the phytic acid molecules have variable pKa values between 1.5 and

>10.0. Beside their well- known effects on digestibility of minerals (Cowieson et al., 2004;

Maenz et al., 1999; Torre et al.; 1991), increasing evidence suggests that phytic acid may affect

digestibility of amino acids (Cowieson et al., 2004). Phytic acid may become negatively charged

when a proton is disassociated attracting positively charged molecules at appropriate pH (Adeola

and Sands, 2003). Phytic acid also interacts with protein and creates binary (phytic acid-protein)

or ternary (phytic acid-mineral-protein) complexes that occur only at optimal pH conditions in

the gastro-intestinal tracts (Cheryan and Rakis, 1980; Kies et al., 2001; Maenz, 2001). The -

NH2 terminal group and the -NH2 group of lysine, the imidazole group of histidine, and the

positively charged guadinyl group of arginine have been implicated as possible protein binding

sites for phytate at low pH (Cosgrove, 1966). Ternary complexes are rare but usually formed in

slightly alkaline environment in the intestine (Riche et al., 2001).

Most studies on dietary fibre from various plant sources dealt mainly with inclusion

levels and their effects on nutrient utilization and growth of an animal such as swine (Laplace et

al., 1989; Mitaru et al., 1984; Souffrant, 2001), poultry (Slominski and Campbell, 1990;

Villamide and San Juan, 1998), or fish (Anderson et al., 1984; Bromley and Adkins, 1984;

4
Hemre et al., 1995; Kraugerud et al., 2007). Recent studies have shown that soluble such as

pectin and insoluble fibre components such as cellulose and hemicellulose affect digestive

processes of an animal (Amirkolaie et al., 2005; Glencross, 2009; Ovrum-Hansen and

Storebakken, 2007; Zahedifar et al., 2002; Zhu et al., 2005).

An ingredient could have low ANF contents, excellent digestible amino-acid profile, and

the availability of the amino acids could be very good but would not be appropriate for inclusion

if it negatively affects the physical quality of the manufactured feed pellets. Physical properties

of feed pellets are of great importance in aquaculture. Bulk density, hardness, fines, undersized

pellets, sinking velocity and water stability are all pellet quality parameters important to

aquaculture feed manufacturers and fish producers (Aas et al., 2011).

Water stability of a pellet is a very important characteristic (Cho and Bureau, 2001;

Obaldo et al., 2002). Water stability is known to be affected by feed composition, notably by

soluble fibre (e.g., guar gum, Brinker, 2007; Brinker and Reiter, 2011; Storebakken, 1985) or

insoluble fibre (e.g., cellulose, Bromley and Adkins, 1984; Hansen and Storebakken, 2007;

Lupatsch et al., 2001) composition playing important roles. Level and composition of the fibre

components in the plant protein ingredients alter the physical properties of the mash and pellets

during processing. However, few studies have systematically examined the effect of different

various fibre components on physical properties of aquaculture feed (e.g., Baeverfjord et al.,

2006; Glencross et al., 2007; Hansen and Storebakken, 2007; Kraugerud et al., 2011), and

criteria for evaluation of these properties are not well-established (e.g., Glencross et al., 2007;

2010).

1.1. Objectives

5
The main objective of this thesis is to set up criteria for comprehensive evaluation of

plant protein sources encompassing amino acid digestibility and bioavailability, effects of anti-

nutritional factors and fibre components on nutrient utilization, and assessment of physical

properties of feed pellets with inclusion of a novel ingredient. Two novel plant protein

ingredients, Indian mustard protein concentrate and Indian mustard protein meal with variable

fibre and phytic acid contents were selected as test ingredients and compared with a commercial

soy protein concentrate commonly used in commercial rainbow trout and Atlantic salmon feeds.

Bioavailability of arginine from these three ingredients to rainbow trout was evaluated because

of the high concentration of this amino acid in these ingredients. Phytic acid and lignin,

respectively were chosen to assess the effects of ANF and a fibre component, and their

interactions on growth, utilization and partitioning of nutrients, and physiological response of

rainbow trout because these two types of ANF were present at variable concentrations (from low

to very high) in the test ingredients. Finally, to establish criteria to assess physical properties of

aquafeed, graded levels of a purified lignin were used to determine its effect on oil and water

absorption of a dietary mash, and physical properties of steam-pressed pellet including the water

stability.

6
CHAPTER 2

REVIEW OF LITERATURE

2.1. Introduction

Aquaculture is the fastest growing food-producing sector in the world supplying more

than half (37 MMT) of the worlds food fish production in 2009. About 56 million metric tons

(MMT) of finfish and shell-fish valued at US$ 105 billion were produced in 2010 (FAO 2011;

Bostock et al., 2010). More than 200 species of fish, crustaceans and molluscs are cultivated

around the world, but only 28 species together contributed more than two-thirds of the world

production in 2010, a majority of which is from six species of carps (about 16 MMT), followed

by tilapia (2.5 MMT) and salmonid (2.4 MMT) species (FAO, 2011).

The share of farmed fish and crustacean production fed compounded feed worldwide has

risen from about 44% in 1995 to 68% in 2010 (Tacon and Metian, 2008). Fishmeal, long been

considered as major component in compounded feed, has transformed from the major protein

source to a strategic protein ingredient. Currently, fish meal is still used as a major protein source

in larval feed and in some marine fish feed, contribution of which gradually decreases or

diminishes as the fish grow. However, despite decreasing inclusion in the feed of marine shrimp,

marine fish and salmonids over the past few decades, they remain the major consumers of the

fish meal (Naylor et al., 2009; Tacon and Metian, 2008).

The rising demand for compounded feed, and limited supply and rising cost of fish meal

have encouraged investigation on the nutritive value to fish of different protein-rich ingredients

over the past few decades. Various factors such as amino acid composition, anti-nutrients, and

non-nutritional components can affect the nutritive value of protein-rich ingredients and the

7
amount of these ingredients that can be used in the diets of various fish species (Bell 1993;

Bureau et al., 1998; Francis et al., 2001; Higgs et al., 1995; Mwachireya et al., 1999; Refstie et

al., 2001; Sanden et al., 2005).

Known nutritional limitations such as poor amino acid profile, heat-induced chemical

damage to some amino acids during processing, the presence of ANF and non-nutritional fibre

components are major challenges for increasing the inclusion of many economical protein

sources. Lack of information on available amino acids and other nutrients in ingredients forces

fish feed formulators to rely on high safety margins increasing the economical and

environmental costs. Accurate information on chemical composition, digestibility and

bioavailability of amino acids would support more economical feed formulation. Another

challenge is associated with the anti-nutritional compounds in plant proteins (Francis et al., 2001;

Tacon, 1997). Although many are destroyed or inactivated during feed processing, some remain

and could negatively affect digestibility of nutrients in fish. These include phytic acid,

glucosinolates, saponins, tannins, soluble NSP, and gossypol. In addition, the physical

characteristics, such as water stability and pellet durability, of feed pellets are very important to

prevent the economical and environmental loss caused by the increased amount of feed needed to

be used and wastes (Cho and Bureau, 2001; Obaldo et al., 2002).

The main focus of this literature review is to provide an overview of the state of

aquaculture nutrition and feed production, introduce some common and novel plant protein

ingredients potentially valuable for aquaculture feed production and to review strategies for

evalution of plant protein ingredients, with a focus on digestibility and bioavailability of

nutrients in fish, and the physical quality of feed pellets.

8
2.2 Review of Aquaculture Feed Production and Aquaculture Nutrition

World industrial feed output for farmed terrestrial and aquatic animals has exceeded 700

MMT in 2009 from 600 MMT in 2003 (Best, 2009). The share of aquafeed to overall production

is very low (~4%) but has increased more than 60 percent during this period. Global industrial

aquafeed output has increased from 14 MMT in 2001 to about 17 MMT in 2008, an estimate that

doesnt account for the production by feed manufactures (feed mills) with annual capacity of less

than 2400 MT (Best, 2009). Considering this output, the real global aquafeed production is

closer to 37 MMT (Rana, 2009). Several factors such as intensification of the existing farming

systems, increasing culture of high trophic level and high market value finfish species, and

growing reliance on compounded aquafeed could be responsible for the phenomenal growth

experience for aquafeed production (Tacon and Metian, 2008; Tacon et al., 2010). This is

particularly evident in that since 1995, the amount of compounded feed used in finfish farming

has been increasing at a higher rate (2.5 times higher) than that of the growth of the aquaculture

industry (Tacon and Metian, 2008; Figure 2.1).

2.2.1. Evolution of salmonid feed formulation

Global output of salmonids in 2009 was 2.5 MMT, an increase of 67% from its 2000

production of 1.5 MMT (FAO, 2011; Figure 2.2). Atlantic salmon, Salmo salar, are the most

widely farmed salmonid species globally (1.44 MMT), followed by rainbow trout, Oncorhynchus

mykiss (0.7 MMT), Chinook salmon, O. tshawytscha and Coho salmon, O. kisutch (0.19 MMT).

More than half (1.4 MMT) of the worlds salmonid production comes from the Europe (mainly

Norway and Scotland) and about a third is produced in the Americas (mainly Chile and Canada)

(Figure 2.3).

9
Figure 2.1 The world production of fish and shrimp (*excluding filter feeding fishes, dark
grey), and the fraction produced fed compounded feed (in %, light grey) since 1995 (data from
Tacon and Metian, 2008).

80 Production of fish and shrimp (MMT)*


% produced fed compounded feed 68%
70
61%
60 54%
50 44%
40
31.6
30 22.3
20 14.9
9.1
10
0
1995 2000 2005 2010
Year

10
Figure 2.2 Global output of farmed salmonids in weight (million metric tons MMT) and

value (billion USD) (FAO 2011).

3 Production (MMT) Value (in billiion dollars) 14


World production of salmonids

World salmonid production in


12
2.5

value (billion $)
10
(MMT)

2 8

6
1.5
4

1 2
2000 2001 2002 2003 2004 2005 2006 2007 2008 2009
Year

11
Figure 2.3 Production of farmed salmonids (in thousand metric tonne) in continental America,

Asia, and Europe in 2008 and 2009 (FAO, 2011).

1600
1400 2008 2009
Production ('000 tonne)

1200
1000

800
600
400
200
0
Americas Asia Europe
Continents

12
In 2006, annual world production of salmonid feeds was estimated at about 2.6 MMT

worldwide (Tacon and Metian, 2009) up from 1.9 MMT in 2000 (Hardy and Tacon, 2006). Over

a period of six decades, formulations of salmonid diets have evolved from wet-mixture of locally

available ingredients such as salmon eggs, fresh or frozen fish, livestock offal, and oilseed meals

(from 1850 to 1950) to the modern compounded extruded feed used today (Grassl, 1956; Hardy

and Barrows, 2002). Table 2.1 provides a chronological overview of the changes in ingredient

composition, advances in manufacturing practices and nutrient composition of salmonid feeds.

Diet formulation and preparation are the process of combining feed ingredients to form a

mixture that will meet the specific goals of production. It is often a compromise between ideal

and practical considerations. The primary objectives are to produce a mixture that (is):

1. Nutritionally balanced (to support maintenance, growth, reproduction, health)

2. Economical, Palatable, Water stable

3. Minimizes waste output & effect on water quality

4. Produces desirable final product (attractive & safe)

5. Practical considerations:

i. Ingredients price and availability

ii. Anti-nutritive factors

iii. Pelletability of mixture

iv. Storage and handling requirements

13
Table 2.1 A chronological account of the evolution of salmonid feed type, feed ingredients,
crude protein and lipid contents since 1920.

Year Feed type Major ingredients Crude Crude Reference


Protein lipid
1920- Moist pellets Salmon eggs; slaughter house <10% <10% Hardy and
1940 (>50% water) waste, beef-hog liver, salmon Barrows, 2002
viscera etc.
1940- Semi-moist Meat-meal mixtures, blend of 10-20% 10-15% Phillips et al.,
1955 pellets slaughter-house wastes and 1940
commercial ingredients.
1956- Dry-sinking Mostly fish meal, animal by- 35-45% <15% Grassl, 1956;
1980 steam pressed products, soybean meal, corn gluten Hilton et al.,
pellets meal, vitamin and/or mineral 1981
premix.
1990- Dry-floating Mostly fish meal, animal by- 35-50% >30% Hardy and
2000 extruded products, plant protein meals, salmon Barrows, 2002
pellets vitamin and mineral premix. >20%
trout
2000- Dry-floating Reduced fishmeal content, soybean 35-50% >35% Rana et al., 2009
extruded meal and protein concentrate, maize salmon
pellets and wheat gluten, blood meal, meat >30%
meal, gelatin, brewers yeast. trout

14
2.2.2. Evolution of fish meal usage in aquaculture feed

Fish meal has long been considered preferred sources of proteins in the feed of most

farmed crustaceans, most marine fishes, and other intensive farming practices that need a high

quality diet (De Silva et al., 2011). Fish meals are high in small bioactive compounds such as

choline, taurine, anserine etc. These compounds play important roles such as stabilizing protein

structures, protecting cells against osmotic stresses or preventing oxidative damage and

stimulating immune system in swine, poultry and fish (Kuzmina et a;., 2010; Yun et al., 2011).

Despite the effort to decrease fishmeal inclusion rate, aquafeeeds shall utilize about 68% (3.7

MM) of reported world production of fish meal (Tacon and Metian, 2008).

Feed for shrimp (27%), marine fish (20%) and salmonids (14%) account for significant

proportion of fish meal consumption worldwide (Tacon and Metian, 2009). Improved

understanding of the nutritional requirement of farmed fish and shrimp, and better

characterization of alternative ingredients have allowed a significant reduction in fish meal and

fish oil usage in the diets of shrimp, salmon and marine fish over the past decade (Naylor et al.,

2009; Tacon and Metian, 2009). Today, fish meal has become more of a strategic ingredient.

Starter and fry feeds for salmonids and other marine species are generally formulated to supply

appropriate feed intake of the animals. However, fish meal levels are progressively reduced in

feed formulation once once fish reach juvenile and grow-out stages (Naylor et al., 2009)

2.2.3. Desired characteristics of a potential feed ingredient

An increasingly wide variety of feed ingredients are playing a role in the formulation of

compounded fish feeds (Gatlin et al., 2007; Hardy, 2010; Naylor et al., 2009; Sinha et al., 2011;

Tacon and Metian, 2008). Inclusion of any feed ingredient demands adequate supply, consistent

15
quality, competitive price per unit protein, accurate characterization of its nutrient and anti-

nutrient profiles, evaluation of the desired nutrients available to an animal (Bureau, 2008;

Glencross, 2009; Hardy, 2010; Kaushik and Seiliez, 2010), and its effect on physical properties

of the manufactured pellets (Sapkota et al., 2007).

2.3. Overview of World Feed Ingredients

Commonly used animal and plant protein products in fish feed are fish meal, poultry by-

product meal (PBM), meat and bone meal (MBM), soybean meal (SBM), soy protein

concentrates (SPC), and corn gluten meal (CGM). Rendered animal proteins such as poultry by-

product meal, meat and bone meal, feather meal and blood meal products are readily available

and economical sources of protein that have shown to be of relatively high nutritive value to fish

(Tacon et al., 2006). Table 2.2 reports the global or North American (NA) output of major plant

or animal protein-rich feedstuffs. A large variety of oilseeds and grains are cultivated worldwide.

Estimated production of protein meals from soybean, corn and Brassica oilseed (canola-mustard-

rape) globally in 2009-10 was about 300 MMT with soybean meal representing about 175 MMT

worldwide. Worldwide production of rendered animal protein ingredients is somewhat limited,

around ~12 MMT (WRO, 2010; Table 2.2).

16
Table 2.2 Major co-products from grain, oilseed and animal production industries available for

compounded feed production (estimated global output, in MMT).

Projected
Ingredient Year Reference
production

(MMT1)

Major Plant Proteins

Soybean meal 174.9 2010-11 USDA, 2011

Canola/rapeseed/mustard meal 33.5 2010-11 USDA, 2011

DDGS2 (in EU3) 6.0 2010 Ziggers, 2007

Corn DDGS (In NA4) 40.0 2010 OBrien, 2010

Corn Gluten Meal 30.5 2009 USDA, 2011

Major Rendered Proteins 13.0 2010 WRO, 2010

Poultry by-product meal (US) 1.5 2010 Swisher, 2007

Meat and bone meal (US) 2.5 2010 Swisher, 2007

Feather meal (US) 0.4 2010 Swisher, 2007

Porcine meal (US) 0.8 2010 Swisher, 2007

Aquatic Animal Protein (World) 11.0 2010 Tacon, 2010


1
MMT Million metric tons; 2DDGS Dried distillers grain solubles; 3EU European Union;
4
NA North America

17
2.3.1. Production of oilseed meals

Oilseeds are an important agriculture commodity that are generally processed in

vegetable oils and high-protein meals for food, feed, and industrial uses. The economic outcome

of oilseeds begins with separating oil and protein meal. The technology of oil separation has

progressed from primitive manual methods to chemical and mechanical methods using organic

solvents and extrusion technologies, respectively.

Solvent extraction is by far the most widely used method to extract oil where seeds are

prepared by cleaning, drying, tempering, cracking, dehulling, conditioning, and flaking. Oil is

extracted from the flakes using an organic solvent, usually a petroleum distillate high in n-

hexane followed (Johnson and Lusas, 1983; Woerfel, J.B., 1995). The advantage of this process

is high extraction efficiency, with the residual oil in the meal potentially reduced to <2%.

Detailed information of the various solvent extraction processes and the types of solvents has

been summarized by Johnson and Lusas (1983) and Woerfel (1995).

Two mechanical processes, extruding-expelling (EE) and screw-pressing methods are

also used separately or in combination for the extraction of oil (Wijeratne et al., 2004). The

extruding process involves shearing, compressing and heating of the flakes where heat is

provided only by the dissipation of the mechanical energy. The average residance time of the

material within the extruder barrel is less than 30 s. In combination, the extrudate is conveyed

continuously into a crew press, where the the residual oil is pressed out (Wijeratne et al., 2004).

2.3.2. Production of protein concentrates and isolates

Protein concentrates and isolates are produced by further processing of the protein meals.

Concentrates (65-70% CP) are primarily produced after removing the water soluble

18
carbohydrates from the meal. Isolates are highly purified form of proteins (>90% CP) made by

removing most of the non-protein components, residual fats, and carbohydtrates from the meal.

Plant proteins are composed mostly of three types of proteins: albumins (water soluble),

globulins (salt soluble) and glutelins (alkali soluble) (Prakash and Rao, 1986). Water soluble

albumins represent about 45-50% of the total proteins in brassica seeds where as in most other

oil seeds, including soybean, the most common protein is the globulins (Appleqvist, 1972).

Albumins can be isolated by an aqueous extraction method by mixing the sample with

water for 30-120 min and at low temperatures (e.g. 4oC) to prevent proteolysis (Neves and

Lourenco, 1995; Sathe et al., 2002). After centrifugation to remove insoluble materials,

supernatant is dialyzed to eliminate residual salt and precipitate salt-soluble proteins (e.g.,

globulins). A low molecular weight cut-off membrane (<8000) may be used during dialysis to

prevent the loss of proteins of interests (Sathe et al., 2002). The precipitate (salt-soluble

globulins) is then extracted using dilute solutions of NaCl, followed by centrifugation and

dialysis (at 4oC) of the supernatant agains watar.

Aqueaous alcoholic solutions (usually 50-60% ethanol) are also used to separate the

solubles e.g., sugars, odour and flavour substances from the insoluble residues (Uzzan, 1988).

The residue is desolventized at temperatures not exceeding 50-55oC in inert gas environment or

under vacuum to maintain protein quality. Proteins are also isolated using buffer solution of their

isoelectric points (pI), at which they are least soluble.

2.3.3. Further improvement of the protein quality

Further processing of plant proteins are often needed for improving their functional

properties such as protein solubility and removing toxic or inhibitory compounds (Panyam and

19
Kilara, 1996). The isolated proteins can be subjected to chemical, enzymatic or thermal

treatments in order to improve their functional and nutritional properties.

Chemical treatments involving alkylation, oxidation, acylation, esterification and amide

formation have shown to improve the solubility, and the emulsifying and foaming properties of

the proteins (Shih, 1987; Shahidi and Wanasundara, 1998; Sitohy et al., 1999). Enzymatic

hydrolysis such as trypsin treatment of protein bonds can improve protein solubility (Jones and

Tung, 1983). In addition, treatments using pepsin, papayin, tripsin or phytase have been reported

to remove some anti-nutritional factors such as protease inhibitors and phytic acid (Kaushik et

al., 1995; Moure et al., 2006; Murai et al., 1989).

2.3.4. Production approaches

Driven by market demands, the current trend in plant protein production is to develop

tailor-made products destined for a specific industry or for a particular species or life stage of a

species. Various techniques such as roasting, toasting, aqueous alcohol extraction, enzymatic

hydrolysis, thermal processing have been employed to produce various products including

protein meals, concentrate, isolate, and various specialty products texturized and low-antigen

protein concentrate, or enzyme modified protein isolate (Figure 2.4).

20
Figure 2.4 Flow-chart of oil-seed processing showing the techniques used to produce various soy products for consumption (using

soybean as an example).

21
2.3.2. Rapeseed/canola/mustard (Brassicacae)

World production of canola/rapeseed/mustard of the Brassicacae family is currently at 58

MMT contributing about 15% to the worlds oilseed production, second only to soybean at 211

MMT (USDA, 2011). Six Brassica species are cultivated worldwide, three of which contain

diploid genomes: Brassica rapa (Turnip), Brassica nigra (Black mustard), and Brassica

oleracea (broccoli, cauliflower, cabbage). The other three species were inter-specifically bred

from the three diploid species. They are: Brassica juncea (Indian mustard) crossed between B.

rapa and B. nigra, Brassica napus (rapeseed) between B. rapa and B. oleracea, and Brassica

carinata (Ethiopian mustard) between B. oleracea and B. nigra.

Combined world production of protein meals from the Brassicacea seeds have recently

reached 34 MMT in 2010-11 (Table 2.2), and mainly utilized in animal feed (Newkirk et al.,

1997; Thacker and Petri, 2010). New cultivars of Brassica napus (dubbed as canola) and B. rapa

are low in glucosinolate (<30 mol g-1 of oil-free dry matter) and erucic acid (<2% of total fatty

acids in the oil) (Higgs et al., 1982).

2.3.2.1. Indian mustard (Brassica juncea)

Cultivation of another Brassica species, Indian mustard (B. juncea), is gaining popularity

worldwide because its drought resistance attributes are superior to those of the canola cultivars

(Chevre et al., 2008; Oram et al., 1999; Wijesundera et al., 2008). Some Indian mustard varieties

yield more seed per pod, produce heavier seeds, and contain higher dry matter (DM) in the seeds

than the canola varieties in drier conditions (Chauhan et al., 2007; Gunasekera et al., 2006;

Wright et al., 1995). Low drought susceptibility index (a combination of growth, maturation

22
(days), number of pods and seed yield per pod, and DM content), and low-glucosinolate and

erucic acid levels of the newer Indian mustard varieties developed in Australia, India and North

America have raised the potential of growing this crop in the drought-prone regions of the world.

2.3.2.2. Nutritive value of mustard products: issues and challenges

The development of B. juncea cultivars with oil and nutritional characteristics similar to

canola has enhanced the potential of the co-products of this oilseed in animal feeds (Burton et al.,

2004). Amino acid profiles of Indian mustard protein meals and concentrates are very good and

comparable to similar soybean products. Macro-minerals such as calcium and phosphorus are

also in high concentration. The nutritive value of low glucosinolate Indian mustard meals (Jia et

al., 2008; Meng et al., 2006; Newkirk et al., 1997) and Indian mustard protein concentrates and

isolates (Thacker and Petri, 2010) have been evaluated in the diet of poultry. These studies

indicated that the nutritional value of meals and concentrates from low glucosinolate Indian

mustard was equal or superior to that of canola meal samples derived from B. napus and B. rapa

cultivars.

The lower protein and high levels of fibre (indigestible carbohydrates) of Brassica

cultivars by-products limit the potential of these ingredients in high digestible nutrient density

salmonid fish feeds (Carter and Hauler, 2000; Hardy, 2010). Different processing techniques

have been developed to reduce the levels of indigestible carbohydrate and to produce high

protein ingredients from a variety of plant co-products (Bureau and Cho, 1997; Glencross et al.,

2010; Hardy, 2010).

23
2.4. Nutritive Value of Protein Ingredients to Fish

Accurate assessment of nutritive value of feed ingredients is extremely important for the

formulation of cost-effective feeds. Better characterization of the nutrient composition of

feedstuffs is essential to improve their valorization by least-cost feed formulation programs.

More efforts need to be invested in systematically investigating the effect of numerous factors

that can affect the nutritive value of feed ingredients.

Diets produced with poor quality raw materials may have inferior nutritive value and

have adverse effects on production and fish health. Quality criteria for the ingredients must be

respected to insure that the final product is of consistent quality and that deleterious effects are

avoided. The chemical composition (nutrients, anti-nutritional factors, contaminants) usually

provides a snapshot of the potential nutritional value of an ingredient, but the biological aspects,

such as digestibility and availability of nutrients are highly important and yet often overlooked.

Nutritional quality of any ingredient can vary significantly by the sources of raw

materials and processing equipment, techniques, and conditions. Processing techniques such as

expeller or solvent extraction (Glencross et al., 2004b), extraction temperature and duration

(Fenwick et al., 1986), and drying conditions (Glencross et al., 2007a; Stewart et al., 2003) may

affect the nutritive values of oilseed processing co-products. For example, during alkaline

treatment, lysinoalanine is created by the addition of the -NH2 group of lysine to the double

bond of a dehydroalanine residue (Friedman, 1992). Racemization of some amino acids also

occurs during alkaline processing. Heat treatment required during processing may damage some

amino acids through oxidation of sulphahydryl bonds (-SH) to disulfide bonds (S-S), non-peptide

cross linking, Maillards reactions, and oxidative degradation (Moughan and Rutherfurd, 1996;

Moughan and Rutherford, 2008).

24
2.4.1. Overview of methodological approaches

Successfully formulating more economical feeds with cheaper protein sources requires

careful consideration of the nutritional quality of these ingredients to ensure that the performance

and health of the animals and final product quality are not being sacrificed. Digestibility of

nutrients is currently the standard in ingredient and diet evaluation. One option is to simply

assay proximate components and their respective digestibility. However proximate components,

such as crude protein, are not nutrients per se. Crude protein is only a measure of the total

nitrogen (N) content of the feed and is not necessarily an indication of that components

usefulness to the animal. A few studies have focused on the composition and digestibility of

amino acids in common fish feed ingredients (Bureau, 2008; De Silva et al., 2000; Kaushik and

Seiliez, 2010).

Digestibility of amino acids is not always a reliable indicator of the availability of these

nutrients to animals, notably for feed ingredients processed under high heat and pressure

conditions, such as rendered animal protein ingredients. More direct methods of assessment of

the available nutrient contribution of different ingredients to the diet should also be used. One

method of assessing this is through the use of one of a number of bioavailability assays

whereby the effect on growth, protein or nutrient deposition, or other parameters of interest, are

examined with inclusion of the test ingredient at a number of levels and compared to the

response of animals fed diets containing corresponding levels of the component of interest (e.g.

the studied amino acid in a free form (crystalline amino acid) in the case of amino acid

bioavailability studies). In this way, a more accurate assessment of the proportion of an

ingredient which is able to be effectively utilised by the animal can be made.

25
2.4.2. Digestibility

Morrison (1936) stated that digestibility trials provide the general basis for our

knowledge concerning the amounts of digestible nutrients furnished by the many different

feeding stuffs. Measurement of digestibility provides, in general, a good indication of the

availability of energy and nutrients, thus providing a rational basis upon which diets can be

formulated to meet specific standards of available nutrient levels.

2.4.2.1. Digestive process related to proteins and amino acids

Proteins are digested, absorbed and metabolized through a series of processes. In

vertebrates, proteins and polypeptides are broken down to mono-, di-peptides and amino acids

through reactions catalyzed by enzymes mostly secreted in the stomach. The polypeptides and

amino acids are then absorbed in the intestinal brush-border membranes by several active

transport systems and then, to the hepatic portal system through passive diffusion mechanisms

(Bender, 2008). Besides dietary sources, some amino acids and peptides in the intestinal chime

are also derived from the degradation of proteins. These degraded proteins are either reabsorbed

or excreted as components in faeces along with the unutilized dietary amino acids (Figure 2.5).

A detailed definition of the digestibility of AA was provided by Fuller (2003) as a

mechanism that reflects enzymatic hydrolysis and microbial fermentation of ingested proteins

and peptides and absorption of AA and peptides from the gastrointestinal lumen. Digestion and

absorption of nutrients follow closely similar pathways in the digestive system of all vertebrates.

All vertebrates have a digestive tract and associated glands but different parts are not necessarily

directly comparable among the groups. In fish, length and structures of the gastro-intestinal tract

26
(GIT) vary from species to species depending on their food and habitat types (Horn and

Gawlicka, 2001).

Functions of the stomach involve grinding and breaking down large feed particles by

mechanical (contraction-expansion) and chemical processes (pepsinogen- and acid-secreting

functions). In the stomach of fishes, only one secretory cell type exists within the gastric glands

performing both pepsinogen- and acid-secreting functions unlike the differentiation into separate

cells by functions observed in mammals (Wilson and Castro, 2011).

In the intestine, polypeptides are digested to smaller peptides and by several enzymes

derived from pancreas or secretory cells of the intestinal epithelium in slightly alkaline

environment achieved by pancreatic secretion of bicarbonates and bile acids from the gall

bladder (Deguara et al., 2003; Fard et al., 2007) (Table 2.3). The absorption of nutrients occurs in

the intestine by optimizing the intestinal surface area within the constraints of the coelomic

cavity. Mucosal folding and amplification of apical plasma membrane aid considerably to

increase the intestinal surface area. The major cell types in intestine of all fishes are enterocytes,

mucous cells, enteroendocrine cells, rodlet cells (in most teleosts), and ciliated and zymogen

cells (in elasmobranchs) (Wilson and Castro, 2011).

27
Figure 2.5 A conceptual diagram of digestion and absorption of dietary proteins in intestine of

fish.

Absorbed amino acids available for synthesis

Peptides Free amino acids

DIETARY FAECES
PROTEINS
Endogenous
Absorption
Loss

Endogenous Amino Acids

Proximal Intestine Distal Intestine

28
Table 2.3 Major enzymes in fish for the digestion of protein, peptides, amino acids and other

non-nitrogenous compounds.

Source Enzyme Substrate Specificity or products


Stomach Pepsins (pepsinogens) Proteins and Peptide bonds adjacent to
polypeptides aromatic AA
Exocrine Trypsin (trypsinogens) Proteins and Peptide bonds adjacent to
pancrease polypeptides arginine or lysine
Chymotrypsins Proteins and Peptide bonds adjacent to
(chymotrypsinogens) polypeptides aromatic AA
Elastase (proelastase) Elastin, some Peptide bonds adjacent to
other proteins aliphatic AA
Carboxypeptidase A Proteins and Carboxy terminal AA with
(procarboxypeptidase A) polypeptides aromatic or branched aliphatic
side chains
Carboxypeptidase B Proteins and Carboxy terminal AA with
(procarboxypeptidase B) polypeptides basic side chains
Ribonuclease RNA Nucleotides
Deoxyribonuclease DNA Nucleotides
Intestinal Enterokinase Trypsinogen Trypsin
mucosa
Aminopeptidases Polypeptides N-terminal AA from peptide
Dipeptidases Dipeptides Two AA
Nucleases Nucleic acids Purine and pyrimidine bases
Cytoplasm of Various peptidases Di-, tri- and AA
mucosal cells tetrapeptides
Adapted from: Ganong, 2009; Krogdahl and Sundby, 1999; Lauff and Hofer, 1984;

29
2.4.2.2. Assessing digestibility in fish: techniques, issues and limitations

The loss of indigestible matter from the diet as faeces is one of the primary reasons for

variations in the nutritional value of feed ingredients. Measurement of digestibility provides, in

general, a good indication of the availability of energy and nutrients, thus providing a rational

basis upon which diets can be formulated to meet specific standards of available nutrient levels.

In simple terms, digestibility is the measure of disappearance of nutrients in the digestive tract

usually estimated by the difference between the ingested nutrient and the amount of nutrients in

faeces. Various in-vivo and in-vitro methods to determine digestibility of nutrients in fish have

been developed over the years. The in-vivo methods mainly differ in faecal collection

approaches, use of indigestible marker, diet composition, level of test ingredients etc. (e.g.

Bureau and Hua, 2006; Bureau et al., 1999; Cho and Slinger, 1979; Cho et al., 1982; Forster,

1999).

2.4.2.2.1. Fecal collection methods

Meaningful estimation of digestibility relies on collection of fecal material using

adequate approaches. The most widely used technique is the collection of faeces from the lower

part of the intestine by stripping, suction or intestinal dissection (Austreng, 1978; Nose, 1967).

Each of these fecal collection methods has its own advantages and disadvantages. Major

shortcomings of stripping or dissection are: (a) removal of chime before complete digestion and

absorption; (b) contamination of faeces with blood and sloughed mucosal cells; and (c) handling

stress. It generally leads to an underestimation of the digestibility of nutrients (protein in

particular) because of contamination of feces with endogenous material that would otherwise be

reabsorbed before the feces are excreted (Cho et al., 1982).

30
Passive methods for the collection of faecal material yield more meaningful estimates of

digestibility of nutrients, if used correctly. These include the system of Ogino et al. (1973) in

which the feces are collected by passing the effluent water from the fish tanks through a filtration

column (TUF column), the system of Cho and Slinger (1979) in which a settling column is used

to separate the feces from the effluent water (Guelph system) and that of Choubert et al. (1979)

in which a mechanically rotating-screen is used to filter out fecal material (St-Pee system).

Vandenberg and de la Noe (2001) compared apparent digestibility coefficients (ADC) in

rainbow trout among three different faecal collection methods that include continuous filtration

and collection (St. Pee system, Choubert et al., 1979), settling column (Cho and Slinger, 1979)

and stripping by massaging the ventral surface from the anal fin towards the anus. The authors

reaffirmed that collection of faeces involving physical interactions with the animal such as

stripping results in lower ADC values vs. both passive faecal collection methods.

2.4.2.2.2. Indigestible markers

The standard approach in the indirect digestibility assessment for any fish species is the

incorporation of one or more inert compound as indigestible marker in the test diets. These

compounds are assumed to be homogeneously incorporated into the feed, do not interfere with

the digestive metabolism of the animals, never absorbed or utilized, analysed accurately in feed

and feces, and non-toxic and harmless to animal and human (Austreng et al., 2000).

Chromium oxide (Cr2O3) has been the most commonly used inert marker, but

digestibility values obtained using Cr2O3 are questionable because of shortcomings such as low

recovery in faeces (Fernandez et al., 1999; Lied et al., 1982), effect on intestinal microflora

(Ring, 1993), and different passage rate than the nutrients or the digesta (Dabrowski and

31
Dabrowska, 1981; Tacon and Rodriguez, 1984). Several other inert markers such as titanium

oxide (TiO2 Lied et al., 1982), polyethylene (Tacon and Rodrigues, 1984), hydrolysis resistant

ash and acid washed sands (Atkinson et al. 1984; Tacon and Rodrigues, 1984), have been

evaluated in fish with variable results.

Rare earth metal oxides such as yttrium oxide (Y2O3) or ytterbium oxide (Yb2O3) are

increasingly preferred in digestibility studies. They do not affect nutrient metabolism in fish

because of the very low concentration needed in the diets (at mg g-1 level) and can be easily

detected by plasma emission spectroscopy and by atomic absorption, thus, are measurable at a

very low concentration (Davies and Gouveia, 2006).

2.4.2.2.3. In-vitro assessment of digestibility

In the last few decades, considerable attention has been given towards in-vitro

digestibility assessment with enzymatic or acid digestion (Bassompierre et al. 1998, Carter et al.,

1999; Dimes and Haard, 1994; Dimes et al., 1994; Haard, 1992; Rungruangsak-Torrisen et al.,

2002). These assays are quick and useful but significant variations exist in the source and

number of enzymes, use of chemicals and environmental factors such as pH and temperature.

Another in-vitro method, the everted sleeve technique developed by Buddington et al.

(1987) and adopted in several studies (Berge et al., 1999; Rosas et al., 2008), have become a

useful technique that attempts to imitate the nutrient transport mechanisms in the gastro-

intestinal tracts. A drawback of this method is as solute concentration increases in the tissue, the

rate of uptake decreases or reaches a plateau and fails to provide accurate digestibility values

when large quantity of nutrients are available. Recently, a two-step hydrolysis method is

developed to address this issue using sequentially closed reactor for acid digestion and semi-

32
permeable membrane reactor for alkaline digestion, and at the same time, continuously removing

the solutes from the reactor (Hamdan et al., 2009; Morales and Moyano, 2010).

2.4.2.2.4. Limitations of digestibility of protein in feed ingredients

Digestibility gives a limited view of the nutritional value of an ingredient, and provides

no indication of the proportion of a nutrient used for growth and other metabolic functions (Stein

et al., 2007). Digestibility is not always a reliable indicator of the availability of proteins to

animals, notably for feed ingredients processed under high heat and pressure conditions.

Digestibility calculated by the difference between intake and fecal excretion doesnt

count for the amount from endogenous loss or microbial production. Endogenous loss of AA

depends on dietary ingredients (e.g., composition of soluble and insoluble fibres in the

ingredients, presence of anti-nutritional factors) and physiological state of the animal. In

addition, obtained values are based on the disappearance of nutrients from the digestive tract that

do not reflect the net breakdown or synthesis in the intestinal lumen or the form in which they

are absorbed. This is specifically an issue with ingredients that have undergone extensive thermal

processing. In heat-treated feed ingredients, some AA, such as lysine, may be present in

chemical forms, like Maillard reaction products, that may be absorbed but are metabolically

inactive and preclude their utilization for protein synthesis (Moughan and Rutherfurd, 1996).

2.4.3. Bioavailability of amino acids

Apparent digestibility of any nutrient is a measure of the disappearance of that nutrient

from the digestive tract, defined as that present in the feed but not in the fecal material (Hepher,

1990). This missing component is then considered to have been absorbed and utilized by the

33
animal which could be a misrepresentation of the true value of the ingredient because the

nutrient may be in a form able to be absorbed by the animal but may remain unavailable for

growth and other metabolic processes (Waibel et al., 1977). This has been observed in some

amino acids, such as lysine, the most chemically reactive amino acid, containing an -amino

group that is susceptible to chemical modification during processing and prolonged storage

(Moughan and Rutherfurd, 1996). These groups likely bind to others, forming structures which

resist chemical reaction, making them less available for metabolism.

Simple growth trials, where test ingredients replace fish meal or other established

ingredients in the diet and the subsequent effect on growth is monitored, have similar problems

with their value and replication due to interactions with other ingredients (Lewis and Bayley,

1995). For example, a diet containing a specific ingredient, or level of an ingredient, may meet

the nutritional requirements of an animal when some nutrients are supplied by certain other

ingredients (such as highly digestible essential amino acid content ingredients like fish meal) but

may not show the same effect if the diet was formulated with alternative ingredients. These

alternative ingredients may provide apparently adequate levels of nutrients but in forms which

are less effectively utilised (Regost et al., 1999).

A more relevant approach is to consider the available nutrient contribution of the test

ingredients to the diet. In this way, a more accurate assessment of the levels of each essential

nutrient in that ingredient actually usable by the animal can be undertaken, leading to more

confidence in the nutritional adequacy of the diet and less wastage of expensive ingredients

currently provided in excess to ensure nutrient requirements are met. In order to assess the

nutrient contribution of an ingredient to the diet, one must first determine the level of that

nutrient able to be used by the animal for metabolic processes (termed bioavailability).

34
Bioavailability is defined as the proportion of ingested dietary AA that is absorbed in a chemical

form that these AA potentially suitable for metabolism or synthesis (Ammerman et al., 1995;

Batterham, 1992; Lewis and Baily, 1995; Stein et al., 2007).

2.4.3.1. Methodological approaches

Traditionally, estimates of bioavailability of AA have been obtained using slope-ratio

assays. Slope-ratio assays provide direct and practical assessment of the availability of individual

amino acids and have been used to assess the bioavailability of individual amino acids for

animals (Lewis and Bailey, 1995). Batterham et al. (1979) first proposed the slope-ratio assay to

determine availability of a single amino acid, lysine in various plant proteins for growing pigs

and rats. Major and Batterham (1981) applied similar approach to determine availability of AA

in protein concentrates in chicks. A properly designed slope-ratio assay based on basal diet and

supplementation with a test protein source usually generate meaningful estimate of the

bioavailability of essential AA and other nutrients (Lewis and Bayley, 1995).

Several statistical approaches such as linear and non-linear regression approaches have

been proposed. For example, relative bioavailability of a nutrient from a test source compared to

the standard is determined from the ratio of the slopes of a linear regression or in the case of an

exponential regression, from the slopes of the steepness coefficients.

2.4.3.1.1. Slope-ratio assay linear regression models

Slope-ratios are usually determined by a linear regression model (y = b1 + b2*S + b3*T,

S - standard and T - test diets) that requires pre-validation of three assumptions: linearity of the

responses, common intercepts for all lines, and the equal basal level response to the common

35
intercept (Batterham et al., 1979; Finney et al., 1951, 1964; Littel et al., 1997). Fundamental

requirements for validity of a linear assay as described by Finney (1971) are: (1) common point

of intersection of the regression lines for standard and tests; and (2) the Y value of the common

point of intersection and the Y value of the mean response measured at X = lowest level or blank

level are not significantly different.

Linear models have been extensively used to determine bioavailability of individual

amino acids or minerals in poultry (Adeola, 1998; Batterham et al., 1986; Lumpkins and Batal,

2005; Major and Batterham, 1981; Parsons et al., 1998), swine (Adeola, 1996, 2009; Batterham

et al., 1990; Fanimo et al., 2006; Mateo et al., 2007; Petersen et al., 2011; Shoveller et al., 2009),

and recently, in fish (El-Haroun and Bureau, 2007; Li et al., 2008, 2009; Poppi et al. 2011;

Zarate and Lovell, 1997).

2.4.3.1.2. Non-linear approaches

In case of non-linear responses, an exponential regression model (y = b1 + b2*(1 exp

(b3*S + b4*T) can be used to estimate bioavailability from the ratio of the steepness coefficients

b3 and b4, where b1 (the response to basal diet) and b2 (difference between the lowest and

highest response) have to be equal (Kim et al., 2006; Littel et al., 1997; Liu et al., 2004).

Broken-line models can also be used to determine the RBV of the nutrient of interest where the

slopes of the linear portion of the responses are used (Coon et al., 2007; Wiltafsky et al., 2009).

2.4.3.2. Limitations of slope-ratio approaches

Like other assays, slope-ratio assays also have some limitations and are sensitive to the

experimental artefacts (Poppi et al., 2011). Two important dietary considerations for slope-ratio

36
assays are that the response to changes in test amino acid intake must be predictable and the

observed response must not be influenced by other dietary nutrients in the test feed ingredient

(Batterham, 1992; Littel et al., 1997). The latter condition is difficult to achieve with most

ingredients since most protein sources are complex ingredients containing variable levels of

various components (non-starch polysaccharides, anti-nutritional factors). Besides, the

composition and source of amino acids in the diets also play a role in feed intake and nutrient

utilization. Meaningful comparison of RBV estimates between different sources can be difficult

when variable composition of other nutrients or anti-nutrients in the sources aside from the

amino acid of interest affect feed intake, nutrient digestion and utilization.

2.4.4. The presence and effects of ANF and other non-nutritive components

2.4.4.1. General overview

Anti-nutritional factors (ANFs) are defined as innate components of an ingredient with a

limiting effect on feed intake, digestion, nutrient absorption, and well-being (Francis et al.,

2001). Plant protein ingredients contain a wide variety of ANF, most commonly protease

inhibitors, phytic acid, lectins, gossypol, antivitamins, allergens, saponin etc. (Tacon, 1997).

Among these, phytic acid (PA) and its salts, phytates have been identified as the major anti-

nutritional factors affecting nutrient utilization in farmed fishes with increasing inclusion of plant

protein ingredients in aquafeed (Francis et al., 2001). Common anti-nutritional factors in soybean

meal and their effects on salmonids are presented in the Table 2.4.

37
Table 2.4 Major anti-nutrients in soybean meals and their effects on salmonid fishes

Anti-nutrient Types of effects References

Trypsin inhibitors - bind with trypsin and Krogdahl and Holm, 1983;

chymotrypsin de Haard et al., 1996;

- smaller fish are more sensitive Murai et al. 1989

Lectin or - effects intestinal function Buttle et al., 2001

hemagluttanins - binds to the enterocytes lining the

intestinal villi

Antigenic proteins - effects non-specific defense Rumsey et al., 1994

(glycinin or glycinin) mechanisms

Saponin (glucosides) - palatability, feed intake Bureau et al., 1998

Phytic acid - overall depression in growth and Numerous studies

nutrient availability

Flavonoids (glycinin, - affects the synthesis of estrogens Pelissero et al., 1991

daidzin, genistin) - endocrine disruption, gamete

quality

- may have some positive effects

38
Other non-nutritive components in plant ingredients include indigestible starch and non-
starch polysaccharides. Dietary fibres are the sum of plant polysaccharides and lignin not
hydrolysed by endogenous enzymes (Kritchevsky, 1988; Wenk, 2001). Polysaccharides are
complex polymers of sugars occur in structural variations differing in side-chain types and the
distribution of glycoside linkages in the macro-molecular backbone (Figure 2.6). For example,
two pectins can differ in structure based on their composition of the smooth region composed of
apigalacturonan, homo-galacturonan, and xylo-galacturonan, and hairy chain ramified regions,
such as rhamno-galacturonan-I, and rhamno-galacturonan-II (Paulsen and Barsett, 2005).

39
Figure 2.6 Dietary fibre components (adapted from Thebaudin et al., 1997).

40
2.4.4.2. Phytic acid

Phytic acid (PA) - myoinositol hexaphosphate (IP6), a natural compound usually formed

during maturation of plant seeds, is a common constituent in plant ingredients (Figure 2.7). In

dormant seeds, phytate-P represents 60 to 90 percent of the total phosphorus, and is the primary

storage of both phosphorus and inositol in plant seeds (Reddy, 2001). However, its location

within the seed differs among plants. For example, ninety percent of the PA in corn is found in

the germ portion of the kernel, whereas in wheat and rice it is concentrated in the aleurone layers

of the kernel and the outer bran (ODell and De Boland, 1976). In most oilseeds and grain

legumes, it is associated with protein and concentrated within subcellular inclusions called

globoids that are distributed throughout the kernel; however, in soybean seeds, there appears to

be no specific location (Ravindran et al., 1995). Table 2.4 reports total phosphorus and phytate-P

content in some commonly used plant ingredients.

2.4.4.2.1. Interactions with minerals

Phytic acid has 12 replaceable protons or reactive sites. Six are strongly acidic, with pKa

of 1.5 to 2.0; two are weakly acidic with a pKa of approximately 6.0, and four are very weakly

acidic with pKa of 9.0 to 11.0 (Costello, et al., 1976; Erdman, 1979). This means that at pH

normally encountered in feeds and in the digestive tract, PA will carry a strong negative charge

and is capable of binding di- and trivalent cations such as Zn2+, Cu2+, Ni2+, Co2+, Mn2+, Fe2+, and

Ca2+ in very stable complexes (Wise, 1983, 1995; Persson et al., 1998; Maenz, et al., 1999;

Kaufman and Kleinberg, 1971; Vohra et al., 1965; Maddaiah, et al., 1964), thus reducing the

availability of these complexed minerals to the animal (Pallauf and Rimbach, 1997).

41
The order of potency of selected minerals to inhibit phytate hydrolysis by phytase at pH 7

is Zn2+ >> Fe2+ > Mn2+ > Fe3+ > Ca2+ > Mg2+ (Maenz et al., 1999). The ability of the different

metal ions to inhibit hydrolysis is related to the stability of the complex, the pH of the solution,

and the phytate to mineral molar ratio. A detail review on the interactions of these minerals with

phytic acid could be found in Lopez et al. (2002).

42
Figure 2.7 The structure of phytic acid (myo-inositol hexaphosphate, IP6) (from Reddy, 2001).

OH OH

OH

OH
OH

OH

43
Table 2.5 Total phosphorus and phyate-P content in the major plant ingredients currently

utilized in compounded aquafeed.

Feed ingredients Total P Phytate P Phytate P Source

(% DM) (% DM) (% total P)

Canola meal 0.88 0.67 59 Selle et al., 2003

Corn gluten meal 0.60 0.42 70 Ravindran et al., 1995

Corn protein concentrate1 0.23 0.16 70 This study

Ground corn1 0.28 0.18 64 This study

Indian mustard meal1 1.14 0.80 70 This study

Indian mustard protein conc.1 1.45 1.17 81 This study

Lupin seeds 5.70 3.50 61 Steiner et al., 2007

Soy protein concentrate1 0.80 0.50 63 This study

Soy protein isolate - 1.4-2.1 Reddy, 2001

Soybean meal1 0.68 0.42 62 This study

Wheat gluten1 0.24 0.09 38 This study

Wheat middlings1 0.63 0.46 73 This study


1
Analyses of total phosphorus and phytate-P content in the samples of these ingredients are

conducted by DANISCO Animal Nutrition, MO.

44
2.4.4.2.2. Interactions with proteins and amino acids

Phytic acid also creates binary (PA-protein) or ternary (PA-mineral-protein) complexes

(Kies et al., 2001). Selle et al. (2000) proposed that the de novo formation of binary protein-PA

complexes in the GIT as the key mechanism by which PA influences the AA digestibility.

Several modes of forming these complexes are suggested: (1) the presence of protein-PA

complexes in feedstuffs, (2) the de novo formation of binary and or ternary complexes in the

digestive tract, and (3) the inhibition of proteolytic enzymes (Kies et al., 2001; Selle et al., 2006).

The de novo formation of loose electrostatic association of PA and proteins in the GIT

occurs only at optimal pH conditions (Cheryan and Rackis, 1980; Maenz, 2001). At pH below

the iso-electric point (pKa) of protein, the anionic phosphate groups of PA binds strongly to the

cationic groups of the protein to form binary complexes that dissolve only at pH below 3.5. The

high pKa basic amino acids (Arg, Lys, His) are highly susceptible to chelation as they become

polar and positively charged at pH below their pKa. There are three possible protein-binding

sites for phytate at low pH: (1) the -NH2 terminal group and the -NH2 group of lysine; (2) the

imidazole group of histidine; and (3) the positively charged guadinyl group of arginine

(Cosgrove 1966). Okubo et al. (1976) reported that above the iso-electric point of protein (pH

4.9), glycinin component of soy protein stops binding to phytin. The binding potential increases

as pH decreases but maximizes at pH 2.5 (Adeola and Sands 2003).

Ternary PA-mineral-protein complexes are usually formed at the intermediate pH levels

(pH 5.0 and above), such as in intestinal environment (pH 8.5-8.8) (Riche et al., 2001). They are

formed via histidine, bound by mineral-PA complexes involving mainly Ca, Mg, and Zn. At pH

above 10, these ternary structures are disrupted, the protein is released, and what precipitates out

are the mineral-phytate complexes (Kies and Selle, 1998).

45
2.4.4.2.3. Interactions with proteolytic enzymes

Singh and Krakorian (1982) reported reduced activity of the proteolytic enzyme trypsin

by phytate at pH 7.5 in an in-vitro study that increases with the increase of Na-phytate

concentration. Trypsinogen, an inactive precursor of trypsin, is known to bind Ca-ions at two

sites, where as trypsin has only one binding site. The authors hypothesized that the possible

binding of Ca in trypsin by phytate inhibits the activity of the enzyme. Later, Caldwell (1992)

confirmed this hypothesis, and reported negative effects of phytic acids bound Ca on the

activation of trypsinogen and the stability of trypsin.

2.4.4.2.4. Effect of PA on nutrient utilization

Laining et al. (2010) reported decreasing vertebral deposition of Ca, Mg and Zn in

Juvenile Japanese Flounder (Paralichthys olivaceus) with increasing dietary PA levels. Among

the minerals, Ca and Mg deposition in fish were statistically similar in all diets except for in

those fed highest PA (2%) diets. However, Zn deposition in vertebrae decreased significantly at

even 0.5% PA, from 115 mg kg-1 in control to 100 mg kg-1. Similar effects on Zn bioavailability

were also reported in channel catfish Ictalurus punctatus (Gatlin and Phillips, 1989), tilapia

Oreochromis niloticus (Nwanna, 2004), and rainbow trout (Vielma et al., 2004).

Interactions of PA with protein in fish showed significant decrease in CP digestibility in

several fish species, such as Atlantic salmon Salmo salar (Sajjadi and Carter, 2004a,b), rainbow

trout Oncorhynchus mykiss (Spinelli et al., 1983; Sugiura et al., 2001) or common carp Cyprinus

carpio (Hossain and Jauncey 1991). PA also reported to affect digestibility of amino acids Phe,

46
Ile and Leu in poultry (Cowieson et al., 2006; Newkirk and Classen, 2001), and swine (Kies,

2005).

2.4.4.3. Fibre components and their effects

Physico-chemical properties of different components of dietary fibre vary in their

solubility, gel formation, viscosity, holding and absorption capacity of water and lipid. Although

the effects of these properties are studied rarely, these properties can significantly affect the

digestive processes in an animal (Souffrant, 2001). In farmed animals, most studies on dietary

fibre from various plant sources have dealt with inclusion levels and their effects on protein and

energy utilization, and growth. For example: pig (Laplace et al., 1989; Mitaru et al., 1984;

Souffrant, 2001), poultry (Slominski and Campbell, 1990; Villamide and San Juan, 1998), or fish

(Anderson et al., 1984; Bromley and Adkins, 1984; Hemre et al., 1995; Kraugerud et al., 2007).

One of the earliest studies in fish was by Bromley and Adkins (1984) where the authors

reported a significant increase in the size of the stomach of rainbow trout from 2% to 3.5% of the

body weight with increasing dietary cellulose level from 30% to 40%. They also reported a

significant reduction in protein deposition rate when fish were fed high fibre diets. Recent

studies have shown that not the dietary fibre, but the composition of its components influences

digestive processes in fish (Amirkolaie et al., 2005; Glencross, 2009; Ovrum-Hansen and

Storebakken, 2007; Zahedifar et al., 2002; Zhu et al., 2005). Glencross (2009) reported that the

ADC of DM, CP, and GE in rainbow trout was affected more by lignin than pectin or cellulose.

Considering the limited activity of intestinal microflora in salmonids, Leenhouwers et al. (2008)

hypothesized that the biological activity of various fibre components on the digestive processes

are governed by the composition of the polymer chains of the fibre components.

47
2.4.4.4. Lignin

Lignin, a phenolic plant polymer responsible to provide structural support and the flow of

nutrients, forms an integral part of the secondary cell walls of most plants and algae (Jung and

Fahey, 1983). They are derived from the aromatic amino acid producing shikimate-chorismate

pathway in vascular plants (Lewis et al., 1998), providing rigidity to the plants and rendering the

walls water-impermeable (Whetten and Sederoff, 1995), and constitute some of the

metabolically most expensive products generated by plants (Lewis and Yamamoto, 1989).

A lignin polymer is composed of three monolignols - sinapyl alcohol, coniferyl alcohol,

and p-coumaryl alcohol, which upon incorporation, are referred as syringyl (S), guaiacyl (G), or

hydroxyphenyl (H) units, respectively. The proportion of S, G, and H varies among different cell

types and among species, can substantially affect the physical properties of the lignin polymer

(Boerjan et al., 2003). It is believed that monolignols are stabilized to lignin by -glucosidases

(from the soluble fibre -glucans) and glucosyl transferases (Dharmawardhana and Ellis, 1998).

2.4.4.4.1. Lignin biosynthesis

The lignification process encompasses the biosynthesis of monolignols, their transport to

the cell wall, and polymerization into the final molecule. The types of lignin formed are

dependent upon the release of monolignols, such as p-coumaryl alcohol, coniferyl alcohol, and

sinapyl alcohol into the wall matrix followed by diffusion to the site of incorporation (Figure

2.8). These monolignols, differing only in the substitution pattern on their aromatic rings, are

polymerized into lignin.

48
Figure 2.8 Structure of three monolignol precursors to lignin (adapted from Hatfield and

Vermerris, 2001).

p-Coumaryl alcohol Coniferyl alcohol Sinapyl alcohol

49
2.4.4.4.2. Effects of purified lignin

In nature, lignins form covalent bonds with cellulose and hemicelluloses, and do not exist

in their pure form. Three major lignin products are currently available: 1) ligno-sulphonates

from the sulphuric acid treatment in the wood-pulping industries; 2) kraft lignin from alkaline

treatments of the wood fibres; and 3) purified lignin produced by disrupting chemical integrity

of native lignin-cellulose complexes using aqueous-ethanol extraction process at temperatures

between 1850C and 1950C (Baurhoo et al., 2008). Lora et al. (1993) reported that purified lignin,

in contrast to other commercially available forms, consists of low-molecular weight phenolic

fragments, and may possess some functional properties unavailable in other forms.

Anti-microbial and pre-biotic effects of the phenolic components of purified lignin and

associated lignans are well-known (Davidson and Brenan, 1981). Phenolic acids (lignans) and

flavonoids are used as preservatives to prevent microbial contaminations in food products (Roller

and Sreedhar, 2002; Ultee et al., 2000). Mono-phenolic compounds in lignin, such as carvacrol,

thymol and cinnamaldehyde possess anti-bacterial properties. Carvacrol and thymol disintegrate

bacterial cell-membranes by lowering the intra-cellular pH. The acidic condition depletes the

intra-cellular ATP concentration eventually killing the bacteria (Baurhoo et al., 2008). Baurhoo

et al. (2007a,b) also suggested some prebiotic effects of lignin increasing the concentration of

beneficial microflora, Lactobacilli and Bifidobacteria in the intestine of broiler chickens at level

<2.5% in the diets.

2.4.4.4.3. Effect of lignin on minerals and amino acid availability

Studies on some terrestrial animals suggest deleterious effects of lignified fibres on

mineral, amino acid and energy digestibility (Wenk et al., 1998). Wenk et al. (1998) reported

50
significant reduction in energy digestibility with the diets containing all three forms of lignins

than the lignin-free control diet.

In vitro studies showed significantly higher binding capacity of lignin with bi-valent ions

of minerals (e.g., Fe+2, Cu+2 and Zn+2) at intestinal pH (e.g., pH ~5) than pectin and cellulose

(Platt and Clydesdale, 1987). In a study, Fernandez and Philipis (1982) reported hydroxyl and

methoxy groups of lignin react with Fe to form highly stable hexadentate complexes, and

concluded that lignin, but not cellulose or low methoxy pectin, is a potential binder of Fe+2 under

pH conditions similar to those in the proximal intestine of animals. The authors suggested two

active sites in lignin molecule that bind with these minerals in the order of Fe > Cu > Zn.

Lignin also forms protein complexes under oxidative conditions that are generally

resistant to microbial fermentation in ruminant (Fernandez and Philipis, 1982; Zahedifar et al.,

2002). In a study, Whitmore (1982) reported high binding capacity of lignin with proline

followed by serine and glycine. The author hypothesized that proteins, specifically those

containing hydroxyproline, could be very susceptible to bonding with dehydrogenation polymer

of lignin under oxidizing conditions. In fish, Glencross et al. (2008) reported significant

deleterious effects of lignin on CP, lipid and GE digestibility in rainbow trout.

2.4.5. Physical characteristics of aquaculture feeds

2.4.5.1. Overview of feed production technology

Feed manufacturing is the physical process of forming ingredient mixtures in to particles

used to feed fish or shrimp. Any aquaculture feed needs to be stable enough to hold the nutrients

together for the time to be ingested. It also needs to withstand stresses occurred during handling

and transportation. Fines or undersized feed particles are not consumed by fish and contribute to

51
increased nitrogen and phosphorus in the water, reduced water quality in the culture units, and

increased feed cost. The basic steps of manufacturing feed involve grinding, mixing,

conditioning, pelleting, cooling, top-dressing, sacking, storing, and shipping (Figure 2.9).

Grinding, mixing, conditioning and expansion all are important critical steps before

pelleting (Hardy and Barrows, 2002). The main purpose of grinding is to reduce the particle size

using various mechanical processes such as impact (e.g., hammer mills), cutting (e.g., rotary

cutters), and crushing (e.g., roller mills). Ground particles are then passed through sieves to

achieve a uniform particle size. All ingredients are mixed to produce a homogenous blend. Some

or all the oil can be added to the mixture at this time.

After mixing, feed mixtures are conditioned in a chamber with steam and physical

agitation. Conditioning prepares a feed mixture for pelleting by increasing the moisture content,

heating the mixture, and adding energy to activate the gluten proteins (NRC, 2011). Feed

mixtures can be conditioned for short time (~30 secs) for compressed pellets or up to 5 minutes

before extrusion pelleting. The longer conditioning period also results in a higher degree of

starch gelatinization improving overall carbohydrate digestibility (Aas et al., 2011).

Fish and shrimp require different types of pellets such as extruded slow sinking or

floating pellets or compressed sinking pellets, respectively, to maximize intake and to reduce

wastes from unused feed. Different types of feed can be produced by compression that produces

feed with very high bulk density (0.5-0.6 g cm-3) or by extrusion, a more versatile process that

can be used to produce floating pellets or slow sinking pellets.

52
Figure 2.9 A generic diagram of the steps in a typical aquaculture feed manufacturing process.

53
2.4.5.2. Physical properties of pellets and their importance

Physical properties of an ingredient are a measure of its effect on the characteristics of a

feed (Aas et al., 2011). An ingredient may have an excellent nutritional profile but its value

diminishes if cannot be incorporated effectively into a feed (Refstie and sgrd, 2009).

Physical properties include fluid absorption capacity, bulk density and hardness, sinking

velocity, viscosity and water stability of the pellets. Sinking or floating properties of the pellets

can also be determined by bulk density (Table 2.6). Besides bulk density, pellets of different

species of fish and shrimp require different floating or sinking properties that also largely

correspond to the final total fat content in the pellet (Table 2.7). Durability of a pellet can be

determined by hardness and water stability. Water stability of a pellet is usually determined by

the time it takes to disintegrate and may include the determination of the dry matter loss. Rapidly

disintegrating feeds are highly undesirable in commercial farming practices causing significant

environmental and economical losses (Cho and Bureau, 2001; Obaldo et al., 2002).

54
Table 2.6 Indication of final pellets bulk density in relation to floating or sinking properties

(Durand, 2011)

Feed characteristics Fast sinking Slow sinking Neutral Floating

floatability

Bulk density (g cm-3) >640 540 - 600 480 540 <450

Table 2.7 Final fat contents (%) in the pellets of different aquaculture species in relation to

floating or sinking properties (Durand, 2011)

Species Salmon Trout Carp/tilapia/catfish Shrimp

Final pellet total fat, % >35 15-35 5-15 <5

Texture bulk density Slow sinking Slow sinking Floating Fast sinking

55
2.4.5.3. Different factors affecting these physical properties

Type of feedstuffs (Olsen et al., 2006; Refstie et al., 2006), feedstuff components

(Hansen and Storebakken, 2007; Hilton et al., 1981), and processing and manufacturing

conditions (Sorensen et al., 2005; Sorensen et al., 2009) have substantial effects on the stability

of a feed requiring to adjust processing and manufacturing parameters that include physical,

chemical and thermal treatments (e.g., moisture, temperature, pressure, flow, torque, screw

speed) (Draganovic et al., 2011; Thomas, 1996).

Besides increasing the waste-outputs (Cho and Bureau, 2001), rapid disintegration of a

feed after ingestion appeared to cause oil-belching in salmonids fed high-fat diets often

coinciding with gastric dilation air sacculitis (GDAS) (Baevefjord et al., 2006; Forgan and

Forster, 2007; Lumsden et al., 2010).

2.4.5.4. Components of plant proteins influencing the pellet characteristics

Pellet water stability is highly vulnerable to the changes in ingredient composition of the

diet, physical and chemical properties of the ingredients, and processing and manufacturing

conditions. Water stability can be improved by adding starch and soluble (e.g., guar gum,

Brinker, 2007; Brinker and Reiter, 2011; Storebakken, 1985) or insoluble (e.g., cellulose,

Bromley and Adkins, 1984; Hansen and Storebakken, 2007; Lupatsch et al., 2001) fibres as

binders in fish feed.

Level and composition of the fibre components in the plant protein ingredients alter the

physical properties of the mash and pellets during processing often requiring modifications of the

manufacturing parameters. Careful attention to modification of processing and manufacturing

conditions is needed for producing highly stable pellets.

56
2.5. Conclusion

Long-term sustainability of aquaculture operations depends on improving cost-

effectiveness of feed and increasing the reliance on economical protein sources to replace or

supplement conventional protein ingredients. Salmonid feeds are formulated with an increasingly

wide variety of ingredients of plant, animal or microbial origins.

Plant protein ingredients significantly differ from their animal counter-parts in terms of

AA compositions, fatty acid and lipid profiles, starch and fibre contents, and presence of some

ANF. They are usually deficient in one or more EAA that need to be supplemented to meet the

requirement of farmed fishes. Further research is needed to investigate the effects of various

ANFs and fibre components, individually and in combinations, on nutrient digestibility,

utilization and metabolism as well as on the intestinal function, structure, defense mechanisms

and gut micro-flora. This will add to the ability to predict how a novel plant ingredient as well as

combinations of plant ingredients may affect the fish and to identify steps needed to

reduce/prevent adverse health effects.

Specific effects of alternative protein sources on the digestive physiology of fish have

been most closely studied in the case of soybean products in feeds for farmed salmonids. The

components in soybeans that lead to reduced nutrient digestibility and the inflammatory response

in the distal intestine are still largely unknown, but affect many specific digestive processes,

from feed intake to enzyme activities to nutrient transport to gut histology. Such details are often

lacking in studies focusing on protein sources that are already in use in formulated feeds for

various fish species. Thus the long-term implications they may have on fish production, health,

57
and product quality are questionable, and this topic deserves substantial investments in further

research to preserve and enhance the sustainability of the aquaculture industry.

Evaluation of a plant-protein ingredient for its inclusion in aquafeed therefore, needs

comprehensive species-specific studies starting with characterization of physical and chemical

(amino acids, fatty acids, and mineral compositions) properties followed by digestibility and

bioavailability of amino acids, and growth and health implications of the ANFs and fibre

components of the ingredient.

58
CHAPTER 3

DIGESTIBILITY OF AMINO ACIDS IN TWO NOVEL INGREDIENTS, INDIAN

MUSTARD PROTEIN CONCENTRATE AND INDIAN MUSTARD MEAL COMPARED TO

THAT OF A SOY PROTEIN CONCENTRATE IN RAINBOW TROUT AND ATLANTIC

SALMON

Abstract

Digestibility trials were carried out to determine the apparent digestibility of amino acids

to rainbow trout (Oncorhynchus mykiss) and Atlantic salmon (Salmo salar) of two novel plant

protein ingredients, Indian mustard protein concentrate (IMC, 62% crude protein) and Indian

mustard meal (IMM, 42% crude protein) produced through a proprietary low temperature

process, and a commercially available soy protein concentrate (SPC, 57% crude protein). The

apparent digestibility coefficient (ADC) of crude protein (CP) in IMC was 90% and 97% for

trout and salmon, respectively. The ADCs of CP in IMM and SPC were also very high (>95%)

for both species. The ADCs of lysine (95%), isoleucine (91%), leucine (86%) and phenylalanine

(90%) in IMC were slightly lower (P<0.05) than those in IMM and SPC for rainbow trout.

Conversely, in Atlantic salmon, ADCs of isoleucine in both IMC (94%) and IMM (92%) were

lower than those found in SPC (97%). Despite small differences, ADC of most AA in these

novel products were very good (>90%). These results suggest that low glucosinolate Indian

mustard protein concentrate and Indian mustard meal are ingredients of high nutritive value.

However, it is hypothesized that the high phytic acid level in the Indian mustard protein

concentrate may negatively affect the digestibility of CP and certain amino acids when this

ingredient is used at high levels in the feed of fast growing fish.

59
3.1. Introduction

Plant protein sources are increasingly used to satisfy the growing demands of the

aquafeed industry for economical protein sources (Hardy, 2010). Meals and protein concentrates

produced from rapeseed (Brassica napus) and canola (Brassica rapa) cultivars with low

glucosinolates (<30 mol g-1 of oil-free dry matter of seed) and erucic acid (<2% of total fatty

acids in the oil) have been widely studied and are used in aquaculture feeds worldwide (Cheng et

al., 2010; Drew et al., 2007b; Shafaeipur et al., 2008; Thiessen et al., 2003; 2004; Zhou and Yue,

2010). Cultivation of another oilseed from the Brassicacae family, Brassica juncea, known as

Indian or oriental or brown mustard, is gaining popularity in some regions of North America,

India and Australia since it is generally more tolerant to heat, drought and fungal infection than

rapeseed or canola (Chevre et al., 2008; Oram et al., 1999; Wijesundera et al., 2008; Woods et

al., 1991). Some Indian mustard varieties yield more seeds per pod, produce heavier seeds, and

contain higher dry matter in the seeds than the canola varieties (Chauhan et al., 2007;

Gunasekera et al., 2006; Wright et al., 1995). Low drought susceptibility index (a combination of

growth, maturation (days), number of pods and seed yield per pod, and dry matter content) of

some Indian mustard varieties is further increasing the potential of growing this crop in the

drought-prone regions of the world.

The development of B. juncea cultivars with oil and nutritional characteristics similar to

canola has enhanced their potential in animal feeds (Burton et al., 2004). The nutritive value of

low glucosinolate Indian mustard meals (Jia et al., 2008; Meng et al., 2006; Newkirk et al.,

1997), and Indian mustard protein concentrates and isolates (Thacker and Petri, 2010) have been

evaluated in the diets of poultry. These studies indicated that the nutritional value of meals and

concentrates from low glucosinolate Indian mustard was equal or superior to that of canola meal

60
samples derived from B. napus and B. rapa cultivars. The lower protein and high levels of fibre

(indigestible carbohydrates) in the Brassica protein meals limit the potential of these ingredients

in highly digestible nutrient dense salmonid fish feeds (Carter and Hauler, 2000; Hardy, 2010).

Different processing techniques have been developed to reduce the levels of indigestible

carbohydrate and to produce high protein ingredients from a variety of plant co-products (Bureau

and Cho, 1997; Glencross et al., 2010; Hardy, 2010). Studies have indicated that protein

concentrates of soybean (Bureau and Cho, 1997; Kaushik et al., 1995; Storebakken et al., 1998a,

b) and canola (Higgs et al., 1994; Mwachireya et al., 1999) are highly digestible and can be

included at high level in the diets of salmonid fish species. Nutritional quality of these plant

products can vary significantly by the sources of raw materials and processing equipment,

techniques, and conditions. Processing techniques such as expeller or solvent extraction

(Glencross et al., 2004b), extraction temperature and time (Fenwick et al., 1986), and drying

conditions (Glencross et al., 2007a; Stewart et al., 2003) may play important roles in modifying

nutritive values of oilseed processing co-products.

Fish, like any other animal, have a dietary requirement for essential amino acids (EAA)

and non-specific amino group sources. Fish feeds are increasingly formulated on a digestible

amino acid basis (El-Haroun et al., 2009; Gaylord et al., 2010; Poppi et al., 2011; Yamamoto et

al., 2005). However, very limited effort has been invested in estimating the digestibility or

availability of individual amino acids in fish feed ingredients (Bureau, 2008; De Silva et al.,

2000; Kaushik and Seiliez, 2010). Poor characterization of the digestibility or bioavailability of

nutrients results in the need to use broader safety margins when formulating feeds, reducing the

ability to formulate feeds on a least-cost basis, and limiting the confidence of feed formulators in

the nutritive value of many ingredients (Bureau, 2006). Accurate information on digestibility of

61
amino acids of traditional and novel ingredients is therefore, essential to improve the

cost-effectiveness of aquaculture feeds.

Subtle differences in the digestibility of some nutrients appear to exist between Atlantic

salmon and rainbow trout - the two most widely cultured salmonid species (Azevedo et al., 2004;

Glencross et al., 2004a; Hua and Bureau, 2009; Krogdahl et al., 2004; Refstie et al., 2000).

However, limited information is available on potential differences in the digestibility of amino

acids between these two species. This study investigated the apparent digestibility of crude

protein and amino acids in novel Indian mustard meal and a novel Indian mustard protein

concentrate produced using a proprietary low temperature (~60oC)extraction process using

florine based solvent. The digestibility of these two novel ingredients was compared to that of a

commercially available soy protein concentrate in rainbow trout (Oncorhynchus mykiss) and

Atlantic salmon (Salmo salar).

3.2. Methods

The digestibility trials were conducted at the UG/OMNR Fish Nutrition Research

Laboratory (Department of Animal and Poultry Science, University of Guelph, Ontario, Canada)

using the standard protocol of Cho et al. (1982) and Bureau et al. (1999). Indian mustard protein

concentrate (IMC) and meal (IMM) were obtained from Bio-Exx Specialty Proteins Ltd.

(Toronto, ON, Canada) and the soy protein concentrate (SPC HP 300) was obtained from

Hamlet Protein A/S (Aarus, Denmark). The Indian mustard protein concentrate and meal are

produced using a proprietary low temperature extraction process while the soy protein

concentrate is produced through a proprietary enzymatic treatment and aqueous extraction

62
process. Digestibility of nutrients in these ingredients for rainbow trout and Atlantic salmon was

determined in two consecutive trials carried out in the same facilities, same aquatic systems,

protocols and staff.

3.2.1. Dietary design

A high quality reference diet (basal diet, Table 3.1) was prepared and combined with each

test ingredient at 70:30 ratio (as is basis) to produce three test diets (one for each ingredient), as

per the indirect digestibility protocol of Cho et al. (1982).

An inert marker, yttrium oxide-Y2O3 (Sigma-Aldrich Inc., MD, USA), was added to the

basal diet at 100 mg kg-1 as digestion indicator. The diets were mixed (using a Hobart mixer

Hobart Ltd., Don Mills, Ontario), pelleted (using a laboratory steam pellet mill California

Pellet Mill Co., San Francisco, California), and dried in forced air at room temperature for 24h.

Dried pellets for each diet were kept at 4oC until used. The amount of diet needed was weighed

weekly and kept at room temperature. The proximate composition, gross energy and amino acid

profiles of three test ingredients and four diets are reported in the Tables 3.2 and 3.3,

respectively.

63
Table 3.1 Ingredient composition of the basal diet (%)

Ingredients %

Fish meal, herring, 70% crude protein 25

Blood meal, porcine, spray-dried 11

Corn gluten meal, 60% crude protein 25

Wheat middlings, 17% crude protein 10

Corn, grain ground 10

Vitamin premix (Martin Mills, PM1-83) 1

Mineral premix (Martin Mills, 9504) 0.5

Dicalcium phosphate 0.5

Fish oil, herring 17

Total 100

64
Table 3.2 Proximate chemical composition, gross energy, essential amino acid profiles, total

phosphorus and phytic acid content of the test ingredients (as is basis)

Ingredients
Proximate composition (as is) Indian mustard Indian mustard Soy protein
concentrate meal concentrate
Dry matter, % 95 95 93
Organic matter, % 87 88 86
Crude protein, % 62 44 57
Lipid, % 0.2 1.9 2.2
Total carbohydrate, % 30 48 33
Neutral detergent fibre, % 13 17 7
Acid detergent fibre, % 11 13 6
Ash, % 8 7 8
Gross energy, kJ.g-1 19 19 19
Essential amino acids (% in the weight of ingredient as is)
Arginine 4.1 2.8 3.9
Histidine 1.6 1.1 1.4
Isoleucine 2.4 1.7 2.5
Leucine 4.2 2.9 4.1
Lysine 2.6 1.9 2.7
Methionine+Cystine* 2.8 1.9 1.3
Phenylalanine 2.3 1.6 2.7
Threonine 2.2 1.6 1.8
Tryptophan* 1.0 0.6 0.6
Valine 3.0 2.1 2.7
Phosphorus and phytate-P (as is)
Total phosphorus, % 1.5 1.1 0.8
Phytate-P, % 1.2 0.8 0.5

Phytate-P is calculated as phytic acid*0.282; ; * manufacturer data


65
Table 3.3 Proximate analysis and, gross energy, essential amino acid profiles, total phosphorus

and phytic acid concentrations of the diets (as is basis)

Diets

Proximate composition (as is) 1 2 3 4


70%-Ref + 70%-Ref + 70%-Ref +
Ref
30%-IMC 30%-IMM 30%-SPC
Dry matter, % 96 95 96 97
Organic matter, % 88 89 89 89
Crude protein, % 51 53 48 53
Lipid, % 19 13 14 14
Total carbohydrates, % 21 23 29 24
Neutral detergent fibre, % 8 9 11 7
Acid detergent fibre, % 4 6 7 4
Ash, % 6 7 6 7
Gross energy, kJ g-1 24 23 23 23
Essential amino acids (% in the weight of diet)
Arginine 2.9 3.4 2.7 3.1
Histidine 1.9 1.8 1.5 1.6
Isoleucine 2.0 2.2 1.8 2.1
Leucine 6.7 5.9 5.2 5.8
Lysine 2.8 2.9 2.4 2.5
Methionine+Cystine 1.6 2.0 1.7 1.6
Phenylalanine 2.9 2.8 2.4 2.8
Threonine 1.9 2.0 1.7 1.8
Valine 3.5 3.3 2.8 3.1
Phosphorus and phytate-P
Total phosphorus, % 1.0 1.2 1.1 1.0
Phytate-P , % 0.3 0.6 0.4 0.3
Ref reference diet, IMC Indian mustard protein concentrate, IMM Indian mustard meal, and SPC soy protein

concentrate. Phytate-P is calculated as phytic acid*0.282.

66
3.2.2. Experimental conditions

Rainbow trout (Oncorhynchus mykiss, Ontario domestic fall-spawning strain) and

Atlantic salmon (Salmo salar, Lahave River strain) fingerlings were obtained from the Alma

Aquaculture Research Station of the University of Guelph (Elora, ON, Canada) and the

Normandale Fish Culture Station of the Ontario Ministry of Natural Resources (Normandale,

ON, Canada), respectively. Fish were stocked in individually aerated fibreglass tanks (60-L),

where water temperature and photoperiod were maintained at 15oC and 12h L 12h D cycle,

respectively. Fish were acclimatized to the experimental conditions for seven days before

starting the experiments. During that time, fish were fed a high quality commercial salmonid diet

(Profishent High Energy Feed, Martin Mills, Elmira, ON, Canada).

Rainbow trout (average weight 71g) and Atlantic salmon (average weight 65g) were

stocked in 24 tanks at 13 and 15 fish tank-1, respectively. The four experimental diets were

randomly allocated to two collection units each. Each unit has three tanks with a column for

faeces collection. The fish were acclimatized to the experimental diets for four days before the

initiation of faeces collection. Fish were hand-fed to near-satiety three times daily between 0930

and 1600 h. Feed residues and faeces were removed everyday an hour after the last daily meal.

One-third of the water in the tanks was drained during the process to ensure complete removal of

faeces and uneaten food. At 0900 h of the following day, the settled faeces and surrounding

water were gently withdrawn into a large centrifuge bottle (250 ml). Collected samples were then

centrifuged at 4,000 x g for 15 min (Eppendorf Centrifuge 5810R, Eppendorf AG, Hamburg,

Germany), separated from supernatant, and stored in a container at -20oC till the end of the trial.

A total of five pooled samples for rainbow trout and four pooled samples for Atlantic salmon

67
were collected for each diet over the experimental periods. At the end of the trials, samples were

lyophilized, finely ground, and stored at -20oC for proximate and amino acid analysis.

3.2.3. Chemical analyses

Diets, ingredients, and faeces samples were analyzed for dry matter (DM) and ash

according to AOAC (1995), crude protein (CP, %N x 6.25, Dumas method) using a

LECO Analyzer (LECO Corporation, St. Joseph, MI), lipids with petroleum ether

extraction at 90oC and 150 PSI for 60 min using ANKOM XT-20 analyzer (ANKOM

Technology, Macedon, NY), and gross energy (GE) using an automated oxygen bomb

calorimeter (Model # 1271, Parr Instruments, Moline, Illinois). Neutral detergent fibre

(NDF) and acid detergent fibre (ADF) contents of the ingredients and the diets were

determined according to the methods of Van Soest et al. (1991) using an Ankom 200

Fibre Analyzer (ANKOM Technology, Macedon, NY). Phosphorus and Y2O3 was

analysed in a contracting laboratory (Laboratory Services Division, University of Guelph,

ON, Canada) using inductively coupled plasma optical emission spectroscopy technique

(ICP-OES). Phytic acid content in the ingredients and diets were analyzed using high

performance liquid chromatography (HPLC, Waters Corporation, Millford, CA)

according to the method of Lehrfeld (1989) at the Danisco Animal Nutrition laboratory,

Danisco, St. Louis, MO. Amino acid profiles of the test ingredients, diets and faeces were

analyzed at the Amino Acid Laboratory of the Toronto Hospital for Sick Children (ON,

Canada) using an acid hydrolysis method. For the quantification of amino acids,

approximately 0.05 g of dried sample was hydrolysed in 15 ml test tube with 450 L of

6N HCl and 1% phenol for 24 hrs at 110oC under pre-purified nitrogen atmosphere. After

68
24 hrs, tubes were removed from oven, cooled, capped, vortex mixed and centrifuged at 2500

rpm. Then, 5 L of the supernatant was derivatized following PICO-Tag method (Cohen et al.,

1989). After derivatization, supernatant were re-dissolved in 150 L phosphate buffer and 2 L

of this solution was injected into the column of an ACQUITY Ultra-Performance LC machine

(Waters Corporation, Millford, CA) for quantification of all amino acids except tryptophan,

methionine and cysteine which are partially destroyed in the hydrolysis process.

3.2.4. Calculations

The apparent digestibility coefficients (ADC) for the nutrients and energy of the test and

reference diets were calculated as follows (Cho et al. 1982):

ADC = 1-(F/D x Di/Fi), where D = % nutrient (or kJ.g-1 GE) of diet, F = % nutrient (or

kJ.g-1 GE) of feces, Di = % digestion indicator of diet (Y2O3), Fi = % digestion indicator of feces

(Y2O3). ADC of the test ingredients (ADCingr) and individual amino acids were calculated based

on the digestibility of the reference diet and the test diets using the equation proposed by Forster

(1999) as mathematically simplified by Bureau and Hua (2006):

ADCtest ingredient = ADCtest diet + [(ADCtest diet - ADCref. diet) x (0.7 x Dref / 0.3 x Dingr)] where,

Dref = % nutrient (or kJ.g-1 GE) of reference diet mash (as is), Dingr = % nutrient (or kJ.g-1 GE) of

test ingredient (as is). Total carbohydrate content is calculated as 100-CP-lipid-ash and organic

matter content is calculated from the difference between dry matter and ash.

3.2.5. Statistical analysis

Data were analysed using SAS v9.2 (Cary, NC, USA). ADC of nutrients and amino acids

of the three test ingredients were compared using Factorial ANOVA under Proc GLM.

Significant differences were determined by Tukey's HSD (honestly significant difference) test at
69
P<0.05 level. Significance of the species-ingredient interaction effects were determined

by the probability values of their least square means.

3.3. Results

3.3.1. Apparent digestibility of energy and nutrients

Table 3.4 reports the ADC of proximate components, GE and EAA of the

reference and the test diets. ADC of most nutrients and GE were significantly different

(P<0.05) between rainbow trout and Atlantic salmon except those for total carbohydrates

(51% for rainbow trout and 49% for Atlantic salmon) and ash (60% for rainbow trout

and 55% for Atlantic salmon) (Table 3.5). ADC of all proximate components,

phosphorus and GE were significantly different among the ingredients (P<0.05). ADC of

CP in the SPC (99%) was higher (P<0.05) than the Indian mustard products (93% and

97% for IMC and IMM, respectively). The ADC of lipid was high, exceeding 100% in

some instances, for all three ingredients because of very low lipid content in the

ingredients (i.e., 0.2, 1.9 and 2.2% in IMC, IMM and SPC, respectively).

Significant species-ingredient interaction effect was observed on the ADC of DM,

OM, CP, P and GE. Least square means of the ADC of DM and GE in IMM for Atlantic

salmon were significantly lower (P<0.0001 and P<0.001, respectively) than those in

IMM for rainbow trout. However, among the six species-ingredient interactions, ADC of

CP in all test ingredients for Atlantic salmon and in IMM and SPC for rainbow trout were

higher than the ADC of CP in IMC for rainbow trout (P<0.001). Lower ADC of CP in

70
IMC for rainbow trout may be attributable to the slightly lower digestibility of four EAA,

isoleucine, leucine, lysine and phenylalanine compared to other amino acids (Table 3.6).

3.3.2. Apparent digestibility of amino acids

ADC of eight EAA and the effects of species, ingredient and species-ingredient

interactions are reported in Table 3.6. ADC of all EAA were not significantly different between

rainbow trout and Atlantic salmon except for those of arginine (P<0.05), histidine (P<0.01) and

valine (P<0.05). Conversely, ADCs of all EAA were significantly different (P<0.05) among the

ingredients except for that of arginine (Table 3.6). None of the six possible species-ingredient

interaction scenarios were significant for the ADC of arginine and threonine. Despite the lack of

significant interaction effects, ADC of histidine and valine in IMC for rainbow trout were

significantly lower than their ADC in SPC for Atlantic salmon (P<0.05 and P<0.01,

respectively). Significant species-ingredient interaction effects were observed in the ADC of four

EAA, lysine, isoleucine, leucine and phenylalanine (P<0.05, Table 3.6).

All NEAA in the ingredients showed high ADC for both the species except for the

glycine in Atlantic salmon. Apparent digestibility of glycine in IMM (93%) was significantly

lower than their ADC in SPC (97%). No species-ingredient interaction effect was observed in the

ADC of any of these NEAA. However, ADC of aspartic acid, glycine, and tyrosine were

significantly different between the two species. Two of these NEAA, aspartic acid and glycine

were also significantly different among the ingredients.

71
Table 3.4 ADC of nutrients, phosphorus, gross energy and essential amino acids in the reference diet (diet 1) and the three test diets
for rainbow trout and Atlantic salmon (mean SD)
Rainbow trout Atlantic salmon
Diet 1 2 3 4 1 2 3 4
70% Ref 70% Ref 70% Ref 70%Ref 70%Ref 70% Ref
ADC (%) Ref Diet 30%IMC 30%IMM 30%SPC Ref Diet 30%IMC 30%IMM 30%SPC
Dry matter 70.3 2.6 71.5 0.4 74.2 1.4 75.4 1.2 76.4 1.4 76.2 1.4 75.3 0.3 78.6 0.4
Organic matter 72.0 2.7 73.8 0.6 75.8 1.3 76.9 1.1 78.3 1.4 79.1 1.1 77.0 0.4 80.2 0.4
Crude protein 87.8 2.6 88.6 0.7 90.7 0.6 91.2 0.6 91.9 1.6 93.6 0.7 93.2 0.7 94.3 0.3
Total carbohydrates 33.6 4.9 34.6 4.5 44.8 4.1 41.6 1.2 30.2 2.1 32.7 2.9 38.9 6.7 40.5 1.8
Ash 43.9 2.5 45.2 3.4 48.8 2.7 54.4 2.9 47.7 1.6 44.9 2.5 48.1 2.3 57.6 1.8
Phosphorus 54.0 1.8 45.9 1.2 46.2 3.1 52.0 2.9 54.6 1.6 44.0 1.6 49.4 2.3 57.9 1.2
Gross energy 72.6 3.1 75.4 0.7 77.3 1.3 78.1 1.0 82.2 1.1 82.6 1.1 81.0 0.4 83.5 1.1
ADC (%) of essential amino acids
Arginine 93.1 1.4 94.8 0.4 94.5 0.3 95.0 0.9 96.1 1.0 97.1 0.5 96.8 0.4 97.4 0.2
Histidine 92.2 1.2 92.6 0.5 93.0 0.3 93.5 1.4 95.2 1.5 95.8 0.8 96.0 0.4 96.6 0.1
Lysine 94.9 0.6 95.0 1.1 96.0 0.3 97.3 0.4 96.2 0.8 97.1 0.5 96.5 0.2 97.4 0.1
Threonine 89.0 1.9 89.4 0.6 89.3 0.3 90.3 1.5 94.4 1.4 94.4 0.8 93.3 0.5 94.8 0.2
Valine 90.0 2.0 89.9 0.6 90.8 0.2 91.3 1.4 94.5 1.5 94.4 0.8 94.1 0.6 95.5 0.3
Isoleucine 86.6 2.8 88.1 0.8 89.3 1.1 90.1 1.3 93.7 1.3 93.8 0.9 93.2 0.8 94.7 0.4
Leucine 88.2 2.7 87.8 1.0 90.2 1.1 90.0 1.6 94.9 1.7 94.7 1.0 94.8 0.6 95.3 0.4
Phenylalanine 88.4 2.5 88.8 1.0 91.5 1.8 92.5 1.0 94.2 1.8 94.8 0.8 94.6 0.7 95.4 0.3
72
Table 3.5 ADC1 of nutrients and gross energy in Indian mustard protein concentrate (IMC), Indian mustard meal (IMM) and soy

protein concentrate (SPC) for two rainbow trout and Atlantic salmon (mean SD).

Apparent Digestibility Coefficients (%)


Species Ingredient Organic Crude Total
Dry matter matter protein carbohydrates Ash Phosphorus Gross energy
(%) (%) (%) (%) (%) (%) (%)
IMC 74.1b 1.5 76.7b 1.5 b
90.1 2.0 a
36.4 12.4 47.7b 9.8 32.2ab 3.3 83.1b 2.6
Trout IMM 83.0a 3.4 85.1a 4.3 98.6a 2.3 56.6a 18.8 59.5ab8.6 28.2b 10.2 90.4a 5.2
SPC 87.1a 3.9 88.5a 3.6 98.2a 2.1 55.0a 8.4 73.3a 8.0 45.6a 12.2 93.4a 3.8

IMC 75.9ab 4.0 80.9a 3.8 96.9a 2.0 37.8b 9.0 39.9b 7.2 25.9b 4.3 83.8ab 4.0
Atlantic
salmon IMM 72.7b 1.0 74.1b 1.3 96.8a 2.4 49.5ab14.8 48.9b 7.2 37.1b 7.8 77.8b 1.5
SPC 83.8a 1.1 84.8a 1.4 99.3a 1.0 60.7a 5.4 75.4a 4.9 68.3a 5.1 87.1a 4.1
Effect (P- value)2
Species P<0.01 P<0.01 P<0.05 NS NS P<0.05 P<0.001
Ingredients P<0.001 P<0.001 P<0.001 P<0.01 P<0.001 P<0.001 P<0.01
Species X Ingredient P<0.01 P<0.01 P<0.01 NS NS P<0.01 P<0.01
1
Means of the ADC of a nutrient for ingredients in a species followed by a common superscript are not significantly different (Tukeys HSD, P>0.05).
2
Effects of species, Ingredient and species-ingredient interactions for individual nutrient are deemed significant at P<0.05 (in bold).

73
Table 3.6 ADC1 of eight essential amino acids in Indian mustard protein concentrate (IMC),

Indian mustard meal (IMM) and soy protein concentrate (SPC) for rainbow trout and Atlantic

salmon (mean SD).

Apparent Digestibility Coefficients (%)


Essential amino acids2
Species Ingredient
Arg His Lys Thr Val Ile Leu Phe
a a b a a b b
97.4 93.5 95.2 90.4 89.5 90.8 86.3 90.0b
IMC 1.0 2.0 2.1 2.2 2.6 2.6 5.7 4.6

97.5a 95.8a 98.0ab 90.3a 92.2a 95.8a 100.0a 100.0a


Trout
IMM 0.9 1.6 4.3 1.5 1.4 4.3 7.9 9.2

97.1a 95.0a 100.0a 91.4a 92.8 95.1a 94.2ab 100.0a


SPC 1.0 3.5 0.6 2.8 2.0a 1.0 2.5 2.1

98.9a 97.3b 99.0a 94.3a 94.oab 94.2ab 93.8a 96.5a


IMC 1.2 2.6 1.6 2.1 3.2 2.7 5.0 3.2

Atlantic 98.6a 98.8b 97.5a 90.1b 92.7b 91.6b 94.3a 96.6a


salmon IMM 1.3 1.6 1.1 2.0 2.9 3.1 4.5 3.6

99.6a 100.0a 99.7a 94.3ab 98.4a 96.6a 97.2a 98.4a


SPC 0.5 0.5 0.5 2.5 1.1 1.1 1.9 1.2
Effect (P-value)3
Species P<0.05 P<0.01 NS NS P<0.05 NS NS NS
Ingredients NS P<0.05 P<0.01 P<0.01 P<0.01 P<0.05 P<0.05 P<0.01
Species*Ingredient NS NS P<0.05 NS NS P<0.05 P<0.05 P<0.05
1
Means of the ADC of a nutrient for ingredients in a species followed by a common superscript letter are not
significantly different (Tukeys HSD test, P>0.05).
2
Arg arginine; His histidine; Lys lysine; Thr threonine; Val valine; Ile isoleucine; Leu leucine; Phe
phenylalanine.
3
Effects of species, Ingredient and species-ingredient interactions for individual nutrient are deemed significant at
P<0.05 (in bold).

74
3.4. Discussion

Apparent digestibility of nutrients and amino acids of two novel ingredients, a meal and a

protein concentrate produced from Indian mustard, compared to those of a commercially

available soy protein concentrate for rainbow trout and Atlantic salmon were assessed in this

study. IMC and IMM used in this study were produced by a proprietary low-temperature (<50oC)

oil-extraction process that prevents protein from denaturing leaving more soluble forms of

protein than the more commonly used high-temperature (>65oC) extraction processes.

Crude protein and the EAA in the three test ingredients were highly digestible (ADC

>90%) for both rainbow trout and Atlantic salmon. The observed ADC of crude protein in IMC

(90%) was similar to the protein digestibility of two rapeseed (canola) protein concentrates (89-

90%) fed to rainbow trout (Higgs et al., 1995; Thiessen et al., 2004). However, ADC of crude

protein (90%) and four essential amino acids i.e. lysine (95%), isoleucine (91%), leucine (86%)

and phenylalanine (90%) in IMC were slightly lower (P<0.05) than their ADC in IMM and SPC

for rainbow trout. Newkirk et al. (1997) examined the digestibility of crude protein of different

oilseed meals (four Indian mustard meals, two canola meals and one low-fibre soybean meal) in

broiler chickens and reported lower CP digestibility for high fibre oilseed meals compared to low

fibre oilseed meals. Total carbohydrate content (calculated by difference) in the IMC (30%) used

in this study was similar to that in the SPC (33%) but much lower than in the IMM (48%). NDF

content was also higher in IMM (17%) than in IMC (13%) and SPC (7%). In this study,

differences in ADC of CP did not appear to be related to the total carbohydrate or to the NDF

and ADF contents of the diets and ingredients. However, the ADC of threonine, valine and

isoleucine of the IMM were significantly lower (P<0.05) than those of the IMC and SPC for

Atlantic salmon, and that could be attributable to the enhanced loss of endogenous proteins in

75
IMM fed fish. Both threonine and valine are reported to present in higher concentration (5.1%

and 4.8%, respectively) than other EAA in the ileal endogenous proteins in swine (Jansman et

al., 2002). Mucins produced by intestinal cells are rich in threonine and high threonine losses

have been reported in swine (De Lange et al. 1989; Zhu et al., 2005) and poultry (Cowieson et

al., 2008) fed high fibre diets.

It can be hypothesized that the low ADC of nutrients in IMC for rainbow trout could be

related to the high phytic acid (PA) content (4.2%) compared to IMM (2.8%) and SPC (1.6%). In

this study, PA intake by rainbow trout fed 30%-IMC diet was significantly higher (37 mg fish-1

d-1) than that of by fish fed 30%-IMM (19 mg fish-1 d-1) or 30%-SPC (25 mg fish-1 d-1) diets. A

relationship may exist between the ADCs of crude protein observed in this study and the PA

intake of the animals. Spinelli et al. (1983) reported a negative effect of PA on the apparent

digestibility of CP in canola meal for rainbow trout. The authors also reported significantly lower

growth and feed efficiency in rainbow trout fed diets containing 0.5% PA compared to those fed

PA-free diets. In this study, feeding the 30% IMC diet (2.0% PA) resulted in significantly lower

apparent digestibility of CP compared to the basal diet (0.9% PA) and the diets containing 30%

IMM (1.5% PA) and 30% SPC (1.1% PA). At pH-values below the isoelectric point of protein,

the anionic phosphate groups of PA bind strongly to the cationic groups of the protein to form

binary insoluble complexes at acidic pH (e.g., in the stomach, pH 2-3), and become unavailable

for intestinal (pH 8-9) absorption (Sugiura et al., 2006).

Among the EAA, the ADC of histidine ( = -0.228; P<0.05) and lysine ( = -0.175;

P<0.05) were inversely correlated with the dietary PA intake in this study (Figure 3.1). The ADC

of aspartic acid ( = -0.352; P<0.01), valine ( = -0.294; P<0.05), leucine ( = -0.465; P<0.05),

and phenylalanine ( = -0.367; P<0.05) were also significantly correlated with the phytic acid

76
intake. Broken-line regression analysis of ADC of Phe and PA intake showed that ADC of the

amino acid was not affected below the PA intake of 17 mg-1 fish d-1 (Figure 3.2). Histidine,

lysine, and arginine are highly susceptible to chelation in both stomach and intestine as they

become polar and positively charged at pH below their pKas (6.1, 10.5 and 12.5, respectively;

Cosgrove, 1966). High dietary PA intake appears to have more significant impact on the ADC of

histidine than lysine and arginine, which agrees well with the differences in pKa between these

amino acids. The -NH2 terminal group and the -NH2 group of lysine and the imidazole group

of histidine and the positively charged guadinyl group of arginine have been implicated as

possible protein binding sites for phytate (Selle et al., 2000). Kinetic and thermodynamic studies

with human serum albumin indicate that PA forms salt-like linkages combining first with the -

NH2 terminal group, followed by the -NH2 of lysine, then histidine and guanidyl groups of

arginine (Cheryan and Rackis, 1980).

No significant differences in the ADC of EAA were observed between rainbow trout and

Atlantic salmon except for those of arginine, histidine and valine. Differences in digestibility of

these three amino acids between the two species were small (less than 5%) and probably

negligible. The large differences in growth rates (TGC between 0.184 and 0.237 for rainbow

trout and between 0,045 and 0.085 for Atlantic salmon) and feed intakes that exist between the

species used in this study make it difficult to meaningfully determine if the small differences in

the digestibility of EAA in different ingredients is determined by species or feed intake

differences. The relationships between PA intake and the digestibility of certain amino acid

observed here are interesting but circumstantial and many other factors may explain these results.

The effect of dietary PA on digestibility of protein and amino acids deserves to be systematically

investigated in future trials.

77
In conclusion, the novel Indian mustard protein concentrate and Indian mustard meal

evaluated in this study appear to be excellent sources of digestible EAA to rainbow trout and

Atlantic salmon tested in this study. The amino acid profile of these ingredients is very good

when compared to essential amino acid requirements of teleost fishes. However, additional

studies are needed to determine if digestibility of CP and certain amino acids can be affected by

dietary PA intake. Small differences in amino acid digestibility were observed between species

but these differences are probably negligible.

78
Figure 3.1 Apparent digestibility coefficients (%) of arginine - Arg (pKa 12.5), lysine Lys
(pKa 10.5) and histidine - His (pKa 6.1) in Indian mustard protein concentrate (IMC), Indian
mustard meal (IMM) and soy protein concentrate (SPC) for rainbow trout and Atlantic salmon
and their relationships to the dietary phytic acid intake. Dietary phytic acid intake in 30%-IMC,
30%-IMM and 30%SPC diets were 29 mg.fish-1.d-1, 15 mg.fish-1.d-1 and 14 mg.fish-1.d-1 for
rainbow trout and 10 mg.fish-1.d-1, 6 mg.fish-1.d-1 and 5 mg.fish-1.d-1 for Atlantic salmon,
respectively.

100
Apparent digestibility coefficient (%)

98
pKa (Arg) = 12.5
R2 = 0.86

96
pKa (Lys) = 10.5
Arginine R2 = 0.78

94 Lysine
pKa (His) = 6.1
Histidine
R2 = 0.94

92
0 5 10 15 20 25 30
Dietary phytic acid intake (mg fish-1 d-1)

79
Figure 3.2 Broken line regression of ADC (%) of phenylalanine in Indian mustard protein

concentrate (IMC), Indian mustard meal (IMM) and soy protein concentrate (SPC) for rainbow

trout and Atlantic salmon with the dietary phytic acid intake (R2=0.83). Dietary phytic acid

intake in 30%-IMC, 30%-IMM and 30%SPC diets were 29 mg.fish-1.d-1, 15 mg.fish-1.d-1 and 14

mg.fish-1.d-1 for rainbow trout and 10 mg.fish-1.d-1, 6 mg.fish-1.d-1 and 5 mg.fish-1.d-1 for Atlantic

salmon, respectively.

100
ADC of phenylalanine (%)

95

90

85
0 5 10 15 20 25 30
Dietary phytic acid intake (mg fish-1 d-1)

80
CHAPTER 4

RELATIVE BIOAVAILABILITY OF ARGININE IN NOVEL INDIAN MUSTARD

PROTEIN CONCENTRATE AND INDIAN MUSTARD MEAL COMPARED TO THAT OF

A SOY PROTEIN CONCENTRATE IN RAINBOW TROUT (ONCORHYNCHYS MYKISS)

AS DETERMINED BY THE SLOPE-RATIO ASSAY

Abstract

Relative bioavailability (RBV) of arginine (Arg) from two novel ingredients, Indian

mustard protein concentrate (IMC, 62% crude protein) and Indian mustard meal (IMM, 42%

crude protein), and a soy protein concentrate (SPC, 57% crude protein) was compared to that of

crystalline L-arginine (L-Arg) in rainbow trout using slope-ratio assay. A basal diet highly

deficient in Arg (1.23%) was formulated using corn gluten meal and skim milk powder as the

main protein sources. Eight other isoproteic and isoenergetic diets were formulated to contain

increasing amount of IMC, IMM, SPC, and L-Arg (~1.35% and ~1.5% Arg). The experimental

diets were fed for 16 weeks using a standard growth assay protocol. Growth rate (expressed as

thermal-unit growth coefficients, TGC), weight gain (g fish-1), protein (PD, g fish-1) and lipid

(LD, g fish-1) of fish increased linearly with increasing level of Arg for all test ingredients. RBV

was determined by linear regression analysis using TGC and PD as the response variables and

Arg intake (g fish-1) and level (%) as predictor variables. In general, ARG availability from

protein-bound sources were higher than those from L-Arg. RBV estimates of Arg from IMC and

IMM were ranged from 123% to 187% and from 116% to 211%, respectively than that from L-

Arg (assumed as 100% bio-available) (P<005). The findings suggest that the RBV of Arg of

IMC and IMM are very high and comparable to that of the commercial SPC used in this study.

81
4.1. Introduction

Cultivation of oilseeds from the Brassicacae family has gained in popularity over the past

several decades. Various protein-rich feedstuffs (meals and protein concentrates) produced from

rapeseed (Brassica napus) and canola (Brassica rapa) are increasingly used in aquaculture feeds

worldwide. The nutritive value of these ingredients has been the focus of significant number of

studies (Cheng et al., 2010; Drew et al., 2007; Shafaeipur et al., 2008; Thiessen et al., 2003;

2004; Zhou and Yue, 2010). Cultivation of another oilseed from the Brassicacae family,

Brassica juncea, known as Indian or oriental or brown mustard, is gaining popularity in some

regions of North America, India and Australia since it is generally more tolerant to heat, drought

and fungal infection than rapeseed or canola (Chvre et al., 2008; Oram et al., 1999; Wijesundera

et al., 2008; Woods et al., 1991). Protein meals and concentrates produced from Indian mustard

cultivars have excellent amino acid (AA) profile and appeared to be of high nutritive value to

poultry and swine (Jia et al., 2008; Meng et al., 2006; Newkirk et al., 1997; Thacker and Petri,

2010).

Accurate characterization of AA composition and bio-availability of feed ingredient is

important to improve cost-effectiveness of feeds. Lack of accurate information force feed

formulators to rely on high safety margins for AA and other nutrients in the feed leading to high

production costs (Bureau, 2006, 2008; Cho and Bureau, 2001). Fish feeds are increasingly

formulated on digestible AA basis (El-Haroun et al., 2009; Gaylord et al., 2010; Poppi et al.,

2011; Yamamoto et al., 2005) but very limited effort has been invested in estimating the

digestibility or availability of individual AA in fish feed ingredients (Bureau, 2008; De Silva et

al., 2000; Kaushik and Seiliez, 2010). Bioavailability is defined as the proportion of ingested

dietary AA absorbed in a chemical form potentially suitable for metabolism or synthesis

82
(Ammerman et al., 1995; Batterham, 1992; Lewis and Baily, 1995; Stein et al., 2007). Among

several approaches, slope-ratio assays offer direct and practical assessment of the availability of

individual AA and have been used to assess the bioavailability of AA for farmed animals (Lewis

and Bailey, 1995). In this assays, graded levels of the test ingredient are included in diets with a

known nutritional composition. These diets, formulated to be first limiting in the nutrient (e.g.,

amino acid) of interest, are then fed to experimental animals and the growth, and other responses

of interest, of the animals are recorded in response to increasing levels of the test ingredient. A

regression line is then fitted to the data and the slope of this line is compared to that of the data

from animals fed diets containing graded levels of a purified form, or well defined source, of the

nutrient of interest.

Relative bio-avalaibility is determined by comparing the slopes of the response of fish to

increasing level of test nutrient supplied by the test ingredients with the slope of the response of

those fed diets where the test nutrient is provided by the reference source of this nutrient. The

slope-ratio is usually determined by linear regression model (y = b1 + b2*S + b3*T, S - standard

and T - test diets) that needs pre-validation of three assumptions: linearity of the responses,

common intercepts for all lines, and the equal basal level response to the common intercept

(Batterham et al., 1979; Finney et al., 1951, 1964; Littel et al., 1997). Linear models have been

extensively used to determine bioavailability of individual AA or minerals in poultry (Adeola,

1998; Batterham et al., 1986; Lumpkins and Batal, 2005; Major and Batterham, 1981; Parsons et

al., 1998), swine (Adeola, 1996, 2009; Batterham et al., 1990; Fanimo et al., 2006; Mateo et al.,

2007; Petersen et al., 2011; Shoveller et al., 2009), and recently, in fish (El-Haroun and Bureau,

2007; Li et al., 2008, 2009; Poppi et al. 2011; Zarate and Lovell, 1997). In case of non-linear

responses, an exponential regression model (y = b1 + b2*(1 exp (b3*S + b4*T) can be used to

83
estimate bioavailability from the ratio of the steepness coefficients b3 and b4, where b1 (the

response to the basal diet) and b2 (the difference between the lowest and highest response) have

to be equal for all sources (Kim et al., 2006; Littel et al., 1997; Liu et al., 2004). Another method

is the broken-line models where the slopes of the linear portion of the responses are used to

determine the RBV of the nutrient of interest (Coon et al., 2007; Wiltafsky et al., 2009). Like

other assays, slope-ratio assays also have some limitations and are sensitive to the experimental

artefacts (Poppi et al., 2011). In these assays, the desired levels of the nutrient in question in the

diets are achieved by adding the test ingredients. This approach allows for the concentration of

various other components such as trace-minerals and fibres to vary, and results can be affected

by the concentration of test nutrient and dietary intake.

Indian mustard protein concentrates and meals are very high in crude protein (62% and

44%, respectively) and in most EAA of which Arg constitutes about 4.1% and 2.8%,

respectively. Accurate estimation of biologically available AA in these high protein ingredients

would allow feed manufactures to further diversify the protein sources while enabling them to

attain greater degree of accuracy in formulation using these feedstuffs. In fish, RBV of Lys (El-

Haroun and Bureau, 2007) and Arg (Poppi et al., 2011) from blood meal and feather meal,

respectively in rainbow trout, and MHA-FA (Li et al., 2009) in juvenile striped bass have been

determined by linear regression method. Rainbow trout has high dietary Arg requirement (1.5%,

NRC, 2011), and responds very well and predictably to small changes in the dietary levels of Arg

(Cho et al., 1992; Rodehutscord et al., 1995). Since Indian mustard protein concentrate and

protein meals are rich in Arg (4.1% and 2.8%, respectively), dietary concentrations of this AA

can be manipulated by simply adding the ingredients to the Arg deficient basal diet. Information

of the bioavailability of Arg could potentially be a reliable indicator of the availability of other

84
AA. The major focus of this study was to determine the bioavailability of Arg from novel Indian

mustard protein concentrate and meal, and a commonly used soy protein concentrate (SPC)

compared to a crystalline source (98.5% pure L-Arg) in rainbow trout using a slope-ratio assay.

The efficacy of different regression approaches such as with two-level versus one-level of Arg

from the test and standard sources, and with level (% diet) and intake (g fish-1) as predictors was

assessed.

4.2. Methods

The growth trial was conducted at the UG/OMNR Fish Nutrition Research Laboratory

(Department of Animal and Poultry Science, University of Guelph, Ontario, Canada) using

standard research protocol. (1999). Indian mustard protein concentrate (IMC) and meal (IMM)

were obtained from Bio-Exx Specialty Proteins Ltd. (Toronto, ON, Canada), and the soy protein

concentrate (SPC HP 300) was obtained from Hamlet Protein A/S (Aarus, Denmark). The

Indian mustard protein concentrate and meal are produced using a proprietary low temperature

extraction process (<60oC) while the soy protein concentrate is produced through a proprietary

enzymatic treatment and aqueous extraction process. The proximate composition, gross energy

(GE), and amino acid profiles of three test ingredients are reported in Table 4.1.

85
Table 4.1 Proximate chemical composition, gross energy, essential amino acid profiles, total

phosphorus and phytic acid content of the test ingredients (as is basis).

Ingredients
Proximate composition (as is) Indian mustard Indian mustard Soy protein
concentrate meal concentrate
Dry matter, % 95 95 93
Organic matter, % 87 88 86
Crude protein, % 62 44 57
Lipid, % 0.2 1.9 2.2
Total carbohydrate, % 30 48 33
Neutral detergent fibre, % 13 17 7
Acid detergent fibre, % 11 13 6
Ash, % 8 7 8
Gross energy, kJ.g-1 19 19 19
Essential amino acids (% in the weight of ingredient as is)
Arginine 4.1 2.8 3.9
Histidine 1.6 1.1 1.4
Isoleucine 2.4 1.7 2.5
Leucine 4.2 2.9 4.1
Lysine 2.6 1.9 2.7
Methionine+Cystine* 2.8 1.9 1.3
Phenylalanine 2.3 1.6 2.7
Threonine 2.2 1.6 1.8
Tryptophan* 1.0 0.6 0.6
Valine 3.0 2.1 2.7
Phosphorus and phytate-P (as is)
Total phosphorus, % 1.5 1.1 0.8
Phytate-P, % 1.2 0.8 0.5

Phytate-P is calculated as phytic acid*0.282; * manufacturers data


86
4.2.1. Experimental diets

A practical basal diet was formulated (Diet 1, Table 4.2) to be highly deficient in arginine

(1.23% Arg) using a combination of corn gluten meal and skim milk powder. The diet contained

approximately 40% crude protein (CP) and 20% crude lipids (Table 4.3). Eight other isoproteic

and isoenergetic (on a digestible energy basis) experimental diets were prepared with graded

levels of arginine (~1.35% and ~1.5% of the diets) with increasing amount of three test

ingredients (IMC, IMM, and SPC) or crystalline L-Arg (98.5% pure; USB Corporation, OH,

USA) (Table 4.3). Levels of all nutrients (except Arg) met the requirements of rainbow trout

according to NRC (2011) recommendations. Ingredients and diets were mixed using a Hobart

mixer (Hobart Ltd., Don Mills, ON, Canada), steam-pelleted using a laboratory scale steam-

pellet mill (California Pellet Mill, San Francisco, CA, USA), and dried under forced-air at room

temperature for 24 hrs. Dried diets were sieved to appropriate size and stored at 4C until used.

The amount of feeds needed were weighed on a weekly basis and kept at room temperature.

4.2.2. Experimental conditions

Rainbow trout (Oncorhynchus mykiss, Ontario domestic fall-spawning strain) fingerlings

were obtained from the Alma Aquaculture Research Station of the University of Guelph (Elora,

ON, Canada). The experimental system, consists of four blocks of nine 60-L fibreglass tanks in

each, were individually aerated and supplied with filtered well water. Water flow, temperature

and photoperiod were maintained at ~3 L min-1, 13oC, and 12h L-12h D cycle, respectively. Fish

were pre-acclimatized to the experimental conditions for seven days, when they were fed a high

quality commercial salmonid diet (Profishent High Energy Feed, Martin Mills, Elmira, Ontario,

Canada).

87
Nine diets were allocated to each experimental block following a completely randomized

design. Fifteen fish, an average weight of 16 g, were distributed to each tank and hand-fed to

near-satiety three times a day on weekdays and once daily on weekends for 112 days. Feed

consumption was recorded weekly and the fish were weighed in every four weeks. The animals

were treated complying with the guidelines of the Canadian Council on Animal Care (CCAC,

1984) and the University of Guelph Animal Care Committee.

88
Table 4.2 Ingredients composition of the diets (g kg-1).

Ingredients (g kg-1 diet) Diet 1 Diet 2 Diet 3 Diet 4 Diet 5 Diet 6 Diet 7 Diet 8 Diet 9
Skim milk powder 250 250 250 235 210 235 210 235 210
Corn gluten meal 260 259 258 215 167 215 167 215 175
Indian mustard protein conc. - - - 70 140 - - - -
Indian mustard meal - - - - - 100 200 - -
Soy protein conc. - - - - - - - 75 150
Blood meal 40 40 40 35 30 35 30 35 30
Gelatinized starch 117 116.5 116 112 120 82 60 107 102
Wheat gluten meal 70 70 70 70 70 70 70 70 70
Fish oil 180 180 180 180 180 180 180 180 180
1
Vitamin premix 10 10 10 10 10 10 10 10 10
1
Amino acid premix 10 10 10 10 10 10 10 10 10
Lysine, BioLys (51% L-Lys) 35 35 35 35 35 35 35 35 35
DL-Met 3 3 3 3 3 3 3 3 3
L-Arg (98.5% pure) - 1.5 3 - - - - - -
Mineral premix3 10 10 10 10 10 10 10 10 10
NaH2PO4 15 15 15 15 15 15 15 15 15
Total (g) 1000 1000 1000 1000 1000 1000 1000 1000 1000
1
Provides per kg of diet: retinyl acetate (vitamin A), 3750 IU; cholescalciferol (vitamin D), 3000 IU; dl-a-tocopheryl
acetate (vitamin E), 75 IU; menadione sodium bisulphite (vitamin K), 1.5 mg; ascorbic acid polyphosphate (Stay
CTM 25% ascorbic acid), 75 mg; choline chloride, 1500 mg; cyanocobalamine (vitamin B12), 0.03 mg; biotin, 0.21
mg; inositol, 450 mg; folic acid, 1.5 mg; niacin, 15 mg; Pantothenic acid, 30 mg; pyridoxineHCl, 7.5 mg;
riboflavin, 9.0 mg; thiaminHCl, 1.5 mg.
2
Provides per kg of diet: tryptophan, 2000 mg; threonine, 3000 mg; histidine, 1500 mg.
3
Provides per kg of diet: sodium chloride (NaCl, 39% Na, 61% Cl), 3077 mg; ferrous sulphate (FeSO47H2O, 20%
Fe), 65 mg; potassium iodide (KI, 24%K), 11 mg; manganese sulphate (MnSO 4, 36% Mn), 89 mg; zinc sulphate
(ZnSO47H2O, 40% Zn), 150 mg; copper sulphate (CuSO45H2O, 25% Cu), 28 mg; di-sodium selenite (Na2SeO3),
10 mg.

89
Table 4.3 Nutrient, gross energy and essential amino acid (EAA) composition of the test diets
(DM basis).
Diet 1 Diet 2 Diet 3 Diet 4 Diet 5 Diet 6 Diet 7 Diet 8 Diet 9
Proximate Composition (Determined)
Dry matter (%) 95 93 94 93 93 96 95 94 94
Crude protein (%) 40 41 40 42 41 40 40 39 40
Lipid (%) 19 20 19 19 19 20 19 20 20
Ash (%) 5 5 5 5 5 5 5 5 6
1
NDF (%) 2.2 2.3 2.3 3.7 4.3 4.1 5.1 3.1 3.1
ADF (%)2 1.9 1.9 2.0 2.6 3.0 3.8 4.7 2.1 2.4
Phosphorus (%) 0.8 0.9 0.9 1.0 1.0 0.9 0.9 0.9 0.9
Phytate-P (%)3 0.1 0.1 0.1 0.2 0.3 0.2 0.3 0.1 0.1
Available P (%) 0.7 0.8 0.7 0.8 0.6 0.7 0.7 0.7 0.7
-1
GE (kJ g ) 22.3 22.3 22.2 22.2 21.9 22.1 22.1 22.2 21.9
Digestible Nutrient Composition (Calculated)
Digestible Protein (%) 95 95 95 95 94 96 96 96 96
Digestible Energy (kJ g-1) 21.5 21.7 21.4 21.4 21.4 20.8 21.1 21.4 21.2
Essential amino acid composition (Analyzed)4
Arginine 1.23 1.37 1.54 1.36 1.50 1.34 1.49 1.38 1.53
Histidine 0.98 1.04 0.99 1.04 1.01 1.04 1.02 1.01 1.04
Isoleucine 1.39 1.45 1.39 1.33 1.39 1.39 1.40 1.46 1.44
Leucine 3.95 4.11 3.85 3.56 3.31 3.57 3.42 3.64 3.37
Lysine 2.70 2.73 2.68 3.03 2.77 2.88 2.77 3.07 3.15
Methionine + Cysteine* 1.46 1.46 1.45 1.50 1.53 1.50 1.52 1.44 1.42
Phenylalanine + Tyrosine 2.98 3.08 2.91 2.79 2.62 2.81 2.78 2.90 2.87
Threonine 1.31 1.36 1.29 1.27 1.30 1.31 1.34 1.40 1.38
Tryptophan* 0.54 0.54 0.53 0.51 0.48 0.51 0.49 0.60 0.65
Valine 1.75 1.79 1.71 1.73 1.71 1.73 1.74 1.75 1.75
1
NDF Neutral Detergent Fibre. 2ADF Acid Detergent Fibre. 3Analyzed as phytic acid (PA). Phytate-P is
calculated as 0.282*PA. 4Methionine, cysteine, and tryptophan were not analyzed. *calculated values.

90
4.2.3. Chemical analysis

At the beginning and end of the experiment, 36 fish (one from each tank) and five fish

from each tank, respectively, were sampled for proximate composition and kept at -20oC until

processed. The pooled whole fish samples were cooked in an autoclave, ground into

homogeneous slurry in a food processor, lyophilized, then finely ground, and stored at -20oC

until analysis.

Diets, ingredients, and whole carcass samples were analyzed for dry matter (DM) and ash

according to AOAC (1995), crude protein (CP, %N x 6.25, Dumas method) using a LECO

Analyzer (LECO Corporation, St. Joseph, MI), lipids with petroleum ether extraction at 90oC

and 150 PSI for 60 min using ANKOM XT-20 analyzer (ANKOM Technology, Macedon, NY),

and gross energy (GE) using an automated oxygen bomb calorimeter (Model # 1271, Parr

Instruments, Moline, Illinois). Neutral detergent fibre (NDF) and acid detergent fibre (ADF)

contents of the ingredients and the diets were determined according to Van Soest et al. (1991)

using an Ankom 200 Fibre Analyzer (ANKOM Technology, Macedon, NY). Phytic acid content

in the ingredients and diets were determined using high performance liquid chromatography

(HPLC, Waters Corporation, Millford, CA) according to Lehrfeld (1989) at the Danisco Animal

Nutrition laboratory, Danisco, St. Louis, MO. Amino acid profiles of the test ingredients and the

diets were analyzed at the Amino Acid Laboratory of the Toronto Hospital for Sick Children

(ON, Canada) using the acid hydrolysis method. For the quantification of amino acids,

approximately 0.05 g of dried sample was hydrolysed in 15 ml test tube with 450 L of 6N HCl

and 1% phenol for 24 hrs at 110oC under purified nitrogen atmosphere. After 24 hrs, tubes were

removed from oven, cooled, capped, vortex mixed and centrifuged at 2500 rpm. Then, 5 L of

the supernatant was derivatized following PICO-Tag method (Cohen et al., 1989). After

91
derivatization, the supernatant were re-dissolved in 150 L phosphate buffer and 2 L of this

solution was injected into the column of an ACQUITY Ultra-Performance LC machine (Waters

Corporation, Millford, CA) for quantification of all amino acids except tryptophan, methionine

and cysteine, which were partially destroyed in the hydrolysis process.

4.2.4. Calculations

Growth rate was expressed as thermal-unit growth coefficient (TGC) and calculated as

[100 x (FBW1/3 IBW1/3) / (Temp (0C) x Number of days)] where FBW is the final body

weight and IBW is the initial body weight of the fish, and expressed as thermal-unit growth

coefficient (TGC) (Cho, 1992). Feed efficiency (FE) was calculated as live weight gain /dry

feed intake. Protein (PD) and lipid (LD) deposition, nitrogen retention efficiency (NRE),

recovered energy (RE), and energy retention efficiency (ERE) were also assessed. PD, LD and

RE were calculated from the difference between final and initial carcass content. NRE and ERE

were calculated as: NRE, % = [[FBW x Nfinal) (IBW x carcass Ninitial)]/DNI] x 100 and ERE, %

= [[FBW x GEcarcassfinal) (IBW x GEcarcassinitial)]/DEI] x 10, where: FBW = final body

weight (g) where IBW = initial body weight (g), DNI = digestible nitrogen intake, and DEI =

digestible energy intake.

4.2.5. Statistical analysis

All data were subjected to Levenes tests (Levene, 1960) of equality of variance followed

by regression analysis against dietary Arg levels using the SAS general linear model procedure

(PROC GLM, SAS v9.2, SAS Institute, Cary, NC, USA). Group means of feed intake, weight

gain, TGC, FE, PD, LD, NRE and ERE were compared using least square means (LSmeans

92
statement in the mixed model analysis) and considered significant at the Bonferroni adjusted

level of P<0.0125. Effects of dietary Arg levels, source, and their interactions were compared

among all sources, and between protein-bound (IMC, IMM, and SPC) and free (L-Arg) sources

of Arg were also analyzed using the PROC GLM procedure and considered significant at

P<0.05.

The responses to Arg level from all four sources were tested for linearity for the

compliance with the criteria necessary for linear slope-ratio model using non-orthogonal

polynomial contrast separately for each source and considered significant at the Bonferroni

adjusted level of P<0.0125. The x, y values for the intercept was calculated from the response to

the basal diet (1.23% Arg) satisfying the requirement for a common intercept (Littel et al., 1997).

RBV estimates of arginine of the test ingredients were derived with one and two levels of

arginine by comparing the slopes of IMC, IMM and SPC with that of L-Arg with either Arg level

(% diet) or intake (g fish-1) as predictors. The difference between the slopes was determined

using 95% confidence intervals and considered significant at P<0.05.

4.3. Results

Fish accepted the basal diet well, and feed intake and growth rate were good for all diets

and comparable to those obtained in other feeding trials at the University of Guelph. Weight gain

of fish fed the basal diet (1.23% Arg) was significantly lower (P<0.05) than those fed high

arginine diets (~1.5% Arg) formulated with IMM, IMC and SPC. Protein, lipid and energy gains

of fish fed these three diets were also significantly higher than those of fish fed all the other

diets.

93
4.3.1. Growth performance and nutrient deposition

Growth rate and protein deposition (PD, g fish-1) in response to intake are presented in

the Figures 4.1 and 4.2, respectively. Fish fed the basal diet had significantly lower (P<0.05)

feed intake, weight gain (Table 4.4), and PD and NRE (Table 4.5) than the fish fed diets with

~1.35% or ~1.5% levels of Arg. Feed intake, weight gain, TGC, PD and LD but not FE, RE,

NRE and ERE were significantly different (P<0.05) between free and protein-bound sources.

Energy retention efficiency also increased linearly (P<0.0125) with increasing level of IMC but

decreased significantly with increasing levels of IMM in the diets (P<0.0125).

4.3.2. Relative bioavailability of arginine

TGC and PD increased linearly with increasing level of Arg and were subjected to the

slope ratio-assay. In general, bioavailability of Arg was higher from all protein bound sources

than those from L-Arg. The RBV of arginine from IMC calculated with PD as response variable

were 178% and 123% higher than that from the standard source, L-Arg (assumed as 100%) for

the predictor variables Arg level (%) and Arg intake (g fish-1), respectively (P<0.05; Tables 4.6,

4.7; Figures 4.3 and 4.4). RBV of Arg from IMM was also significantly higher than that from L-

Arg for Arg level as predictor variable but not for the Arg intake (Tables 4.6, 4.7). On the other

hand, RBV of Arg from SPC was slightly higher but not significantly different than that from the

L-Arg (P>0.05).

94
Table 4.4 Weight gain, TGC and feed intake and feed efficiency of rainbow trout fed diets with
increasing levels of arginine supplied by L-Arg or three protein-bound sources (mean SD).
Feed
Arginine Feed intake Weight gain TGC
Source efficiency
% DM g fish-1 g fish-1 gain:feed
L-Arginine
Diet 1 1.23 85.7 13.7 87.3 14.8 0.168 0.018 1.02 0.05
Diet 2 1.37 102.3 14.5 105.4 13.4 0.187 0.012 1.03 0.05
Diet 3 1.54 103.1 12.4 117.0 16.6 0.200 0.014 1.13 0.04
SEM1 5.7 5.4 0.05 0.04
Linear NS P<0.0125 P<0.0125 P<0.0125
Quadratic NS NS NS NS
Indian mustard protein concentrate
Diet 4 1.36 105.2 201 118.4 19.6 0.201 0.018 1.13 0.03
Diet 5 1.50 115.5 16.0 130.3 11.9 0.212 0.009 1.13 0.06
SEM 8.7 6.9 0.01 0.04
Linear P<0.0125 P<0.0125 P<0.0125 P<0.0125
Quadratic NS NS P<0.0125 NS
Indian mustard meal
Diet 6 1.34 108.7 6.6 117.4 5.3 0.200 0.004 1.08 0.02
Diet 7 1.49 116.5 7.3 124.4 6.4 0.209 0.009 1.07 0.02
SEM 9.3 5.5 0.006 0.02
Linear P<0.0125 P<0.0125 P<0.0125 NS
Quadratic NS NS NS NS
Soy protein concentrate
Diet 8 1.38 108.5 10.7 115.2 13.5 0.197 0.011 1.06 0.04
Diet 9 1.53 118.8 5.3 127.7 6.5 0.210 0.006 1.07 0.01
SEM 9.8 6.0 0.006 0.02
Linear P<0.0125 P<0.0125 P<0.0125 NS
Quadratic NS NS NS NS
Effects (P-value, all sources)
Ingredient NS NS NS NS
Level ** * * NS
Ingredient*Level NS NS NS *
Effects (P-value, free & protein-bound)
Ingredient * * ** NS
Level ** * ** **
Ingredient*Level NS NS NS **
1
S.E.M. standard error mean; NS = not statistically significant, * = P<0.05, ** = P<0.01.

95
Table 4.5 Protein (PD) and lipid (LD) deposition, recovered energy (RE), N- (NRE) and
energy (ERE) retention efficiencies for rainbow trout fed diets with increasing levels of arginine
(mean SD).
Arg PD LD RE NRE ERE
Source
%DM g fish-1 g fish-1 kJ fish-1 % %
L-Arginine
Diet 1 1.23 13.1 2.3 14.1 3.2 8.7 1.6 40.4 2.7 30.2 2.7
Diet 2 1.37 16.5 1.4 17.4 3.0 10.9 1.4 42.2 3.3 33.7 3.8
Diet 3 1.54 17.9 2.6 18.9 3.8 11.7 1.3 45.5 2.1 38.3 1.3
1
SEM 0.7 0.8 0.5 1.6 2.3
Linear P<0.0125 P<0.0125 NS NS NS
Quadratic NS NS NS NS NS
Indian mustard concentrate
Diet 4 1.36 18.1 2,8 19.8 3.9 12.2 2.1 44.1 1.5 38.8 1.3
Diet 5 1.50 20.8 2.4 21.6 2.8 13.3 1.2 46.0 3.4 39.7 2.6
SEM 1.2 1.2 0.7 1.0 0.5
Linear P<0.0125 P<0.0125 P<0.0125 NS P<0.0125
Quadratic NS NS P<0.0125 NS NS
Indian mustard meal
Diet 6 1.34 18.4 0.2 19.8 1.4 12.2 0.9 43.6 2.6 40.0 1.6
Diet 7 1.49 19.5 1.3 19.8 1.4 12 0.8 42.9 1.4 36.1 1.9
SEM 0.9 1.0 5.7 0.4 2.0
Linear P<0.0125 P<0.0125 P<0.0125 NS NS
Quadratic NS NS NS NS P<0.0125
Soy protein concentrate
Diet 8 1.38 17.7 2.3 19.0 2.7 11.3 1.3 44.0 2.6 34.4 1.2
Diet 9 1.53 20.1 1.2 20.8 1.2 12.6 0.5 45.6 1.3 35.6 1.3
SEM 1.0 1.0 0.6 0.8 1.1
Linear P<0.0125 P<0.0125 P<0.0125 NS NS
Quadratic NS NS NS NS NS
Effects (P-value, all sources)
Ingredient NS NS NS NS **
Level * NS NS NS NS
Ingredient*Level NS NS NS NS **
Effects (P-value, free & protein-bound)
Ingredient * * NS NS NS
Level * NS NS * NS
Ingredient*Level NS NS NS NS NS
1
S.E.M. standard error mean; NS = not statistically significant, * = P<0.05, ** = P<0.01.

96
Table 4.6 Slopes (b) and confidence intervals (CI) of thermal-unit growth coefficients (TGC)
and protein deposition (PD) in fish fed diets from different sources of arginine (L-arginine,
Indian mustard protein concentrate - IMC, Indian mustard protein meal - IMM, soy protein
concentrate - SPC) as determined by linear regressions using two and one level of arginine (%)
and arginine intake (g fish-1).
Thermal-unit growth coefficient Protein deposition (g fish-1)

Source Two levels One level Two levels One level


Slope Slope Slope Slope
CI CI CI CI
Arginine (% diet)
L-Arg 0.110 0.139 17.2 24.8
0.077 to 0.143 0.071 to 0.208 11.2 to 23.1 15.0 to 34.7

IMC 0.183* 0.260 30.6* 39.0


0.138 to 0.229 0.153 to 0.366 22.3 to 38.8 20.2 to 57.7

IMM 0.179* 0.293* 28.4* 48.2*


0.142 to 0.229 0.265 to 0.322 22.0 to 34.8 43.5 to 52.8

SPC 0.152 0.195 24.9 31.1


0.125 to 0.178 0.138 to 0.252 20.6 to 29.3 20.5 to 41.6
Arginine intake (g fish-1)
L-Arg 0.054 0.05 8.6 8.2
0.048 to 0.061 0.038 to 0.061 7.3 to 10.0 6.1 to 10.4

IMC 0.063 0.076* 10.6* 11.5*


0.055 to 0.070 0.069 to 0.082 9.6 to 11.6 11.0 to 12.1

IMM 0.063 0.077* 10.0 12.4*


0.056 to 0.070 0.071 to 0.084 8.7 to 11.2 11.1 to 13.7

SPC 0.054 0.061 8.9 9.65


0.050 to 0.058 0.056 to 0.066 8.3 to 9.5 8.5 to 10.8
IMC = Indian mustard protein concentrate; IMM = Indian mustard protein meal; SPC = Soy protein concentrate.

Linear model = a+b1*x1+b2*x2+b3*x3+b4*x4; a (TGC) = 0.1675; a (PD) = 13.1. CI = Confidence Interval (lower

limit to upper limit). *Significantly different than L-Arg (P<0.05).

97
Table 4.7 Bioavailability of arginine from Indian mustard protein concentrate - IMC, Indian
mustard protein meal - IMM, soy protein concentrate - SPC in rainbow trout in compare to L-
arginine based on thermal-unit growth coefficients (TGC) and protein deposition (PD) as
determined by linear regressions using two and one level of arginine (%).
Thermal-unit growth coefficient Protein deposition (g fish-1)
Two levels One level Two levels One level
Source Arg availability Arg availability Arg availability Arg availability
(%) (%) (%) (%)
R2 R2 R2 R2
Arginine (% diet)
L-Arg 100 100 100 100
65 62 56 72

IMC 166 187 178 157


71 70 69 63

IMM 163 211 165 194


79 98 76 98

SPC 138 140 145 125


84 82 84 78
Arginine intake (g fish-1)
L-Arg 100 100 100 100
93 90 89 86

IMC 117 152 123 140


94 99 95 100

IMM 117 154 116 151


94 98 91 97

SPC 100 122 103 118


97 99 97 97
IMC = Indian mustard protein concentrate; IMM = Indian mustard protein meal; SPC = Soy protein concentrate.

Linear model = a+b1*x1+b2*x2+b3*x3+b4*x4; a (TGC) = 0.1675; a (PD) = 13.1.

CI = Confidence Interval (lower limit to upper limit) .

98
Figure 4.1 Weight gain (g fish-1) and thermal-unit growth coefficient (TGC) of rainbow trout in

response to dietary arginine intake.

0.220
Thermal-unit growth coefficient

0.200

0.180

0.160

TGC = 0.06x + 0.11, R = 0.86


0.140
0.50 1.00 1.50 2.00 2.50
Arginine intake (g fish-1)

99
Figure 4.2 Protein (PD) and lipid (LD) deposition in rainbow trout in response to dietary

arginine intake (g fish-1).

25
Protein deposition (g fish-1)

20

15

PD = 9.3x + 2.9, R = 0.86


10
0.5 1.0 1.5 2.0 2.5
Arginine intake (g fish-1)

100
Figure 4.3 Thermal-unit growth coefficient (TGC) and protein deposition (PD) of rainbow

trout in response to arginine level (%) in diets from four arginine sources (L-Arg, IMC - Indian

mustard protein concentrate, IMM -Indian mustard protein meal, SPC -soy protein concentrate).

Solid and dotted lines represent response based on two and one level of arginine, respectively.

0.240 L-Arg IMC IMM SPC


L-Arg IMC IMM SPC
Thermal-unit growth

0.220
coefficient

0.200

0.180
(A)

0.160
1.23 1.27 1.31 1.35 1.39 1.43 1.47 1.51

23.0
Protein deposition (g fish-1)

21.0

19.0

17.0

15.0
(B)
13.0
1.23 1.27 1.31 1.35 1.39 1.43 1.47 1.51
Arginine (% in diet)

101
Figure 4.4 Thermal-unit growth coefficient (TGC) and protein deposition (PD) of rainbow

trout in response to arginine intake (g fish-1) from four arginine sources (L-Arg, IMC - Indian

mustard protein concentrate, IMM -Indian mustard protein meal, SPC -soy protein concentrate).

Solid and dotted lines represent response based on two and one level of arginine, respectively.

0.22 L-Arg IMC IMM SPC


L-Arg IMC IMM SPC
0.21
Thermal-unit growth coefficient

0.20

0.19

0.18

0.17

0.16
1.10 1.30 1.50 1.70 1.90
Arginine intake (g fish-1)

L-Arg IMC IMM SPC


22.0 L-Arg IMC IMM SPC
Protein deposition (g fish-1)

20.0

18.0

16.0

14.0

12.0
1.10 1.20 1.30 1.40 1.50 1.60 1.70 1.80 1.90
Arginine intake (g fish-1)

102
4.4. Discussion

Relative bioavailability of Arg from Indian mustard protein concentrate and meal, and a

soy protein concentrate compared to that from crystalline L-Arg in rainbow trout was assessed in

this study. The increase in weight gain, FE and TGC (P<0.05) with increasing dietary Arg

supplied by crystalline L-Arg confirmed that the basal diet was deficient in Arg and the high Arg

levels in the test diets were around the requirement for maximum weight gain currently estimated

to be about 1.5% of dietary DM (NRC, 2011).

Fish fed the crystalline L-Arg based diets showed considerably lower growth

performance such as weight gain, feed efficiency, and nutrient utilization than the fish fed the

protein-bound Arg diets. Various studies earlier have reported lower amino acid utilization by

fish fed crystalline amino acids than those fed protein bound sources (Ambardekar et al., 2009;

El-Haroun and Bureau, 2007; Peres and Oliva-Teles, 2005; Poppi et al., 2011; Zarate and Lovell,

1997). Reasons for the low AA utilization suggested are nutrient leaching and catabolism of free

amino acids being absorbed faster than the protein-bound amino acids (Ambardekar et al., 2009;

Chi et al., 2011; Zarate and Lovell, 1997; Zarate et al., 1999). Earlier studies in growing swine

indicated that increasing the frequency of feeding could improve utilization of free amino acids

(Batterham and Murrison, 1981). Recent studies in fish and shrimp suggested that using coated

amino acids could also possibly improve the utilization of supplemental amino acids (Alam et

al., 2004, 2005; Chi et al., 2011; Lemme, 2010).

Conditions for a valid slope-ratio assay, such as linearity of the responses, common

intercept for standard and tests, and the similar Y values of the common point of intersection,

were met by confirming the linearity of TGC and PD, forcing a common intercept, and the single

data point at the lowest Arg level, respectively. TGC and PD increased linearly with increasing

103
Arg levels in all four protein sources (Tables 4.4 and 4.5). When Arg level is the predictor, RBV

of Arg in IMC and IMM were significantly higher in one-level assay than those in two-level

assays suggesting possible curvilinear or levelling of the responses at higher Arg levels. One

level assays are rare compare to two or more levels (e.g., Batterham et al., 1978; Batterham,

1987). Batterham et al. (1978) compared bioavailability of lysine from cottonseed meal, soybean

meal, meat and bone meal and fish meal for growing swine determined by one (0.3% of the

diets) and seven (0.3% and six other levels with 0.05% increments) levels of Lys and reported no

difference in the availability of lysine among the test ingredients except soybean meal. This

could potentially be a particularity of the experimental design where the maximum level of

dietary Lys experiment (0.6%) was still much lower than the requirement (0.8%) of growing

swine providing similar slopes in both experiments (Martinez and Knabe, 1990).

Recently, Levesque et al. (2010) suggested that maximum intake of the test amino acid

should be set at least 2 SD below the requirement based on the 10% CV estimated by Bertolo et

al. (2005) because of the inter-animal variability of AA requirement. Restricting test amino acid

intake to 80% of its requirement ensures a linear response to the changes in the test amino acid

intake. In this study, the fish were fed to satiety resulting in a variable Arg intake that was

significantly correlated with the protein deposition (R2 = 0.85; P<0.05). In such scenario, AA

intake (g fish-1) appeared to be a better predictor for the determination of bioavailability than the

dietary level (%).

Two important dietary considerations for slope-ratio assays are that the response to

changes in test amino acid intake must be predictable and the observed response must not be

influenced by other dietary nutrients in the test feed ingredient (Batterham, 1992; Littel et al.,

1997). The latter condition is difficult to achieve since most protein sources are complex

104
ingredients with variable levels of non-starch polysaccharides, anti-nutritional factors etc. Like

other farmed animals, non-starch polysaccharides appear to effect nutrient digestion and

absorption by modulating digesta viscosity and rate of passage and by altering gut physiology

and morphology in fish (Brinker and Reiter, 2011; Glencross et al., 2008; Leenhouwers et al.,

2007; Refstie et al., 1999; Storebakken, 1985). Besides, the composition and source of amino

acids in the diets also play a role in feed intake and nutrient utilization. For example, Nang thu et

al. (2007) observed that required amount of lysine intake for protein deposition in rainbow trout

differed significantly between WG and CGM diets suggesting that ingredient matrix may play a

role in amino acid composition. High Arg:Lys ratio or high leucine content in CGM may have

resulted in a reduced feed intake and lysine utilization efficiencies in fish compared to those fed

WG diets. Efficiency of amino acid utilization is usually affected by dietary composition and

source of amino acids, dietary energy level and source, and other components (Encarnao et al.,

2004; Rodehutscord et al., 2000). Meaningful comparison of RBV estimates between different

sources therefore, can be difficult when variable composition of other nutrients or anti-nutrients

in the sources aside from the amino acid of interest affecting the feed intake, nutrient digestion

and utilization.

It can be concluded that the bioavailability of arginine from novel Indian mustard protein

concentrate (IMC) and protein meal (IMM) for rainbow trout is comparable to that of the soy

protein concentrate (SPC) used in this study. Similar performances of fish fed these two novel

plant protein sources compared to those fed commercially available SPC suggest that the quality

of these ingredients is very good, and can be readily utilized in nutrient-dense fish feeds.

Practical basal diets in bioavailability studies in fish have been formulated with either CGM or

WG as the major protein source (Li et al., 2008, 2009; Nang thu et al., 2007; Poppi et al., 2011).

105
Nangthu et al. (2007) reported that imbalance amino acid composition in CGM induced by

excess leucine and high Arg-Lys ratio, may have affected the feed intake suggesting WG could

be a superior candidate as major protein ingredient in basal or reference diets than the CGM. The

difference in the lysine utilization efficiencies observed by Nang thu et al. (2007) could also be

an effect of the variations in digesta viscosity and passage rate because of variable composition

of NSPs in these two ingredients. Besides the variable compositions of other nutrients or anti-

nutrients in the test sources, variations in ingredient compositions of the basal diets may also

hinder from the meaningful comparison of the results among the slope-ratio studies.

Applications of the slope-ratio assay to determine bioavailability of an amino acid could

be very useful tool to assess the quality of novel protein sources in aquafeed. Determination of

appropriate statistical approach (e.g., linear, polynomial or broken-line methods) is essential

before designing a bioavailability trial. Some important points to be considered are: common

intercept of the responses for individual sources, appropriate feeding regime to maintain the level

of test nutrient intake reflecting its level in the diets, and adjusting composition of other

ingredient components such as fibre or anti-nutritional factors by supplementation to nullify their

effects on indicator responses. Further studies are also needed to standardize the practical basal

diets and as well as, to establish a set of criteria for the level of the nutrients of interest in the test

diets for valid, robust and useful slope-ratio responses.

106
CHAPTER 5

EFFECTS OF DIETARY PHYTIC ACID AND PURIFIED LIGNIN AND

THEIR INTERACTIONS ON MORPHOLOGICAL INDICES, GROWTH

PERFORMANCE, NUTRIENT AND ENERGY PARTITIONING OF

RAINBOW TROUT, ONCORHYNCHUS MYKISS

Abstract

This study examined the effects of dietary phytic acid (PA) and purified lignin, and their

interactions on growth performance, physiological indices, nutrient deposition and partitioning in

rainbow trout in a 12 week trial. Nine isoproteic and isoenergetic diets were formulated differing

only in their PA and lignin concentrations. Diets were formulated to be marginally adequate in

five essential amino acids, histidine, lysine, methionine (+cysteine), threonine, and tryptophan.

Fish in six dietary treatments (2 PA X 3 lignin levels) were pair-fed with the treatment with the

lowest feed intake. Fish in three other high dietary PA treatments (2.25%) were fed to satiety,

since their feed intake was much lower (48-58 g fish-1) than those in the pair-feeding treatments

(78 g fish-1). Dietary PA levels affected only the Fultons condition index (FCI) and whole

carcass nitrogen retention efficiency (NRE) (P<0.05) while lignin levels in the diets significantly

affected the whole carcass lipid deposition (LD) and lipid retention efficiency (LRE) among the

pair-fed fish. On the other hand, when fed to satiety, feed intake by fish fed the lignin diets was

significantly reduced contributing to the differences in whole carcass recovered energy (RE) and

ERE and dressed carcass PD, RE, NRE and ERE between the fish fed diets with and without

lignin (P<0.01).

107
5.1. Introduction

Plant protein sources are increasingly used in aquaculture feed to satisfy the growing

need for more economical ingredients. These ingredients contain wide variety of compounds,

some with anti-nutritional properties, such as protease inhibitors, phytic acid, glucosinolates,

saponins, polyphenols, oligosaccharides and non-starch polysaccharides (e.g., mannan, galactan,

pectin, arabinose), lignin, haemaglutanin (e.g., lectins), antivitamins (e.g., anti-pyridoxine,

oxidase, anti-nicotinic acid, cyanogens), or antigenic compounds (e.g., glycinin and beta-

conglycinin (Francis et al., 2001; Tacon, 1997). Various factors such as amino acid composition,

anti-nutrients, and fibre components of a plant protein ingredient can affect its nutritive value

and the amount that can be used in the diets of various fish species (Bell 1993; Bureau et al.,

1998; Francis et al., 2001; Higgs et al., 1995; Mwachireya et al., 1999; Refstie et al., 2001;

Sanden et al., 2005).

The scope of most digestibility and bioavailability studies has been limited to

determining the effect of incorporation of various plant protein ingredients at the expense of fish

meal and other feedstuffs. These diets are formulated to the essential amino acid levels greatly in

excess of the requirement of the animal and result in poor characterization of the test ingredients.

The poor knowledge on the nutrients available to an animal necessitates the use of broader safety

margins when formulating feeds, reducing the ability to formulate on a truly least cost basis, and

forcing feed formulators to impose severe restrictions on the use of poorly characterized

ingredients (Bureau, 2006). Accurate characterization should extend beyond the assessment of

digestibility and bioavailability of nutrients to the effect of anti-nutritional factors and fibre

components on nutrient utilization in an animal.

108
In salmonids, studies with ANF mostly dealt with the effect of their inclusion levels on

growth performance and nutrient utilization (Burel et al., 2000; Cheng and Hardy, 2003; Kaushik

et al., 1995; Pratoomyot et al., 2010; Santigosa et al. 2010; Storebakken et al., 1998b). Growth

and dietary protein and mineral utilization in rainbow trout have been reported to be affected by

various anti-nutritional factors. For example, phytic acid (Kaushik et al., 1995; Spinelli, 1993;

Storebakken et al., 1998b; Teskeredzic et al., 1995; Wang et al., 2009a,b), soyasaponin (Bureau

et al., 1998; Knudsen et al., 2008), glucosinolate (Burel et al., 2000), or soluble and insoluble

non-starch polysaccharides (Amirkolaie et al., 2005; Enes et al., 2008; Glencross et al., 2008;

2011). Phytic acid is probably one of the most studied ANF in fish nutrition because of effects on

availability of phosphorus, and other valuable minerals and amino acids. Proton disassociation

sites on the PA molecules have variable pKa values between 1.5 and >10.0. They become

negatively charged when the proton is disassociated attracting positively charged molecules at

favourable pH (Adeola and Sands, 2003). Bivalent cations of minerals (e.g., Zn2+, Fe2+/3+, Ca2+,

Mg2+, Mn2+ and Cu2+) are particularly vulnerable to this process, creating phytic acid-mineral

binary complexes indigestible to the animal(Denstadli et al., 2006; Laining et al., 2010). Phytic

acid also interacts with protein creating binary (PA-protein) or ternary (PA-mineral-protein)

complexes that occur only at optimal pH conditions in the GIT (Cheryan and Rakis 1980; Kies et

al., 2001; Maenz, 2001). The -NH2 terminal group and the -NH2 group of Lys, the imidazole

group of His, and the positively charged guadinyl group of Arg have been implicated as possible

protein binding sites for phytate at low pH (Cosgrove, 1966). Ternary complexes are rare but

usually formed in slightly alkaline environment in the intestine (Riche et al., 2001).

Dietary fibre is defined as the sum of plant polysaccharides and lignin that are not

hydrolysed by endogenous digestive enzymes (Kritchevesky, 1988; Wenk, 2001). Dietary fibres

109
can also be divided into two physico-chemical groups soluble (pectins, beta-glucans and gums)

and insoluble fibres (cellulose, hemi-cellulose and lignin). Most studies on dietary fibre from

various plant sources dealt mainly with inclusion levels and their effects on nutrient utilization

and growth of an animal such as swine (Laplace et al., 1989; Mitaru et al., 1984; Souffrant,

2001), poultry (Slominski and Campbell, 1990; Villamide and San Juan, 1998), or fish

(Anderson et al., 1984; Bromley and Adkins, 1984; Hemre et al., 1995; Kraugerud et al., 2007).

Recent studies have shown that not the dietary fibre but the type and composition of its

components influences digestive processes in an animal (Amirkolaie et al., 2005; Glencross,

2009; Hansen and Storebakken, 2007; Zahedifar et al., 2002; Zhu et al., 2005).

Lignin, a highly polymerized hydrophobic phenolic plant compound, forms an integral

part of the secondary cell walls of most plants and algae. These polymers are composed mainly

of three different monolignols - sinapyl, coniferyl, and p-coumaryl alcohols, upon incorporation,

which are referred to as syringyl (S), guaiacyl (G), or hydroxyphenyl (H) units of lignin,

respectively. The proportion of S, G, and H varies among plant species, and can substantially

impact the physical properties of the polymer (Baurhoo et al., 2008; Boerjan et al., 2003). It is

hypothesized that monolignols are stabilized to lignin by the enzymes -glucosidases and

glucosyl transferases (Dharmawardhana and Ellis, 1998).

In vitro studies showed a significantly greater binding capacity of lignin with bi-valent

minerals such as Fe+2, Cu+2 and Zn+2 at intestinal pH (e.g., pH ~5) compared to that of pectin and

cellulose (Platt and Clydesdale, 1987). Hydroxyl and methoxy groups of lignin molecules under

pH conditions similar to those in the proximal intestine of animals form highly stable

hexadentate complexes with iron (Fernandez and Philipis, 1982). Affinity of these two binding

sites to minerals are in the order of Fe > Cu > Zn.

110
Under oxidative conditions, lignin also forms highly stable protein complexes resistant to

degradation by microbial fermentation or by acid hydrolysis (Zahedifar et al., 2002). In a study

involving three amino acids, proline, serine and glycine, Whitmore (1982) reported high binding

capacity of lignin with proline followed by serine and glycine. The author hypothesized that

proteins, especially those containing hydroxyproline, could be very susceptible to bonding with

dehydrogenation polymer of lignin under oxidizing conditions. Despite evidences of lignin-

mineral or lignin-protein bonding from in-vitro studies, there is a very limited effort to

understand the effect of lignin on performance of commercially important farmed animals.

A comprehensive understanding of the effect of PA, lignin and lignin-PA interactions on

nutrient utilization and performance of commercially important fishes would provide a strong

basis for the evaluation of novel plant protein ingredients in aquafeed. This study examined the

effects of graded levels of PA and purified lignin inclusion in a practical diet formulated to be

marginally adequate in some essential amino acids, and their interactions on nutrient utilization

in rainbow trout in a 12 weeks growth trial.

5.2. Methods

The growth trial was conducted at the UG/OMNR Fish Nutrition Research Laboratory

(Department of Animal and Poultry Science, University of Guelph, Ontario, Canada) for 12

weeks using standard research protocol. Phytic acid (50 wt% in aqueous solution) was obtained

from Sigma-Aldrich Co. (St. Louis, MO). Purified lignin was obtained from Pure Lignin

Environmental Technology, Kelowna, BC, Canada. The lignin, produced through a proprietary

close-loop process with low-pressure steam, was composed of 76% lignin and 24% cellulose. A

basal mix was prepared composed mainly of animal proteins i.e., herring fish meal, meat and

111
bone meal, regular poultry by-product meal, and spray-dried blood meal and a plant protein, corn

protein concentrate (Empyreal 75, Cargill Corn Milling, Blair, NB) (Table 5.1). Dextrin,

vitamin and mineral premixes, L-lysine, DL-methionine, di-calcium phosphate, and NaCl were

also added to reflect the composition of a typical trout feed. All the ingredients were mixed using

a large Marion mixer (Marion, IA).

5.2.1. Experimental diets

Nine isoproteic and isoenergetic experimental diets were prepared by adding fish oil and

appropriate amount of kaolin, cellufil, phytic acid, and purified lignin to the basal mix (Table

5.2). All diets were formulated to contain 2% cellulose. The concentration of all EAA but Arg,

Ile, Leu, Phe, and Val were formulated to be marginally above the requirement of rainbow

trout in all nine diets (NRC, 2011; Table 5.3). Basal mix and other ingredients were mixed using

a Hobart mixer (Hobart Ltd., Don Mills, ON, Canada), steam-pelleted using a laboratory scale

steam-pellet mill (California Pellet Mill, San Francisco, CA, USA), and dried under forced-air at

room temperature for 24 hrs. Dried diets were sieved to appropriate size and stored at 4C until

used. The amount of feeds needed were weighed on a weekly basis and kept at room

temperature.

5.2.2. Experimental conditions and feeding management

Rainbow trout (Oncorhynchus mykiss, Ontario domestic fall-spawning strain) fingerlings

were obtained from the Alma Aquaculture Research Station (Elora, ON, Canada). The

experimental system consists of three blocks of nine 60-L fibreglass tanks in each. Tanks were

supplied with filtered well water and individually aerated. Water flow, temperature and

112
photoperiod were maintained at ~3 L min-1, 13oC and 12h L 12h D cycle, respectively. Fish

were pre-acclimatized to the experimental conditions for seven days, fed a high quality

commercial salmonid diet (Profishent High Energy Feed , Martin Mills, Elmira, ON, Canada).

Nine diets were allocated to each block following a completely randomized block design.

Fifteen fish, an average weight of 40 g, were distributed to each tank. All treatments were fed to

satiety for the first two weeks to determine the lowest feed intake diet. After two weeks, fish

assigned diets 1, 2, 4, 5, and 7 were pair-fed to those fed diet 8, while continuing satiation

feeding for fish fed diets 3, 6, and 9. Feed consumption was recorded weekly and the fish were

weighed in every four weeks. Canadian Council on Animal Care (CCAC, 1984) and the

University of Guelph Animal Care Committee guidelines were followed to ensure ethical

treatment of the animal.

5.2.3. Morphological indices

At the beginning and the end of the experiment, fork length (mm) and weight (g) of 27

fish and 8 fish tank-1, respectively were measured to determine Fultons condition index (FCI;

Fulton, 1902). Five of these fish were then gutted, and viscera and liver were weighed for

viscera-somatic index (VSI) and hepato-somatic index (HSI), respectively. For the carcass-bones

ratio, eight fish were cooked at 900C for 20 minutes; all bones including the head were carefully

separated from muscle and skins, and weighed in g.

5.2.4. Chemical analysis

At the beginning of the experiment, 27 fish (one from each tank) were euthanized for

proximate composition and kept at -20oC until processed, and at the end of the experiment, a

113
total of ten fish tank-1 (five for whole carcass and five for dressed carcass and visceral

composition) from each tank were sacrificed. The pooled carcass and visceral samples were

autoclaved, ground into homogeneous slurry in a food processor, freeze-dried, then finely ground

and stored at -20oC until analysis.

Diets, ingredients, and carcass samples were analyzed for dry matter (DM) and ash

according to AOAC (1995), crude protein (CP, %N x 6.25, Dumas method) (Leco analyzer,

LECO Corporation, St. Joseph, MI), lipids with petroleum ether extraction at 90oC and 150 PSI

for 60 min using ANKOM XT-20 analyzer (ANKOM Technology, Macedon, NY), and gross

energy (GE) using an automated oxygen bomb calorimeter (Model # 1271, Parr Instruments,

Moline, Illinois). Visceral samples were analyzed only for DM, CP and lipid contents.

Neutral (NDF) and acid detergent fibre (ADF) contents of the ingredients and the diets

were determined according to Van Soest et al. (1991) using an ANKOM 200 Fibre Analyzer

(ANKOM Technology, Macedon, NY). Lignin was determined by reacting ADF residue with

cold 72% sulphuric acid. The sample was then ashed at 5500C for 4 hours and the residue was

measured gravimetrically (Van Soest and Reobertson, 1981). Phytic acid content in the

ingredients and diets were determined colorimetrically using a commercial assay kit (Megazyme

International, Ireland). Amino acid profiles of the test ingredients and the diets were analyzed at

the Amino Acid Laboratory of the Toronto Hospital for Sick Children (ON, Canada) using the

acid hydrolysis method. For the quantification of amino acids, approximately 0.05 g of dried

sample was hydrolysed in 15 ml test tube with 450 L of 6N HCl and 1% phenol for 24 hrs at

110oC under pre-purified nitrogen atmosphere. After 24 hrs, tubes were removed from oven,

cooled, capped, vortex mixed and centrifuged at 2500 rpm. Then, 5 L of the supernatant was

derivatized following PICO-Tag method (Cohen et al., 1989). After derivatization, supernatant

114
were re-dissolved in 150 L phosphate buffer and 1.5 L of this solution was injected into the

column of an ACQUITY Ultra-Performance Liquid Chromatography - UPLC machine (Waters

Corporation, Millford, CA) for quantification of all amino acids except tryptophan, methionine

and cysteine which are partially destroyed in the hydrolysis process.

5.2.5. Calculations

FCI (K) was calculated using the Fultons (1902) formula, where N = 5, BW

= body weight (g), L = fork length (mm). VSI and HSI were calculated as 100 x [viscera weight

(g)/BW (g)] and 100 x [liver weight (g)/BW (g)], respectively. Carcass:bone ratio was calculated

as whole carcass/bone weight.

Growth rate was calculated as thermal-unit growth coefficient (TGC), [100 x (FBW1/3

IBW1/3) / (Temp (0C) x Number of days)], where FBW is the final body weight and IBW is the

initial body weight of the fish (Cho, 1992). Feed efficiency (FE) was calculated as live weight

gain /dry feed intake. Protein (PD) and lipid (LD) deposition, nitrogen retention efficiency

(NRE), recovered energy (RE), and energy retention efficiency (ERE) were also determined. PD

and LD in whole carcass, dressed carcass, and viscera were calculated from the difference

between final and initial values. NRE (%) and ERE (%) were calculated as [[FBW x Nfinal)

(IBW x carcass Ninitial)]/DNI] x 100, [[FBW x carcass GE final) (IBW x carcass GEinitial)]/DEI] x

10, respectively, where DNI = digestible nitrogen intake, and DEI = digestible energy intake.

5.2.6. Statistical analysis

All data were analyzed using SAS V9.2, SAS Institute, Cary, NC, USA. Pair-feeding and

satiation-feeding treatments were analyzed separately. The difference among the treatments were

115
determined by Tukeys HSD test at P<0.05. The differences in responses of fish in treatments

pair-fed with diet 8 were assessed with Dunetts t-test.

The effects of phytic acid (0%, 1.13%) and lignin (0%, 1.5%, 3%) levels in the diets and

their interactions on final body weight, TGC, feed intake, feed efficiency, fish physiological

indices (VSI, HSI, FCI, and bone:carcass ratio), and nutrient deposition and retention

efficiencies, and PD:LD ratio were analyzed using SAS general linear model procedure (Proc

GLM) for the pair-feeding treatments, while only the effects of lignin were assessed for the

treatments fed high phytic acid diets. Significance of the effects was assessed using non-

orthogonal polynomial (linear and quadratic) contrasts for the pair-feeding treatments.

116
Table 5.1 Basal mix for the test diets.

Ingredient (% in diet)
Herring fish meal, 64% CP 6.8
Meat-bone meal, porcine, 47% CP 10.7
Poultry by-products meal, 66% CP 4.9
Feather meal, 85% CP 2.0
Spray dried blood meal, porcine, 93% CP 2.0
Corn protein concentrate, 77% CP 15.6
Gelatin 9.8
Gelatinized starch 20.1
Wheat middlings 2.0
Vitamin premix1 1.0
Lysine (Bio-lys) 1.3
DL-Met 0.1
Mineral premix2 1.0
Ca(H2PO4)2 0.5
NaCl 0.5
Total (% diet) 78.0
1
Provides per kg of diet: retinyl acetate (vitamin A), 3750 IU; cholescalciferol (vitamin D), 3000 IU; dl-a-tocopheryl
acetate (vitamin E), 75 IU; menadione sodium bisulphite (vitamin K), 1.5 mg; ascorbic acid polyphosphate (Stay C
25% ascorbic acid), 75 mg; Choline chloride, 1500 mg; cyanocobalamine (vitamin B12), 0.03 mg; biotin, 0.21 mg;
inositol, 450 mg; folic acid, 1.5 mg; niacin, 12 mg; Pantothenic acid, 30 mg; pyridoxineHCl, 7.5 mg; riboflavin, 9.0
mg; thiaminHCl, 1.5 mg.
2
Provides per kg of diet: sodium chloride (NaCl, 39% Na, 61% Cl), 3077 mg; ferrous sulphate (FeSO47H2O, 20%
Fe), 65 mg; potassium iodide (KI, 24%K), 11 mg; manganese sulphate (MnSO4, 36% Mn), 89 mg; zinc sulphate
(ZnSO47H2O, 40% Zn), 150 mg; copper sulphate (CuSO45H2O, 25% Cu), 28 mg; di-sodium selenite (Na2SeO3),
10 mg.

117
Table 5.2 Ingredient and proximate chemical composition of the test diets (DM basis).

Pair-feeding Satiation Feeding


Diet 1 Diet 2 Diet 4 Diet 5 Diet 7 Diet 8 Diet 3 Diet 6 Diet 9

0%PA 1.13%PA 0%PA 1.13%PA 0%PA 1.13%PA 2.25%PA 2.25%PA 2.25%PA


0%lignin 0%lignin 1.5%lignin 1.5%lignin 3%lignin 3%lignin 0%lignin 1.5%lignin 3%lignin

Basal mix + composition of ingredients


Basal mix, % 78 78 78 78 78 78 78 78 78
Fish oil, herring, % 16 16 16 16 16 16 16 16 16
Kaolin, % 4.0 3.6 3.3 2.5 2.1 1.8 1.1 0.7 0.4
Cellufil, % 2.0 2.0 2.0 1.5 1.5 1.5 1.0 1.0 1.0
Phytic acid, % 0.00 0.36 0.00 0.36 0.00 0.36 0.75 0.75 0.75
Purified lignin (76% lignin, 24%
0.00 0.00 0.00 1.97 1.97 1.97 3.95 3.95 3.95
cellulose), %
Proximate chemical composition
Phytic acid, % diet 0.06 1.14 0.07 1.11 0.07 1.14 2.21 2.29 2.29
Lignin, % diet 0.0 0.0 1.4 1.4 3.0 3.0 0.0 1.5 3.0

Dry matter, % 95 95 95 97 95 96 95 96 96
Crude protein, % 43 44 45 43 44 42 42 43 43
Lipid, % 19 19 19 19 19 19 18 18 19
Total carbohydrate, % 26 25 27 28 28 30 27 29 28
Neutral detergent fibre, % 5.7 6.2 6.7 6.1 6.2 6.0 5.6 6.4 5.6
Acid detergent fibre, % 5.2 5.3 5.4 6.0 5.6 5.8 5.5 6.3 5.5
Ash, % 11 12 10 10 9 9 13 10 10
Gross energy, kJ g-1 23 22 23 22 23 23 22 22 23

118
Table 5.3 Analyzed essential and non-essential amino acid composition of the test diets (digestible AA, % DM).

Pair-feeding Satiation feeding


Diet 1 Diet 2 Diet 4 Diet 5 Diet 7 Diet 8 Diet 3 Diet 6 Diet 9
EAA requirement of rainbow
trout (%), dry matter basis 0%PA 1.13%PA 0%PA 1.13%PA 0%PA 1.13%PA 2.25%PA 2.25%PA 2.25%PA
NRC, 2011 0%lignin 0%lignin 1.5%lignin 1.5%lignin 3%lignin 3%lignin 0%lignin 1.5%lignin 3%lignin

EAA, % in diet (DM basis)


Arginine 1.5 2.9 2.9 3.0 2.8 2.9 3.1 2.9 3.0 3.0
Histidine 0.8 0.9 1.0 1.0 0.9 0.9 0.9 1.0 1.0 0.9
Isoleucine 1.1 1.5 1.5 1.5 1.5 1.5 1.5 1.5 1.5 1.5
Leucine 1.5 4.6 4.7 4.7 4.5 4.5 4.5 4.7 4.6 4.4
Lysine 2.4 2.4 2.5 2.4 2.5 2.4 2.4 2.5 2.4 2.5
Methionine+cystine1 1.1 1.1 1.1 1.1 1.1 1.1 1.1 1.1 1.1 1.1
Phenylalanine 0.7 2.1 2.1 2.1 2.0 2.1 2.1 2.1 2.1 2.0
Threonine 1.1 1.2 1.3 1.3 1.3 1.2 1.3 1.3 1.2 1.2
1
Tryptophan 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2
Valine 1.2 2.1 2.2 2.2 2.1 2.0 2.2 2.2 2.1 2.1

NEAA, % in diet(DM basis)


Alanine 4.1 4.1 4.3 3.9 4.0 4.1 4.1 3.9 4.0
Aspartic acid 2.8 2.9 3.3 2.4 3.1 3.2 2.8 3.2 3.2
Cysteine1 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4
Glutamic acid 7.2 7.4 7.5 7.0 7.3 7.5 7.3 7.0 7.1
Glycine 5.3 5.2 5.7 4.9 5.2 5.4 5.2 5.1 5.2
Proline 4.7 4.6 4.9 4.4 4.6 4.8 4.6 4.4 4.6
Serine 1.9 1.9 2.0 1.9 1.9 2.1 1.9 1.9 1.9
Tyrosine 1.3 1.3 1.3 1.3 1.3 1.3 1.3 1.3 1.3
1
Calculated value (digestible AA, % in diet)

119
5.3. Results

No mortality occurred over the 12-week trial but a reduced feed intake with increasing

PA levels in the diets was observed during the first two weeks when fish in all treatments were

fed to satiety (Figure 5.1). After that, diets 1, 2, 4, 5, and 7 were pair-fed to the diet 8 while

satiation-feeding continued for fish fed 2.25% PA diets (3, 6, 9).

Final body weight, growth rate, FE, VSI, HSI, FCI, and whole-carcass:bone ratio of fish

in the pair-feeding treatments were slightly higher but not statistically different than those fed the

diet 8 as determined by the Dunnets t-test. None but the FCI among the performance attributes

and physiological indices was affected by PA or lignin or their interactions. Among the pair-

feeding treatments, FCI was significantly higher in fish fed the PA diets (1.05-1.07) than the PA-

free diets (1.02-1.05. In the satiation-feeding treatments, feed intake, TGC and FCI but not FE

and other physiological indices were affected by lignin (Table 5.4).

Among the pair-feeding treatments, whole carcass PD (130 mg fish-1 d-1, Dunetts t test -

P>0.05) and NRE (40%, Dunetts t test - P<0.05) in fish fed the control diet was slightly higher

than the other diets. In contrast, PD and LD in the dressed carcass were not affected but the NRE

of fish fed the control diet was significantly higher than those fed diet 8 (34%, Dunetts t test -

P<0.05). PA levels in the diets significantly affected the whole carcass NRE (P<0.05) (Table

5.5), but did not affect the protein or lipid deposition in the dressed carcass and viscera (Table

5.6). Dietary lignin affected only LD and LRE in the whole carcass (P<0.05).

120
Figure 5.1 Total feed intake of the fish fed experimental diets when fed to satiety during the

first two weeks of the trial. Diets 1, 2, 3 (0% lignin), diets 4, 5, 6 (1.5% lignin) and diets 7, 8, 9

(3% lignin).

14 0% phytic acid
1.13% phytic acid
12
2.25% phytic acid
Feed intake (g fish-1)

10

0
1 4 7 2 5 8 3 6 9
Diets

121
Table 5.4 Performance and physiological indices of fish (initial weight = 40 g fish-1) fed the experimental diets for 84d at 15C.

Thermal-unit
Description of Final body growth Feed Feed
treatment weight coefficient intake efficiency Viscero- Hepato- Fulton's Whole
Phytic acid (PA) somatic somatic condition carcass :
g fish-1 g fish-1
Lignin index index index Bone ratio
Pair-feeding

Diet 1 0.0% PA 0.0% Lignin 104.1 1.2 0.120 0.001 77.5 0.2 0.87 0.01 10.0 0.2 1.17 0.03 1.05 0.02 4.7 0.4
Diet 2 1.1% PA 0.0% Lignin 100.7 3.1 0.118 0.003 77.5 1.2 0.83 0.03 12.0 0.6 1.50 0.25 1.05 0.02 4.6 0.2
Diet 4 0.0% PA 1.5% Lignin 100.4 1.0 0.118 0.003 77.5 0.0 0.83 0.02 10.6 0.3 1.27 0.03 1.02 0.01 4.5 0.1
Diet 5 1.1% PA 1.5% Lignin 102.5 4.2 0.118 0.007 77.5 1.9 0.83 0.05 10.8 0.4 1.33 0.09 1.07 0.02 4.4 0.4
Diet 7 0.0% PA 3.0% Lignin 101.1 1.5 0.117 0.003 77.5 0.3 0.83 0.02 10.7 0.3 1.30 0.06 1.02 0.01 4.4 0.1
Diet 8 1.1% PA 3.0% Lignin 95.9 5.8 0.112 0.008 77.5 6.6 0.77 0.02 10.4 0.8 1.27 0.03 1.06 0.01 4.4 0.2
Effects
PA NS NS NS NS NS NS P<0.05 NS
Lignin NS NS NS NS NS NS NS NS
PA x
Lignin NS NS NS NS NS NS NS NS
Satiation feeding
Diet 3 2.2% PA 0.0% Lignin 84.1 2.9 0.092 0.003b 57.7 1.9b 0.81 0.01 11.8 0.4 1.22 0.06 1.10 0.02b 4.7 0.2
Diet 6 2.2% PA 1.5% Lignin 76.6 3.0 0.079 0.004a 50.5 0.5a 0.77 0.04 13.6 1.3 1.36 0.08 1.06 0.00a 4.5 0.3
Diet 9 2.2% PA 3.0% Lignin 78.3 1.2 0.078 0.001a 48.2 0.4a 0.80 0.04 11.7 0.2 1.29 0.03 1.02 0.00a 4.4 0.1
Effects
Lignin NS P<0.05 P<0.01 NS NS NS P<0.01 NS
Note: * Significant difference between pair-fed treatments and diet 8 were determined by Dunetts t tests. Difference between other diets was determined by
Tukey's HSD test (P0.05). Different letters denote significant differences; responses with no notations mean no differences among the diets. Effects of PA,
lignin and their interactions were determined by multiple linear regression (Y = b + PA + lignin + PA x lignin) and deemed significant at P<0.05.

122
Table 5.5 Whole carcass nutrient deposition and retention efficiency of rainbow trout fed test diets with various levels of phytic acid

and lignin.

Whole Carcass
Lipid Energy Protein :
Protein Lipid Recovered N retention retention retention lipid
Description of treatment deposition deposition energy efficiency efficiency efficiency deposition
Phytic acid (PA) Lignin mg fish-1 d-1 mg fish-1 d-1 j fish-1 d-1 % % %

Pair-feeding
Diet 1 0.0% PA 0.0% Lignin 130.2 1.5 72.6 6.2 58.9 1.0 40.0 0.5* 47.5 4.0ab 42.5 1.1 1.8 0.2ab
Diet 2 1.1% PA 0.0% Lignin 120.2 4.2 79.7 3.7* 59.2 1.8 36.4 1.0 54.2 1.9*b 45.4 0.7* 1.5 0.0*a
Diet 4 0.0% PA 1.5% Lignin 126.2 0.8 79.7 1.8* 58.6 2.7 37.7 0.3 53.1 1.1*b 43.4 1.1 1.6 0.1ab
Diet 5 1.1% PA 1.5% Lignin 119.6 6.9 72.7 7.4 58.0 5.0 35.5 1.4 46.6 5.1ab 42.5 1.7 1.7 0.2ab
Diet 7 0.0% PA 3.0% Lignin 123.6 3.9 66.2 5.8 58.2 3.6 37.0 1.1 43.7 3.7ab 43.0 1.6 1.9 0.1ab
Diet 8 1.1% PA 3.0% Lignin 115.8 14.1 56.8 6.7 50.9 5.9 35.6 1.4 37.1 2.1a 40.4 0.9 2.0 0.1b

Effects
PA NS NS NS P<0.05 NS NS NS
Lignin NS P<0.05 NS NS P<0.01 NS P<0.05
PA x Lignin NS NS NS NS NS NS NS

Satiation feeding
Diet 3 2.2% PA 0.0% Lignin 87.1 3.1 43.2 6.3 38.3 2.2b 37.2 1.0 40.0 5.3 40.8 1.0b 2.1 0.3
Diet 6 2.2% PA 1.5% Lignin 78.9 4.4 33.7 1.5 31.3 1.3ab 37.0 1.8 35.1 1.4 39.0 0.8ab 2.4 0.0
Diet 9 2.2% PA 3.0% Lignin 77.3 1.9 26.0 7.1 27.5 2.6a 37.8 1.2 27.77.6 33.7 2.7a 3.7 1.3

Effects
Lignin NS NS P<0.01 NS NS P<0.05 NS
Note: * Significant difference between pair-fed treatments and diet 8 were determined by Dunetts t tests. Difference between other diets was determined by
Tukey's HSD test (P0.05). Different letters denote significant differences; responses with no notations mean no differences among the diets. Effects of PA,
lignin and their interactions were determined by multiple linear regression (Y = b + PA + lignin + PA x lignin) and deemed significant at P<0.05.

123
Table 5.6 Nutrient deposition and retention efficiencies in dressed carcass and viscera of rainbow trout fed the experimental diets.

Dressed carcass Viscera


Protein Lipid Recovered
Description of treatment deposition deposition energy Protein deposition Lipid deposition
Phytic acid (PA) Lignin mg fish-1 d-1 mg fish d j fish-1 d-1
-1 -1
mg fish-1 d-1 mg fish-1 d-1
Pair-feeding
Diet 1 0.0% PA 0.0% Lignin 110.7 2.8 40.4 4.1 43.5 2.3 14.8 0.4 22.8 3.1a
Diet 2 1.1% PA 0.0% Lignin 99.4 4.4 48.3 7.0 43.6 3.5 17.6 1.4 34.5 3.8*b
Diet 4 0.0% PA 1.5% Lignin 102.3 3.0 44.4 3.6 42.7 1.7 16.6 0.5 23.2 1.0a
Diet 5 1.1% PA 1.5% Lignin 107.0 9.7 55.3 6.7 49.1 5.1* 15.2 1.2 23.1 1.4a
Diet 7 0.0% PA 3.0% Lignin 100.0 3.2 38.9 4.1 39.9 1.0 14.1 0.2 28.9 5.0ab
Diet 8 1.1% PA 3.0% Lignin 88.6 11.4 34.5 4.5 35.5 4.1 16.2 1.2 22.1 1.7a
Effects
Phytic acid NS NS NS NS NS
Lignin NS NS NS NS NS
Phytic acid x Lignin NS NS NS NS P<0.05
Satiation feeding
Diet 3 2.2% PA 0.0% Lignin 62.6 5.1b 20.7 2.3 23.7 1.8b 21.8 1.8 35.1 6.8
Diet 6 2.2% PA 1.5% Lignin 35.4 1.1a 9.5 1.6 12.3 0.9a 25.3 1.6 38.3 1.8
Diet 9 2.2% PA 3.0% Lignin 39.3 3.4a 12.5 3.7 12.5 2.1a 21.3 0.6 37.8 5.0
Effects
Lignin P<0.01 NS P<0.01 NS NS
Note: * Significant difference between pair-fed treatments and diet 8 were determined by Dunetts t tests. Difference between other diets was determined by
Tukey's HSD test (P0.05). Different letters denote significant differences; responses with no notations mean no differences among the diets. Effects of PA,
lignin and their interactions were determined by multiple linear regression (Y = b + PA + lignin + PA x lignin) and deemed significant at P<0.05.

124
5.4. Discussion

Effects of dietary PA and purified lignin on growth performance, physiological indices,

and nutrient deposition in rainbow trout were assessed in this study. Overall FE and growth of

fish in all treatments including those fed the control diet were low (between 0.77 and 0.87)

compared to those observed in similar trials. In this study, five EAA: His (0.9-1.0% DM), Lys

(2.4-2.5% DM), Met+Cys (1.1% DM), Thr (1.2-1.3% DM), and Trp (0.2% DM) were formulated

to be marginally adequate to the requirement of rainbow trout. It is likely that a portion of these

AA were not available resulting in lower than usual feed efficiency. Most protein sources used in

this study are animal by-products, which have subjected to various pressure and temperature

conditions during processing and manufacturing. Bureau et al. (1999) showed that manufacturing

characteristics such as duration, pressure and temperature during hydrolysis of feather meals,

during cooking of MBM or PBM, and coagulation and drying methods of blood meals, affect the

digestibility of crude proteins of these feed stuffs in rainbow trout. Proteins may react with fats

and their oxidation products, and chemicals used during processing. Amino acids involved in

these reactions are structurally altered and often, the reaction is irreversible resulting in the AA

being biologically unavailable. Lys, with its -amino group, is particularly susceptible to such

damage, and so are Met, Cys, and Trp (Batterham, 1992; Moughan and Rutherfurd, 1996).

In this study, when fish in all treatments in the first two weeks, and in the high PA

(2.25%) treatments throughout the trial were fed to satiety, feed intake appeared to be affected by

dietary PA (Figure 5.1) and lignin (Table 5.4), respectively. It is plausible that one or more of the

five marginally adequate EAA in the diets may not be completely available to fish resulting in an

imbalanced AA profiles. Despite slightly lower performances of the fish fed PA-high lignin diets

(diet 8) than the other diets, PA or lignin level did not affect TGC, FE or any of the physiological

125
indices among the pair-feeding treatments. Only FCI was affected, where fish fed PA diets

showed higher FCI (1.06-1.07) than the PA free diets (1.02). Dabrowski et al. (2007) reported

that significant differences in growth of a cichlid species (Midas, Amphilophus citrinellum) but

not in FE the fish fed a protein-bound EAA diet supplemented with free AA (FAA) and two

AA deficient diets, (-) Lys (deficient in Lys, His, Ile, Phe) and (-) Arg (deficient in Trp, and Arg,

Thr, Val, Leu, and Met). One of the earliest studies on the effect of PA on fish was conducted by

Spinelli et al. (1983) where the authors reported significant differences in weight gain between

rainbow trout fed diets with 0.5% sodium or calcium phytate (72 and 74 g, respectively) and

without PA (86 g) but observed no effects on FE. Similar observations were also reported in

salmon (Sajjadi and Carter, 2004a, b) and carps (Usmani and Jafri, 2002). In another study,

Denstadli et al. (2006) reported insignificant effect on the ADC of CP and FE (between 0.68 and

0.72) in Atlantic salmon with maximum inclusion rate of 2% PA of the diet. But feed intake was

severely affected at the dietary PA of 1% or higher (11.9-15.5 g fish-1) in the first 39 days but

after that, low feed intake were observed in the fish fed only the high PA (2%) diet (28.8 g fish-1)

compared to that of the fish fed the control diet (19.0 and 50.1 g fish-1, respectively). These

results are in agreement with the present study in that animals control their PA intake by

controlling the total feed intake reducing the effect of PA on the deposition and partitioning of

nutrients.

In this study, dietary PA levels affected only the FCI and the whole body NRE of the

fish. FCI is a measure of energy reserve where values >1 represent fish with high fat while fish

with <1 CI, represent the fish that suffered from adverse environmental or poor feeding

conditions (Rtz and Llroet, 2003). Although LD tend to decrease with the addition of lignin,

both LD and whole body LRE were not significantly affected but slightly higher in fish fed PA

126
diets than the control diets resulting in higher FCI. LD (35 mg fish-1 d-1) in viscera of fish fed

lignin free PA diets was significantly higher than the control diets (23 mg fish-1 d-1) (Table 5.6)

suggesting that much of the excess lipid were deposited as intra-peritoneal fat (IPF) rather than

as muscle fat.

Dietary PA have been reported to significantly affect CP digestibility in several fish

species, for example, Atlantic salmon Salmo salar (Sajjadi and Carter, 2004a,b), rainbow trout

Oncorhynchus mykiss (Spinelli et al., 1983; Sugiura et al., 2001), and common carp Cyprinus

carpio (Hossain and Jauncy, 1989). Recently, Laining et al. (2010) reported a decreasing ADC of

CP from 95% in the fish fed the control diet (0% PA) to 90% in those fed 1% PA diets. Selle et

al. (2000) proposed that the de novo formation of binary protein-PA complexes in the

gastrointestinal tract as the key mechanism by which PA negatively influences the AA

digestibility. Several modes of action of forming these complexes are: (1) the presence of

protein-PA complexes in feedstuffs, (2) the de novo formation of binary and ternary protein-PA

complexes in the digestive tract and (3) the inhibition of proteolytic enzymes (Kies et al., 2001;

Selle et al., 2006).

Significantly lower NRE in fish fed the PA diets observed in this study are in agreement

with several studies (Nangthu et al., 2010; Storebakken et al., 1998b; Usmani and Jafri, 2002).

Nangthu et al. (2010) reported that Lys utilization efficiency was severely affected when

formulated to be marginally adequate in a plant protein based diet resulting in a significantly

lower NRE. In an earlier study, ADC of His and Lys appeared to be negatively influcenced by

the PA intake (Chapter 3). All experimental diets in this study were marginally adequate in these

two EAA. The inclusion of PA could have further reduced their availability resulting in low NRE

in fish fed the PA diets (P<0.05).

127
None of the growth performance and physiological indices of rainbow trout was affected

by lignin in the pair-feeding treatments but when fed to satiety, lignin appeared to significantly

suppress feed intake, which likely affected the TGC and FCI of rainbow trout. No such effects on

feed intake have been demonstrated in studies with swine and broiler chickens fed purified lignin

(Schedle et al., 2008; Baurhoo et al., 2007). A synergistic interaction of lignin with PA in this

case can not be ruled out because of the very high PA concentration (2.25%) in these diets.

Schedle et al. (2008) investigated the effects of insoluble dietary fibre in the variable

lignin diets on the performance of piglets. The authors used Chinese Mason Pine pollen (29%

lignin) as the source of lignin. Higher feed intake by the piglets fed lignin diets than those fed the

lignin-free diet resulted in a significantly higher weight gain. However, apparent crude protein

digestibility was significantly lower only in piglets fed high lignin (1.7%) diets than those fed the

other diets. In another study, Wang et al. (2009b) fed feedlot lambs diets containing 1.5% and

3% purified lignin, and reported no differences in the weight gain between lambs fed diets with

and without lignin.

In this study, lignin did not affect the whole carcass PD and NRE but the LD (P<0.05),

LRE (P<0.01), and PD:LD (P<0.05) in the pair-feeding treatments, where as in the satiation-

feeding treatments, only whole carcass RE (P<0.01) and ERE (P<0.05) appeared to be

significantly affected. Glencross (2009) reported a comparable effect of lignin where not the

ADC of N but those of GE and DM were significantly reduced from 92% and 84%, respectively

in rainbow trout fed the control diet to 68% and 77%, respectively in those fed the high lignin

(3%) diets. In this study, unlike the whole carcass, NRE (P<0.05) and ERE (P<0.01) in dressed

carcass were significantly affected by lignin in the pair-feeding treatments. Although not

128
significant, LD (P=0.061) and RE (P=0.096) in the dressed carcass of fish pair-fed PA-high

lignin treatment was lower than those in the other treatments.

The lack of notable interactions between PA and lignin in this study suggests that their

mechanisms and regions of actions are different. In nature, hydrophobic lignin molecules co-

exist with the hydrophilic cellulose, and together, they act as nutrient transporter in various plant

tissues. It can be speculated that a portion of dietary lipid might have bound by the lignin

molecules affecting digestibility of lipids. This is the first detailed study elucidating effects of

lignin and its interaction or not with PA on performance, physiological indices and nutrient

deposition of rainbow trout. It can be concluded that fish, like any other animal, controls intake

of the feed when fed diets with imbalanced AA profile or deficient in one or more EAA, or when

one or more ANF are present in order to maintain their homeostasis. Neuro-gastric mechanisms

to control feed intake in such situations is not well-understood and should be explored further to

improve our understanding on anti-nutritional components of plant ingredients and their effects

on nutrient dynamics in farmed animals.

129
CHAPTER 6

ASSESSING THE EFFECTS OF LIGNIN ON PHYSICAL PROPERTIES OF STEAM-

PELLETED EXPERIMENTAL TROUT FEED

Abstract

A series of trials were conducted to evaluate the effects of lignin on the physical properties

of a dietary-mash and a steam-pelleted fish feed. Oil and water absorption of dietary mash and

physical properties of the pellets were assessed with graded levels of purified lignin. Absorption

of both vegetable oil (VO) and fish oil (FO) decreased significantly from 91% to 80% and from

84% to 76% in the mash with 0% and 3% lignin, respectively, but did not change thereafter.

Water absorption capacity was not affected (P>0.05) with increasing levels of lignin. Bulk

density, sinking velocity, hardness and water stability (i.e. dry matter, nitrogen and lipid loss) of

the steam-pelleted feed were examined at 0%, 1.5% and 3% lignin levels. Bulk density increased

significantly with increasing levels of lignin but hardness was not affected with inclusion of up

to 1.5% (75 and 73 Newton for 0% and 1.5% lignin pellets, respectively), that reduced

significantly (P<0.01) at 3% inclusion (51 Newton). Water stability of the pellets was

significantly reduced by increasing duration in water and smaller size but not affected by lignin

(P<0.05). Loss of nutrients increased linearly with increasing time where 2-mm pellets lose

nutrients faster than the 3.35-mm pellets. Overall, lipid loss was higher (P<0.05) than dry matter

and nitrogen loss for all treatments. This study suggests that at level up to 1.5%, lignin had no

effect on the physical quality of the steam-pelleted aquaculture feed.

130
6.1. Introduction

Physical properties are a measure of the effect an ingredient has on the characteristics of

a feed (Aas et al., 2011). These properties include fluid absorption capacity of mash and pellets,

and bulk density, hardness, sinking velocity, viscosity and water stability of the pellets. Water

stability of a pellet is usually determined by the time it takes to disintegrate and may include the

determination of the dry matter loss. Rapidly disintegrating feeds cause significant

environmental and economical losses and are highly undesirable in commercial farming

practices (Cho and Bureau, 2001; Obaldo et al., 2002). Water stability can be improved by

adding starch and soluble (e.g., guar gum, Brinker, 2007; Brinker and Reiter, 2011; Storebakken,

1985) or insoluble (e.g., cellulose, Bromley and Adkins, 1984; Hansen and Storebakken, 2007;

Lupatsch et al., 2001) fibres as binders in fish feed. Level and composition of the fibre

components in plant protein ingredients alter the physical properties of mash and pellets during

processing often requiring modifications of the manufacturing parameters. Until today, few

studies have assessed physical properties of aquaculture feed affected by the fibre components of

an ingredient (e.g., Baeverfjord et al., 2006; Glencross et al., 2007b; Hansen and Storebakken,

2007; Kraugerud et al., 2011), and the criteria for evaluation of these properties are not well-

established (e.g., Glencross et al., 2007b; 2010).

The insoluble lignin, a highly polymerized phenolic plant polymer responsible for

providing structural support and the flow of nutrients, is the second most abundant compound in

plant kingdom after cellulose (Jung and Fahey, 1983). Lignosulfonates and Kraft lignins,

products of pulping, are primarily used as binders to improve durability of the animal feed pellets

(Gargulak and Lebo, 2000). These lignin derivatives are composed mostly of high-molecular-

weight fragments and have shown some toxicity in rainbow trout (Pessala et al., 2010; Roald,

131
1977). Purified lignin, in contrast, consists mostly of low-molecular-weight phenolic fragments

with putative anti-microbial and prebiotic effects (Davidson and Brenen, 1981; Kleinert and

Barth, 2008; Lora et al., 1993; Roller and Seedhar, 2002; Ultee et al., 2000). Prebiotic effects of

purified lignin are reported recently in broiler chickens (Baurhoo et al., 2007a), and in feedlot

lambs (Wang et al., 2009b) at low inclusion level. Baurhoo et al. (2008) hypothesized that lignin,

in its purified form, possesses functional properties not available in other lignin compounds. In

addition, apparent digestibility of nutrient, gut morphology, growth and feed intake in poultry

(Baurhoo et al., 2007b) and rabbits (Chiou et al., 1994) remain unaffected when fed diets with

<4% of lignin suggesting the possibility of using purified lignin in limited quantity in

aquaculture feed both as binder and as pre-biotics.

In this study, an attempt was made to standardize the assessment criteria of some physical

properties of compounded aquaculture feed with inclusion of purified lignin, as a model

ingredient. The physical properties assessed are the oil and water absorption in the mash and

bulk density, hardness, sinking velocity, and water stability of the steam-pressed pellets

containing various levels of lignin.

6.2. Methods

Purified lignin, extracted from pine (Pinus sylvestris L.) bark using a proprietary process,

was obtained from the Pure Lignin Environmental Technology, Kelowna, BC, Canada. A dietary

mash consisting of fish meal, animal by-products (meat and bone meal, feather meal, blood meal

and gelatin), and corn protein concentrate as protein source was prepared (Table 6.1). Dextrin,

vitamin and mineral premixes, L-lysine, DL-methionine, di-calcium phosphate, and NaCl were

also added to reflect the composition of a typical trout feed (Table 6.1). The basal mash,

132
produced by mixing the ingredients in a Marion mixer (Marion, IA), was used to determine oil

and water absorption with five levels of lignin (0%, 1.5%, 3%, 6% and 12%). Fish oil (16%), and

appropriate amount of kaolin (an inert product with low shrink-swell capacity) and lignin were

added to the mash to prepare three diets of 0%, 1.5% and 3% lignin. The diets were thoroughly

mixed (using a Hobart mixer Hobart Ltd., Don Mills, Ontario), pelleted (using a laboratory

steam pellet mill California Pellet Mill Co., San Francisco, California), and dried in forced air

at ambient temperature for 24 h. Pellets of 2.00-mm and 3.35-mm diameter were produced to test

the size effects on bulk density and pellet water stability (loss of dry matter (DM), nitrogen (N)

and lipid). Another set of pellets of 3.35-mm diameter were produced by mixing mash diet with

0%, 1.5% and 3% lignin using a laboratory built pellet gun for assessing the effects of lignin on

sinking velocity and hardness of the pellets.

133
Table 6.1 Ingredient composition of the basal mix prepared for assessing oil and water

absorption capacity with the addition of lignin. 16.0% fish oil was added before mixing with

kaolin and lignin to prepare three diets of 0%, 1.5% and 3% lignin.

Ingredient (% in diet)
Herring fish meal, 64% CP 6.8
Meat-bone meal, porcine, 47% CP 10.7
Poultry by-products meal, 66% CP 4.9
Feather meal, 85% CP 2.0
Spray dried blood meal, porcine, 93% CP 2.0
Corn protein concentrate, 77% CP 15.6
Gelatin 9.8
Gelatinized starch 20.1
Wheat middlings 2.0
Vitamin premix1 1.0
Lysine (Bio-lys) 1.3
DL-Met 0.1
Mineral premix2 1.0
Ca(H2PO4)2 0.5
NaCl 0.5
Total (% diet) 78.0
1
Provides per kg of diet: retinyl acetate (vitamin A), 3750 IU; cholescalciferol (vitamin D), 3000 IU; dl-a-tocopheryl
acetate (vitamin E), 75 IU; menadione sodium bisulphite (vitamin K), 1.5 mg; ascorbic acid polyphosphate (Stay
CTM 25% ascorbic acid), 75 mg; choline chloride, 1500 mg; cyanocobalamine (vitamin B12), 0.03 mg; biotin, 0.21
mg; inositol, 450 mg; folic acid, 1.5 mg; niacin, 12 mg; Pantothenic acid, 30 mg; pyridoxineHCl, 7.5 mg;
riboflavin, 9.0 mg; thiaminHCl, 1.5 mg.
2
Provides per kg of diet: sodium chloride (NaCl, 39% Na, 61% Cl), 3077 mg; ferrous sulphate (FeSO 47H2O, 20%
Fe), 65 mg; potassium iodide (KI, 24%K), 11 mg; manganese sulphate (MnSO4, 36% Mn), 89 mg; zinc sulphate
(ZnSO47H2O, 40% Zn), 150 mg; copper sulphate (CuSO45H2O, 25% Cu), 28 mg; di-sodium selenite (Na2SeO3),
10 mg.

134
6.2.1. Oil and water absorption of the mash

Absorption of FO, VO, and water was determined according to a method slightly

modified from Sathivel et al. (2005). 500-mg of sample mash with appropriate amount of lignin

was placed into a 50-ml centrifuge tube, 10-ml of oil or de-ionized water was addred, then

thoroughly mixed using a steel spatula for 2 minutes, and kept for 30 minutes at ambient

temperature mixing intermittently in every 10 minutes. The mixture was centrifuged at 2560 x g

for 25 minutes using an Eppendorf centrifuge model 5810R (Eppendorf AG, Hamburg,

Germany). After decanting the free oil or water, absorption was determined from the weight

difference and expressed as percent of the sample weight.

6.2.2. Physical properties of the pellets

Bulk density, sinking velocity, hardness and water stability of the pellets with 0%, 1.5%

and 3% lignin were assessed. For bulk density, pellets of each size were filled in a 50-ml

measuring cylinder up to the mark, and weighed in g. The bulk density of the pellets was then

calculated as the weight of this volume and expressed as g L-1. Sinking velocity was determined

by placing a pellet 1-cm below the surface in a 500-mL measuring cylinder filled with water.

Time taken by a pellet to travel to the bottom (26.7 cm) was measured using a digital stop-watch

and expressed as cm s-1. Ten randomly selected pellets were used to determine average sinking

velocity for each treatment.

Hardness was determined as the force to crush a pellet across diameter using a texture

analyzer, Instron Universal Testing Machine, Model 1122 (Instron, Norwood, MA) equipped

with a load-cell and a flat disk. Ten pellets of each treatment were randomly selected for the

analysis. The analyzer was set with a test speed of 0.17 mm s-1. The probe was set to pass a

135
maximum distance of 3 mm and trigger at a contact pressure of 5 g. Crush strengths were defined

as the peak force at breaking of the pellet and expressed in Newton (N).

A horizontal shaking method, slightly modified from Obaldo et al. (2002), was employed

using a shaking water bath (Thermo Scientific Haake SWB25, Thermo Fisher Scientific, MA)

for assessing water stability of pellets measured as the amount of nutrients leached over time.

Representative feed samples of 10 g for each of the 2-mm and 3.35-mm pellets were taken into

250-ml Erlenmeyer flasks containing 100-ml de-ionized water, and were subjected to horizontal

shaking (at 100 rpm) for 1, 10, and 20 min at 20oC. Remaining solids were then separated using

a 1-mm mesh sieve, carefully poured into pre-weighed aluminum pans with minimal disturbance,

and dried at 105oC for 24 hr for dry matter loss (AOAC, 1995). Nitrogen (N) and lipid-contents

of the dried samples were determined using a LECO Analyzer (LECO Corporation, St. Joseph,

MI) and by petroleum ether extraction at 90oC and 150 PSI for 60 min using ANKOM XT-20

analyzer (ANKOM Technology, Macedon, NY), respectively.

6.2.3. Statistical analysis

All data were subjected to GLM procedure using SAS 9.2, SAS Institute, Cary, NC,

USA. Effects of lignin, pellet size, time and their interactions were assessed for pellet water

stability. For the bulk density, effects of lignin level, pellet size and their interactions were

assessed while for sinking velocity, hardness, and water and oil absorption, a simple linear

regression (a + bx) model was used. A parameter was deemed significant at P<0.5. Differences

among the mash or pellets with various levels of lignin were assessed by Tukeys-b test and

considered significant at P<0.05.

136
6.3. Results

6.3.1. Oil and water absorption

The absorption of VO (average 82%) was significantly higher (P<0.01) than the FO

(average 76%). Absorption of both VO and FO decreased significantly (P<0.05) from the mash

containing 0% lignin to those with 3% lignin (from 84% to 76% for FO and from 91% to 80%

for VO, respectively), but did not change at higher concentration (Table 6.2). Water absorption

was not affected by lignin (P>0.05).

6.3.2. Physical properties

Sinking velocity of the pellets decreased significantly with increasing levels of lignin

(Table 6.3), from 5.1 cm s-1 to 4.5 cm s-1 to 3.2 cm s-1 with the 0%, 1.5% and 3% lignin,

respectively. Hardness of the pellets did not differ (P>0.05) between the pellets without (75 N)

and those with 1.5% lignin (73 N), but decreased significantly (P<0.05) with 3% lignin (47 N,

Table 6.3). Lignin levels (P<0.05), pellet size (P<0.01), and their interactions (P<0.05), all

affected the bulk density (Table 6.4). Among the water stability parameters, loss of DM from the

pellets ranged between 7% and 13%, respectively within the first minute in the water and

increased significantly for both 2-mm (between 45% and 63%) and 3.35-mm (between 23% and

28%) pellets at the 10th minute in water (Figure 6.1). Duration in water affected nutrient loss the

most (P<0.01) followed by size x time interactions (affecting only the DM and N loss, P<0.05),

where as lignin and pellet size did not affect the loss of any nutrient (P>0.05, Table 6.4).

137
Table 6.2 Oil and water absorption capacity (in %) of the diet mash with various levels of

lignin (Mean SE). The difference in the absorption capacity of oil and water among different

levels of lignin was assessed by Tukeys b test and considered significant at P<0.05.

Absorption capacity of mash (%)

Lignin (% in mash) Fish oil, herring Mixed vegetable oil Water

(%) (%) (%)

0% 84 1b 90 4b 134 3

1.5% 77 1ab 86 2ab 134 2

3% 75 3a 80 1a 140 3

6% 75 0a 78 0a 139 2

12% 74 2a 78 0a 134 2

138
Table 6.3 Sinking velocity (cm s-1) at water temp. 22OC, hardness (Newton) and bulk density

of the diet pellets with various levels of lignin (mean SE). The difference among the pellets of

three lignin levels (0%, 1.5% and 3%) was assessed by Tukeys-b test and considered significant

at P<0.05.

Sinking Hardness Bulk density


Lignin level velocity
(%)
cm s-1 Newton g L-1
2.0 mm 3.35 mm
c b
0% 5.1 0.3 753 594 12b 522 5a

1.5% 4.50.2b 733b 581 5a 537 13b

3% 3.20.1a 513a 580 2a 540 12b

139
Figure 6.1 Effect of lignin inclusion rate (0%, 1.5% and 3%) on dry matter loss and nitrogen

leaching from pellets of two sizes (2-mm and 3-mm) following immersion in shaking water bath

at 100 rpm min-1 for >1-min, 10-min and 20-min of durations at 22oC.

Pellet size (2.00 mm) Pellet size (3.35 mm)

80 80
Dry matter loss (%)

Dry matter loss (%)


60 60

40 40

20 20

0 0
0 10 20 0 10 20
Time (min) Time (min)

60 60
50 50
N loss (%)

40 40
N loss (%)

30 30
20 20
10 10
0 0
0 10 20 0 10 20
Time (min) Time (min)

80 80
60
Lpid loss (%)
Lpid loss (%)

60
40 40
0% Lignin 0% Lignin
20 1.5% Lignin 20 1.5% Lignin
3% Lignin 3% Lignin
0 0
0 10 20 0 10 20
Time (min) Time (min)

140
Table 6.4 Parameter estimates (b) of the effect of lignin levels (0%, 1.5% and 3%), pellet size, duration in water (time) and their

interactions on pellet water stability (%), parameter estimates (b) of the effect of the three lignin levels on sinking velocity and

hardness of the pellets, and oil and water absorption of the mash (lignin levels 0%, 1.5%, 3%, 6%, and 12%). Pellet water stability was

measured as dry matter (DM), nitrogen (N), and lipid loss.

Pellet quality Diet mash


Pellet water stability Oil & water absorption
N- Bulk Sinking Vegetable
DM loss loss Lipid loss density velocity Hardness Fish oil Oil Water
% % % g L-1 cm s-1 Newton % % %
Intercept 18.8 13.4 40.7** 423.7** 5.2** 78.2** 79.8** 86.7* 135.4**

Lignin 5.0 3.5 5.6 21.6** -0.6** -7.9** -0.6* -0.9* 0.5
Pellet size -2.3 -1.2 -5.7 50.3** - - - - -
Time 3.8** 3.6** 3.0** - - - - - -
Lignin x size -1.8 -1.8 -2.0 -7.9* - - - - -
Lignin x time -0.0 0.1 -0.0 - - - - - -
Pellet size x time -0.7* -0.7* -0.5 - - - - - -
Lignin x size x time 0.1 0.0 0.1 - - - - - -
2
Adj. R 0.88 0.88 0.92 0.90 0.61 0.49 0.40 0.52 0.33
**P < 0.01; * P < 0.05

141
6.4. Discussion

Effects of lignin on oil and water absorption by a dietary mash and on some physical

properties of steam-pressed pellets of an experimental salmonid feed were investigated in this

study. The focus of this study was to develop protocols for assessing the quality of steam-pressed

pellets. Type of feedstuffs (Olsen et al., 2006; Refstie et al., 2006), feedstuff components

(Hansen and Storebakken, 2007; Hilton et al., 1981), and processing and manufacturing

conditions (Sorensen et al., 2005; Sorensen et al., 2009) have substantial effects on the stability

of a feed requiring to adjust processing and manufacturing parameters including physical,

chemical and thermal treatments (e.g., moisture, temperature, pressure, flow, torque, screw

speed) (Draganovic et al., 2011; Thomas, 1996).

Modern farming practices involve storage and transportation of large volume of feed that

are mechanically or manually distributed to the rearing vessels (e.g., ponds, tanks, pens, cages

etc.) demanding highly durable and water-stable feeds. The feeds also need to be stable enough

to avoid rapid disintegration after ingestion. Besides increasing the waste-outputs (Cho and

Bureau, 2001), rapid disintegration of a feed after ingestion appeared to cause oil-belching in

salmonids fed high-fat diets often coinciding with gastric dilation air sacculitis (GDAS)

(Baevefjord et al., 2006; Forgan and Forster, 2007; Lumsden et al., 2010).

6.4.1. Oil and water absorption

Sathivel et al. (2005) defined absorption of any fluid as the ability of a mash to absorb

and retain water and oil against a centrifugal force. Glencross et al. (2009) adopted a similar

method to assess water absorption by dry mash but for oil absorption, the authors proposed a

vacuum infused oil uptake method where oven-dry pellets placed in a mixer with excess oil,

142
thoroughly mixed for 1-min, transferred to a beaker, then placed in the vacuum chamber of a

freeze drier. After reaching the vacuum, the chamber was re-equilibrated to the atmospheric

pressure and oil infused to the pellets was then weighed by difference. This method deemed

suitable for extruded pellets where pellets are expanded during production and have air pockets

but may not be feasible for steam-pressed hard pellets.

In this study, oil absorption decreased significantly with lignin levels of up to 3%.

Significant differences were observed between the mashes with 0% and 3% or more lignin for

both FO and VO (Table 6.2). Approximately 7% more VO was absorbed than the FO

irrespective of the lignin levels. FO used in this study is North American herring oil composed of

26% SFA, 57% MUFA and 17%PUFA (Bureau et al., 2008). Significantly higher density of the

VO (specific gravity, 0.874 0.01) than that of the FO (specific gravity, 0.846 0.02) used in this

study may explain the differences in the absorption of these oils.

6.4.2. Pellet quality

The bulk density, sinking velocity, and hardness of the pellets were significantly affected

by lignin in this study. Bulk density were affected by both lignin (P<0.05) and pellet size

(P<0.01) (Table 6.4). Larger steam-pressed pellets are usually denser than the smaller pellets

resulting in higher bulk density. Other factors that may affect the density of pellets are: condition

temperature (C), screw speed (rpm), die temperature (C), pressure (bar), torque (N m), and

specific mechanical energy SME (Wh kg-1), and the composition of non-starch polysaccharides

and level of starch gelatinization (Kraugerud et al., 2011; Sorensen et al., 2009).

The common method to measure sinking velocity is to drop a pellet into a measuring

cylinder filled with water and recording the time to reach the bottom (Glencross et al., 2011;

143
Kannadhason et al., 2009), where some pellets tend to float because of the surface tension or oils

around the pellets or air-bubbles, producing erroneous results (e.g., Glencross et al., 2011). To

avoid this, in this study, individual pellets were dipped into 1-cm below the surface for 2 seconds

using a forcep and then released. This has eliminated the risk of floatation, and should be

preferred to determine sinking velocity.

Hardness of the pellets was reduced significantly from 75 and 72 N (0% and 1.5% lignin,

respectively) to 51 N (3% lignin), suggesting that up to 1.5% of purified lignin can easily be

added without affecting the durability of the pellets. Durability is usually measured as hardness

(the force to disintegrate) and as water stability (Refstie and sgrd, 2009). Obaldo et al. (2002)

defined water stability as the retention of physical integrity of the pellets with minimal

disintegration and nutrient leaching. The authors tested static water method, horizontal and

vertical shaking methods to test water stability of shrimp feed, and concluded that both shaking

methods are suitable for routine laboratory analysis as they allow pellet and water movement like

normal farming conditions.

The leaching of nutrients is generally determined as dry matter loss assuming that protein

ingredients and oils are uniformly distributed in the pellets and therefore, estimating the dry

matter loss could represent the protein and lipid loss from the pellets (e.g., Aas et al., 2011;

Baeverfjord et al., 2006; Lim and Dominy, 1990; Obaldo et al., 2002). Findings from this study

suggested otherwise where the dry matter loss differed significantly from the lipid loss, and did

not represent the loss of individual nutrients. Not the levels of lignin but the duration in water

and pellet size significantly affected the nutrient loss where smaller pellets lost some of the

nutrients at faster rate than the larger pellets. While DM loss closely corresponded to the N-loss

144
(e.g., 46% and 42%, respectively, after 10-min in water for 2-mm pellets), loss of lipid was

significantly higher for all treatments (e.g., 58% after 10-min in water for 2-mm pellets).

6.4.3. Conclusion

This study provided general guidelines and methods for assessing quality of mash and

steam-pressed pellets of an experimental aquaculture feed. Significant changes in oil absorption

of dietary mash and in bulk density, sinking velocity, and hardness of steam-pressed pellets

occurred with increasing level of purified lignin. Dispersing, binding and emulsifying properties

of lignin are different than other binders such as cellulose or guar gum contributing to the

changes in the physical properties of mash and pellets. Pellet water stability is highly vulnerable

to the changes in ingredient composition of the diet, physical and chemical properties of the

ingredients, and processing and manufacturing conditions. While highly durable and water stable

pellets that are able to withstand shear handling and distribution stresses, and disturbances in the

water are desirable to the farmers, they may not be suitable for digestion and absorption of

nutrients (Hilton et al., 1981). Careful attention to modification of processing and manufacturing

conditions is therefore needed to maintain the delicate balance of producing highly stable pellet.

145
CHAPTER 7
GENERAL DISCUSSION

7.1. Evaluation of A Novel Plant Protein Ingredient: Why is It Important?

Modern compounded aquaculture feeds include protein sources ranging from fish meal to

variety of animal and plant products. In the past few decades, several animal by-products such as

poultry meal and blood meal have been tested and effectively utilized in salmonid diets (Bureau

et al., 1999; Gaylord and Barrow, 2008). However, world production of the animal by-products

(13 MMT) is limited and much less than the production of plant proteins (Table 2.1). The advent

of novel processing techniques and avaialability of high quality products with excellent amino

acid profiles comparable to fish meal have caused plant protein sources to become more

prominent ingredients in aquatic animal diets (Gaylord et al., 2010). These novel techniques have

also allowed manufacturers to remove some of the inherent anti-nutrients that limited usage of

plant protein sources. Many high-quality plant ingredients are a complex mixtures of proteins,

fibre and starch components that require thorough evaluation in order to determine their

nutritional value and appropriate use levels in prospective diets (Glencross 2007b). Gatlin et al.

(2007) recently reviewed the use of sustainable plant products in aquafeeds and highlighted the

need for better characterization of these ingredients. Glencross et al. (2007b) also reviewed

strategies generally employed in ingredient evaluation of aquaculture feeds and suggested

several general but key components of the ingredient evaluation process, starting with

characterization of macro- and micro-nutrients (Figure 7.1), digestibility and utilization of

nutrients, effects of anti-nutritional components on nutrient utilization and health condition of the

target animals, and finally, functional properties.

146
An important evalution criterion for any protein source has traditionally been to

determine the digestibility of crude protein. Fish feeds are increasingly formulated digestible

amino acid basis and the paucity of information on available amino acid content in ingredients

forced feed formulators to use broader safety margins when formulating feeds (Bureau, 2006).

Accurate information on the available amino acid contents in traditional and novel ingredients is,

therefore, crucial for cost-effective formulation of modern aquaculture feeds.

7.2. Digestibility of Amino Acids from Plant Protein Sources

In fish, digestibility of nutrients is usually assessed as total G-I tract digestibility, using

indirect methods by adding indigestible markers to the diet, and expressed as apparent

digestibility coefficients (ADC). When evaluating a protein ingredient, typically a portion of a

standard reference diet is partially replaced with the test ingredient, and then the ADC of

nutrients in the test ingredient is calculated using the indirect method from ADC values obtained

for the reference diet and the test diets (Bureau and Hua, 2006; Bureau et al, 1999; Cho et al.,

1982; Forster, 1999). Several in-vitro methods, such as enzymatic- or acid-digestion methods or

a combination of both have also been developed (Baasompierre et al., 1998; Carter et al., 1999;

Dimes and Haard, 1994). Various methods of assessing digestibility of nutrients in fish are

described in the Chapter 2.

Several factors may affect the digestibility values in replacement studies and they are: the

amount of test ingredient included in a test diet (20-40%), the types of indigestible markers,

faecal collection methods, and composition of fibre components and anti-nutritional factors.

Fibre components and anti-nutritional factors constitute a significant portion of the plant

proteins, composition of which vary considerably based on species, genetics, source, processing

147
techniques and conditions. These components in plant proteins may substantially interfere with

nutrient utilization, growth and intestinal morphology of salmonids (Bureau et al., 1998; Francis

et al., 2001; Kaushik et al., 1995; Krogdahl et al., 2010; Wang et al., 2009a). Fractionation of

different components to produce high quality proteins helps to isolate some fibre components but

may increase the concentration of some ANFs. Some of the ANF can be removed by further

mechanical, chemical or biological treatments.

The key assumptions when using indirect methods are that there are no interactions

among ingredients that differentially affect digestibility, and altering the level of an ingredient in

a diet does not change the digestibility values of that ingredient. It has been shown that ADC of

nutrients of some ingredients, particularly ingredients with high carbohydrate content, varies

considerably (Stone, 2003). A critical review of various assumptions in a typical digestibility

study, and their implications for derived ADC values can be found in Glencross et al. (2007b).

In this study, ADC of crude protein and amino acids of two novel plant protein

ingredients, Indian mustard protein concentrate and Indian mustard protein meal, were compared

to those of a commonly used soy protein concentrate in rainbow trout and Atlantic salmon using

the indirect method, where 30% of a high quality reference diet was replaced with the three test

ingredients (Chapter 3). High ADC values of most essential amino acids (>90%) of these novel

ingredients are comparable to those of other commonly used protein sources indicating the

possibility of their utilization as excellent protein sources in salmonid diets.

Among the essential amino acids, ADC of lysine, isoleucine and phenylalanine were

significantly lower in fish fed Indian mustard protein concentrate than in those fed the Indian

mustard protein meal and soyprotein concentrate. It is hypothesized that low ADC of crude

protein and some amino acids in IMC for rainbow trout could be related to its high PA content

148
(4.2%) compared to IMM (2.8%) and SPC (1.8%). PA intake by rainbow trout fed 30%-IMC

diet was significantly higher (37 mg fish-1 d-1) than those fed 30%-IMM (19 mg fish-1 d-1) and

30%-SPC (25 mg fish-1 d-1) diets. A relationship may exist between the ADCs of crude protein

observed in this study and the PA intake of the animals. High dietary PA intake appears to have

more significant impact on the ADC of the lowest pKa (dissociation constant) His than the

higher pKa AA, Lys and Arg (Figure 3.1). Histidine, lysine, and arginine are highly susceptible

to chelation in both stomach and intestine as they become polar and positively charged at pH

below their pKas (6.1, 10.5 and 12.5, respectively; Cosgrove, 1966). Broken-line regression

analysis of ADC of Phe and PA intake showed that ADC of the AA was not affected below the

PA intake of 17 mg-1 fish d-1 (Figure 3.2). The results suggested that the level of phytate intake

may also influence the AA utilization in rainbow trout besides the formation of indigestible PA-

AA binary or PA-mineral-AA ternary complexes.

7.3. Bio-availability of Amino Acid and Slope-ratio Assay

In Chapter 4, bioavailability of arginine for fish from two novel plant proteins, Indian

mustard protein concentrate and Indian mustard protein meal, and a commercially available soy

protein concentrate relative to those from a crystalline L-Arg source was explored. Both TGC

and protein deposition were significantly linear in relation to both Arg level and intake, and used

as response variables for the assay.

The findings indicated a significantly higher Arg availability from novel Indian mustard

protein concentrate and protein meals than that from the crystalline L-Arg. The responses tended

to plateau in regression analysis with dietary Arg level as predictor resulting in reduced

bioavailability of Arg with increasing dietary Arg level as determined by two (~1.23 and ~1.35%

149
Arg) or three (~1.23, ~1.35% and ~1.5% Arg)point assays. In contrast, responses were

significantly linear with Arg intake and did not affect the slopes in either assay. A similar

quadratic effect was also observed by El-Haroun and Bureau (2007) in three-point analyses to

determine bioavailability of Lys from various sources in rainbow trout, where slope-ratios were

determined with Lys level (%) as predictor variable.

Several important inferences can be made from this study for a statistically valid slope-

ratio assay. The foremost criterion for a valid slope-ratio assay should be an appropriate

experimental design relevant to the chosen statistical model. For example, a linear regression

model can be performed with two levels of the target AA from the test sources but care should be

taken to maintain the AA intake below 80% of the requirement to ensure a linear response. In

such scenario, feed intake need to be controlled and equalized for all treatments. On the other

hand, there should be more than two levels of target AA for a non-linear or broken line approach,

where at least two levels must be below the requirement of the animal including the diet highly

deficient in the target amino acid.

Variations in the basal diet composition among experiments may also result in significant

variations in RBV of an AA from the same test ingredient. Practical basal diets in bioavailability

studies in fish have been formulated with either corn gluten meal or wheat gluten meal as the

major protein source (Li et al., 2008, 2009; Nangthu et al., 2007; Poppi et al., 2011). Nangthu et

al. (2007) compared lysine utilization efficiency in rainbow trout fed CGM and WG based diets

and reported that the imbalanced amino acid profile in CGM induced by excess leucine and high

Arg-Lys ratio, may have affected the feed intake, suggesting WG to be better source of AA than

the CGM. The difference in the lysine utilization efficiencies observed by Nangthu et al. (2007)

could also be an effect of the variations in digesta viscosity and passage rate because of variable

150
composition of NSPs in these two ingredients. Besides the variable compositions of other

nutrients or anti-nutrients in the test sources, variations in ingredient compositions of the basal

diets may also hinder the meaningful comparison of the results among the slope-ratio studies

illustrating the need for a standardized basal diet for bioavailability studies along with the

standardized experimental designs and feeding protocols.

7.4. Effects of Phytic Acid and Lignin on Nutrient Utilization (except minerals)

Effects of dietary phytic acid (PA) and purified lignin on growth performance,

physiological indices, and nutrient deposition in rainbow trout are described in chapter 5. Overall

feed efficiency and growth rate of fish in all treatments including those fed the control diet were

low compare to those observed in similar trials. It is likely that a portion the amino acids

formulated to be marginally adequate, were not completely available resulting in lower than

usual growth rate and feed efficiency.

Several studies suggested that reduced availability of one or more dietary AA negatively

influences feed intake in animals including fish (Aja et al., 1999; Koehnle et al., 2004;

Yamamoto et al., 2001; 2004). Feed intake in fish as in other animals is regulated by

neuropeptides that vary by origin, mode and regions of actions depending on external

environment, life-stage and composition of the diets. Several neuropeptides such as ghrelin,

neuropeptide-Y family of peptides (NPY-Y), galanin, orexins (hypocretin-1 and hypocretin-2)

and agouti-related proteins have been identified in fish as appetite stimulators (Nieminen et al.,

2004; Silverstein et al., 1998; Silverstein and Plisetskaya, 2000; Volkoff and Peter, 2001a;

Volkoff et al., 2005).

151
Among the pair-feeding treatments, lipid deposition in fish fed PA diets was slightly

higher than those fed PA-free diets. Imbalanced profile of available AA originated from the

reduced digestibility of some AA as reported in the chapter 3 may have resulted in more AA

available for gluconeogenesis resulting in higher lipid deposition by increasing fatty acid

synthetase (FAS) activity. Wang et al (2009a) also reported higher lipase activity in rainbow

trout fed untreated soybean meal based diets compared to those fed the phytase-treated diets.

FA and TAG synthesis is a highly regulated cellular process crucial to the metabolic

homeostasis of organisms. Figueiredo-Silva et al. (2010) reported significantly higher FAS

activity in the liver of blackspot seabream fed plant protein based diets (30.0 mIU mg-1 protein)

than those fed fishmeal based diets (20.5 mIU mg-1 protein) resulting in high total lipid

deposition. Lipid retention exceeded the intake in both treatments but deposited at much higher

amount in the fish fed the plant protein based diets than those fed the fish meal based diets

(188% and 127% of the intake, respectively). In contrast, activity of the enzyme lipoprotein

lipase (LPL) in adipose tissue were lower (P>0.05) in fish fed the plant protein diets (1.3 mIU

mg-1 protein) than in those fed fish meal based diets. Similar effects were also observed in the

adipose tissue of pigs fed high protein diets, where significantly lower FAS activity was

observed when fed a high protein diet than those fed low protein diets (Zhao et al., 2010). Dias et

al. (2005) suggested that AA imbalance, particularly a deficiency of exclusively ketogenic AA

such as lysine, leads to increase catabolism of other AA, leaving more carbon chains and

glucogenic substrates available for lipogenesis.

None of the growth performance and physiological indices was affected by lignin in the

pair-feeding treatments but when fed to satiety, lignin appeared to suppress the feed intake

contributing to the differences in growth and Fultons condition index. No such effects on feed

152
intake have been demonstrated in studies with pigs and broiler chickens fed purified lignin

(Schedle et al., 2008; Baurhoo et al., 2007). Diets of the satiation feeding treatment are very high

in PA contents, and a possible synergistic interaction of lignin with PA may have further reduced

the feed intake.

7.5. Beyond Tradition

Protein ingredients for aquaculture feed are usually evaluated in digestibility trials by

assessing the ADC of crude protein. Diets are increasingly formulated on a digestible amino acid

basis that requires accurate information on digestible amino acid values of the protein

ingredients. A significant amount of efforts have recently been invested to determine digestible

amino acid values of protein ingredients of plant origin (Gatlin et al., 2007; Hardy 2010).

However, digestibility is a measure of disappearance of nutrients that doesnt reflect the

utilization of nutrients by the animal for deposition and metabolism. A more practical approach

is to determine the bioavailability of amino acids using growth trials, and protein deposition as

an indicator variable.

Bioavailability assays are attractive because the response criteria are similar to those

under practical conditions. A potential advantage of these assays is that metabolic costs to the

animal that are induced by feeding a particular ingredient and that are associated with digestion

and absorption are reflected in this measure of availability. On the other hand, these assays are

tedious and costly, and can only determine the availability of one amino acid at a time. These are

the main reasons that there are so few studies on the bioavailability of amino acids in various fish

species. In addition, most of these studies are not properly designed for a statistically valid slope-

ratio response. There is a serious need for standardizing the experimental design, research

153
protocols and ingredient composition of the basal diets for valid determination of the

bioavailability of amino acids that is comparable among laboratories and the research groups.

Plant protein ingredients also contain antinutritional compounds and fibrous components

that usually interfere with nutrient utilization in animals. In the last two decades, studies on these

components have mainly focussed on nutrient digestibility, feed intake, growth, nutrient

deposition and retention efficiencies providing some very useful information on the effects of

these anti-nutrients on various economically important fish species including salmonids. There

has been increasing number of studies on various hormones and their effects on the regulation of

enzymes affecting feed intake, digestibility, and nutrient utilization and partitioning dynamics in

various fish species in recent years. Plant proteins now represent the majority of the dietary

proteins in fish feed and therefore, there is a serious need to improve our understanding on how

anti-nutrients and various fibre components affect regulation of these hormones and enzymatic

activities in fish species.

A framework for evaluation of plant protein ingredients is provided in Figure 7.2. It can

be concluded that a comprehensive evaluation of a plant protein ingredient for their inclusion in

compound aquaculture feed should include series of studies encompassing digestibility and bio-

availability of amino acids, effects of anti-nutrients and fibre components, effects of these

ingredients on physical quality of the pellets (chapter 6), and probably, how inclusion of these

ingredients and their various antinutrients and fibre components affect the nutrient utilization

dynamics and welfare of the fishes.

154
Figure 7.1 A flow-diagram for characterization of the components of a plant protein ingredient.

155
Figure 7.2 A frame-work for evaluation of a plant protein ingredient.

156
Bibliography

Aas, T. S., Terjesen, B. F., Sigholt, T., Hillestad, M., Holm, J., Refstie, S., Baeverfjord, G.,

Rorvik, K.A., Sorensen, M., Maikeoehme, sgrd, T., 2011. Nutritional responses in

rainbow trout (Oncorhynchus mykiss) fed diets with different physical qualities at stable

or variable environmental conditions. Aquac. Nutr. x-x, doi: 10.1111/j.1365-

2095.2011.00868.x.

Adeola O., Sands J.S., 2003. Does supplemental dietary phytase improve amino acid utilization?

A perspective that it does not. Anim. Sci. 81, 7885.

Adeola, O., 1996. Bioavailability of tryptophan in soybean meal for 10-kg pigs using slope-ratio

assay. Anim. Sci. 74, 24112419.

Adeola, O., 1998. Bioavailability of tryptophan in soybean meal and tryptophan retention in the

carcasses of four-week old ducks. Poult. Sci. 77, 1312-1319.

Adeola, O., 2009. Bioavailability of threonine and tryptophan in peanut meal for starter pigs

using slope-ratio assay. Animal 3, 677-684.

Aja, S.M., Barrett, J.A., Gietzen, D.W., 1999. CCKA and 5-HT3 receptors interact in anorectic

responses to amino acid deficiency. Phar. Biochem. Behav. 62, 487-491.

Alam, M.S., Teshima, S., Koshio, S., Ishikawa, M., 2004. Effects of supplementation of coated

crystalline amino acids on growth performance and body composition of juvenile kuruma

shrimp Marsupenaeus japonicas. Aquac. Nutr. 10, 309-316.

Alam, M.S., Teshima, S., Koshio, S., Ishikawa, M., Uyan, O., Hernandez, L.H.H., Michael, F.R.,

2005. Supplemental effects of coated methionine and/or lysine to soy protein isolate diet

for juvenile kuruma shrimp, Marsupenaeus japonicas. Aquaculture 248, 13-19.

157
Ambardekar, A.A., Reigh, R.C., Williams, M.B., 2009 Absorption of amino acids from intact

dietary proteins and purified amino acid supplements follows different time-courses in

channel catfish (Ictalurus punctatus). Aquaculture 291, 179-187.

Amirkolaie, A.K., Leenhouwers, J.I., Verreth, J.A.K., Schrama, J.W., 2005. Type of dietary fibre

(soluble versus insoluble) influences digestion, faeces characteristics and faecal waste

production in Nile tilapia (Oreochromis niloticus L.). Aquac. Res. 26, 1157-1166.

Ammerman, C.B., Baker, D.H., Lewis, A.J., 1995. Bioavailability of Nutrients for Animals:

Amino Acids, Minerals, and Vitamins. Academic Press, San Diego, CA.

Anderson, J., Jackson, A.J., Matty, A.J., Capper, B.S., 1984. Effects of dietary carbohydrate and

fibre on the tilapia Oreochromis niloticus (Linn.). Aquaculture 37, 303314.

Anderson, J.W., Chen, W-J.L., 1979. Plant fiber, carbohydrate and lipid metabolism. Am. J Cl.

Nutr. 32, 346-363.

AOAC, 1995. Official Methods of Analysis of AOAC International. Vol. I. Agriculture

chemicals; contaminants, drugs. 16th ed. AOAC International, Arlington.

Applqvist, L.A., 1972. Chemical constituents of rapeseed. In: Applqvist, L.A., Ohlson, R. (Eds)

Rapeseed Cultivation, Composition, Processing and Utilization, Elsevier, New York, pp.

123-173.

Atkinson, J.L., Hilton, J.W., Slinger, S.J., 1984. Evaluation of acid-insoluble ash as an indicator

of feed digestibility in rainbow trout (Salmo gairdneri). Can. J Fish. Aquat. Sci. 41, 1384-

1386.

Austreng, E., 1978. Digestibility determination in fish using chromic oxide marker and analysis

of contents from different segments of the gastrointestinal tract. Aquaculture 13, 265-272.

158
Austreng, E., Storebakken, T., Thomassen, M.S., Refstie, S., Thomassen, Y., 2000. Evaluation of

selected trivalent metal oxides as inert markers used to estimate apparent digestibility in

salmonids. Aquaculture 188, 65-78.

Azevedo, P.A., Leeson, S., Cho, C.Y., Bureau, D.P., 2004. Growth and feed utilization of large

size rainbow trout (Oncorhynchus mykiss) and Atlantic salmon (Salmo salar) reared in

freshwater: diet and species effects, and responses over time. Aquac. Nutr. 10, 401-411.

Baeverfjord, G., Krogdahl, . (1996) Development and regression of soybean meal induced

enteritis in Atlantic salmon, Salmo salar L., distal intestine: a comparison with the

intestines of fasted fish. Fish Dis. 19, 375-387.

Baeverfjord, G., Refstie, S., Krogdahl, P., sgrd, T., 2006. Low feed pellet water stability and

fluctuating water salinity cause separation and accumulation of dietary oil in the stomach

of rainbow trout (Oncorhynchus mykiss). Aquaculture 261, 1335-1345.

Baker, D.M., Larsen, D.A., Swanson, P., Dickhoff, W.W., 2000. Long-term peripheral treatment

of immature coho salmon (Oncorhynchus kisutch) with human leptin has no clear

physiologic effect. Gen. Comp. Endocrinol. 118, 134-138.

Bakke, A.M., Glover, C., Krogdahl, A., 2011. Feeding, digestion, and absorption of nutrients. In:

Grosell, M., Farrell, A.P., Brauner, C.J. (Eds.) The Multifunctional Gut of Fishes, Fish

Physiology Series 30, Elsevier, USA, pp. 57-112.

Bakke-McKellep, A.M., Nordrum, S., Krogdahl, A., Buddington, R.K., 2000. Absorption of

glucose, amino acids, and dipeptides by the intestines of Atlantic salmon (Salmo salar L).

Fish Physiol. Biochem. 22, 33-44.

Bassompiere, M., Kristiansen, H.R., McLean, E., 1998. Influence of weight upon in vitro protein

digestion in rainbow trout. Fish Biol. 52, 213-216.

159
Batterham, E.S., 1987. Availability of lysine in for grower pigs. Proceedings of 1st Conference of

the Australasian Pig Science Association, Werribee, Australia, pp. 121-123.

Batterham, E.S., 1992. Availability and utilization of AA for growing pigs. Nutr. Res. Rev. 5, 1-

18.

Batterham, E.S., Andersen, L.M., Baigent, D.R., White, E., 1990. Utilization of ileal digestible

amino acids by growing pigs: Effect of dietary lysine concentration on efficiency of

lysine retention. Br. J. Nutr. 64, 8194.

Batterham, E.S., Lowe, R.F., Darnell, R.E., Major, E.J., 1986. Availability of lysine in meat

meal, meat and bone meal and blood meal as determined by slope ratio assay with

growing pigs, rats, and chicks and by chemical techniques. Br. J. Nutr. 55, 427440.

Batterham, E.S., Murison, R.D., Lewis, C.E., 1978. An evaluation of total lysine as a predictor of

lysine status in protein concentrates for growing pigs. Br. J. Nutr., 40, 23-28.

Batterham, E.S., Murison, R.D., Lewis, C.E., 1979. Availability of lysine in protein concentrates

as determined by the slop-ratio assay with growing pigs and rats by chemical techniques.

Br. J. Nutr. 97, 383-391.

Baurhoo, B., Letellier, A., Zhao, X., Ruiz-Feria, C.A., 2007a. Cecal populations of lactobacilli

and bifidobacteria and Escherichia coli populations after in vivo Escherichia coli

challenge in birds fed diets with purified lignin or mannanoligosaccharides. Poult. Sci.

86, 2509-2516.

Baurhoo, B., Phillip, L., Ruiz-Feria, C.A., 2007b. Effects of purified lignin and mannan

oligosaccharides on intestinal integrity and microbial populations in the ceca and litter of

broiler chickens. Poult. Sci. 86, 1070-1078.

160
Baurhoo, B., Ruiz-Feria, C.A., Zhao, X., 2008. Purified lignin: Nutritional and health impacts on

farm animalsA review. Anim. Feed Sci. Tech. 144, 175-184.

Bell, J.G., Henderson, R.J., Tocher, D.R., McGhee, F., Dick, J.R., Porter, A., Smullen, R.P.,

Sargent, J.R., 2002. Substituting fish oil with crude palm oil in the diet of Atlantic salmon

(Salmo salar) affects muscle fatty acid composition and hepatic fatty acid metabolism.

Nutrition 132, 222-230.

Bell, J.M., 1993. Factors affecting the nutritional value of canola meal: a review. Can. J Anim.

Sci, 73, 679697.

Bender, D.A., 2008. Introduction to Nutrition and Metabolism. CRC Press, NY, 416p.

Berge, G.E., Bakke-McKellep, A.M., Lied, E., 1999. In vitro uptake and interaction between

arginine and lysine in the intestine of Atlantic salmon (Salmo salar). Aquaculture 179,

181-193.

Bertolo, R.F., Moehn, S., Pencharz, P.B., Ball, R.O. 2005. Estimate of the variability of the

lysine requirement of growing pigs using the IAAO technique. Anim. Sci. 83, 2535-2542.

Best, P., 2009. World Feed Panorama: Global feed volumes grow again. 2009(1), 12-15.

Boerjan, W., Ralph, J., Baucher, M., 2003. Lignin biosynthesis. An. Rev. Plant Biol. 54, 519-

546.

Bostock, J., McAndrew, B., Richards, R., Jauncy, K., Telfer, T., Lorenzen, K., Little, D., Ross,

L., Handisyde, N., Fatward, I., Corner, R., 2010. Aquaculture: global status and trends.

Phil Trans. R. Soc. 365, 2897-2912.

Brinker, A., 2007. Guar gum in rainbow trout (Oncorhynchus mykiss) feed: The influence of

quality and dose on stabilisation of faecal solids. Aquaculture 267, 315-327.

161
Brinker, A., Reiter, R. 2011. Fish meal replacement by plant protein substitution and guar gum

addition in trout feed, Part I: Effects on feed utilization and fish quality. Aquaculture 310,

350-360.

Bromley, P.J., Adkins, T.C., 1984. The influence of cellulose filler on feeding, growth and

utilization of protein and energy in rainbow trout, Salmo gairdnerii Richardson. Fish

Biol. 24, 235-244.

Buddington, R.K., Chen, J.W., Diamond, J., 1987. Genetic and phenotypic adaptation of

intestinal nutrient transport in fish. Physiology 393, 261-281.

Bureau, D.P. and Cho, C.Y., 1997. Ingredients quality: An essential factor in the formulation of

cost-effective diets in aquaculture. In: Duerr, E.O., Johnson, L., Reagen, M., (Eds),

Expanding Agriculture Co-Product Uses in Aquaculture Feeds. Workshop Proceedings.

December 5-7, 1994, Des Moines, Iowa, 234-274.

Bureau, D.P., 2006. Rendered products in fish aquaculture feeds. In: Meeker, D.L. (Ed)

Essential Rendering: All about The Animal By-products Industry. National Renderers

Association, Arlington, Virginia, pp. 179-194.

Bureau, D.P., 2008. Meaningful characterization of the nutritive value of processed animal

products. Intl. AquaFeed 11(5), 18-22.

Bureau, D.P., Harris, A.M., Cho, C.Y., 1998. The effects of purified alcohol extracts from soy

products on feed intake and growth of Chinook salmon (Oncorhynchus tshawytscha) and

rainbow trout (Oncorhynchus mykiss). Aquaculture 161, 27-43.

Bureau, D.P., Harris, A.M., Cho, C.Y., 1999. Apparent digestibility of rendered animal protein

ingredients for rainbow trout (Oncorhynchus mykiss). Aquaculture 180, 345-358.

Bureau, D.P., Hua, K., 2006. Letter to the editor of Aquaculture. Aquaculture, 252:103-105.

162
Bureau, D.P., Hua, K., 2010. Towards effective nutritional management of waste outputs in

aquaculture, with particular reference to salmonid aquaculture operations. Aquac. Res.

41, 777-792.

Bureau, D.P., Hua, K., Cho, C.Y., 2006. Effect of feeding level on growth and nutrient

deposition in rainbow trout (Oncorhynchus mykiss Walbaum) growing from 150 to 600 g.

Aquac. Res. 37, 1090-1098.

Bureau, D.P., Hua, K., Harris, A.M., 2008. The effect of dietary lipid and long-chain n-3 PUFA

levels on growth, energy utilization, carcass quality, and immune function of rainbow

trout, Oncorhynchus mykiss. J. World. Aquac. Soc. 39, 1-21.

Bureau, D.P., Meeker, D.L., 2009. Terrestrial animal fats. In: Turchini, G.M., Ng, W-K., Tocher,

D.R. (Eds), Fish Oil Replacement and Alternative Lipid Sources in Aquaculture Feeds,

CRC Press, NY, pp. 245-266.

Burel, C., Boujard, T., Escaffre, A.M., Kaushik, S.J., Boeuf, G., Mol, K.A., Van der Geyten, S.,

Kuhn, E.R., 2000. Dietary low-glucosinolate rapeseed meal affects thyroid status and

nutrient utilization in rainbow trout (Oncorhynchus mykiss). Br. J Nutr. 83, 653-664.

Burton, W.A., Ripley, V.L., Potts, D.A., Salisbry, P.A., 2004. Assessment of genetic diversity in

selected breeding lines and cultivars of canola quality Brassica juncea and their

implication for canola breeding. Euphytica 136, 181-192.

Caldwell, R.A., 1992. Effect of calcium and phytic acid on the activation of the trypsinogen and

the stability of trypsin. Agri. Food Chem. 40, 43-46.

Carter, C.G., Bransden, M.P., Van Barneld, R.J., Clarcke, S.M., 1999. Alternative methods for

nutrition research on the southern bluefin tuna, Thunnus maccoyii: in vitro digestibility.

Aquaculture 179, 59-70.

163
Carter, C.G., Hauler, R.C., 2000. Fish meal replacement by plant meals in extruded feeds for

Atlantic salmon, Salmo salar L. Aquaculture 185, 299-311.

CCAC, 1984. Wild vertebrates in the field and in the laboratory. In: Guide to the care and use of

experimental animals. Vol. 2., Canadian Council on Animal Care, Ottawa, Ont.: CCAC

1984, pp. 191-208.

Chauhan, J.S., Tyagi, M.K., Kumar, A., Nashaat, N.I., Singh, M., Singh, N.B., Jakhar, M.L.,

Welham, S.J., 2007. Drought effects on yield and its components in Indian mustard

(Brassica juncea L.). Plant Breed. 126, 399-402.

Cheng, Z., Ai, Q., Mai, K., Xu, W., Ma, H., Li, Y., Zhang, J., 2010. Effects of dietary canola

meal on growth performance, digestion and metabolism of Japanese seabass, Lateolabrax

japonicas. Aquaculture 305, 102-108.

Cheng, Z.J., Hardy, R.W., 2003. Effect of microbial phytase on apparent nutrient digestibility of

barley, canola meal, wheat and wheat middlings, measured in vivo using rainbow trout

(Oncorhynchus mykiss). Aquac. Nutr. 8, 271-277.

Cheryan M., Rackis, J., 1980. Phytic acid interactions in food systems. Crit. Rev. Food Sci. Nutr.

13, 297335.

Chevre, A.M., Brun, H., Eber, F., Letanneur, J.C., Vallee, P., Ermel, M., Glais, I., Hua, L.,

Sivasithamparam, K., Barbetti, M.J., 2008. Stabilization of resistence to Leptosphaeria

maculans in Brassica napus-B.juncea recombinant lines and its introgression into spring-

type Brassica napus. Plant Dis. 92, 1208-1214.

Chi, S.Y., Tan, B.P., Lin, H.Z., Mai, K.S., Ai, Q.H., Wang, X.J., Zhang, W.B., Xu, W., Liufu,

Z.G., 2011. Effects of supplementation of crystalline or coated methionine on growth

164
performance and feed utilization of the pacific white shrimp, Litopeneaeus vannamei.

Aquac. Nutr. 17, e1-e9.

Chiou, P.W., Yu, B., Lin, C., 1994. Effect of different components of dietary fiber on the

intestinal morphology of domestic rabbits. Comp. Biochem. Physiol. 108, 629-638.

Cho, C.Y., 1992. Feeding systems for rainbow trout and other salmonids with reference to

current estimates of energy and protein requirements, Aquaculture 100, 107123.

Cho, C.Y., Bureau, D.P., 1995. Determination of dietary energy requirement of fish with

particular reference to salmonids. Appl. Ich. 11, 41-165.

Cho, C.Y., Bureau, D.P., 2001. A review of diet formulation strategies and feeding systems to

reduce excretory and feed wastes in aquaculture. Aquac. Res. 32, 349-360.

Cho, C.Y., Kaushik, S., Woodward, B., 1992. Dietary arginine requirement of young rainbow

trout (Oncorhynchus mykiss). Comp. Biochem. Physiol. 102A, 211216.

Cho, C.Y., Slinger, S.J., 1979. Apparent digestibility measurement in feedstuffs for rainbow

trout. In: Halver, J.E., Tiews, K., (Eds.), Finfish Nutrition and Fishfeed Technology,

Heenemann-Verkagsgesellschaft, Berlin, pp 239-247.

Cho, C.Y., Slinger, S.J., Bayley, H.S., 1982. Bioenergetics of salmonid fishes: Energy intake,

expenditure and productivity. Comp. Biochem. Physiol. 73B, 25-41.

Choubert, G. jr., de la Noue, J., Luquet, P., 1982. Digestibilty in fish: improved device for the

automatic collection of faeces. Aquaculture 29, 185-189.

Cohen, A., Meys, M., Tarvin, T.L., 1989. The Pico-Tag Method. A Manual of Advanced

Techniques for Amino Acid Analysis, Waters Division of Millipore, Milford, MA.

165
Coon, C.N., Seo, S., Manangi, M.K., 2007. The determination of retainable phosphorus, relative

biological availability, and relative biological value of phosphorus sources for broilers.

Poult. Sci. 86, 857-868.

Cosgrove, D.J., 1966. The chemistry and biochemistry of inositol polyphosphates. Rev. Pure

Appl. Chem. 16, 209-215.

Costello, A.J. R., Glonek, T., Myers, T.C., 1976. 31P nuclear magnetic resonance-pH titrations

of myo-inositol hexaphosphate. Carbohydr. Res. 46, 159171.

Cowieson, A.J., Acamovic, T., Bedford, M.R., 2006. Phytic acid and phytase: implications for

protein utilization by poultry. Poult. Sci. 85, 878-885.

Cowieson, A.J., Ravindran, V., Selle, P.H., 2008. Influence of dietary phytic acid and source of

microbial phytase on ileal endogenous amino acid flows in broiler chickens. Poult. Sci.

87, 2287-2299.

Dabrowska H., Wojno, T., 1977. Studies on the utilization by rainbow trout (Salmo gairdneri) of

feed mixtures containing soybean meal and an addition of amino acids. Aquaculture, 10,

297310.

Dabrowski, K., Arslan, M., Terjesen, B.F., Zhang, Y., 2007. The effect of dietary indispensible

amino acid imbalances on feed intake: Is there a sensing of deficiency and neural

signalling present in fish? Aquaculture 268, 136-142.

Dabrowski, K., Dabrowska, H., 1981. Digestion of protein by rainbow trout (Salmo gairdneri

Rich.) and absorption of amino acids within alimentary tract. Comp. Biochem. Physiol.

69A, 99-111.

166
Dabrowski, K., Poczyczynski, P., Kock, G., Berger, B., 1989. Effect of partially or totally

replacing fish meal protein by soybean meal protein on growth, food utilization and

proteolytic enzyme activities in rainbow trout (Salmo gairdneri). New in vivo test for

exocrine pancreatic secretion. Aquaculture 77, 2949.

Davidson, P.M., Brenen, A.L., 1981. Antimicrobial activity of non-halogenated phenolic

compounds. Food Prot. 44, 623632.

Davies, S.J., Gouveia. A., 2006. Comparison of yttrium and chromic oxide as inert dietary

markers for the estimation of apparent digestibility coefficient in mirror carp Cyprinus

carpio fed on diets containing soybean, maize and fish derived proteins. Aquac. Nutr.12,

451458.

De Lange, C.F.M., Sauer, W.C., Mosenthin, R., Souffrant, W.B., 1989. The effect of feeding

different protein-free diets on the recovery and amino acid composition of endogenous

protein collected from the distal Ileum and feces in pigs. Anim. Sci. 67, 746-754.

De Silva, S.S., Francis, D.S., Tacon, A.G.J., 2011. Fish oils in aquaculture in retrospect. In:

Turchini, G.M., Ng., W-K., Tocher, D.R. (Eds) Fish oil replacement and alternative lipid

sources in aquaculture feeds, pp. 1-20.

De Silva, S.S., Gunasekera, R.M., Gooley, G., 2000. Digestibility and amino acid availability of

three protein-rich ingredient-incorporated diets by Murray cod Maccullochella peelii

peelii (Mitchell) and the Australian shortfin eel Anguilla australis Richardson. Aquac.

Res. 31, 195-205.

Deguara, S., Jauncey, K., Agius, C.. 2003. Enzyme activities and pH variations in the digestive

tract of gilthead sea bream. J. Fish Biol. 62, 10331043.

167
Denstadli, V., Skrede, A., Krogdahl, A., Sahlstrom, S., Storebakken, T., 2006. Feed intake,

growth, feed conversion, digestibility, enzyme activities and intestinal structure in

Atlantic salmon (Salmo salar L.) fed graded level of phytic acid. Aquaculture 256, 365-

376.

Dharmawardhana, D.P., Ellis, B.E., 1998. -Glucosidase and glucosyltransferases in lignifying

tissues. In: Lewis, N.G., Sarkanen, S. (Eds.) Lignin and Lignan Biosynthesis, ACS

Symposium Series 697. American Chemical Society, Washington, DC, USA. pp. 7683.

Dias, J., Alvarez, M.J., Arzel, J., Corraze, G., Diez, A., Bautista, J.M., Kaushik, S.J., 2005.

Dietary protein source affects lipid metabolism in the European sea bass (Dicentrarchus

labrax). Comp. Biochem. Physiol. 142A, 19-31.

Dimes, L.E., Garcia-Carreno, F.L., Haard, N.F., 1994. Estimation of protein digestibility. III.

Studies on the digestive enzymes from the pyloric ceca of rainbow trout and salmon.

Comp. Biochem. Physiol. 109, 349-360.

Dimes, L.E., Haard, N.F., 1994. Estimation of protein digestibility I. Development of an In-

vitro method for estimating protein digestibility in salmonids (Salmo gairdneri). Comp.

Biochem. Physiol. 108, 349-362.

Draganovic, V., Goot, A.J., Boom, R.V-D., Jonkers, J., 2011. Assessment of the effects of fish

meal, wheat gluten, soy protein concentrate and feed moisture on extruder system

parameters and the technical quality of fish feed. Anim. Feed Sci. Tech. 165, 238-250.

Drew, M.D., Borgeson, T.L., Thiessen, D.L., 2007. A review of feed ingredients to enhance diet

digestibility in finfish. Anim. Feed Sci. Tech. 138, 118-136.

Drew, M.D., Ogunkoya, A.E., Janz, D.M., Van Kessel, A.G., 2007. Dietary influence of

replacing fish meal and oil with canola protein concentrate and vegetable oils on growth

168
performance, fatty acid composition and organochlorine residues in rainbow trout

(Oncorhynchus mykiss). Aquaculture 267, 260-268.

Durand, D., 2011. Aquafeed twin-screw extrusion processing : versatile and ideal for aquafeed.

Intl. Aquafeed 14(3), 30-33.

El-Haroun, E.R., Azevedo, P.A., Bureau, D.P., 2009. High dietary incorporation levels of

rendered animal protein ingredients on performance of rainbow trout Oncorhynchus

mykiss (Walbaum, 1972). Aquaculture 290, 269-274.

El-Haroun, E.R., Bureau, D.P. 2007. Comparison of the bioavailability of lysine in blood meals

of various origins to that of l-lysine HCL for rainbow trout (Oncorhynchus mykiss).

Aquaculture 262, 402-409.

Encarnao, P., De Lange, C., Rodehutscord, M., Hoehler, D., Bureau, W., Bureau, D.P., 2004.

Diet digestible energy content affects lysine utilization, but not dietary lysine equirements

of rainbow trout (Oncorhynchus mykiss) for maximum growth. Aquaculture 235, 569

586.

Enes, P., Panserat, S., Kaushik, S.J., Oliva-Teles, A., 2008. Growth performance and metabolic

utilization of diets with native and waxy maize starch by gilthead seabream (Sparus

aurata) juveniles. Aquaculture 274, 101-108.

Erdman, J. W., Jr., 1979. Oilseed phytates: nutritional implications. J. Am. Oil Chem. Soc. 56,

736-741.

Fanimo, A., Susenbeth, A., Sudekum, K., 2006. Protein utilisation, lysine bioavailability and

nutrient digestibility of shrimp meal in growing pigs. Anim. Feed Sci. Tech. 129, 196-

209.

169
FAO., 2011. Food and Agriculture Organization of the United Nations, FISHSTAT Plus:

Universal Software for Fishery Statistical Time-Series (Food and Agriculture

Organization), Rome, Version 2.32.

Fard, M.R.S., Weisheit, C., Poynton, S.L., 2007. Intestinal pH profile of rainbow trout

Oncorhynchus mykiss and microhabitat preference of the flagellate Spironucleus

salmonis (Diplomonadida). Dis. Aquat. Org. 76, 241-249.

Fenwick, G.R., Spinks, E.A., Wilkinson, A.P., Heaney, R.K., Legoy, M.A., 1986. Effect of

processing on the antinutrient content of rapeseed. J Sci. Food Agri. 37, 735-741.

Fernandez, F., Miquel, A.G., Martinez, R. Serra, E., Guinea, J., Narbaiza, F.J., Caseras, A.,

Baanante, I.V., 1999. Dietary chromic oxide does not affect the utilization of organic

components but can alter the utilization of mineral salts in gilthead seabream Sparus

aurata. Nutrition 129, 1053-1059.

Fernandez, R., Phillips, S.F., 1982. Components of fiber bind iron in vitro. Am. J Cl. Nutr. 35,

100-106.

Figueiredo-Silva, A.C., Corraze, G., Borges, P., Valente, L.M.P., 2010. Dietary protein/lipid

level and protein source effects on growth, tissue composition and lipid metabolism of

blackspot seabream (Pagellus bogaraveo). Aquacut. Nutr. 16, 173-187.

Finney, D.J., 1951. The statistical analysis of slope-ratio assays. J. Gen. Microbiol. 5, 223-230.

Finney, D.J., 1964. The meaning of bioassay. Biometr. 21, 785-798.

Finney, D.J., 1971. Statistical Method in Biological Assay, 2nd Edition, Griffin Publishers,

London.

170
Fonseca-Madrigal, J., Karalazos, V., Campbell, P.J., Bell, G.J., Tocher, D.R., 2005. Influence of

dietary palm oil on growth and tissue fatty acid compositions, and fatty acid metabolism

in liver and intestine in rainbow trout (Oncorhynchus mykiss). Aquac. Nutr. 11, 241-250.

Forgan, L.G., Forster, M.E., 2007. Development and physiology of gastric dilation air sacculitis

in Chinook salmon, Oncorhynchus tshawytscha (Walbaum). Fish Dis. 30, 459-469.

Forster, I. 1999. A note on the method of calculating digestibility coefficients of nutrients

provided by single ingredients to feeds of aquatic animals. Aquac. Nutr. 5, 143145.

Francis, D.S., Turchini, G.M., Jones, P.L., De Silva, S.S., 2007. Effects of fish oil substitution

with a mix blend vegetable oil on nutrient digestibility in Murray cod, Maccullochella

peelii peelii. Aquaculture 269, 447-455.

Francis, G., Makkar, H.P.S., Becker, K., 2001. Antinutritional factors present in plant derived

alternate fish feed ingredients and their effects in fish. Aquaculture 199, 197-227.

Friedman, M., 1992. Dietary impact of food processing. Ann. Rev. Nutr. 12, 119-137.

Fuller, M., 2003. AA bioavailability a brief history. In: Ball, R.O. (Ed) Digestive Physiology in

Pigs. Proc. Of 9th Intl. Symp. Vol. 1, University of Alberta, Alberta, Canada, pp 183-198.

Fulton, T., 1902. Rate of growth of sea fishes. Fish. Scotl. Sci. Invest. Rep., Scotland, 20 pp.

Ganong, W.F., 2009. Digestion and absorption. In: Ganong, W.F. (Ed) Reviews in Medical

Biology, 14th Edition, Appelton & Lange, London, pp. 398-407.

Gargulak, J.D., Lebo, S.E., 2000. Commercial use of lignin-based materials, in: Glasser, W.G.,

Northey, R.A., Schultz, T.P. (Eds.), Lignin: Historical, Biological, and Materials

Perspectives. ACS Symposium Series, Vol. 742, American Chemical Society,

Washington DC, pp. 304-320.

171
Gatlin, D.M. III., Phillips, H.F., 1989. Dietary calcium, phytate, and zinc interactions in channel

catfish. Aquaculture 79, 259-266.

Gatlin, D.M., Barrows, F.T., Brown, P., Dabrowski, K., Gaylord, T.G., Hardy, R.W., Herman,

E., Hu, G., Krogdahl, ., Nelson, R., Overturf, K., Rust, M., Sealey, W., Skonberg, D.,

Souza, E.J., Stone, D., Wilson, R., Wurtle, E., 2007. Expanding the utilization of

sustainable plant products in aquafeeds: a review. Aquac. Res. 38, 551-579.

Gaylord, T.G., Barrows, F.T., 2008. Apparent digestibility of gross nutrients from feedstuffs in

extruded feeds for rainbow trout, Oncorhynchus mykiss. J. World Aquacult. Soc. 39, 827-

834.

Gaylord, T.G., Barrows, F.T., Rawles, S.D., 2010. Apparent amino acid availability from

feedstuffs in extruded diets for rainbow trout Oncorhynchus mykiss. Aquac. Nutr. 16,

400-406.

Glencross, B., 2009. The influence of soluble and insoluble lupin non-starch polysaccharides on

the digestibility of diets fed to rainbow trout (Oncorhynchus mykiss). Aquaculture 294,

256-261.

Glencross, B., Carter, C.G., Duijster, N., Evans, D.R., Dods, K., McCafferty, P., Hawkins, W.E.,

Maas, R., Sipsas, S., 2004a. A comparison of the digestibility of a range of lupin and

soybean protein products when fed to either Atlantic salmon (Salmo salar) or rainbow

trout (Oncorhynchus mykiss). Aquaculture 237, 333-346.

Glencross, B., Hawkins, W., Curnow, J., 2004b. Nutritional assessment of Australian canola

meals. I. Evaluation of canola oil extraction method and meal processing conditions on

the digestible value of canola meals fed to the red seabream (Pagrus auratus, Paulin).

Aquac. Nutr. 35, 15-24.

172
Glencross, B., Hawkins, W., Evans, D. Rutherford, N., Dods, K., McCafferty, P., Sipass, S.,

2007b. Evaluation of the influence of drying process on the nutritional value of lupin

protein concentrates when fed to rainbow trout (Oncorhynchus mykiss). Aquaculture 265,

218-229.

Glencross, B., Hawkins, W., Evans, D., Rutherford, N., McCafferty, P., Dods, K., Hauler, R.

2011. A comparison of the effect of diet extrusion or screw-press pelleting on the

digestibility of grain protein products when fed to rainbow trout (Oncorhynchus mykiss).

Aquaculture 312, 154-161.

Glencross, B., Hawkins, W., Evans, D., Rutherford, N., McCafferty, P., Dods, K., Karopoulos,

M., Veitch, C., Sipsas, S., Buirchell, B., 2008. Variability in the composition of lupin

(Lupinus angustifolius) meals influences their digestible nutrient and energy value when

fed to rainbow trout (Oncorhynchus mykiss). Aquaculture 277, 220-230.

Glencross, B., Hawkins, W., Maas, R., Karopoulos, M., Hauler, R., 2010. Evaluation of the

influence of different species and cultivars of lupin kernel meal on the extrusion process,

pellet properties and viscosity parameters of salmonid feeds. Aquac. Nutr. 16, 13-24.

Glencross, B.D., Booth, M., Allan, G.L., 2007a. A feed is only as good as its ingredients ? a

review of ingredient evaluation strategies for aquaculture feeds. Aquac. Nutr. 62, 100-

134.

Gonzalez-Nuez, V., Gonzalez-Sarmiento, R., Rodriguez, R.E., 2003.Identification of two

proopiomelanocronin genes in zebrafish (Danio rerio). Brain. Res. 120, 1-8.

Grassl, E.F., 1956. Pelleted dry rations for trout propagation in Michigan hatcheries. Trans. Am.

Fish. Soc. 86, 307-322.

173
Gunasekera, C.P., Martin, L.D., Siddique, K.H.M., Walton, G.H., 2006. Genotype by

environment interactions of Indian mustard (Brassica juncea L.) and Canola (B. Napus

L.) in Mediterranean-type environments 1. Crop growth and seed yield. Eur. J. Agr. 25,

1-12.

Haard, N.F., 1992. A review of proteolytic enzymes from marine organism and their applications

in the food industry. Aquat. Food Prod. Tech. 1, 17-35.

Haard, N.F., Dimes, L.E., Arndt, R.E., Dong, F.M., 1996. Estimation of protein digestibility-IV.

Digestive proteinases from the pyloric caeca of coho salmon (Oncorhynchus kisutch) fed

diets containing soybean meal. Comp. Biochem. Physiol. 115, 533-540.

Hamdan, M., Moyano, F.J., Schuhardt, D., 2009. Optimization of a gastrointestinal model

applicable to the evaluation of bio-accessibility in fish feeds. Sci. Food Agri. 89, 1195-

1201.

Hansen, J., Storebakken, T., 2007. Effects of dietary cellulose level on pellet quality and nutrient

digestibilities in rainbow trout (Oncorhynchus mykiss). Aquaculture 272, 458-465.

Hardy, R.W. 2010. Utilization of plant proteins in fish diets: effects of global demand and

supplies of fishmeal. Aquac. Res. 41, 770-776.

Hardy, R.W., Barrows, F.T., 2002. Diet formulation and manufacture. In: Halver, J.E., Hardy,

R.W. (Eds) Fish Nutrition, 3rd Edition, Academic Press, San Diego, California, pp. 506-

601.

Hardy, R.W., Tacon, A.G.J., 2002. Fish meal: historical uses, production trends and future

outlook for supplies. In: Stickney, R.R., MacVey, J.P., (Eds). Responsible Marine

Aquaculture, CABI Publishing, New York, pp. 311-325.

174
Hemre, G.I., Sandnes, K., Lie, ., Torrisen, O., Waagb, R., 1995. Carbohydrate nutrition in

Atlantic salmon, Salmo salar, growth and feed utilisation. Aquac. Res. 26, 149154.

Hepher, B., 1990. Ingestion, digestion and absorption of food. Hepher, B. (Ed) Nutrition of

Pond Fishes, Academic Press, Cambridge, pp. 1663.

Higgs, D.A., Dosanjh, B.S., Prendergast, A.F., Beames, R.M., Hardy, R.W., Riley, W., Deacon,

G., 1995. Use of rapeseed/canola protein products in finfish diets. In: Lim, C.E., Sessa,

D.J., (Eds) Nutrition and Utilization Technology in Aquaculture, AOCS Press,

Champaign, IL, pp. 130-156.

Higgs, D.A., McBride, J.R., Markert, J.R., Dosanjh, B.S., Plotnikoff, M.D., Clarke, W.C., 1982.

Evaluation of Tower and Candle rapeseed (canola) meal and Bronowski rapeseed protein

concentrate as protein supplements in practical dry diets for juvenile chinook salmon

(Oncorhynchus tshawytscha). Aquaculture 29, 131.

Higgs, D.A., Prendergast, A.F., Dosanjh, B.S., Beames, D.M., Deacon, G., Hardy R.W., 1994.

Canola protein offers hope for efficient salmon production. In: Mackinlay, D.D. (Ed.)

High Performance Fish, Fish Physiology Association, Vancouver, BC, pp. 377382.

Hilton, J.W., Cho, C.Y., Slinger, S.J., 1981. Effect of extrusion processing and steam pelleting

diets on pellet durabilify, pellet water absorption, and the physiological response of

rainbow trout (Salmo gairdneri R.). Aquaculture 25, 185-194.

Holmgren, S., 1995. Neuropeptide control of the cardiovascular system in fish and reptiles. Braz.

J. Med. Biol. Res. 28, 1207-1216.

Hooft, J.M., Elmor, A.E.I., Encarnao, P., Bureau, D.P., 2011. Rainbow trout (Oncorhynchus

mykiss) is extremely sensitive to the feed-borne Fusarium mycotoxin deoxynivalenol

(DON). Aquaculture 311, 224-232.

175
Horn, M.H., Gawlicka, A., 2001. Digestion in fish. Encyclopedia of Life Sciences, John Wiley

and Sons Ltd., NY.

Hossain, M.A., Jauncey, K., 1989. Studies on protein, energy and amino acid digestiblility of

fish meal, mustard oil cake, linseed amd sesame meal for common carp (Cyprinus carpio

L.) Aquaculture 83, 59-72.

Hua, K., Bureau, D.P., 2009. A mathematical model to explain variations in estimates of starch

digestibility and predict digestible starch content of salmonid fish feeds. Aquaculture

294, 282-287.

Jansman, S.W., van Leeuwen, P., Rademacher, M., 2002. Evaluation through literature data of

the amount and amino acid composition of basal endogenous crude protein at the

terminal ileum of pigs. Anim. Feed Sci. Tech. 98, 49-60.

Jia, W., Slominski, B.A., Guenter, W., Humphreys, A., Jones, O., 2008. The effect of enzyme

supplementation on egg production parameters and omega-3 fatty acid deposition in

laying hens fed flaxseed and canola seed. Poult. Sci. 87, 2005-2014.

Johnson, L.A., Lusas, E.W., 1983. Comparison of alternative solvents for oils extraction. J. Am.

Oil Chem. Soc. 60, 181A-193A.

Jones, L.J., Tung, M.A., 1983. Functional properties of modified oilseed protein concentrates

and isolates. Can. Inst. Food Sci. Tech. J. 16, 57-62.

Jung, H.G., Fahey Jr. J.C., 1983. Nutritional implications of phenolic monoers and lignin: a

review. Anim. Sci. 57, 206-219.

Kannadhason, S., Muthukumarappan, K., Rosentrater, K.A., 2009. Effects of ingredients and

extrusion parameters on aquafeeds containing DDGS and tapioca starch. Aquacult. Feed

Sci. Nutr. 1, 6-21.

176
Kaufman, H.W.,Kleinberg, I., 1971. Effect of pH on Calcium Binding by Phytic Acid and its

Inositol Phosphoric Acid Derivatives and on the Solubility of their Calcium Salts. Arch.

Oral. Biol. 16, 445-460.

Kaushik, S.J., Cravedi, J.P., Lalles, J.P., Sumpter, J., Fauconneau, B., Laroche, M., 1995. Partial

or total replacement of fish meal by soybean protein on growth, protein utilization,

potential estrogenic or antigenic effects, cholesterolemia and flesh quality in rainbow

trout, Oncorhynchus mykiss. Aquaculture 133, 257-274.

Kaushik, S.J., Seiliez, I., 2010. Protein and amino acid nutrition and metabolism in fish: current

knowledge and future needs. Aquac. Res. 41, 322-332.

Ketola, H.G., 1975. Mineral supplementation of diets containing soybean meal as a source of

protein for rainbow trout. Prog. Fish-Cult. 37, 73-75

Kies, A.K., 2005. Phytase studies in pigs and poultry. Effect on protein digestion and energy

utilization. Doctoral thesis. Wageningen University, The Netherlands.

Kies, A.K., Selle, P.H., 1998. A review of the antinutritional effects of phytic acid on protein

utilisation by broilers. Proc. Aus. Poult. Sci. Symp. 10, 128-131.

Kies, A.K., Van Hemert, H.F., Sauer, W.C., 2001. Effect of phytase on protein and amino acid

digestibility and energy utilization. World Poult. Sci. 57, 109-125.

Kim, B.G., Lindemann, M.D., Rademacher, M., Brennan, J.J., Cromwell, G.L., 2006. Efficacy of

dl-methinine hydroxyl analogue free acid and dl-methionine as methionine sources for

pigs. Anim. Sci. 84, 104-111.

Kleinert, M., Barth, T., 2008. Phenols from lignin. Chem. Eng. Tech. 31, 736-745.

177
Knudsen, D., Jutfelt, F., Sundh, H., Sundell, K., Koppe, W., Frokiaer, H., 2008. Dietary soya

saponins increase gut permeability and play a key role in the onset of soyabean-induced

enteritis in Atlantic salmon (Salmo salar L.). Br. J Nutr. 100, 120-129.

Koehnle, T., Russel, M., Gietzen, D., 2004. Rats rapidly reject diets deficient in essential amino

acids. J. Nutr. 133, 23312335.

Kraugerud, O.F., Jrgensen, H.Y., Svihus, B., 2011. Physical properties of extruded fish feed

with inclusion of different plant (legumes, oilseeds, or cereals) meals. Anim. Feed Sci.

Tech. 163, 244-254.

Kraugerud, O.F., Penn, M., Storebakken, T., Refstie, S., Krogdahl, ., Svihus, B., 2007. Nutrient

digestibilities and gut function in Atlantic salmon (Salmo salar) fed diets with cellulose

or non-starch polysaccharides from soy. Aquaculture 273, 98-107.

Kritchevsky, D., 1988. Dietary fiber. Annual Rev. Nutr. 8, 301-328.

Krogdahl, ., Bakke-McKellep, A.M., Baeverfjord, G., 2003. Effects of graded levels of

standard soybean meal on intestinal structure, mucosal enzyme activities, and pancreatic

response in Atlantic salmon (Salmo salar L.). Aquac. Nutr. 9, 361-371.

Krogdahl, ., Holm, H., 1983. Pancreatic proteinases from man, trout, rat, pig, cow, chicken,

mink and fox. Enzyme activities and inhibition by soybean and lima bean proteinase

inhibitors. Comp. Biochem. Physiol. 74, 403409.

Krogdahl, ., Penn, M., Thorsen, J., Refstie, S., Bakke, A.M., 2010. Important antinutrients in

plant feedstuffs for aquaculture: an update on recent findings regarding responses in

salmonids. Aquac. Res. 41, 333-344.

178
Krogdahl, ., Sundby, A., 1999. Characteristics of pancreatic function in fish. In: Pierzyinowski,

S.G., Zabielski, R. (Eds) Biology of The Pancreas in Growing Animals, Elsevier Science,

Amsterdam, pp. 437-458.

Krogdahl, ., Sundby, A., Olli, J.J., 2004. Atlantic salmon (Salmo salar) and rainbow trout

(Oncorhynchus mykiss) digest and metabolize nutrients differently. Effects of water

salinity and dietary starch level. Aquaculture 229, 335-360.

Kuzmina, V.V., Gavroskaya, L.K., Ryzhova, O.V., 2010. Taurine. Effect on exotrophica and

metabolissm in mammals and fish. J. Evol. Biochem. Physiol. 46, 19-27.

Laining, A., Traifalgar, R.F., Thu, M., Komilus, C.F., Kader, A., Koshio, S., Ishikawa, M.,

Yokoyama, S., 2010. Influence of dietary phytic acid on growth, feed intake, and nutrient

utilization in juvenile Japanese Flounder, Paralichthys olivaceus. J World Aquac. Soc.

41, 746-754.

Laplace, J.P., Darcy-Vrillon, B., Perez, J.M., Henry, Y., Giger, S., Sauvant, D., 1989.

Associative effects between two fibre sources on ileal and overall digestibilities of amino

acids, energy and cell-wall components in growing pigs. Br. J. Nutr. 61, 75-78.

Lauff, M., Hofer, R., 1985. Proteolytic enzymes in fish development and the importance of

dietary enzymes. Aquaculture 37, 335-346.

Leenhouwers, J., Ortega, R., Verreth, J., Schrama, J., 2007. Digesta characteristics in relation to

nutrient digestibility and mineral absorption in Nile tilapia (Oreochromis niloticus L.) fed

cereal grains of increasing viscosity. Aquaculture 273, 556-565.

Leenhouwers, J.I., Pellikaan, W.F., Huizing, H.F.A., Coolen, R.O.M., Verreth, J.A.J., Schrama,

J.W., 2008. Fermentability of carbohydrates in an in vitro batch culture method using

179
inocula from Nile tilapia (Oreochromis niloticus) and European sea bass (Dicentrarchus

labrax). Aquac. Nutr. 14, 523-532.

Lehrfeld, J., 1989. High-performance liquid chromatography analysis of phytic acid on a pH-

stable, macroporous polymer column. Cereal Chem. 66, 510-515.

Lemme, A., 2010. Availability and effectiveness of free amino acids in aquaculture. In: Cruz-

Suarez, L.E., Ricque-Marie, D., Tapia-Salazar, M., Nieto-Lpez, M.G., Villarreal-

Cavazos, D. A., Gamboa-Delgado, J. (Eds), Avances en Nutricin Acucola X -

Memorias del Dcimo Simposio Internacional de Nutricin Acucola, 8-10 de

Noviembre, San Nicols de los Garza, N. L., Mxico, pp. 264-275.

Levesque, C.L., Moehn, S., Pencharz, P.B., Ball, R.O., 2010. Review of advances in metabolic

bioavailability of amino acids. L. Sci. 133, 4-9.

Lewis, A.J., Bayley, H.S., 1995. AA bioavailability. In: Ammerman, C.B., Baker, D.H., Lewis,

A.J., (Eds), Bioavailability of Nutrients for Animals, AA, Minerals, and Vitamins.

Academic Press, New York, NY, pp. 35-65.

Lewis, N.G., Davin, L.R., Sarkanen, S., 1998. Lignin and lignin biosynthesis: distinctions and

reconciliations. In: Lewis N.G., Sarkanen, S., (Eds) Lignin and Lignan Biosynthesis,

ACS Symposium Series 697, American Chemical Society, Washington, DC, pp1-28.

Lewis, N.G., Yamamoto, E., 1989. Tannins-their place in plant metabolism. In: Hemingway

R.W., Karchesy, J.J., (Eds) Chemistry and Significance of Condensed Tannins, Plenum

Press, New York and London, pp. 23-46.

Levene, H., 1960. Contributions to Probability and Statistics: Essays in Honor of Harold

Hotelling, Olkin I. (Ed.), Stanford University Press, pp. 278-292.

180
Li, P., Burr, G., Wen, Q., Goff, J., Murthy, H., Gatlin III, D., 2009. Dietary sufficiency of sulfur

amino acid compounds influences plasma ascorbic acid concentrations and liver

peroxidation of juvenile hybrid striped bass (Morone chrysops M. saxatilis).

Aquaculture 287, 414-418.

Li, P., Mai, K., Trushenski, J., Wu, G., 2009. New developments in fish amino acid nutrition:

towards functional and environmentally oriented aquafeeds. Amino Acids 37, 43-53.

Li, P., Wang, X., Gatlin III, D., 2008. RRR--Tocopheryl succinate is a less bioavailable source

of vitamin E than all-rac--tocopheryl acetate for red drum, Sciaenops ocellatus.

Aquaculture 280, 165-169.

Lied, E., Julshamn, K., Brackkan, O.R., 1982. Determination of protein digestibility in Atlantic

cod (Gadus morhua) with internal and external indicators. Can. J Fish. Aquat. Sci. 39,

854-861.

Lim, C., Dominy, W., 1990. Evaluation of soybean meal as a replacement for marine animal

protein in diets for shrimp (Penaeus vannamei). Aquaculture 87, 53-63.

Lin, X., Volkoff, H., Narnaware, Y., Bernier, N.J., Peyon, P., Peter, R.E., 2000. Brain regulation

of feeding behaviour and food intake in fish. Comp. Biochem. Physiol. 126, 415-434.

Littel, R.C., Henry, P.R., Lewis, A.J., Ammerman, C.B., 1997. Estimation of relative

bioavailability of nutrients using SAS procedures. Anim. Sci. 75, 2672-2683.

Liu, Z., Bateman, A., Bryant, M., Abebe, A., Roland, D., 2004. Estimation of DL-methionine

hydroxyl analogue relative to DL-methionine in layers with exponential and slope-ratio

models. Poult. Sci. 83, 1580-1586.

Lopez, H.W., Leenhardt, F., Coudray, C., Remesy, C., Minerals and phytic acid interactions: is it

a real problem for human nutrition? Int. J. Food Sci. Technol. 37, 727739.

181
Lora, J.H., Creamer, A.W., Wu, L.C.F., Goyal, G.C., 1993. Industrial scale production of

organosolv lignins: characteristics and applications. In: Kennedy, J.F., Phillips, G.O.,

Williams, P.A., (Eds), Cellulosics: Chemical, Biochemical and Material Aspects. Ellis

Horwoods Ltd., Sussex, England, pp. 252256.

Lumpkins, B., Batal, A., 2005. Bioavailability of lysine and phosphorus in distillers dried grains

with solubles. Poult. Sci. 84, 581-586.

Lumsden, J.S., Wybourne, B., Minamikawa, M., Tubbs, L., 2010. Gastric dilation and air

sacculitis in Chinook salmon, Oncorhynchus tshawytscha (Walbaum); correlation of

macroscopic and microscopic lesions, and relationship of the syndrome to

glomerulonephritis and serum biochemistry. Fish Dis. 33, 737-747.

Lupatsch, I., Kissil, G.W., Sklan, D., Pfeffer, D., 2001. Effects of varying dietary protein and

energy supply on growth, body composition and protein utilization in gilthead seabream

(Sparus aurata L.). Aquac. Nutr. 7, 71-80.

Lusas, E.W., Riaz, M.N., 1995. Soy protein products: processing and use. J. Nutr. 125, 973S-

580S.

Maddaiah, V.T., Kurnick, A.A., Reid, B.L., 1964. Phytic acid studies. Proc. Soc. Exp. Biol. Med.

115, 391-393.

Maenz, D.D., 2001. Enzymatic characteristics of phytases as they relate to their use in animal

feeds. In: Bedford, M.R., Partridge, G.G. (Eds), Enzymes in Farm Animal Nutrition,

CABI Publishing, NY 6184pp.

Maenz, D.D., Engele-Schan, C.M., Newkirk, R.W., Classen, H.L., 1999. The effect of minerals

and mineral chelators on the formation of phytase-resistant and phytase-susceptible forms

of phytic acid in solution of canola meal. Anim. Feed Sci. Technol. 81, 177192.

182
Major, E.J., Batterham, E.S., 1981. Availability of lysine in protein concentrates as determined

by the slope-ratio assay with chicks and comparisons with rat, pig, and chemical assays.

Br. J. Nutr. 46, 513519.

Mambrini, M., Roem, A.J., Cravdi, J.P., Lalls, J.P., Kaushik, S.J., 1999. Effects of replacing

Fish Meal with Soy Protein Concentrate and of DL-Methionine Supplementation in

High-Energy, Extruded Diets on the Growth and Nutrient Utilization of Rainbow Trout,

Oncorhynchus mykiss. Anim. Sci. 77, 2990-2999.

Marcone, M.F., 1999. Biochemical and biophysical properties of plant storage proteins: a current

understanding with emphasis on 11S seed globulins. Food Res. Intl. 32, 79-92.

Martinez, G.M., Knabe, D.A., 1990. Digestible lysine requirement of starter and grower pigs.

Anim. Sci. 68, 2748-2755.

Mateo, R.D., Spallholz, J.E., Elder, E., Yoon, I., Kim, S.W., 2007. Efficacy of dietary selenium

sources on growth and carcass characteristics of growing-finishing pigs fed diets

containing high endogenous selenium. Anim. Sci. 85, 1177-1183.

Meng, X., Slominski, B.A., Campbell, L.D., Guenter, W., Jones, O., 2006. The use of enzyme

technology for improved energy utilization from full-fat oilseeds. Part I: canola seed.

Poult. Sci. 85, 1025-1030.

Mitaru, B.N., Blair, R., Reichert, R.D., Roe, W.E., 1984. Dark and yellow rapeseed hulls,

soybean hulls and a purified fiber source: Their effects on dry matter, energy, protein and

amino acid digestibilities in cannulated pigs. Anim. Sci. 59, 1510-1518.

Morales, G.A., Moyano, F.J., 2010. Application of an in vitro gastrointestinal model to evaluate

nitrogen and phosphorus bioaccessibility and bioavailability in fish feed ingredients.

Aquaculture 306, 244-251.

183
Morrison, F.B., 1936. Feeds and Feeding. A Handbook for The Student and Stockman.

Twentieth Edition, The Morrison Publishing Company, Ithaca, Newyork.

Moughan, P.J., Rutherford, S.M., 1996. A new method for determining digestible reactive lysine

in foods. Agri. Food Chem. 44, 2202-2209.

Moughan, P.J., Rutherford, S.M., 2008. Available lysine in foods: a brief historical overview.

AOAC Intl. 91(4), 901-906.

Moure, A., Sineiro, J., Domnguez, H., Paraj, J.C., 2006. Functionality of oilseed protein

products: A review. Food Res. Intl. 39, 945-963.

Murai, T., Ogata, H., Kosutarak, P., Arai, S., 1989. Effects of amino acid supplementation and

methanol treatment on utilization of soy flour in fingerling carp. Aquaculture 56, 197-

206.

Mwachireya, S.A., Beames, R.M., Higgs, D.A., Dosanjh, B.S., 1999. Digestibility of canola

protein products derived from the physical, enzymatic and chemical processing of

commercial canola meal in rainbow trout Oncorhynchus mykiss (Walbaum) held in fresh

water. Aquac. Nutr. 5, 7382.

Nangthu, T.T., Bodin, N., De Saeger, S., Larondelle, Y., Rollin, X., 2010. Substitution of fish

meal by sesame oil cake (Sesamum indicum L.) in the diet of rainbow trout

(Oncorhynchus mykiss W.). Aquac. Nutr. 17, 80-89.

Nang Thu, T.T.N., Parkouda, C., de Saeger, S., Larondelle, Y., Rollin, X., 2007. Comparison of

the lysine utilization efficiency in different plant protein sources supplemented with L-

lysineHCl inn rainbow trout (Oncorhynchus mykiss) fry. Aquaculture 272, 477-488.

NRC, 2011. Nutrient Requirements of Fish and Shrimp, Animal Nutrition Series, National

Research Council, National Academies Press, Washington, DC.

184
Naylor, R.L., Hardy, R.W., Bureau, D.P., Chiu, A., Elliott, M., Farell, A.P., Forster, I., Gatlin,

D.M., Goldburg, R.J., Hua, K., Nichols, P.D., 2009. Feeding aquaculture in an era of

finite source. PNAS 106, 15103-15110.

Neves, V.A., Lourenco, E.J., 1995. Isolation and in vitro hydrolysis of globulin G1 from lentils

(Lens culinaris Medik). J. Food Biochem. 19, 109-120.

Newkirk, R.W., Classen, H.L., 2001. The non-mineral nutritional impact of phytate in canola

meal fed to broiler chicks. Anim. Feed Sci. Tech. 91, 115-128.

Newkirk, R.W., Classen, H.L., Tyler, R.T., 1997. Nutritional evaluation of low glucosinolate

mustard meals (Brassica juncea) in broiler diets. Poult. Sci. 76, 1272-1277.

Nieminen, P., Mustonen, A.M., Hyvarian, H., 2003. Fasting reduces plasma leptin- and ghrelin-

immunoreactive peptide concentrations of the burbot (Lota lota) at 20C but not at 100C.

Zool. Sci. 20, 1109-1115.

Nose, T. 1967. On the metabolic faecal nitrogen in young rainbow trout. Bull. Japan Soc. Fish.

Sci. 17, 97-105.

Nwanna, L.C., 2004. Growth and mineral deposition in Nile tilapia (Oreochromis niloticus) fed

untreated soybean meal supplemented with phytase. J Food Agri. Env. 2, 51-56.

OBrien, D.M., 2010. US distillers grains supply-use under E-10, E12 and E-15. AgMRC

Renewable Energy & Climate Change Newsletter, July: 6-11.

ODell, B.L., de Boland, A., 1976. Complexation of phytate with proteins and cations in corn

germ and oilseed meals. J. Agric. Food Chem. 24, 804808.

Obaldo, L.G., Divakaran, S., Tacon, A.G., 2002. Method for determining the physical stability of

shrimp feeds in water. Aquac. Res. 33, 369-377.

185
Ogino, C., Kakino J., Chen, M.S., 1973 . Determination of metabolic fecal nitrogen and

endogenous nitrogen excretion of carp. Bull. Jpn. Soc. Sci. Fish. 39, 519523.

Okubu, K., Myers, D.V., Iacobucci, G.A., 1976. Binding of phytic acid to glycinin. Cereal

Chem. 53, 513.

Olsen, R.E., Suontama, J., Langmyhr, E., Mundheim, H., Ringo, E., Melle, W., Malde, M.K.,

Hemre, G.-I., 2006. The replacement of fish meal with Antarctic krill, Euphausia superba

in diets for Atlantic salmon, Salmo salar. Aquac. Nutr. 12, 280-290.

Olsson, C., Aldman, G., Larsson, A., Holmgren, S., 1999. Cholecystokinin affects gastric

emptying and stomach motility in the rainbow trout Oncorhynchus mykiss. J. Exp. Biol.

202, 161-170.

Oram, R.N., Salisbury, P.A., Kirk, J.T.O., Burton, W.A., 1999. Brassica juncea breeding. In:

Salisbury, P.A., Potter, T.D., McDonald G., Green, A.G. (Eds), Canola in Australia: The

First Thirty Years. Proceedings of the 10th International Rapeseed Congress, Canberra,

Australia, pp. 3740.

vrum-Hansen, J., Storebakken, T., 2007. Effects of cellulose inclusions on pellet quality and

digestibility of main nutrients and minerals in rainbow trout (Oncorhynchus mykiss).

Aquaculture 272, 458-465.

Page, J.W., Andrews, J.W., 1973: Interactions of dietary levels of protein and energy on channel

cat fish (Ictalurus punctatus). J. Nutr. 103, 13391346.

Pallauf, J., Rimbach, G., 1997 Nutritional significance of phytic acid and phytase. Arch. Anim.

Nutr. 50, 301-319.

Panyam, D., Kilara, A., 1996. Enhancing the functionality of food proteins by enzymatic

modifcation. Trends Food Sci. Tech. 7, 120-125.

186
Parsons, C.M., Zhang, Y., Araba, M., 1998. Availability of amino acids in high-oil corn. Poult.

Sci. 77, 1016-1019.

Paulsen, B.S., Barsett, H., 2005. Bioactive pectic polysaccharides. Adv. Poly. Sci. 186, 69-101.

Pelissero, C., Le Menn, F., Kaushik, S.J., 1991. Estrogenic effect of dietary soya bean meal on

vitellogenesis in cultured Siberian sturgeon Acipenser baeri. Gen. Comp. Endocrinol. 83,

447457.

Peres, H., Oliva-Teles, A., 2005. The effect of dietary protein replacement by crystalline amino

acid on growth and nitrogen utilization of turbot Scophthalmus maximus juveniles.

Aquaculture 250, 755-764.

Persson, H., Trk, M., Nyman, M. Sandberg, A.-S., 1998. Binding of Cu2+, Zn2+, and Cd2+ to

inositol tri-, tetra-, penta-, and hexaphosphates. J. Agric. Food Chem.,46, 31943200.

Pessala, P., Schultz, E., Kukkola, J., Nakari, T., Knuutinen, J., Herve, S., Paasivirta, J., 2010.

Biological effects of high molecular weight lignin derivatives. Ecotox. Env. Safety 73,

1641-1645.

Petersen, G.I., Pedersen, C., Lindemann, M.D., Stein, H.H., 2011. Relative bioavailability of

phosphorus in inorganic phosphorus sources fed to growing pigs. Anim. Sci. 89, 460-466.

Phillips, A.M., Tunison, A.V., Fenn, A.H., Mitchell, C.R., McCay, C.M., 1940. The nutrition of

trout. Fish. Res. Bull. 9, 2-32.

Platt, S.R., Clydesdale, F.M., 1987. Mineral binding characteristics of lignin, guar gum,

cellulose, pectin and neutral detergent fiber under simulated duodenal pH conditions.

Food Sci. 52, 1414-1419.

187
Poppi, D., Quinton, V.M., Hua, K., Bureau, D.P., 2011. Development of a test diet for assessing

the bioavailability of arginine in feather meal fed to rainbow trout (Oncorhynchus

mykiss). Aquaculture 314, 100-109.

Prakash, V., Rao, N.M.S., 1986. Physical chemical properties of oilseed proteins. CRC Crit. Rev.

Biochem. 20, 265-363.

Pratoomyot, J., Bendiksen, E.A., Bell, J.G., Tocher, D.R., 2010. Effects of increasing

replacement of dietary fishmeal with plant protein sources on growth performance and

body lipid composition of Atlantic salmon (Salmo salar L.). Aquaculture 305, 124-132.

Rana, K., 2009. Feed Management: Global Challenges. Intl. AquaFeed 2009(4), 40-44.

Rana, K.J., Siriwardena, S., Hasan, M.R., 2009. Impact of rising feed ingredient prices on

aquafeeds and aquaculture production. FAO Fish. Aquac. Tech. Paper No. 541.

Rtz, H-J., Llroet, J. 2003. Variation in fish condition between Atlantic cod (Gadus morhua)

stocks, the effect on their productivity and management implications. Fish. Res. 60, 369-

380.

Ravindran, V., Bryden, W.L., Kornegay, E.T., 1995. Phytates: Occurrence, bioavailability and

implications in poultry nutrition. Poult. Avian Biol. Rev. 6, 125-143.

Ravindran, V., Cabahug, S., Ravindran, G., Bryden, W.L., 1999. Influence of Microbial Phytase

on Apparent Ileal Amino Acid Digestibility of Feedstuffs for Broilers. Poult. Sci. 78,

699-706.

Reddy, N.R., 2001. Occurrence, distribution, content, and dietary intake of phytate. In: Reddy,

N.R., Sathe, S.K., (Eds.) Food Phytates, CRC Press, Florida, pp. 25-51.

188
Refstie, S., sgrd, T., 2009. Advances in aquaculture feeds and feeding: salmonids. In: Burnell,

G., Allan, G., (Eds) New Technologies in Aquaculture, CRC Press, New York, pp. 498-

541.

Refstie, S., Korsen, .J., Storebakken, T., Baeverfjord, G., Lein, I., Roem, A.J., 2000. Differing

nutritional responses to dietary soybean meal in rainbow trout (Oncorhynchus mykiss)

and Atlantic salmon (Salmo salar). Aquaculture 190, 49-63.

Refstie, S., Landsverk, T., Bakke-McKellep, A.M., Ring, E., Sundby, A., Shearer, K.D.,

Krogdahl, ., 2006. Digestive capacity, intestinal morphology, and microflora of 1-year

and 2-year old Atlantic cod (Gadus morhua) fed standard or bioprocessed soybean meal.

Aquaculture 261, 269-284.

Refstie, S., Storebakken, T., Baeverfjord, G., Roem, A.J., 2001. Long-term protein and lipid

growth of Atlantic salmon (Salmo salar) fed diets with partial replacement of fish meal

by soy protein products at mediumor high lipid level. Aquaculture 193, 91-106.

Refstie, S., Svihus, B., Shearer, K., Storebakken, T., 1999. Nutrient digestibility in Atlantic

salmon and broiler chickens related to viscosity and non-starch polysaccharide content in

different soya bean products. Anim. Feed Sci. Tech. 79, 331345.

Riche, M., Trottier, N.L., Ku, P.K., Garling, D.L., 2001. Apparent digestibility of crude protein

and apparent availability of individual amino acids in tilapia (Oreochromis niloticus) fed

phytase pretreated soybean meal diets. Fish Physiol. Biochem. 25, 181-194.

Ring, E., 1993. The effect of chromic oxide (Cr2O3) on faecal lipid and intestinal microflora of

sea-water reared Arctic charr, Salvelinus alpinus (L.). Aquac. Fish. Man. 24, 341-344.

Roald, S.O., 1977. Acute toxicity of lignosulphonates on rainbow trout (Salmo gairdneri). Bull.

Env. Con. Toxicol. 17, 702-706.

189
Rodehutscord, M., Borchert, F., Gregus, Z., Pack, M., Pfeffer, E., 2000. Availability and

utilisation of free lysine in rainbow trout (Oncorhynchus mykiss). 1, Effect of dietary

crude protein level. Aquaculture 187, 163176.

Rodehutscord, M., Jacobs, S., Pack, M., Pfeffer, E., 1995. Response of rainbow trout

(Oncorhynchus mykiss) growing from 50 to 170 g to supplements of either L-arginine or

L-threonine in a semipurified diet. Nutrition 125, 970-975.

Rodrigues, I.M., Coelho, J.F.J., Carvalho, M.G.V.S., 2011. Isolation and valorization of

vegetable proteins from oilseed plants: methods, limitations and potential. J. Food Eng.,

doi:10.1016/j.jfoodeng.2011.10.027.

Roller, S., Seedhar, P., 2002. Carvacrol and cinnamic acid inhibit microbial growth in fresh-cut

melon and kiwifruits at 4 C and 8 C. Lett. Appl. Microbiol. 35, 390394.

Rosas, A., Vazquez-Duhalt, R., Tinoco, R., Shimada, A., Dabramo, L.R., Viana, M.T., 2008.

Comparative intestinal absorption of amino acids in rainbow trout (Oncorhynchus

mykiss), totoaba (Totoaba macdonaldi) and Pacific bluefin tuna (Thunnus orientalis).

Aquac. Nutr. 14, 481-489.

Rumsey, G.L., Siwicki, A.K., Anderson, D.P., Bowser, P.R., 1994. Effect of soybean protein on

serological response, non-specific defense mechanisms, growth, and protein utilization in

rainbow trout. Vet. Immunol. Immunopathol. 41, 323-339.

Rungruangsak-Torrissen, K., Rustad, A., Sunde, J., Eiane, S.A., Jensen, H.B., Opstvedt, J.,

Nygard, E., Samuelsen, T.A., Mundheim, H., Luzzana, U., Venturini, G., 2002. In vitro

digestibility based on fish crude enzyme extract for prediction of feed quality in growth

trials. J Sci. Food Agri. 82, 644-654.

190
Sajjadi, M., Carter, C., 2004a. Effect of phytic acid and phytase on feed intake, growth,

digestibility and trypsin activity in Atlantic salmon (Salmo salar, L.). Aquac. Nutr. 10,

135-142.

Sajjadi, M., Carter, C., 2004b. Dietary phytase supplementation and the utilisation of phosphorus

by Atlantic salmon (Salmo salar L.) fed a canola-meal-based diet. Aquaculture 240, 417-

431.

Sanden, M., Berntssen, M.H.G., Krogdahl, A., Hemre, G.-I., Bakke-McKellep, A.M., 2005. An

examination of the intestinal tract of Atlantic salmon (Salmo salar L.) parr fed different

varieties of soy and maize. Fish Dis. 28, 317330.

Santigosa, E., Rodrigez, M..S., Rodiles, A., Barroso, F.G., Alarcn, F.J., 2010. Effect of

diets containing a purified soybean trypsin inhibitor on growth performance, digestive

proteases and intestinal histology in juvenile sea bream (Sparus aurata L.). Aquac. Res.

41, 187-198.

Sapkota, A.R., Lefferts, L.Y., McKenzie, S., Walker, P., 2007. What do we feed to food-

production animals? A review of animal feed ingredients and their potential impacts on

human health. Env. Health Pers. 115, 663-670.

Sathe, S.K., Hamaker, B.R., Sze-Tao, K.W.C., Venkatachalam, M., 2002. Isolation, purification,

and biochemical characterization of a novel water soluble protein from Inca peanut

(Plukenetia volubilis L.). J.Agric. Food Chem. 50, 4096-4908.

Sathivel, S., Smily, S., Witoon, P., Bechtel, P.J., 2005. Functional and nutritional properties of

red salmon (Oncorhynchus nerka) enzymatic hydrolysates. Food Chem. Toxicol. 70, 401-

406.

191
Schedle, K., Plitzner, C., Ettle, T., Zhao, L., Domig, K., Windisch, W., 2008. Effects of insoluble

dietary fibre differing in lignin on performance, gut microbiology, and digestibility in

weanling piglets. Arch. Anim. Nutr. 62, 141-151.

Selle P.H., Ravindran V., Bryden W.L., Scott T., 2006. Influence of dietary phytate and

exogenous phytase on amino acid digestibility in poultry: a review. Poult. Sci. 43, 89

103.

Selle, P.H., Ravindran, V., Caldwell, R.A., Bryden, W.L., 2000. Phytate and phytase:

consequences for protein utilization. Nutr. Res. Rev. 13, 255-278.

Selle, P.H., Walker, A.R., Bryden, W.L., 2003. Total and phytate-phosphorus contents and

phytase activity of Australian-sources feed ingredients for pigs and poultry. Aus J. Exp.

Agri. 43, 475-479.

Shafaeipour, A., Yavari, V., Falahatkar, B., Maremmazi, J. Gh., Gorjipour, E., 2008. Effects of

canola meal on physiological and biochemical parameters in rainbow trout

(Oncorhynchus mykiss). Aquac. Nutr. 14, 110-119.

Shahidi, F., Wanasundra, P.K.J.P.D., 1998. Effect of acylationon flax protein functionility. In:

Whitaker, J.R., Shahidi, F., Lpez-Mungua, A., Yada, R.Y., Fuller, G., (Eds) Functional

Properties of Proteins and Lipids, ACS Symposium Series 708, Washington DC, pp. 96-

120.

Shih, F.F., 1987. Deamidation of protein in a soy extract by ion exchange resins catalysis. Food

Sci. 52, 1529-1531.

Shoveller, A.K., Moehn, S., Rademacher, M., Htoo, J.K., Ball, R.O., 2009. Methionine-hydroxy

analogue was found to be significantly less bioavailable compared to dl-methionine for

protein deposition in growing pigs. Animal 4, 61-66.

192
Silverstein, J.T., Breninger, J., Baskin, D.G., Plisetskaya, E.M., 1998. Neuropeptide-Y like gene

expression in the salmon brain increases with fasting. Gen. Comp. Endocrinol. 110, 157-

165.

Silverstein, J.T., Plisetskaya, E.M., 2000. The effects of NPY and insulin on food intake

regulation in fish. Am Zool. 40, 296-308.

Singh, M., Krakorian, A.D., 1982. Inhibition of trypsin activity in vitro by phytase. Agri. Food

Chem. 30, 799-800.

Sinha, A.K., Kumar, V., Makkar, H.P.S., De Boeck, G., Becker, K., 2011. Non-starch

polysaccharides and their role in fish nutrition A review. Food Chem. 127, 1409-1426.

Sitohy, M., Popimeau, Y., chobert, J.M., Haertl, T., 1999. Mild method of simultaneous

methionine grafting and phosphorylation of soybean globulins improves their functional

properties. Nahrung 43, 3-8.

Slominski, B.A., Cambell, L.D., 1990. Non-starch polysaccharidesof canola meal: quantification,

digestibility in poultry and potential benefit of dietary enzyme supplementation. J Sci.

Food Agri. 53, 175-184.

Sorensen, M., Stjepanovic, N., Romarheim, O., Krekling, T., Storebakken, T., 2009. Soybean

meal improves the physical quality of extruded fish feed. Anim. Feed Sci. Tech. 149,

149-161.

Sorensen, M., Storebakken, T, Shearer, K.D., 2005. Digestibility, growth and nutrient retention

in rainbow trout (Oncorhynchus mykiss) fed diets extruded at two different temperatures.

Aquac. Nutr. 11, 251-256.

Souffrant, W.B., 2001. Effect of dietary fibre on ileal digestibility and endogenous nitrogen

losses in the pig. Anim Feed Sci. Tech. 90, 93-102.

193
Spinelli, J., Houle, C., Wekell, J.C., 1983. The effect of phytates on the growth of rainbow trout

(Salmo gairdneri) fed purified diets containing varying quantities of calcium and

magnesium. Aquaculture 30, 71-83.

Stein, H. H., Seve, B., Fuller, M.F., Moughan, P.J., de Lange, C.F.M., 2007. Invited review:

Amino acid bioavailability and digestibility in pig feed ingredients: Terminology and

application. Anim. Sci. 85, 172180.

Stewart, O.J., Raghavan, G.S.V., Orsat, V., Golden, K.D., 2003. The effect of drying on

unsaturated fatty acids and trypsin inhibitor activity in soybean. Process Biochem. 39,

483-489.

Stone, D.A.J., 2003. Dietary carbohydrate utilization by fish. Rev. Fish. Sci. 11, 337-369.

Storebakken, T., 1985. Binders in fish feeds I. Effect of alginate and guar gum on growth,

digestibility, feed intake and passage through the gastrointestinal tract of rainbow trout.

Aquaculture 47, 11-26.

Storebakken, T., Kvien, I.S., Shearer, K.D., Grisdale-Helland, B., Helland, S.J., Berge, G.M.,

1998a. The apparent digestibility of diets containing fish meal, soybean meal or bacterial

meal fed to Atlantic salmon (Salmo salar): evaluation of different faecal collection

methods. Aquaculture 169, 195-210.

Storebakken, T., Shearer, K.D., Roem, A.J., 1998b. Availability of protein, phosphorus and other

elements in fish meal, soy-protein concentrate and phytase-treated soy-protein-

concentrate-based diets to Atlantic salmon, Salmo salar. Aquaculture 161, 365-379.

Storebakken, T., Shearer, K.D., Roem, A.J., 2000. Growth, uptake and retention of nitrogen and

phosphorus, and absorption of other minerals in Atlantic salmon Salmo salar fed diets

194
with fish meal and soyprotein concentrate as the main sources of protein. Aquac. Nutr.

6, 103-108.

Sugiura, S.H., Gabaudan, J., Dong, F.M., Hardy, R.W., 2001. Dietary microbial phytase

supplementation and the utilization of phosphorus, trace minerals and protein by rainbow

trout (Oncorhynchus mykiss Walbaum) fed soybean meal-based diets. Aquac Res. 32,

583-592.

Sugiura, S.H., Roy, P.K., Ferraris, R.P., 2006. Dietary acidification enhances phosphorus

digestibility but decreases H+/K+- ATPase expression in rainbow trout. Exp. Biol. 209,

3719-3728.

Swisher, K., 2007. World Production and Trade of Animal Proteins. URL:

http://www.fprf.org/UserFiles/File/Chile%20%20NRA%20Latin%20America/8%20Kent

%20Swisher.pdf. Accessed on October 17, 2010.

Tacon, A.G.J, Hasan, M.R., Subasinghe, R.P., 2006 Use of fisheries resources as feed inputs to

aquaculture development: Trends and policy implications (Food and Agriculture

Organization, Rome) FAO Fisheries Circular 1018.

Tacon, A.G.J., 1997. Fish meal replacers: review of anti-nutrients within oil seeds and pulses a

limiting factor for the aquafeed green revolution? In: Tacon A., Basurca B. (Eds.)

Feeding Tomorrows fish. Cahiers options Mediterraneans 22, 154-182.

Tacon, A.G.J., 2010. Growth of global aquaculture and the feed ingredients available. In:

Seminar on Can A Growing Aquaculture Industry Continue to Use Fishmeal and Fish Oil

in Feeds and Remain Sustainable? EP INTERGROUP on Climate Change, Biodiversity

and Sustainable Development, March 3, 2010.

195
http://www.ebcd.org/website%2010/Fishmeal/PPT%20fishmeal/Dr%20ALbert%20Taco

n.pdf. Accessed on October 15, 2010.

Tacon, A.G.J., Metian, M., 2008. Global overview on the use of fish meal and fish oil in

industrially compounded aquafeeds: trends and future prospects. Aquaculture 285, 146-

158.

Tacon, A.G.J., Metian, M., 2009. Fishing for feed or fishing for food: increasing global

competition for small pelagic forage fish. Ambio 38, 294-302.

Tacon, A.G.J., Rodriguez, A.M.P. 1984. Comparison of chromic oxide, crude fibre, polyethylene

and acid-insoluble ash as dietary markers for the estimation of apparent digestibility

coefficients in rainbow trout. Aquaculture, 43:391-399.

Teskerdi, Z., Higgs, D.A., Dosanjh, B.S., McBride, J.R., Hardy, R.W., Beames, R.M., Jones,

J.D., Simell, M., Vaara, T., Bridges, R.B., 1995. Assessment of undephytinized and

dephytinized rapeseed protein concentrate as sources of dietary protein for juvenile

rainbow trout (Oncorhynchus mykiss). Aquaculture 131, 261-277.

Thacker, P.A., Petri, D., 2010. The effects of mustard protein concentrate on nutrient

digestibility and performance of broiler chickens. J. Anim. Sci. Biotech. 1, 85-92.

Thebaudin, J.Y., Lefebvre, A.C., Harrington, M., Bourgeois, C.M., 1997. Dietary fibres:

nutritional and technological interest. Trends Food Sci. Tech. 8, 41-47

Thiessen, D.L., Campbell, G.L., Adelizi, P.D., 2003. Digestibility and growth performance of

juvenile rainbow trout (Oncorhynchus mykiss) fed with pea and canola products. Aquac.

Nutr. 9, 65-75.

196
Thiessen, D.L., Maenz, D.D., Newkirk, R.W., Classen, H.L., Drew, M.D., 2004. Replacement of

fishmeal by canola protein concentrate in diets fed to rainbow trout (Oncorhynchus

mykiss). Aquac. Nutr. 10, 379-388.

Thomas, M., 1996. Physical quality of pelleted animal feed 1. Criteria for pellet quality. Anim.

Feed Sci. Tech. 61, 89-112.

Torre, M., Rodriguez, A.R., Saura-Calixto, F., 1991. Effects of dietary fiber and phytic acid on

mineral availability. Cr. Rev. Food Sci. nutr. 30, 1-22

Turchini, G.M., Torstensen, B.M., Ng, W-K. 2009. Fish oil replacement in finfish nutrition. Rev.

Aquacult. 1, 10-57.

Ultee, A., Slump, R.A., Steging G., Smid, E.J., 2000. Antimicrobial activity of carvacrol toward

Bacillus cereus on rice. Food Prot. 63, 620624.

USDA., 2011. World Agricultural Supply and Demand Estimates, WASDE- 492 March 10,

2011. United States Department of Agriculture, World Agriculture Outlook Board, WA,

D.C. 20250-3812, 39 pp.

Usmani, N., Zafri, A.K., 2002. Influence of dietary phytic acid on the growth, conversion

efficiency, and carcass composition of Mrigal Cirrhinus mrigala (Hamilton) Fry. J.

World Aquac. Soc. 33, 199-204.

Uzzan, A., 1988. Vegetable protein products from seeds: technology and uses in the food

industry. In: Hudson, B,J,F.. (Ed) Developments in Food Proteins, New York, Elsevier

Applied Science 6, 73-118.

Van Soest, P.J., Robertson, J.B., Lewis, B.A., 1991. Methods for dietary fiber, neutral detergent

fiber, and nonstarch polysaccharides in relation to animal nutrition. Dairy Sci. 74, 3583

3597.

197
Van Soest, P.J., Robertson, P.J., 1981. The detergent system of analysis and its application to

human foods. In: James, W.P.T., Theander, O. (Eds), The Analysis of Dietary Fibre in

Food. Marcel Dekker Inc, New York, p. 123.

Vandenberg, G.W., de la Noe, J., 2001. Apparent digestibility comparison in rainbow trout

(Oncorhynchus mykiss) assessed using three methods of faeces collection and three

digestibility markers. Aquac. Nutr. 7, 237-245.

Vielma, J., Ruohonen, K., Gabaudan, J., Vogel, K., 2004. Top-spraying soybean meal-based

diets with phytase improves protein and mineral digestibilities but not lysine utilization in

rainbow trout, Oncorhynchus mykiss (Walbaum). Aquac. Res. 35, 955-964.

Villamide, M.J., San Juan, L.D., 1998. Effect of chemical composition of sunflower seed meal

on its true metabolizable energy and amino acid digestibility. Poult. Sci. 77, 1884-1892.

Vohra, P., Gray, G.A., Kratzer, F.H., 1965. Phytic acid-metal complexes. Proc. Soc. Exp. Biol.

Med. 120, 447-449.

Volkoff, H., Canosa, L.F., Unniappan, S., Cerd-Reverter, J.M., Bernier, N.J., Kelly, S.P. and

Peter, R.E., 2005. Neuropeptides and the control of food intake in fish. Gen. Comp.

Endocrinol. 142, 3-19.

Volkoff, H., Peter, R.E., 2001a. Interactions between orexin A, NPY and galanin in the control

of feed intake in gold fish, Carassius auratus. Regul. Pept. 101, 59-72.

Volkoff, H., Peter, R.E., 2001b. Characterization of two forms of cocaine- and amphetamine-

regulated transcript (CART) peptide precursors in goldfish: molecular cloning and

distribution, modulation of expression by nutritional status, and interactions with leptin.

Endocrinology 142, 5076-5088.

198
Wang, F., Yang, Y-H., Han, Z-Z., Dong, H-W., Yang, C-H., Zou, Z-Y., 2009a. Effects of

phytase pretreatment of soybean meal and phytase-sprayed in diets on growth, apparent

digestibility coefficient and nutrient excretion of rainbow trout (Oncorhynchus mykiss

Walbaum). Aquac. Intl. 17, 143-157.

Wang, Y., Marx, T., Lora, J., Phillip, L.E., McAllister, T.A., 2009b. Effects of purified lignin on

in vitro ruminal fermentation and growth performance, carcass traits and fecal shedding

of Escherichia coli by feedlot lambs. Anim. Feed Sci. Tech. 151, 21-31.

Wee, K.L., Shu, S.W., 1989. The alternative values of boiled full fat soybean in pelleted feed for

Nile tilapia. Aquaculture 81, 303-314.

Wenk, C., 2001. The role of dietary fibre in the digestive physiology of the pig. Anim. Feed Sci.

Tech. 90, 21-33.

Wenk, C., Messikommer, R., Bee, G., 1998. Effect of lignin from different origins in

combination with carbohydrases in pig diets on digestion processes. Anim. Sci. Suppl.

76:179.

Whetten, R., Sederoff, R., 1995. Lignin biosynthesis. Plant Cell 7, 1001-1013.

Whitmore, F.W., 1982. 6 Lignin-protein complex in cell walls of Pinus elliottii: Amino acid

constituents. Phytochemistry 21, 315-318.

Wijesundera, C., Ceccato, C., Fagan, P., Shen, Z., Burton, W., Salisbury, P., 2008. Canola

quality Indian mustard oil (Brassica juncea) is more stable to oxidation than convention

canola oil (Brassica napus). J Am Oil Chem. Soc. 85, 693-699.

Wilson, J.M., Castro, L.F.C., 2011. Morphological diversity of the gastro-intestinal tract in

fishes. In: Grosell, M., Farrell, A.P., Brauner, C.J., (Eds) The Multifunctional Gut of

Fishes, Fish Physiology Series 30, Elsevier, USA, pp. 2-56

199
Wiltafsky, M.K., Bartelt, J., Relandeau, C., Roth, F.X., 2009. Estimation of the optimum ratio of

standardized ileal digestible isoleucine to lysine for eight- to twenty-five-kilogram pigs in

diets containing spray-dried blood cells or corn gluten feed as a protein source. Anim.

Sci. 87, 2554-2564.

Wise, A., 1983. Dietary factors determining the biological activities of phytate. Nutr. Abstr. Rev.

53, 791806.

Wise, A., 1995. Phytate and zinc bioavailability. Intl. J. Food Sci. Nutr. 46, 5363.

Wijeratne, W.B., Wang, T., Johnson, L.A., 2004. Extrusion based oilseed processing methds. In:

Dunford, N.T., Dunford, H.B., (Eds.) Nutritionally Enhanced Edible Oil Processing,

AOCS Publishing, Chapter 4, pp. 1-18.

Woerfel, J.B., 1995. Extraction. In; Erickson, D.R., (Ed.) Practical Handbook of Soybean

Processing and Utilization, AOCS Press, Champaign, Illinois, pp. 65-92.

Woods, D.L., Capcara, J.J., Downey, R.K., 1991. The potential of mustard (Brassica juncea (L.)

Coss) as an edible oil crop on the Canadian prairies. Can. J. Plant Sci. 71, 195198.

Wright, P.R., Morgan, J.M., Jessop, R.S., Cass, A. 1995. Comparative adaptation of canola

(Brassica napus) and Indian mustard (B. juncea) to soil water deficits: yield and yield

components. Field Corps Res. 42, 1-13.

WRO 2010. World Renderers Organization. Global rendered product production estimate. URL:

http://www.worldrenderers.org, accessed on October 17, 2010.

Yamammoto, T., Shima, T., Furuita, H., 2004. Antagonistic effects of branched-chain amino

acids induced by excess protein bound leucine in diets for rainbow trout (Oncorhynchus

mykiss). Aquaculture 232, 539-550.

200
Yamammoto, T., Shima, T., Furuita, H., Suzuki, N., Sanchez-Vazquez, F.J., Tabata, M., 2001.

Self-selection and feed consumption of diets with a complete amino acid composition and

a composition deficient in either methionine or lysine by rainbow trout, Oncorhynchus

mykiss (Walbaum). Aquac. Res. 32, 83-81.

Yamamoto, T., Sugita, T., Furuita, H., 2005. Essential amino acid supplementation to fish meal-

based diets with low protein to energy ratios improves the protein utilization in juvenile

rainbow trout Oncorhynchus mykiss. Aquaculture 246, 379-391.

Yun, B., Mai, K., Zhang, W., Xu, W., 2011. Effects of dietary cholesterol on growth

performance, feed intake and cholesterol metabolism in juvenile turbot (Scophthalmus

maximus L.) fed high plant protein diets. Aquaculture 319, 105-110.

Zahedifar, M., Castro, F.B., rskov, E.R., 2002. Effect of hydrolytic lignin on formation of

proteinlignin complexes and protein degradation by rumen microbes. Anim. Feed Sci.

Tech. 95, 8392.

Zarate, D.D., Lovell, R.T., 1997. Free lysine (L-lysine HCL) is utilized for growth less

efficiently than protein-bound lysine (soybean meal) in practical diets by young channel

catfish (Ictalurus punctatus). Aquaculture 159, 87-100.

Zarate, D.D., Lovell, R.T., Payne, M., 1999. Effects of feeding frequency and rate of stomach

evacuation on utilization of dietary free and protein-bound lysine for growth by channel

catfish Ictalurus punctatus. Aquac. Nutr. 5, 17-22.

Zhao, S., Wang, J., Song, X., Zhang, X., Ge, C., Gao, S., 2010. Impact of dietary protein on lipid

metabolism-related gene expression in porcine adipose tissue. Nutr. Metabolism 7, 1-13.

201
Zhou, Q-C., Yue, Y-R. 2010. Effect of replacing soybean meal with canola meal on growth, feed

utilization and haematological indices of juvenile hybrid tilapia, Oreochromis niloticus

Oreochromis aureus. Aquac. Res. 41, 982-990.

Zhu, C. L., Rademache, M., de Lange, C.F.M., 2005. Increasing dietary pectin level reduces

utilization of digestible threonine, but not lysine intake, for body protein deposition in

growing pigs. Anim. Sci. 83, 10441053.

Ziggers, D., 2007. Ethanol production and its co-products in Europe. All About Feed. URL:

http://www.allaboutfeed.net/article-database/ethanol-production-and-its-co-products-in-

europe-id1251.html, accessed on October 15, 2010.

202

Anda mungkin juga menyukai