Hostname: page-component-7479d7b7d-fwgfc Total loading time: 0 Render date: 2024-07-11T02:04:04.190Z Has data issue: false hasContentIssue false

The role of food structure in gastric-emptying rate, absorption and metabolism

Published online by Cambridge University Press:  06 September 2023

Alan Mackie*
Affiliation:
School of Food Science and Nutrition, University of Leeds, Leeds LS2 9JT, UK
*
Corresponding author: Alan Mackie, email a.r.mackie@leeds.ac.uk
Rights & Permissions [Opens in a new window]

Abstract

The high levels of non-communicable diseases such as CVD and type 2 diabetes mellitus are linked to obesity and poor diet. This continuing emphasis on health in relation to food is proving a powerful driver for the development of cheap but palatable and more functional foods. However, the efficacy of such foods is often hard to prove in human subjects. Thus, a suite of tools has been developed including in silico and in vitro simulations and animal models. Although animal models offer physiologically relevant platforms for research, their use for experimentation is problematic for consumers. Thus, in vitro methods such as Infogest protocols have been developed to provide digestion endpoints or even an indication of the kinetics of digestion. These protocols have been validated for a range of food systems but they still miss the final absorption step. This review discusses the use of such in vitro models and what further steps need to be included to make the bioaccessibility determination more relevant to bioavailability and human health.

Type
Conference on ‘Architecture of food: Processing, structure and health’
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press on behalf of The Nutrition Society

There is increasing evidence that the food we eat needs to be healthier and more sustainable. Henry Dimbleby in his National Food Strategy (https://www.nationalfoodstrategy.org/) highlighted the issue. The same range of drivers is also pushing the trend for more minimally processed food, while concern about additives and a lack of understanding of food ingredients is leading to classification of some foods as ‘ultra processed’(Reference Levine and Ubbink1). At the same time, climate concerns associated with animal production are driving trends in lower consumption of animal-based foods: meat, dairy, eggs, etc. in favour of more plant-based foods. Despite these concerns about nutritional quality and sustainability, the food production system must continue to feed everyone all of the time. These concerns may have arisen because as a population, we do not value our food sufficiently or in a way that balances nutrition and sustainability with cost, safety and palatability, which are the primary drivers of consumer choice. For much of the population, food choice is driven by palatability and cost so consequently these have become the main drivers for retailers and fast-food outlets(Reference Drewnowski and Specter2). It has been suggested that in addition to providing dietary advice, the food suppliers should be making more of health by stealth approaches(Reference Jackson, Cameron and Rolfe3). There are already a number of examples where the food industry has been able to make significant changes to formulations of staple products such as the reduction of salt in bakery products(Reference Regan, Kent and Raats4). There is similar work ongoing to increase very gradually the amount of fibre in a range of food products. Although there may be evidence correlating a low-fibre intake with disease, we can take a more objective measure such as glycaemic index (GI) as a measure of dietary quality rather than fibre content. Not least because the current definitions of dietary fibre are unhelpful(Reference Slavin5). In a recent article, Jenkins et al. showed that the risk of CVD among the study participants increased with dietary GI. This was accentuated in those with a BMI over 25 kg/m2(Reference Jenkins, Dehghan and Mente6). Specifically, for those with a BMI less than 25 kg/m2 the hazard ratio for the top GI quintile was 1⋅14(sd 0⋅14), while for those with a BMI over 25 kg/m2 the hazard ratio for the top quintile was 1⋅38(sd 0⋅16). It is noteworthy that the average BMI of middle-aged (55–64) UK adults is 28⋅1 kg/m2(Reference Pai and Gulliford7). Similarly, there is a strong correlation between dietary GI and risk of type 2 diabetes mellitus. Local dose dependence of the relative risk of type 2 diabetes mellitus on the GI in prospective cohort studies combined showed that the dose–response type 2 diabetes mellitus–GI risk relation rose by 32 % per 0⋅1 increment in the GI(Reference Livesey, Taylor and Livesey8).

Although we have evidence that dietary GI correlates with non-communicable diseases, the mechanism is less clear. What we know is that the GI is linked to digestion kinetics(Reference Jenkins, Kendall and Augustin9). The main site of control of digestion kinetics is the gastric compartment because the stomach acts as a container for the food that is consumed in a meal and releases it in a controlled way into the small intestine, which is the primary site of digestion and absorption. The rate of gastric emptying is governed by a number of factors, specifically the textural properties of the gastric chyme and the nutrients being delivered to the duodenum(Reference Goyal, Guo and Mashimo10). Thus, liquids empty faster than more solid meals and low-energy foods empty faster than high-energy foods. In this context water has a short gastric residence time while a nutrient-dense solid meal will have a long residence time. In addition, there is a link between the blood glucose concentrations and gastric emptying, with higher concentrations decreasing gastric-emptying rates(Reference Phillips, Deane and Jones11). The control of gastric emptying is done through a number of mechanisms. Glucose absorption in the small intestine induces a feedback loop via cholecystokinin, peptide YY, glucagon-like peptide 1 (GLP-1) and glucose-dependent insulinotropic polypeptide (GIP) which are secreted from the intestine in response to nutrient exposure. GLP-1 and GIP induce the release of insulin and GLP-1 inhibits glucagon secretion, which attenuates postprandial glycaemic excursions. At the same time, the blood glucose concentration modulates gastric emptying, such that acute elevations of blood glucose levels slow gastric emptying (effects are evident even within the physiological range) and emptying is accelerated during hypoglycaemia.

Food structure and gastric emptying

As an illustration of some of these effects, we undertook a study to determine the extent to which oat particle size in a porridge could alter glucose absorption, gastric emptying, gastrointestinal hormone response and subjective feelings of appetite and satiety(Reference Mackie, Bajka and Rigby12). In a crossover design, eight participants were fed porridge prepared from either oat flakes or oat flour with the same protein, fat, carbohydrate and mass. Subjective appetite ratings, gastric contents and plasma glucose, insulin and gastrointestinal hormones were determined over a period of 3 h post-consumption. The use of MRI provides direct visualisation of gastric content: changes in gastric emptying and also what is emptied. As an example, Fig. 1A and 1B shows MRI images for 5 and 25 min, respectively, post-consumption of oat flake porridge. The abdominal cross-sections show that most of the 175 ml liquid consumed with the 264 g porridge was emptied within 25 min of consumption. This highlights that structure is important in defining what is emptied. Regardless of the early emptying of gastric liquid from the oat flake porridge, its structure meant that after 3 h post-consumption the oat flake porridge had an average 25 % greater volume remaining in the stomach than the starch porridge. Despite the limited differences in the rate of gastric emptying, significant differences were seen in plasma GIP and insulin and minor differences in GLP-1. The peak in GIP and GLP-1 was at 20 min post-consumption for the starch porridge and 35 min for the flake porridge. The peak timing was reversed for insulin and no differences were seen in plasma glucose. Thus, highlighting that in healthy individuals blood glucose concentrations are tightly constrained even when very different amounts are being absorbed as suggested by the differences in GIP.

Fig. 1. Axial FIESTA (Fast Imaging Employing Steady-state Acquisition) MRI images of the stomach (outlined) taken 5 min (A) and 25 min (B) post-consumption of oat flake porridge. Image (A) shows a layer above the oat flake porridge that is not apparent after 25 min (B). Axial TrueFISP (True Fast Imaging with steady state precession) MRI images of the stomach (outlined) taken 5 min after consumption of a semi-solid (C) or liquid (D) version of the same meal.

Similarly, structural effects can be seen in the digestion and absorption of protein and lipids. Indeed food structures can be tailored to alter the timing of the delivery of specific macronutrients. Making use of density differences to drive creaming or sedimentation of components can be particularly effective(Reference Mackie, Rafiee and Malcolm13). When participants consumed a liquid or semi-solid meal with the same fat, protein, carbohydrate and energy as shown in Fig. 1C and 1D, differences in emptying behaviour were seen. The dark shapes in the stomach in Fig. 1C are boluses of high fat and protein cheese formed in the mouth and the swallowed. These remained visible for up to an hour before dispersing, trapping the protein and fat at the bottom of the stomach. In contrast, the liquid meal already showed evidence of creaming of the fat to the top of the stomach after 5 min. The food boluses were measured and compared with particle-size distributions from other meals, which highlighted the influence of the different meal structural properties on gastric chyme(Reference Hornby, Collado-González and Zhang14). Thus in the first hour the composition of the gastric chyme emptied into the small intestine would have been very different. This was confirmed rather circumstantially as the gastric-emptying rate for the liquid meal at 35 min post-consumption was significantly faster. The differences in emptying then led to differences in gastric volume and subjective appetite scores after 3 h. In another study using MRI to compare the effects of energy density and viscosity on gastric-emptying rate(Reference Camps, Mars and de Graaf15), the authors found that increasing the viscosity was less effective at slowing gastric emptying than increasing the energy density. However, the viscosity was more important in increasing the perception of fullness. The results highlighted the lack of satiation from ‘empty energy’ in quickly ingested drinks such as fizzy drinks. Although these examples highlight the role of food structure in digestion kinetics, it is also apparent that measuring the bioactive concentrations in peripheral blood cannot tell the whole story because we do not know what was being absorbed. In order to fill this gap, a number of simulations of digestion have been developed to mimic human physiology and enable the digestive fate of bioactives to be more closely followed.

In vitro simulations of the influence of food structure on digestive fate

The Infogest network has been a key player in developing physiologically relevant and widely usable in vitro simulations of food digestion. The network was originally assembled through a COST Action (FA1005) that had protein digestion as a core consideration in relation to food allergy(Reference Dupont, Bordoni and Brodkorb16). However, this was soon broadened to include all food and macronutrient types. The most widely cited simulation protocol is the static model that uses fixed conditions to simulate upper gastrointestinal tract digestion(Reference Minekus, Alminger and Alvito17,Reference Brodkorb, Egger and Alminger18) . This simulation has been used in hundreds of studies and comprises three phases, oral, gastric and intestinal. It can provide valuable information on digestion endpoints for any bioactives where microbial fermentation does not play a role. The oral phase includes simulated salivary fluid, salivary amylase and recommendations for how the chewing of food should be simulated. The oral phase generates the food bolus that is passed into the stomach for further digestion. The gastric phase of the simulation includes simulated gastric fluid as well as the protease pepsin and a recommendation for the use of gastric lipase. The recommendation is for this phase to last for 2 h at pH 3, which is an estimate of mean pH values achieved in vivo when half of a meal has been emptied from the stomach into the duodenum. Although this is the recommendation, there is some evidence that the pH should be higher, perhaps closer to 5(Reference Sams, Paume and Giallo19) depending on the nature of the meal. This is important because it determines which gastric enzymes are active. Salivary amylase is active in the stomach at higher pH, gastric lipase in the mid-range of pH and the protease pepsin at low pH. After 2 h, chyme from the gastric phase is emptied into the intestinal phase for digestion to be completed. The intestinal simulation includes simulated intestinal fluid, bile and pancreatic enzymes incubated at pH 7 for 2 h. This kind of approach allows a range of different endpoints to be determined, depending on the nutrients or bioactives of interest. These might include determination of peptide profiles or free amine groups for protein digestion, maltose concentrations for starch and NEFA profiles for lipid digestion. It should be highlighted that this is a static model using fixed parameters and thus is unlikely to give physiologically relevant kinetic data. However, it can be used for assessing digestibility of nutrients such as protein(Reference Sousa, Recio and Heimo20) and is a strong alternative to replace animal models for determining protein nutritional quality. The key requirements of this simulation of the upper gastrointestinal tract are reproducibility and physiological relevance. The reproducibility has been confirmed in a number of ring trials(Reference Egger, Ménard and Delgado-Andrade21) and the physiological relevance has been demonstrated in a number of different systems(Reference Egger, Schlegel and Baumann22Reference Miralles, Sanchon and Sanchez-Rivera24). However, it should be noted that this type of simulation is only able to model luminal events and it lacks brush border enzymes and absorptive elements.

As highlighted earlier, it has become apparent that it is not just the extent of digestion that is important for disease but also the rate of digestion and nutrient release. Thus, it is important to have ways of determining digestion kinetics and this can be achieved using simple semi-dynamic simulations of digestion(Reference Mulet-Cabero, Egger and Portmann25) or more sophisticated computer-controlled simulations(Reference Li and Kong26). The most important phase of digestion to simulate in relation to digestion kinetics is the gastric phase. Thus, kinetic simulations tend to concentrate on this phase and control factors such as gastric loading, gastric emptying, acidification rate, enzyme and simulated gastric fluid secretion rate and physical processing of the gastric chyme. The examples shown in this article draw on the Infogest semi-dynamic protocol that has been used to determine the impact of food structure on digestion kinetics but there are many other similar models in the literature.

In a study undertaken in 2012 and published in 2013, the effect of the structure of dairy products was initially investigated(Reference Mackie, Rafiee and Malcolm13). In this study, we were able show that the semi-solid structure persisted in the stomach and suppressed gastric emptying over the first hour compared to the liquid meal. This difference then led to persistent differences in volume of gastric contents and associated feelings of fullness for up to 3 h post-ingestion. Although this study hints at the role of food structure in digestion kinetics in support of previous work(Reference Marciani, Faulks and Wickham27), it does not directly show that the rate of appearance of the products of digestion varied. In order to determine that kind of information, studies must either use intubation(Reference Armand, Borel and Pasquier28) or resort to animal or in vitro models. In a subsequent study the Infogest semi-dynamic model was used to follow the digestion of the same two meals in more detail(Reference Mulet-Cabero, Rigby and Brodkorb29). In those simulations, the detailed analysis of protein and lipid digestion was able to show that gastric behaviour was affected by the initial structure with creaming and sedimentation observed in the case of liquid and semi-solid samples, respectively. Lipid and protein digestion profiles showed clear differences in the amount of nutrients reaching the simulated small intestine and, consequently, the likely bioaccessibility after digestion. The semi-solid sample generated higher nutrient released into the small intestine at an early stage of digestion whereas nutrient accessibility from liquid sample was delayed due to the formation of a cream layer in the gastric phase. This shows the strong effect of the matrix on gastric behaviour, proteolysis and lipolysis.

In two similar studies, dairy processing (heating and/or homogenisation) was used to alter the microstructure of cows' milk prior to simulated digestion(Reference Mulet-Cabero, Mackie and Wilde30,Reference Ye, Cui and Dalgleish31) . Both studies showed the typical clotting behaviour of whole milk in the gastric phase of digestion. They also showed differences in clot consistency depending on the processing applied to the milk. Unprocessed ‘raw’ milk had the firmest clot while homogenised and heated and homogenised presented weaker clots. In particular, the study by Mulet-Cabero et al. showed that the clot from raw or heated milk was dense enough to sediment, while the homogenised and heated and homogenised curd entrapped sufficient lipid to cream to the top of the gastric compartment. These structural changes occurring during the gastric phase resulted in different nutrient emptying, with significant differences between ‘raw’ and heated and homogenised, and more extensive digestion of milk proteins in the heat-treated samples due to the drastic denaturation of the proteins.

This research has shown that in vitro digestion can provide a platform for linking food structure to digestion(Reference Hiolle, Lechevalier and Floury32,Reference Bornhorst, Singh, Doyle and Klaenhammer33) . The use of in vitro simulations of digestion driven by human study data provides a powerful tool to improve the nutritional quality of food. Most recently, both MRI and in vitro simulation have been combined to non-invasively follow the influence of food structure on digestion(Reference Deng, Janssen and Vergeldt34Reference Deng, Seimys and Mars36). In all cases, it is the combination of a range of approaches linking human study data to physiologically relevant in vitro studies that has broadened understanding of the role of food structure in digestion kinetics.

Validation of simulations of digestion

It is clearly important for any model to be validated against data from the system that it is replicating. In the case of in vitro models of human digestion, this can often be problematic(Reference Dupont, Alric and Blanquet-Diot37). The rationale for using in vitro simulations of digestion is that they can provide direct information about the breakdown of food in the gastrointestinal tract that is often hard to gather in human studies. This can in turn improve interpretation of results from nutritional epidemiological studies that necessarily generate correlative outcomes. There are numerous examples of the use of in vitro digestion to determine the fate of specific macronutrients but much of the research focuses on protein and starch. The move to more plant-based diets has led to research on protein quality using the FAO-recommended digestible indispensable amino acid score system(Reference Rutherfurd, Fanning and Miller38,Reference Wolfe, Rutherfurd and Kim39) . The Infogest simulation of digestion had also been used to provide the source data for digestible indispensable amino acid score(Reference Sousa, Recio and Heimo20), with the benefit no animals are involved. However, the in vitro results were validated against results from experiments in vivo. The validation was made by comparison of seven different protein sources with data collected both in pigs and in human subjects(Reference Hodgkinson, Stroebinger and van der Wielen40). The results showed good agreement between the results in vivo and the Infogest in vitro results but as the authors note, the comparison was only made with a limited set of food sources, so no general conclusions about the efficacy of the Infogest approach can be made for assessing all protein quality. The digestion of protein has been investigated in relation to food allergy, in particular because it has been suggested that stability to digestion may be a key parameter for a protein to be an allergen(Reference Astwood, Leach and Fuchs41). However, the outcomes are highly dependent on the nature of the in vitro digestion model being used and its relevance to specific allergens(Reference Torcello-Gómez, Dupont and Jardin42,Reference Torcello-Gómez, Dupont and Jardin43) . In particular, the European Food Safety Authority currently uses a late phase gastric model with low pH and high protease activity but this is very different from an infant simulation(Reference Menard, Bourlieu and De Oliveira44,Reference Menard, Cattenoz and Guillemin45) . As a result, there have been calls for European Food Safety Authority to review the methodology for novel protein risk assessment(Reference Verhoeckx, Bogh and Dupont46). The examples given here are for the Infogest-recommended simulations but there are many others in the literature with increasing levels of sophistication and focusing on different bioactives(Reference Duque-Soto, Quintriqueo-Cid and Rueda-Robles47,Reference Faubel, Cilla and Alegria48) .

In addition to protein, there has been a lot of interest in understanding the role of processing and structure of starch in digestion relating to glycaemic response(Reference Pautong, Anonuevo and de Guzman49,Reference Fernandes, Madalena and Pinheiro50) . Although comparison can be made with in vivo data, as highlighted by Phillips et al. earlier(Reference Phillips, Deane and Jones11), there are many individual factors that can make such a comparison problematic, not least the fact that glucose in peripheral blood is not a good indicator of bioaccessible glucose in the gut(Reference Priyadarshini, Moses and Anandharamakrishnan51). Because starch is the most important digestible polysaccharide in human nutrition usually accounting for 20–50 % of the total energy intake, it has been studied extensively(Reference Bohn, Carriere and Day52). The apparent health benefits of a low-GI diet led the Carbohydrate Quality Consortium to state ‘an urgent need to communicate information on GI and GL’(Reference Augustin, Kendall and Jenkins53). Consequently, it is important to note that a number of studies have shown the validity of the in vitro determination of the GI(Reference Fernandes, Madalena and Pinheiro50,Reference Monro and Mishra54,Reference Argyri, Athanasatou and Bouga55) . The ability of resistant starch to pass through the upper gastrointestinal tract and into the colon has also highlighted the need for more sophisticated models of human digestion that include colonic fermentation.

What more is needed?

As the pressure to reduce animal experiments increases there a drive to find suitable models to replace them(Reference Langley, Evans and Holgate56,Reference Mak, Evaniew and Ghert57) . As a result, there has been a proliferation of in vitro models of digestion, many of which are more sophisticated in the way that they mimic the physical and biochemical environment of the gut(Reference Hur, Lim and Decker58). These include simulations of the gastric phase(Reference Kong and Singh59,Reference Wickham, Faulks and Mann60) where the biochemical and physical environment of the stomach are replicated, or the gastric and intestinal phases(Reference Minekus, Marteau and Havenaar61) or the gastric, intestinal and colonic phases of digestion(Reference Chaikham, Apichartsrangkoon and Jirarattanarangsri62). However, these simulations lack the final stages of digestion and any absorption steps. Although some attempts have been made to include brush border enzymes(Reference Di Stasio, Picascia and Auricchio63), these have not been widely accepted because difficulties in defining activity and exposure time.

There are many in vitro models focusing on just the colonic phase of digestion often based on the early work of MacFarlane et al.(Reference Macfarlane, Macfarlane and Gibson64). These are becoming increasingly important in enabling research to understand the role of dietary fibre and plant bioactives in relation to gut microbiota and the gut–brain axis(Reference Silva, Bernardi and Frozza65). However, the issue with investigating the role of colonic fermentation in the digestion of complex foods using in vitro approaches is the carry-over of compounds from the small intestinal phase to the large intestinal fermentation phase. Transporters in the apical membrane of enterocytes often specifically control absorption in the small intestine. Thus, both highly nutritive molecules such as simple sugars, amino acids and peptides as well as fatty acids are largely removed from intestinal chyme. Additionally, potentially toxic bile acids are reabsorbed in the distal ileum. These and similar absorption mechanisms are not simulated well by passive dialysis so improvements need to be found in presenting realistic digesta to simulations of colonic fermentation.

In silico models of digestion are also becoming more widely available(Reference Le Feunteun, Verkempinck and Floury66,Reference Del Rio, Van der Wielen and Gerrits67) but many of them are focused only on protein digestion using specific rules for modelling the action of proteases. However, more generally applicable in silico models of digestion will only become reliable when more human study data are made available to refine them(Reference Le Feunteun, Mackie and Dupont68). With the rise of machine-learning approaches, this is likely to become a more tractable approach in the future.

Conclusions

A number of non-communicable diseases have digestion kinetics as underlying risk factors. Thus, disease is undoubtedly related to dietary quality suggesting that consumers need to build a better relationship with their food. The evidence presented here shows that food structure can affect gastric-emptying rate and digestion kinetics. This confirms that food structure as well as composition is important in risk factors for non-communicable diseases. The link between food and digestion kinetics can be studied in more detail using in vitro simulations validated using human data. Such models can help provide preliminary data on the slower digesting, more functional foods that are needed to decrease the prevalence of non-communicable diseases.

Acknowledgements

The author acknowledges Dr Bernadette Moore for helpful discussion.

Financial Support

This article was funded by the University of Leeds and draws upon a number of studies undertaken by my team over a number of years. The funding of those studies is given in the articles cited.

Conflict of Interest

None.

Authorship

A. Mackie is the sole author of this article and no authors who would reasonably be considered an author have been excluded.

References

Levine, AS & Ubbink, J (2023) Ultra-processed foods: processing versus formulation. Obes Sci Pract 9, 436439.CrossRefGoogle ScholarPubMed
Drewnowski, A & Specter, SE (2004) Poverty and obesity: the role of energy density and energy costs. Am J Clin Nutr 79, 616.CrossRefGoogle ScholarPubMed
Jackson, P, Cameron, D, Rolfe, S et al. (2021) Healthy soil, healthy food, healthy people: an outline of the H3 project. Nutr Bull 46, 497505.CrossRefGoogle Scholar
Regan, A, Kent, MP, Raats, MM et al. (2017) Applying a consumer behavior lens to salt reduction initiatives. Nutrients 9, 901.CrossRefGoogle ScholarPubMed
Slavin, J (2013) Fiber and prebiotics: mechanisms and health benefits. Nutrients 5, 14171435.CrossRefGoogle ScholarPubMed
Jenkins, DJA, Dehghan, M, Mente, A et al. (2021) Glycemic index, glycemic load, and cardiovascular disease and mortality. N Engl J Med 384, 13121322.CrossRefGoogle ScholarPubMed
Pai, H & Gulliford, MC (2022) Body mass index trajectories and mortality in community-dwelling older adults: population-based cohort study. BMJ Open 12, e062893.CrossRefGoogle ScholarPubMed
Livesey, G, Taylor, R, Livesey, HF et al. (2019) Dietary glycemic index and load and the risk of type 2 diabetes: assessment of causal relations. Nutrients 11, 1436.CrossRefGoogle ScholarPubMed
Jenkins, DJ, Kendall, CW, Augustin, LS et al. (2002) Glycemic index: overview of implications in health and disease. Am J Clin Nutr 76, 266s273s.CrossRefGoogle ScholarPubMed
Goyal, RK, Guo, YM & Mashimo, H (2019) Advances in the physiology of gastric emptying. Neurogastroenterol Motil 31, e13546.CrossRefGoogle ScholarPubMed
Phillips, LK, Deane, AM, Jones, KL et al. (2015) Gastric emptying and glycaemia in health and diabetes mellitus. Nat Rev Endocrinol 11, 112128.CrossRefGoogle ScholarPubMed
Mackie, AR, Bajka, BH, Rigby, NM et al. (2017) Oatmeal particle size alters glycemic index but not as a function of gastric emptying rate. Am J Physiol – Gastrointest Liver Physiol 313, G239G246.CrossRefGoogle Scholar
Mackie, AR, Rafiee, H, Malcolm, P et al. (2013) Specific food structures suppress appetite through reduced gastric emptying rate. Am J Physiol – Gastrointest Liver Physiol 304, G1038G1043.CrossRefGoogle Scholar
Hornby, H, Collado-González, M, Zhang, X et al. (2021) Size and number of food boluses in the stomach after eating different meals: magnetic resonance imaging insights in healthy humans. Nutrients 13, 3636.CrossRefGoogle ScholarPubMed
Camps, G, Mars, M, de Graaf, C et al. (2016) Empty calories and phantom fullness: a randomized trial studying the relative effects of energy density and viscosity on gastric emptying determined by MRI and satiety. Am J Clin Nutr 104, 7380.CrossRefGoogle ScholarPubMed
Dupont, D, Bordoni, A, Brodkorb, A et al. (2011) An international network for improving health properties of food by sharing our knowledge on the digestive process. Food Dig 2, 2325.CrossRefGoogle Scholar
Minekus, M, Alminger, M, Alvito, P et al. (2014) A standardised static in vitro digestion method suitable for food – an international consensus. Food Funct 5, 11131124.CrossRefGoogle ScholarPubMed
Brodkorb, A, Egger, L, Alminger, M et al. (2019) INFOGEST static in vitro simulation of gastrointestinal food digestion. Nat Protoc 14, 9911014.CrossRefGoogle ScholarPubMed
Sams, L, Paume, J, Giallo, J et al. (2015) Relevant pH and lipase for in vitro models of gastric digestion. Food Funct 7, 3045.CrossRefGoogle Scholar
Sousa, R, Recio, I, Heimo, D et al. (2023) In vitro digestibility of dietary proteins and in vitro DIAAS analytical workflow based on the INFOGEST static protocol and its validation with in vivo data. Food Chem 404, 134720.CrossRefGoogle ScholarPubMed
Egger, L, Ménard, O, Delgado-Andrade, C et al. (2016) The harmonized INFOGEST in vitro digestion method: from knowledge to action. Food Res Int 88, 217225.CrossRefGoogle Scholar
Egger, L, Schlegel, P, Baumann, C et al. (2017) Physiological comparability of the harmonized INFOGEST in vitro digestion method to in vivo pig digestion. Food Res Int 102, 567574.CrossRefGoogle ScholarPubMed
Egger, L, Menard, O, Baumann, C et al. (2019) Digestion of milk proteins: comparing static and dynamic in vitro digestion systems with in vivo data. Food Res Int 118, 3239.CrossRefGoogle ScholarPubMed
Miralles, B, Sanchon, J, Sanchez-Rivera, L et al. (2021) Digestion of micellar casein in duodenum cannulated pigs. Correlation between in vitro simulated gastric digestion and in vivo data. Food Chem 343, 128424.CrossRefGoogle ScholarPubMed
Mulet-Cabero, AI, Egger, L, Portmann, R et al. (2020) A standardised semi-dynamic: in vitro digestion method suitable for food – an international consensus. Food Funct 11, 17021720.CrossRefGoogle ScholarPubMed
Li, YW & Kong, FB (2022) Simulating human gastrointestinal motility in dynamic in vitro models. Compr Rev Food Sci Food Saf 21, 38043833.CrossRefGoogle ScholarPubMed
Marciani, L, Faulks, R, Wickham, MSJ et al. (2009) Effect of intragastric acid stability of fat emulsions on gastric emptying, plasma lipid profile and postprandial satiety. Br J Nutr 101, 919928.CrossRefGoogle ScholarPubMed
Armand, M, Borel, P, Pasquier, B et al. (1996) Physicochemical characteristics of emulsions during fat digestion in human stomach and duodenum. Am J Physiol – Gastrointest Liver Physiol 271, G172G183.CrossRefGoogle ScholarPubMed
Mulet-Cabero, AI, Rigby, NM, Brodkorb, A et al. (2017) Dairy food structures influence the rates of nutrient digestion through different in vitro gastric behaviour. Food Hydrocolloids 67, 6373.CrossRefGoogle Scholar
Mulet-Cabero, AI, Mackie, AR, Wilde, PJ et al. (2019) Structural mechanism and kinetics of in vitro gastric digestion are affected by process-induced changes in bovine milk. Food Hydrocolloids 86, 172183.CrossRefGoogle Scholar
Ye, A, Cui, J, Dalgleish, D et al. (2017) Effect of homogenization and heat treatment on the behavior of protein and fat globules during gastric digestion of milk. J Dairy Sci 100, 3647.CrossRefGoogle Scholar
Hiolle, M, Lechevalier, V, Floury, J et al. (2020) In vitro digestion of complex foods: how microstructure influences food disintegration and micronutrient bioaccessibility. Food Res Int 128, 108817.CrossRefGoogle ScholarPubMed
Bornhorst, GM & Singh, RP (2014) Gastric digestion in vivo and in vitro: how the structural aspects of food influence the digestion process. In Annual Review of Food Science and Technology, vol. 5, pp. 111132 [Doyle, MP and Klaenhammer, TR, editors]. Palo Alto, CA: Annual Reviews.Google Scholar
Deng, RX, Janssen, AEM, Vergeldt, FJ et al. (2020) Exploring in vitro gastric digestion of whey protein by time-domain nuclear magnetic resonance and magnetic resonance imaging. Food Hydrocolloids 99, 105348.CrossRefGoogle Scholar
Smeets, PAM, Deng, RX, van Eijnatten, EJM et al. (2021) Monitoring food digestion with magnetic resonance techniques. Proc Nutr Soc 80, 148158.CrossRefGoogle ScholarPubMed
Deng, RX, Seimys, A, Mars, M et al. (2022) Monitoring pH and whey protein digestion by TD-NMR and MRI in a novel semi-dynamic in vitro gastric simulator (MR-GAS). Food Hydrocolloids 125, 107393.CrossRefGoogle Scholar
Dupont, D, Alric, M, Blanquet-Diot, S et al. (2019) Can dynamic in vitro digestion systems mimic the physiological reality? Crit Rev Food Sci Nutr 59, 15461562.CrossRefGoogle ScholarPubMed
Rutherfurd, SM, Fanning, AC, Miller, BJ et al. (2015) Protein digestibility-corrected amino acid scores and digestible indispensable amino acid scores differentially describe protein quality in growing male rats. J Nutr 145, 372379.CrossRefGoogle ScholarPubMed
Wolfe, RR, Rutherfurd, SM, Kim, IY et al. (2016) Protein quality as determined by the digestible indispensable amino acid score: evaluation of factors underlying the calculation. Nutr Rev 74, 584599.CrossRefGoogle ScholarPubMed
Hodgkinson, SM, Stroebinger, N, van der Wielen, N et al. (2022) Comparison of true ileal amino acid digestibility between adult humans and growing pigs. J Nutr 152, 16351646.CrossRefGoogle ScholarPubMed
Astwood, JD, Leach, JN & Fuchs, RL (1996) Stability of food allergens to digestion in vitro. Nat Biotechnol 14, 12691273.CrossRefGoogle ScholarPubMed
Torcello-Gómez, A, Dupont, D, Jardin, J et al. (2020) Human gastrointestinal conditions affect: in vitro digestibility of peanut and bread proteins. Food Funct 11, 69216932.CrossRefGoogle ScholarPubMed
Torcello-Gómez, A, Dupont, D, Jardin, J et al. (2020) The pattern of peptides released from dairy and egg proteins is highly dependent on the simulated digestion scenario. Food Funct 11, 52405256.CrossRefGoogle ScholarPubMed
Menard, O, Bourlieu, C, De Oliveira, SC et al. (2018) A first step towards a consensus static in vitro model for simulating full-term infant digestion. Food Chem 240, 338345.CrossRefGoogle ScholarPubMed
Menard, O, Cattenoz, T, Guillemin, H et al. (2014) Validation of a new in vitro dynamic system to simulate infant digestion. Food Chem 145, 10391045.CrossRefGoogle ScholarPubMed
Verhoeckx, K, Bogh, KL, Dupont, D et al. (2019) The relevance of a digestibility evaluation in the allergenicity risk assessment of novel proteins. Opinion of a joint initiative of COST action ImpARAS and COST action INFOGEST. Food Chem Toxicol 129, 405423.CrossRefGoogle ScholarPubMed
Duque-Soto, C, Quintriqueo-Cid, A, Rueda-Robles, A et al. (2023) Evaluation of different advanced approaches to simulation of dynamic in vitro digestion of polyphenols from different food matrices – a systematic review. Antioxidants 12, 101.CrossRefGoogle Scholar
Faubel, N, Cilla, A, Alegria, A et al. (2022) Overview of in vitro digestion methods to evaluate bioaccessibility of lipophilic compounds in foods. Food Rev Int, 122.Google Scholar
Pautong, PA, Anonuevo, JJ, de Guzman, MK et al. (2022) Evaluation of in vitro digestion methods and starch structure components as determinants for predicting the glycemic index of rice. LWT-Food Sci Technol 168, 113929.CrossRefGoogle Scholar
Fernandes, JM, Madalena, DA, Pinheiro, AC et al. (2020) Rice in vitro digestion: application of INFOGEST harmonized protocol for glycemic index determination and starch morphological study. J Food Sci Technol 57, 13931404.CrossRefGoogle ScholarPubMed
Priyadarshini, SR, Moses, JA & Anandharamakrishnan, C (2022) Determining the glycaemic responses of foods: conventional and emerging approaches. Nutr Res Rev 35, 127.CrossRefGoogle ScholarPubMed
Bohn, T, Carriere, F, Day, L et al. (2018) Correlation between in vitro and in vivo data on food digestion. What can we predict with static in vitro digestion models? Crit Rev Food Sci Nutr 58, 22392261.CrossRefGoogle ScholarPubMed
Augustin, LSA, Kendall, CWC, Jenkins, DJA et al. (2015) Glycemic index, glycemic load and glycemic response: an international scientific consensus summit from the international carbohydrate quality consortium (ICQC). Nutr Metab Cardiovasc Dis 25, 795815.CrossRefGoogle ScholarPubMed
Monro, JA & Mishra, S (2010) Glycemic impact as a property of foods is accurately measured by an available carbohydrate method that mimics the glycemic response. J Nutr 140, 13281334.CrossRefGoogle ScholarPubMed
Argyri, K, Athanasatou, A, Bouga, M et al. (2016) The potential of an in vitro digestion method for predicting glycemic response of foods and meals. Nutrients 8, 209.CrossRefGoogle Scholar
Langley, G, Evans, T, Holgate, ST et al. (2007) Replacing animal experiments: choices, chances and challenges. Bioessays 29, 918926.CrossRefGoogle ScholarPubMed
Mak, IWY, Evaniew, N & Ghert, M (2014) Lost in translation: animal models and clinical trials in cancer treatment. Am J Transl Res 6, 114118.Google ScholarPubMed
Hur, SJ, Lim, BO, Decker, EA et al. (2011) In vitro human digestion models for food applications. Food Chem 125, 112.CrossRefGoogle Scholar
Kong, FB & Singh, RP (2010) A human gastric simulator (HGS) to study food digestion in human stomach. J Food Sci 75, E627E635.CrossRefGoogle ScholarPubMed
Wickham, MJS, Faulks, RM, Mann, J et al. (2012) The design, operation, and application of a dynamic gastric model. Dissolution Technol 19, 1522.CrossRefGoogle Scholar
Minekus, 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
Chaikham, P, Apichartsrangkoon, A, Jirarattanarangsri, W et al. (2012) Influence of encapsulated probiotics combined with pressurized longan juice on colon microflora and their metabolic activities on the exposure to simulated dynamic gastrointestinal tract. Food Res Int 49, 133142.CrossRefGoogle Scholar
Di Stasio, L, Picascia, S, Auricchio, R et al. (2020) Comparative analysis of in vitro digestibility and immunogenicity of gliadin proteins from durum and einkorn wheat. Front Nutr 7, 56.CrossRefGoogle ScholarPubMed
Macfarlane, GT, Macfarlane, S & Gibson, GR (1998) Validation of a three-stage compound continuous culture system for investigating the effect of retention time on the ecology and metabolism of bacteria in the human colon. Microb Ecol 35, 180187.CrossRefGoogle ScholarPubMed
Silva, YP, Bernardi, A & Frozza, RL (2020) The role of short-chain fatty acids from gut microbiota in gut-brain communication. Front Endocrinol 11, 25.CrossRefGoogle ScholarPubMed
Le Feunteun, S, Verkempinck, S, Floury, J et al. (2021) Mathematical modelling of food hydrolysis during in vitro digestion: from single nutrient to complex foods in static and dynamic conditions. Trends Food Sci Technol 116, 870883.CrossRefGoogle Scholar
Del Rio, AR, Van der Wielen, N, Gerrits, WJJ et al. (2022) In silico modelling of protein digestion: a case study on solid/liquid and blended meals. Food Res Int 157, 111271.CrossRefGoogle Scholar
Le Feunteun, S, Mackie, AR & Dupont, D (2020) In silico trials of food digestion and absorption: how far are we? Curr Opin Food Sci 31, 121125.CrossRefGoogle Scholar
Figure 0

Fig. 1. Axial FIESTA (Fast Imaging Employing Steady-state Acquisition) MRI images of the stomach (outlined) taken 5 min (A) and 25 min (B) post-consumption of oat flake porridge. Image (A) shows a layer above the oat flake porridge that is not apparent after 25 min (B). Axial TrueFISP (True Fast Imaging with steady state precession) MRI images of the stomach (outlined) taken 5 min after consumption of a semi-solid (C) or liquid (D) version of the same meal.