Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-19T23:10:53.274Z Has data issue: false hasContentIssue false

In vitro determination of dietary protein and amino acid digestibility for humans

Published online by Cambridge University Press:  01 August 2012

Christine A. Butts*
Affiliation:
The New Zealand Institute for Plant & Food Research Limited, Batchelar Road, Palmerston North4474, New Zealand
John A. Monro
Affiliation:
The New Zealand Institute for Plant & Food Research Limited, Batchelar Road, Palmerston North4474, New Zealand Riddet Institute, Massey University, Tennent Drive, Palmerston North4474, New Zealand
Paul J. Moughan
Affiliation:
Riddet Institute, Massey University, Tennent Drive, Palmerston North4474, New Zealand
*
*Corresponding author: Dr C. A. Butts, fax +64 6 3517050, email Chrissie.butts@plantandfood.co.nz
Rights & Permissions [Opens in a new window]

Abstract

The development, refinement and validation of in vitro digestibility assays for dietary protein and amino acids for single stomached mammals are reviewed. The general principles of in vitro digestibility assays and their limitations are discussed. In vitro protein digestibility assays must be accurate, rapid, cheap, simple, robust, adaptable and relevant to the processes of digestion, absorption, and metabolism. Simple in vitro methods have the potential to give useful measures of in vivo amino acid and protein digestibility for humans. In vitro methods, including the complex multi-component models of digestion simulating the various physical and chemical processes, require independent validation with in vivo data from the target species or an acceptable animal model using the most appropriate in vivo measure of digestibility. For protein sources devoid of anti-nutritional factors or plant fibre, true ileal digestibility is the recommended in vivo baseline, while for plant proteins the recommended in vivo assay is real ileal digestibility. More published comparative studies are required to adequately validate in vitro digestibility assays.

Type
Full Papers
Copyright
Copyright © The Authors 2012

Humans require many nutrients including dietary indispensable amino acids, vitamins, minerals and fatty acids. The dietary indispensable amino acids required by humans are provided to the body mostly as intact dietary proteins which require digestion to release their component amino acids and small peptides. Proteins vary in their content of constituent amino acids and can be devoid of or low in one or more dietary indispensable amino acids, and when fed as the sole source of amino acids cannot sustain life. The biological utilisation of a protein is influenced by its composition, particularly with respect to the dietary indispensable amino acids, its digestion in the gastrointestinal tract, and the absorption and transport of amino acids and peptides from the gastrointestinal tract. Processing and storage conditions of a food affect the interactions between the components, with effects on digestibility that can have both beneficial and detrimental effects on protein nutritional quality.

Digestion and absorption processes in the live animal are complex, highly integrated and regulated, and are adaptable processes that have evolved to efficiently release nutrients for the body's growth, maintenance and reproduction. These dynamic processes are under both neural and hormonal control and respond to various stimuli. To simulate such a complex system in its entirety using static unresponsive in vitro methods is very difficult if not impossible(Reference Fuller1) and therefore how in vitro assays are applied should be tempered accordingly. Also, the effects of the gut microbiota are particularly difficult to simulate as is the diverse impact anti-nutritional factors and dietary fibre have on the digestive tract and its processes. Yet in vivo measures are expensive and time consuming. Therefore there is a need for a rapid reproducible in vitro digestibility bioassay that provides a reliable estimation of digestibility for a wide range of foods.

The development and practical details of in vitro procedures have been previously reviewed(Reference Boisen2Reference Eggum and Boisen5). In vitro protein digestibility assays and their validation with in vivo measures have been reviewed by Moughan(Reference Moughan6). An overview of in vitro digestion models for food applications has recently been published by Hur et al. (Reference Hur, Lim and Decker7). The present review summarises recent developments in in vitro digestibility bioassays and their validation with in vivo measures.

In vitro methods

In vitro digestibility methods began as simple one step incubations with pepsin or other proteases such as trypsin, papain, pronase or rennin(Reference Boisen2). These early single enzyme methods appeared to be satisfactory for comparing the effects of various treatments on a single foodstuff but gave lower digestible protein values than those obtained in vivo (faecal nitrogen digestibility). Other methods have simulated gastric and intestinal digestion using a 2-stage in vitro digestion involving digestion in a pepsin-HCl mixture followed by neutralisation, and then digestion in pancreatin(Reference Akeson and Stahmann8, Reference Büchmann9), trypsin(Reference Saunders, Connor and Booth10), or intestinal fluid from pigs(Reference Furuya, Sakamoto and Takahashi11). In general, there has been good agreement between these in vitro results and in vivo rat true faecal nitrogen digestibility.

Multi-enzyme processes that measure the pH drop after a 10 minute digestion have been developed(Reference Hsu, Vavak and Satterlee12, Reference Satterlee, Marshall and Tennyson13). Good correlations (r>0·8) were found with in vivo digestibility (rat true faecal) particularly when the protein sources were analysed by plant or animal origin(Reference Pedersen and Eggum14, Reference Pedersen and Eggum15), and this method was found to be highly reproducible across six laboratories(Reference McDonough, Sarwar and Steinke16). This approach assumes that the rate of change in pH is correlated with protein digestibility and that there is a direct relationship between the observed pH drop and the extent of protein hydrolysis(Reference Barbeau and Kinsella17, Reference Mozersky and Panettieri18). The components of some food materials, however, interfere with the pH drop due to their buffering capacity(Reference Pedersen and Eggum14, Reference Moughan, Schrama and Skilton19, Reference Urbano, Lopez-Jurado and Frejnagel20).

A modification of this approach is the pH stat method where the pH is kept constant by automatic titration with 0·1 M NaOH and the total amount of alkali used at the end of the incubation is recorded. This modification improved the prediction of in vivo protein digestibility, was reproducible, and gave high correlations (r>0·90) with in vivo (rat faecal) digestibility for highly digestible animal protein sources(Reference Pedersen and Eggum15).

The immobilised digestive enzyme assay (IDEA) system(Reference Porter, Swaisgood and Catignani21) uses a bioreactor containing enzymes immobilised on glass beads, eliminating contamination of the digest with digestive enzymes and preventing autolysis. However, this assay takes two and a half days to complete so is not as rapid as other in vitro techniques. Predicted digestibility values from the IDEA assay have been found to correlate (r = 0·8, 0·83) with rat faecal protein digestibility for a range of foods(Reference Thresher, Swaisgood and Catignani22, Reference Chang, Catignani and Swaisgood23).

Gauthier et al. (Reference Gauthier, Vachon and Jones24) developed an in vitro digestion under constant dialysis (molecular weight cut off 1000 Daltons) with specialised apparatus (dialysis cell method) to address the concern that enzyme activity is suppressed by the products of digestion. Continuous dialysis, however, needs extra and complex equipment, and some researchers do not regard it as necessary(Reference Boisen and Eggum3, Reference Büchmann9). The proteins undigested following a pepsin-pancreatin digestion were determined using polyacrylamide gel electrophoresis (PAGE), allowing the estimation of the molecular weight of the polypeptides and proteins remaining after hydrolysis(Reference Kim and Barbeau25). This procedure was time consuming, taking 5 days to complete, and is likely too labour intensive to be used in a rapid routine assay.

Near-infrared spectroscopy (NIRS) is a very rapid technique with low maintenance costs that shows great promise in evaluating nutritive value for human foods. It has been used routinely for determining the chemical composition, organic matter digestibility, energy digestibility of grains(Reference Hervera, Baucells and Gonzalez28, Reference Pujol, Pérez-Vendrell and Torrallardona29) and oil seeds(Reference Losada, Garcia-Rebollar and Alvarez26) for monogastric animals. There is a requirement, however, for a large number of reference samples to calibrate the instrument and these reference data must come from in vivo digestibility assays. There have been only a few published studies comparing NIRS and in vivo nitrogen or protein digestibility values. Prediction of wheat ileal crude protein digestibility for broilers(Reference Owens, McCann and McCracken27) was found to be highly variable (r 2 = 0·23–0·76). Faecal digestible protein of commercial dry extruded foods for dogs showed good predictions using the pH drop (r 2 = 0·78) method, and in vitro two-step digestion (r 2 = 0·81), but not for NIRS (r 2 = 0·53)(Reference Hervera, Baucells and Gonzalez28). Pig ileal nitrogen digestibility of barley (r 2 = 0·97, 0·72)(Reference Pujol, Pérez-Vendrell and Torrallardona29, Reference McCann, McCracken and Agnew30) but not wheat (r 2 = 0·22)(Reference Garnsworthy, Wiseman and Fegeros31) has been accurately predicted with NIRS.

A complex biochemical model of human adult and infant gastro-duodenal digestion has been developed(Reference Dupont, Mandalari and Molle32, Reference Dupont, Mandalari and Molle33) to investigate the allergenic potential of protein digestion products. Electrophoresis, immunochemical and mass spectrophotometry techniques were used to characterise the digestion products. This study found that heat treatment of milk increased the resistance of casein to the simulated digestion processes. An inter-laboratory study(Reference Mandalari, Adel-Patient and Barkholt34), comparing the digestion of β-casein and β-lactoglobulin using a simulated human gastro-duodenal in vitro assay, found it to be reproducible.

A computer-controlled, dynamic, multi-compartmental model (TIM) that seemingly closely simulates the conditions of the in vivo gastrointestinal tract of humans and monogastric animals has been developed at the TNO Nutrition & Food Research Institute, The Netherlands(Reference Minekus, Marteau and Havenaar35). It comprises compartments with flexible walls heated by water jackets with computer controlled pH adjustment, and rotary and peristaltic pumps to provide mechanical movement of the chyme. It simulates gastric pH change, peristaltic movement, gastric emptying rates, intestinal transit times, gastric, biliary and pancreatic secretion and their activities, small intestinal absorption (TIM-1), and the absorption from the large bowel (TIM-2) of volatile fatty acids and water(Reference Minekus, Marteau and Havenaar35, Reference Havenaar and Minekus36). This model has been applied in one published protein digestion study to date(Reference Venema and Dohnalek37) in which the model was modified to simulate the infant gut and to measure the degree of protein hydrolysis of whey protein hydrolysate infant formulas. The hydrolysis of the proteins was found to be only one factor in determining the digestibility of these food products. TIM has been used to investigate nutraceutical delivery vehicles(Reference Chen, Hebrard and Beyssac38), gut transit time and iron absorption(Reference Salovaara, Alminger and Eklund-Jonsson39), and the availability of heterocyclic aromatic amines(Reference Krul, Luiten-Schuite and Baan40), and has been used in pharmaceutical studies(Reference Blanquet, Zeijdner and Beyssac41). One of the strengths of these models is their flexibility and ability to model different gastrointestinal conditions such as those found in infants (Reference Dupont, Mandalari and Molle32, Reference Dupont, Mandalari and Molle33, Reference Venema and Dohnalek37), or other animal species(Reference Venema, Havenaar and Minekus42, Reference Smeets-Peeters, Minekus and Havenaar43).

A similar system to TIM is the Institute of Food Research, United Kingdom, dynamic gastric model (DGM) comprising two parts that mimic the fundus and antrum of the stomach, and this is integrated with a simulated intestine that mixes the digesta and adds bicarbonate, bile and digestive enzymes(Reference Wickham, Faulks and Mills44) and has also been used in pharmaceutical studies. Model stomach(Reference Kong and Singh45Reference Ferrua and Singh47) and small intestine(Reference Tharakan, Norton and Fryer48) systems have also been developed to better understand the mechanical forces and mass transfer kinetics of food digestion and absorption. To date these systems have not been used for the routine evaluation of protein or amino acid digestibility of foods.

These complex computer controlled systems are expensive to set up and maintain and therefore may not be a useful tool for the routine evaluation of food. Although being far more sophisticated than simple digestion assays, they do not provide an entirely accurate reproduction of the physiological system whereby the digesta interact with the gut cells leading to active transport of nutrients; nor do they simulate the neural and hormonal feedback mechanisms affecting digestion and absorption(Reference Yoo and Chen49). Their validity for predicting the in vivo digestion of food proteins has not yet been fully tested. The choice of a standard method will be a balance between accuracy and ease of use, and the endpoint may simply be to rank foods in order of nutrient digestibility. Multiple regression equations combining in vitro digestibility coefficients and chemical constituent measures may have application where greater accuracy is required.

Limitations of the in vitro approach

In general, there are some key requirements for the development of in vitro digestibility assays: matching in vivo enzymes in presence, sequence, enzyme:substrate ratios; standardising enzyme activities and specificities; controlling co-enzymes and co-factors, pH and temperature; separating digested from undigested material while considering the inhibition of end products on digestion; and allowing for the effects of sample size, particle size and particle size distribution(Reference Moughan6). Digestion and absorption involve complex physiological processes that are virtually impossible to reproduce in vitro (Reference Boisen2, Reference Savoie50). The effects of anti-nutritional factors, dietary dry-matter and fibre, endogenous protein secretions, activity of gut enzymes, and gut bacteria are not able to be mimicked in vitro (Reference Moughan6). The sensitivity of an assay is a function of both the time of the reaction and the enzyme-substrate ratio(Reference Parsons51, Reference Drake, Fuller and Chesson52). In vitro assays may need to include lipases, carbohydrases, and elastases to better mimic in vivo digestion and the release of proteins from the food matrix. There is an assumption that all soluble material is digestible whereas small peptides may not be absorbed in vivo particularly in heat-treated proteins.

An in vitro method may be precise and reliable in the laboratory but the resulting data must be correlated with in vivo data to provide a physiologically relevant measure of digestible protein. While it is not necessarily realistic to expect absolute agreement between in vitro and in vivo observations, in vitro assays should not be ruled out as useful tools. They can be used to rank food proteins according to their protein digestibility, to further our understanding of protein structure under digestion conditions(Reference Monogioudi, Faccio and Lille53), and with further refinement could be useful for the prediction of in vivo nutritive value. The development of multiple regression equations including in vitro digestibility measures combined with chemical constituent measures for individual foods has been recommended(Reference Moughan6). In addition, in vitro digestion techniques have an important role to play in extending our understanding of the release of amino acids and peptides during digestion and the influence protein structure plays on dietary protein quality(Reference Savoie, Agudelo and Gauthier54).

Evaluation and validation with in vivo assays

Many of the studies evaluating in vitro protein digestibility assays have compared the in vitro data with in vivo faecal digestibility. The influence of the microflora in the large intestine however, does not make this comparison useful. The ileal method for determining crude protein digestibility is more accurate(Reference Moughan and Smith55) than the faecal approach, and it is more relevant to in vitro measures given that these methods do not simulate microbial digestion in the large intestine(Reference Moughan56). It has been recommended previously that protein sources that do not contain fibre or anti-nutritional factors should be evaluated against true ileal protein digestibility(Reference Moughan, Schrama and Skilton19), while plant proteins including grains should be evaluated against real ileal protein digestibility(Reference Moughan6). Further details of these in vivo assays are given elsewhere in this supplement. There are only a few studies that have used these recommended comparisons and it is therefore difficult to properly evaluate the accuracy of current in vitro assays.

The determination of true or real ileal amino acid and nitrogen digestibility requires access to the terminal ileum, which is usually obtained via surgery, as well as a specific feeding or labelling technique to measure endogenous losses. The determination of in vivo protein and amino acid digestibility in humans is generally not possible so researchers use animal models. The pig(Reference Rowan, Moughan and Wilson57, Reference Deglaire, Bos and Tome58) and the rat(Reference Bodwell, Satterlee and Hackler59, Reference Rich, Satterlee and Smith60) have been found to be suitable animal models for humans. There is very little published information on ileal amino acid digestibility determined for humans directly(Reference Fuller and Tome61), making the evaluation of in vitro digestibility data even more difficult. One of the few studies to compare human digestibility (albeit faecal digestibility) and in vitro digestibility values on the same foods is that of Bodwell et al. (Reference Bodwell, Satterlee and Hackler59). They found there was no significant correlation between in vitro, human and rat digestibilities when all the proteins were considered together. However, when the data were analysed with the food proteins identified separately as of animal or plant origin the correlations were high (r>0·90).

There have been mixed results in studies comparing in vitro and in vivo ileal digestibilities in pigs on the same protein sources ranging from no significant correlation(Reference Thomaneck, Hennig and Souffrant62), low correlation(Reference Cone and van der Poel63), to good correlations(Reference Pujol, Torrallardona and Boisen64, Reference Jezierny, Mosenthin and Sauer65). A comprehensive study of a wide range of feedstuffs(Reference Jaguelin, Fevrier and Seve66) used standardised true ileal nitrogen digestibility determined in the growing pig and a two-step pepsin/pancreatin in vitro digestion. High correlations were obtained within feedstuffs when the prediction equations also included terms for specific chemical components of the foods. As would be expected, true ileal digestibility gave better correlations than when apparent ileal digestibility was used. The authors concluded that due to differences in endogenous protein loss induced by the different foods a single prediction equation based on the in vitro assay would not be appropriate.

Methodology needs to be standardised across research organisations(Reference Mosenthin, Jansman and Eklund67, Reference Stein, Fuller and Moughan68) and reference proteins should be included in every assay for comparison across studies. In vivo digestibility is influenced by the experimental conditions, and the way digestibility is calculated such as the presence of endogenous protein, and methods of endogenous loss correction(Reference Boisen and Moughan69, Reference Boisen and Moughan70). It is now widely accepted that protein and amino acid digestibility must be determined using accurate measurements of dietary intake and terminal ileal digesta flow corrected for basal endogenous protein loss to give ‘true’ digestibility values. However, some foods also contain components (dietary fibre, anti-nutritional factors) that induce higher endogenous protein losses than the basal losses measured with purified highly digestible protein-free diets. There have been only a few studies(Reference Hur, Lim and Decker7, Reference Pujol and Torrallardona71) that have attempted standardisation of methodology and more need to published(Reference Hur, Lim and Decker7), and correlations determined both within and across food proteins.

Conclusions and recommendations

Simple in vitro digestion methods have the potential to give useful measures of in vivo amino acid and protein digestibility for humans. An in vitro method to measure the extent of digestion of protein must be accurate, rapid, cheap, simple, robust, adaptable and relevant to the processes of digestion, absorption, and metabolism. The more complex in vitro methods including the computer controlled in vitro models of the digestive tract that are attempting to more closely mimic the processes of digestion and absorption may be too expensive and time consuming for the rapid and routine analysis of food protein quality but are useful tools in expanding our understanding of digestion processes. All these methods require independent validation with in vivo data from the target species or an acceptable animal model.

The literature contains many papers describing the development of in vitro protein digestibility assays, but all too often no validation results are given or if they are, an inappropriate in vivo digestibility baseline has been used. Very few studies have included a full developmental process where the sensitivity of important assay variables is tested and then the variables optimised. Repeatability is seldom examined or at least reported. Promising methods have often been published and applied, only to fall into obscurity over time, presumably because they do not provide consistently accurate results.

Over the last few years, the application of in vitro digestibility assays is becoming common, particularly in food science research to examine effects of food physical and chemical structures on nutrient digestibility. There is a pressing need for the thorough development of an agreed (i.e. standardised) protein digestibility assay with application to human foods. Such an assay should be fully tested independently of the developer in a number of different laboratories around the world, and should be evaluated against relevant observations of in vivo digestibility.

Acknowledgements

C. A. B. wrote the manuscript and had responsibility for the final content. P. J. M. and J. A. M. reviewed manuscript. All authors read and approved the final manuscript. The authors declare no conflict of interest. This research received no specific grant from any funding agency in the public, commercial or not-for-profit sectors.

References

1Fuller, MF (1991) Methodologies for the measurement of digestion. In Digestive Physiology in Pigs, pp. 273288 [, editors]. Wageningen: Pudoc.Google Scholar
2Boisen, S (2000) In vitro digestibility methods: history and specific approaches. In Feed Evaluation: Principles and Practice, pp. 153168 [, editors]. Wageningen: Wageningen Pers.Google Scholar
3Boisen, S & Eggum, BO (1991) Critical Evaluation of in vitro Methods for Estimating Digestibility in Simple-Stomach Animals. Nutr Res Rev 4, 141162.CrossRefGoogle ScholarPubMed
4Swaisgood, HE & Catignani, GL (1993) Protein digestibility: in vitro methods of assessment. Adv Food Nutr Res 35, 185236.CrossRefGoogle Scholar
5Eggum, BO & Boisen, S (1991) In vitro techniques of measuring digestion. In Digestive physiology in pigs: Proceedings of the Vth International Symposium on Digestive Physiology in Pigs, pp. 213225 [, editors]. Wageningen, The Netherlands: Pudoc Wageningen.Google Scholar
6Moughan, PJ (1999) In vitro techniques for the assessment of the nutritive value of feed grains for pigs: a review. Aust J Agric Res 50, 871879.CrossRefGoogle Scholar
7Hur, SJ, Lim, BO, Decker, EA, et al. (2011) In vitro human digestion models for food applications. Food Chem 125, 112.CrossRefGoogle Scholar
8Akeson, WR & Stahmann, MA (1964) A pepsin pancreatin digest Index of protein quality evaluation. J Nutr 83, 257261.CrossRefGoogle ScholarPubMed
9Büchmann, NB (1979) Variation in in vitro digestibility of barley protein. J Sci Food Agric 30, 590596.CrossRefGoogle ScholarPubMed
10Saunders, RM, Connor, MA, Booth, AN, et al. (1973) Measurement of digestibility of alfalfa protein concentrates by an in vivo and in vitro methods. J Nutr 103, 503535.CrossRefGoogle Scholar
11Furuya, S, Sakamoto, K & Takahashi, S (1979) A new in vitro method for the estimation of digestibility using the intestinal fluid of the pig. Br J Nutr 41, 511520.CrossRefGoogle ScholarPubMed
12Hsu, HW, Vavak, DL, Satterlee, LD, et al. (1977) A multienzyme technique for estimating protein digestibility. J Food Sci 42, 12691273.CrossRefGoogle Scholar
13Satterlee, LD, Marshall, HF & Tennyson, JM (1979) Measuring protein quality. J Am Oil Chem Soc 56, 103109.CrossRefGoogle ScholarPubMed
14Pedersen, B & Eggum, BO (1981) Prediction of protein digestibility by in vitro procedures based on two multi-enzyme systems. Zeitschrift fur Tierphysiologie, Tierernahrung und Futtermittelkunde 45, 190200.CrossRefGoogle Scholar
15Pedersen, B & Eggum, BO (1983) Prediction of protein digestibility by an in vitro enzymatic pH-stat procedure. Zeitschrift Fur Tierphysiologie Tierernahrung Und Futtermittelkunde-Journal of Animal Physiology and Animal Nutrition 49, 265277.CrossRefGoogle Scholar
16McDonough, FE, Sarwar, G, Steinke, FH, et al. (1990) In vitro assay for protein digestibility: Interlaboratory study. J Assoc Off Anal Chem 73, 622625.Google ScholarPubMed
17Barbeau, WE & Kinsella, JE (1985) Effects of free and bound chlorogenic acid on the in vitro digestibility of ribulose bisphosphate carboxylase from spinach. J Food Sci 50, 10831087.CrossRefGoogle Scholar
18Mozersky, SM & Panettieri, RA (1983) Is pH drop a valid measure of extent of protein hydrolysis? J Agric Food Chem 31, 13131316.CrossRefGoogle Scholar
19Moughan, PJ, Schrama, J, Skilton, GA, et al. (1989) In-vitro determination of nitrogen digestibility and lysine availability in meat and bone meals and comparison with in-vivo ileal digestibility estimates. J Sci Food Agric 47, 281292.CrossRefGoogle Scholar
20Urbano, G, Lopez-Jurado, M, Frejnagel, S, et al. (2005) Nutritional assessment of raw and germinated pea (Pisum sativum L.) protein and carbohydrate by in vitro and in vivo techniques. Nutrition 21, 230239.CrossRefGoogle ScholarPubMed
21Porter, DH, Swaisgood, HE & Catignani, GL (1984) Characterization of an immobilized digestive enzyme system for determination of protein digestibility. J Agric Food Chem 32, 334339.CrossRefGoogle Scholar
22Thresher, WC, Swaisgood, HE & Catignani, GL (1989) Digestibilities of the protein in various foods as determined in vitro by an immobilized digestive enzyme assay (IDEA). Plant Food Hum Nutr 39, 5965.CrossRefGoogle ScholarPubMed
23Chang, HI, Catignani, GL & Swaisgood, HE (1990) Protein digestibility of alkali- and fructose-treated protein by rat true digestibility assay and by the immobilized digestive enzyme assay system. J Agric Food Chem 38, 10161018.CrossRefGoogle Scholar
24Gauthier, SF, Vachon, C, Jones, JD, et al. (1982) Assessment of protein digestibility by in vitro enzymatic hydrolysis with simultaneous dialysis. J Nutr 112, 17181725.CrossRefGoogle ScholarPubMed
25Kim, YA & Barbeau, WE (1991) Evaluation of SDS-PAGE method for estimating protein digestibility. J Food Sci 56, 10821086.CrossRefGoogle Scholar
26Losada, B, Garcia-Rebollar, P, Alvarez, C, et al. (2010) The prediction of apparent metabolisable energy content of oil seeds and oil seed by-products for poultry from its chemical components, in vitro analysis or near-infrared reflectance spectroscopy. Anim Feed Sci Technol 160, 6272.CrossRefGoogle Scholar
27Owens, B, McCann, MEE, McCracken, KJ, et al. (2009) Prediction of wheat chemical and physical characteristics and nutritive value by near-infrared reflectance spectroscopy. Br Poult Sci 50, 103122.CrossRefGoogle ScholarPubMed
28Hervera, M, Baucells, MD, Gonzalez, G, et al. (2009) Prediction of digestible protein content of dry extruded dog foods: comparison of methods. J Anim Physiol Anim Nutr 93, 366372.CrossRefGoogle ScholarPubMed
29Pujol, S, Pérez-Vendrell, AM & Torrallardona, D (2007) Evaluation of prediction of barley digestible nutrient content with near-infrared reflectance spectroscopy (NIRS). Livest Sci 109, 189192.CrossRefGoogle Scholar
30McCann, MEE, McCracken, KJ & Agnew, RE (2006) The use of near infrared reflectance spectroscopy (NIRS) for prediction of the nutritive value of barley for growing pigs. Irish J Agr Food Res 45, 187195.Google Scholar
31Garnsworthy, PC, Wiseman, J & Fegeros, K (2000) Prediction of chemical, nutritive and agronomic characteristics of wheat by near infrared spectroscopy. J Agric Sci, Cambridge 135, 409417.CrossRefGoogle Scholar
32Dupont, D, Mandalari, G, Molle, D, et al. (2010) Comparative resistance of food proteins to adult and infant in vitro digestion models. Mol Nutr Food Res 54, 767780.CrossRefGoogle ScholarPubMed
33Dupont, D, Mandalari, G, Molle, D, et al. (2010) Food processing increases casein resistance to simulated infant digestion. Mol Nutr Food Res 54, 16771689.CrossRefGoogle ScholarPubMed
34Mandalari, G, Adel-Patient, K, Barkholt, V, et al. (2009) In vitro digestibility of beta-casein and beta-lactoglobulin under simulated human gastric and duodenal conditions: A multi-laboratory evaluation. Regul Toxicol Pharmacol 55, 372381.CrossRefGoogle ScholarPubMed
35Minekus, M, Marteau, P, Havenaar, R, et al. (1995) A multicompartmental dynamic computer-controlled model simulating the stomach and small intestine. ATLA-Altern Lab Anim 23, 197209.CrossRefGoogle Scholar
36Havenaar, R & Minekus, M (1996) Simulated assimilation. Dairy Ind Int 61, 1720.Google Scholar
37Venema, K & Dohnalek, MH (2006) Digestibility and availability of nitrogen for whey protein hydrolysate products assessed using a dynamic in vitro gastrointestinal model. J Allergy Clin Immunol 117, S304.CrossRefGoogle Scholar
38Chen, L, Hebrard, G, Beyssac, E, et al. (2010) In vitro study of the release properties of soy-zein protein microspheres with a dynamic artificial digestive system. J Agric Food Chem 58, 98619867.CrossRefGoogle ScholarPubMed
39Salovaara, S, Alminger, ML, Eklund-Jonsson, C, et al. (2003) Prolonged transit time through the stomach and small intestine improves iron dialyzability and uptake in vitro. J Agric Food Chem 51, 51315136.CrossRefGoogle ScholarPubMed
40Krul, C, Luiten-Schuite, A, Baan, R, et al. (2000) Application of a dynamic in vitro gastrointestinal tract model to study the availability of food mutagens, using heterocyclic aromatic amines as model compounds. Food Chem Toxicol 38, 783792.CrossRefGoogle Scholar
41Blanquet, S, Zeijdner, E, Beyssac, E, et al. (2004) A dynamic artificial gastrointestinal system for studying the behavior of orally administered drug dosage forms under various physiological conditions. Pharm Res 21, 585591.CrossRefGoogle ScholarPubMed
42Venema, K, Havenaar, R & Minekus, M (2009) Improving in vitro simulation of the stomach and intestines [, editors]. Cambridge UK: Woodhead Publishing Ltd.CrossRefGoogle Scholar
43Smeets-Peeters, MJE, Minekus, M, Havenaar, R, et al. (1999) Description of a dynamic in vitro model of the dog gastrointestinal tract and an evaluation of various transit times for protein and calcium. ATLA-Altern Lab Anim 27, 935949.CrossRefGoogle Scholar
44Wickham, M, Faulks, R & Mills, C (2009) In vitro digestion methods for assessing the effect of food structure on allergen breakdown. Mol Nutr Food Res 53, 952958.CrossRefGoogle ScholarPubMed
45Kong, F & Singh, RP (2009) Digestion of raw and roasted almonds in simulated gastric environment. Food Biophys 4, 365377.CrossRefGoogle Scholar
46Kong, F & Singh, RP (2008) A model stomach system to investigate disintegration kinetics of solid foods during gastric digestion. J Food Sci 73, E202E210.CrossRefGoogle Scholar
47Ferrua, MJ & Singh, RP (2010) Modeling the fluid dynamics in a human stomach to gain insight of food digestion. J Food Sci 75, R151R162.CrossRefGoogle Scholar
48Tharakan, A, Norton, IT, Fryer, PJ, et al. (2010) Mass Transfer and Nutrient Absorption in a Simulated Model of Small Intestine. J Food Sci 75, E339E346.CrossRefGoogle Scholar
49Yoo, JY & Chen, XD (2006) GIT Physicochemical Modeling - A Critical Review. Int J Food Eng 2, 4.CrossRefGoogle Scholar
50Savoie, L (1994) Digestion and absorption of food: usefulness and limitations of in vitro models. Can J Physiol Pharmacol 72, 407414.CrossRefGoogle ScholarPubMed
51Parsons, CM (1991) Use of pepsin digestibility, multienzyme pH change and protein solubility assays to predict in vivo protein quality of foodstuffs. In In Vitro Digestion for Pigs and Poultry, pp. 105115 [, editor]. Wallingford, United Kingdom: CAB International.Google Scholar
52Drake, AP, Fuller, MF & Chesson, A (1991) Simultaneous estimations of precaecal protein and carbohydrate digestion in the pig. In In Vitro Digestion for Pigs and Poultry, pp. 162176 [, editor]. Wallingford, United Kingdom: CAB International.Google Scholar
53Monogioudi, E, Faccio, G, Lille, M, et al. (2011) Effect of enzymatic cross-linking of beta-casein on proteolysis by pepsin. Food Hydrocolloids 25, 7181.CrossRefGoogle Scholar
54Savoie, L, Agudelo, RA, Gauthier, SF, et al. (2005) In vitro determination of the release kinetics of peptides and free amino acids during the digestion of food proteins. J AOAC Int 88, 935948.CrossRefGoogle ScholarPubMed
55Moughan, PJ & Smith, WC (1985) Determination and assessment of apparent ileal amino acid digestibility coefficients for the growing pig. New Zeal J Agri Res 28, 365370.CrossRefGoogle Scholar
56Moughan, PJ (2003) Amino acid availability: aspects of chemical analysis and bioassay methodology. Nutr Res Rev 16, 127141.CrossRefGoogle ScholarPubMed
57Rowan, AM, Moughan, PJ, Wilson, MN, et al. (1994) Comparison of the ileal and faecal digestibility of dietary amino acids in adult humans and evaluation of the pig as a model animal for digestion studies in man. Br J Nutr 71, 2942.CrossRefGoogle Scholar
58Deglaire, A, Bos, C, Tome, D, et al. (2009) Ileal digestibility of dietary protein in the growing pig and adult human. Br J Nutr 102, 17521759.CrossRefGoogle ScholarPubMed
59Bodwell, CE, Satterlee, LD & Hackler, LR (1980) Protein digestibility of the same protein preparations by human and rat assays and by in vitro enzymic digestion methods. Am J Clin Nutr 33, 677686.CrossRefGoogle ScholarPubMed
60Rich, N, Satterlee, LD & Smith, JL (1980) A comparison of in vivo apparent protein digestibility in man and rat to in vitro protein digestibility as determined using human and rat pancreatins and commercially available proteases. Nutr Rep Int 21, 285300.Google Scholar
61Fuller, MF & Tome, D (2005) In vivo determination of amino acid bioavailability in humans and model animals. J AOAC Int 88, 923934.CrossRefGoogle ScholarPubMed
62Thomaneck, D, Hennig, U & Souffrant, WB (1991) In vitro determination of protein digestibility of feeds using a “digestion cell” (preliminary results). In Digestive Physiology in Pigs, pp. 345348 [, editors]. Wageningen, The Netherlands: Pudoc.Google Scholar
63Cone, JW & van der Poel, AFB (1993) Prediction of apparent ileal protein digestibility in pigs with a Two-step in-vitro method. J Sci Food Agric 62, 393400.CrossRefGoogle Scholar
64Pujol, S, Torrallardona, D & Boisen, S (2001) In vivo and in vitro ileal digestibility of protein and amino acids in barleys. In Digestive Physiology of Pigs: Proceedings of the 8th Symposium, pp. 344346 [, editors]. Wallingford, UK: CABI Publishing.Google Scholar
65Jezierny, D, Mosenthin, R, Sauer, N, et al. (2010) In vitro prediction of standardised ileal crude protein and amino acid digestibilities in grain legumes for growing pigs. Animal 4, 19871996.CrossRefGoogle ScholarPubMed
66Jaguelin, Y, Fevrier, C, Seve, B, et al. (1994) Assessment of the apparent and true N digestibility in pig for several classes of feedstuffs through an in vitro determination. In Digestive Physiology in Pigs, pp. 114117 [, editors]. Bad Doberan, Germany: EAAP.Google Scholar
67Mosenthin, R, Jansman, AJM & Eklund, M (2007) Standardization of methods for the determination of ileal amino acid digestibilities in growing pigs. Livest Sci 109, 276281.CrossRefGoogle Scholar
68Stein, HH, Fuller, MF, Moughan, PJ, et al. (2007) Definition of apparent, true, and standardized ileal digestibility of amino acids in pigs. Livest Sci 109, 282285.CrossRefGoogle Scholar
69Boisen, S & Moughan, PJ (1996) Dietary influences on endogenous ileal protein and amino acid loss in the pig - a review. Acta Agric Scandinavica A Animal Science 46, 154164.Google Scholar
70Boisen, S & Moughan, PJ (1996) Different expressions of dietary protein and amino acid digestibility in pig feeds and their application in protein evaluation: a theoretical approach. Acta Agric Scand Sect Ani Sci 46, 165172.Google Scholar
71Pujol, S & Torrallardona, D (2007) Evaluation of in vitro methods to estimate the in vivo nutrient digestibility of barley in pigs. Livest Sci 109, 186188.CrossRefGoogle Scholar