Nutrients and intestinal adaptation

Alan B.R. Thomson, MD, PhD
Monika Keelan, PhD
Gary E. Wild, MD, PhD

Clin Invest Med 1996; 19 (5): 331-45.

[résumé]


The authors are members of the Cell and Molecular Biology Collaborative Network in Gastrointestinal Physiology. Drs. Thomson and Keelan are with the Nutrition and Metabolism Research Group, Division of Gastroenterology, University of Alberta, Edmonton, Alta., and Dr. Wild is with the Department of Medicine, Division of Gastroenterology, and the Department of Anatomy and Cell Biology, McGill University, Montreal, Que.

Reprint requests to: Dr. Alan B.R. Thomson, Division of Gastroenterology, University of Alberta, 519 Robert Newton Research Bldg., Edmonton AB T6G 2C2; tel 403 492-6490, fax 403 492-7964; alan.thomson@ualberta.ca


Contents


Abstract

The authors review the physiological, cellular and molecular aspects of the patterns, mechanisms and signals of the adaptation of intestinal transport of sugars and lipids, especially in response to manipulations of dietary lipid content. In models of intestinal adaptation, nutrient uptake is enhanced by an up- or down-regulation of the maximal rate of carrier-mediated transport or by alterations in the passive permeability properties (Pd) of the intestinal brush border membrane (BBM). The importance of unstirred water layers has been demonstrated. Alterations in the Pd for lipid uptake are due to changes in the lipid content of the BBM, which in turn are associated with alterations in the activity of lipid-metabolizing enzymes in the enterocyte microsomal membrane (EMM), and, therefore, alterations in the lipid composition of the EMM. Lipid uptake is also mediated by at least two proteins in the BBM, the sodium­hydrogen exchangers and the membrane-fatty-acid-binding protein. Alterations in the maximal transport rate for glucose and fructose transporters are associated with variations in the abundance of their transporters (including sodium-dependent glucose transporter, glucose and fructose transporter and fructose transporter) in the basolateral membrane sodium­potassium adenosine triphosphatase, and in the abundance of the messenger RNA of the transporters. Isocaloric changes in dietary lipids, such as switching from a saturated to a polyunsaturated diet, within the range seen in human consumption, leads to major alterations in passive and active transport processes. In a proposed model, changes in dietary lipids stimulate intracellular second messengers, modifying gene expression of the transporter carriers and of the EMM lipid-metabolizing enzymes. Thus, an understanding of the mechanisms of intestinal adaptation lays the groundwork for future studies of dietary manipulations. It may also lead to dietary interventions to prevent unwanted or to enhance desirable intestinal adaptation, thereby preventing disease.


Résumé

Les auteurs passent en revue les aspects physiologiques, cellulaires et moléculaires de l'adaptation intestinale du transport des sucres et des lipides en réponse aux manipulations des lipides dans la diète. Dans les modèles d'adaptation intestinale, la captation des nutriments est accrue soit par modulation du taux maximal du transport médié par vecteur, soit par des modifications dans les propriétés de perméabilité passive de la membrane de la bordure en brosse intestinale (MBB). L'importance des couches d'eau non agitées a été démontrée. Les modifications dans les propriétés de perméabilité passive pour la captation des lipides sont dues à des changements dans le contenu lipidique de la MBB. Ces derniers sont associés à des modifications dans l'activité des enzymes métabolisant les lipides au niveau de la membrane microsomiale des entérocytes (MME), et sont donc associés à des changements dans la composition lipidique de la MME. De plus, la captation des lipides est médiée par au moins deux protéines de la MBB : les protéines échangeuses de sodium et d'hydrogène et la protéine liant les acides gras membranaires. Des modifications dans le taux maximal de transport des transporteurs du glucose et du fructose sont associées à des variations dans l'abondance de ces transporteurs (y compris le transporteur du glucose dépendant du sodium, le transporteur du glucose et du fructose, et le transporteur du fructose) au niveau de l'adénosine triphosphatase échangeuse de sodium et de potassium dans la membrane basolatérale. Ces modifications sont également associées à une variation dans l'abondance de l'ARNm de ces transporteurs. Les changements iso-caloriques dans les lipides de la diète, tels que le passage d'acides gras saturés à une diète en acides gras poly-insaturés, d'une ampleur comparable à la consommation humaine, conduit à des changements majeurs dans le transport passif et actif. Il est proposé que les changements dans les lipides de la diète induisent des seconds messagers intracellulaires induisant une modification de l'expression génique des vecteurs des transporteurs de même que des enzymes métabolisant les lipides au niveau de la MME. La compréhension des mécanismes de l'adaptation intestinale prépare donc le terrain à l'étude future des manipulations diététiques; de plus, elle conduit au concept d'interventions sur la diète dans le but de prévenir des maladies, soit en empêchant une adaptation intestinale indésirable, ou par l'accroissement d'une adaptation souhaitable.

[Table of contents]


Introduction

Intestinal adaptation occurs in diverse physiological and pathological states and is associated with a characteristic array of changes in structure and function in response to a new set of nutritional demands. These alterations occur along the length of the villus and the intestine. They may also occur during the lifetime of the animal. The changes in intestinal morphology as well as digestive and transport functions may be physiological, occurring with age or in response to dietary changes.[1] They may also result from pathological processes, such as diabetes mellitus, short bowel syndrome, abdominal irradiation or long-term ethanol intake. Although adaptation is sometimes potentially detrimental (e.g., in diabetes mellitus), there are instances in which the adaptive responses are beneficial (e.g., after resection of the small bowel). In this case, the adaptation may enhance the survival potential of the animal, improving nutrition and reducing secretory and malabsorptive losses.

A growing body of experimental evidence assigns a central role to the presence of food in the intestinal lumen as an important mediator of intestinal adaptation.[2­5] The purpose and function of the regulation of nutrient transport adaptation is to maintain a safety margin among nutrient intake, uptake and requirements:[6,7] up-regulation of transporter activity maintains absorptive capacity at levels modestly above intake, whereas down-regulation saves biosynthetic energy by eliminating unneeded transport capacity. It may be possible to change nutrient absorption through dietary manipulation to achieve a desired therapeutic goal, for example, by varying macronutrient composition or by adding such "gut fuels" as glutamine or short-chain fatty acids.

The patterns of adaptation have traditionally been studied primarily from a physiological and morphological perspective.[8,9] Over the past decade, the powerful tools of recombinant DNA technology have been used to characterize the structure and function of a diverse assortment of transport proteins. This approach has begun to provide us with an unprecedented view of how intestinal-nutrient transporters work and with a framework to define the cellular processes that occur during intestinal adaptation. This review highlights the evolution that has revolutionized our understanding of the effects of luminal nutrients on the mechanisms and signals of the adaptation of intestinal transport of sugars and lipids.

[Table of contents]

Structural and physiological considerations

The small-bowel mucosa exhibits a tremendous capacity to adapt to changes in luminal demand. Mucosal hypoplasia develops after starvation or feeding by parenteral nutrition,[10] whereas hyperplasia develops in the distal intestinal remnant after partial resection of the proximal small bowel.[11] Furthermore, small-intestinal mucosal hyperplasia has been observed during lactation,[12] in streptozocin-induced diabetes mellitus,[13] and after surgical diversion of pancreaticobiliary secretions to the distal ileum.[14]

The structural changes that occur in models of intestinal adaptation are accompanied by adaptation of transport processes. For example, an earlier study[15] showed that the mucosal hyperplasia that occurs after partial small-bowel resection is accompanied by increased absorption of nutrients, water and electrolytes in the intestinal remnant. In studies of carrier-mediated transport, increased nutrient uptake usually results from an increase in the maximal transport rate (Vmax), which may reflect increased carrier activity per enterocyte or more transporting enterocytes. The Michaelis affinity constant (Km) is not usually altered in models of adaptation. The measurement of these kinetic parameters is influenced, however, by the presence of an intestinal unstirred-water layer (UWL), and appropriate corrections must be made to obtain valid estimates of these constants.[16,17]

A genetic regulatory program within the intestine provides developmental patterns of cell differentiation along the crypt-to-villus axis. The stem cells are encoded with a specific "positional address" that allows for appropriate spatial differentiation along the crypt-to-villus and the duodenal-to-colonic axes in the gut.[18,19] This reference point may be intrinsic to the stem cell or programmed by interactions between endodermal and stromal or mesenchymal components.

The dynamics of cell turnover are key determinants of the state of cellular differentiation and, consequently, of the number of phenotypically mature absorptive cells on the villi.[15] The crypt-cell production rate and the enterocyte migration rate are altered in various physiological and pathological states, and these changes in cell dynamics ultimately dictate where along the villus enterocytes of a specific age express transporter activity.[20] Thus, absorption may change as a result of an alteration in the number of transporting enterocytes and not solely as a result of variation in the total mucosal surface area.

Environmental factors such as diet play an important secondary modulatory role in determining nutrient-transporter gene expression. The factors responsible for the down- or up-regulation of transporter activity in the upper third of the villus are important to know: about 70% of the uptake of sugars, amino acids and lipids in rats occurs in the upper 30% of the villus, when enterocytes are about 30 hours old.[21,22] Since the messenger RNA (mRNA) levels of, for example, sodium-dependent glucose transporter (SGLT1) do not necessarily parallel this variation in transporter activity,[23] there may be positional factors influencing this apparent post-transcriptional control. Furthermore, the elegant studies of Ferraris and Diamond[6,24,25] suggest that the crypt cells sense the signal for alteration in sugar transport after a change in the carbohydrate content of the diet. In regard to dietary lipid changes and the uptake of amino acids, the signalling process does not appear to influence the production rate of crypt cells or the position along the villus or age of the enterocyte when transport is modified as a result of a change in diet.[21]

The enterocyte transport pool (ETP) comprises the enterocytes whose brush border membrane (BBM) is actively engaged in nutrient transport.[26] The effect of an increase in mucosal surface area on nutrient uptake depends on whether there is a concomitant alteration in the size of the ETP. Recruitment of enterocytes that express SGLT1 in diabetic ileum has been described.[27] A small change in the size of the ETP may result in a large variation in nutrient transport, even in the absence of large changes in the mucosal surface area. It is noteworthy that few studies of adaptation have examined the profile of transporter expression and activity or the changes in the BBM microenvironment along the crypt-to-villus axis. In addition, the denominator used in measuring transport capacity is important: an absorptive mechanism that shows an overall decline when expressed per unit of weight of intestine may display a modest increase when expressed per unit of length or of mass of mucosa, or even a dramatic increase when expressed as total absorptive capacity.[28]

[Table of contents]

Physiological, cellular and molecular aspects of nutrient uptake

The major roles of the small-intestinal epithelium are providing vectorial transport (e.g., absorption and secretion) and acting as a barrier.[29] Sugar and lipid absorption have been covered in earlier reviews.[28­30] Although the paracellular route contributes to sugar transport, the physiological significance of this transport pathway has been challenged.[31,32] Novel expression cloning has led to the molecular characterization of a variety of nutrient transporters.[33,34] SGLT1 was the first nutrient transporter to be cloned by this approach. The SGLT1 complementary DNA (cDNA) obtained by expression cloning encodes a 662-residue protein (74kD) which is a member of the family of sodium-coupled organic solute transporters.[33] The junctional complexes between neighbouring enterocytes and other structural and functional features associated with passive solute movement through the paracellular pathway have been reviewed.[28]

The current model of intestinal hexose transport is as follows. SGLT1 and the sodium-independent fructose carrier (GLUT5) are located in the BBM of the enterocyte.[34­37] These sugars are then transported across the basolateral membrane (BLM) by the facilitative hexose transporter (GLUT2).[38] The sodium­potassium adenosine triphosphatase (ATPase) in the BLM establishes the transcellular sodium gradient required for the BBM sodium-dependent nutrient cotransporters such as SGLT1.[39]

Molecular studies have demonstrated two patterns of adaptation: in one, changes in the activity of BBM digestive enzymes or transporter proteins are paralleled by alterations in the corresponding level of mRNA; in the other, there is discordance between protein and mRNA. SGLT1 expression is greater in the more proximal regions of the small intestine and parallels the proximal-to-distal gradient in glucose absorption.[40,41] In most species, SGLT1 mRNA levels are most abundant in the differentiated midvillus cells, SGLT1 protein appears to be uniformly distributed along the villus, and only low levels of SGLT1 expression are detected in the BBM of crypt enterocytes and at the villus tips.[40,42] An exception appears to be the small intestine of rabbits, in which there are increases in SGLT1 mRNA abundance and immunodetectable protein levels as cells migrate toward the villus tips.[43] It is noteworthy that Smith, Turvey and Freeman[44] demonstrated increased SGLT1 activity along the length of the villus accompanied by relatively constant levels of SGLT1 mRNA. Freeman and colleagues[39] also observed high levels of SGLT1 mRNA at the crypt-villus junction. Taken together, these findings suggest that the discrepancy between SGLT1 expression and functional activity reflects a post-transcriptional factor or factors that influence the transporter.

Both GLUT5 and GLUT2 are expressed predominantly in differentiated villus cells in the proximal regions of the small intestine.[45,46] The functional expression of GLUT2 and GLUT5 have been defined in the oocyte system of Xenopus species.[37] The GLUT2 cDNA encodes a 524-residue protein (61kD) with no homology to SGLT1. The GLUT5 clone codes for a 501-residue protein (50kD) with 41% amino-acid homology with GLUT2.

In preruminant lambs, BBM glucose transport is highest after birth, and there is a greater than 100-fold decline in glucose transport as lambs are weaned from a milk to a grass diet.[47] The developmental changes in BBM glucose-transport rate in the postnatal intestine of lambs correlate with alterations in BBM SGLT1 protein but not with SGLT1 mRNA.[40,48] These findings suggest that the principal level of SGLT1 regulation is translational or post-translational. In postnatal rats, increases in glucose and fructose uptake have been observed,[7] and increases in hexose uptake in rats after birth are paralleled by similar increases in SGLT1, GLUT2 and GLUT5 expression.[49,50] Thus, the relative importance of transcriptional and post-transcriptional events in determining the expression of these nutrient transporters remains unclear.

In the developmental pattern of sucrase-isomaltase (SI) expression in postnatal rats, the regulation of SI expression appears to be transcriptional.[51­56] SI and amino-oligopeptidase are two examples of concordant patterns: studies in rats,[51­53] rabbits[54] and humans[55,56] demonstrate that the increase in SI enzyme activity is accompanied by a parallel increase in steady-state levels of SI mRNA, suggesting regulation at a pretranslational level. There are also examples of discordant patterns: in early life, a rise in lactase activity is accompanied by an increase in lactase mRNA, implying transcriptional control, although there may also be a post-transcriptional component.[57] Indeed, in humans there appears to be mosaic expression of lactase protein,[58] and in adult rats lactase mRNA is confined to the lower halves of the villi, whereas lactase enzyme is present along the full length of the villi.[59]

Lipid absorption has been reviewed.[60­62] The uptake of lipids is thought to occur largely through a process of passive diffusion. Cholesterol absorption is a passive, diffusionally-mediated process from mixed micelles. Cholesterol exists in the BBM in several kinetically distinct pools,[63] and there is evidence that cytosolic cholesterol is metabolically compartmentalized.[1] Intracellular cholesterol is derived from the lumen and from de novo synthesis; as well, approximately 10% of circulating low-density lipoprotein is catabolized by the enterocyte after BLM uptake, and most of this uptake is mediated by a receptor-dependent mechanism.[64] Distinct BBM-associated proteins may be involved: second-order reaction kinetics and saturation have been demonstrated with the use of BBM vesicles, and a 14-kDa protein has been demonstrated by SDS-Page.[65] Pancreatic cholesterol esterase in the BBM and enterocyte cytosol also plays a role in the uptake of cholesterol,[66] as does acylcoenzyme A cholesterol acyltransferase.[1,67,68]

The passive permeation properties of the intestine are influenced by BBM fluidity and variations in the BBM lipid composition. There may be a dissociation between the BBM lipid composition and fluidity, especially if the fluidity of the inner and outer leaflet of the BBM is not measured separately.[69,70] BBM lipids are synthesized in the enterocyte microsomal membranes (EMM), and the EMM traffic to the BBM, where they are inserted. BBM lipid permeation is influenced, at least in part, by the activity of EMM lipid-metabolizing enzymes, with alterations in EMM lipids being reflected, at least in part, by changes in BBM lipids. Because failure to correct for UWL resistance may lead to an underestimate of the true permeability properties of the BBM,[16] correlation between changes in BBM fluidity or lipid composition and lipid permeation must be attempted only after appropriate corrections for the effect of this diffusion barrier. In addition, there may be protein-mediated components of lipid uptake,[71,72] and these must be taken into account when correlating changes in lipid uptake and BBM lipid composition. Finally, because lipid uptake is greater in the upper than in the lower portion of the villus,[22] correlations between EMM and BBM lipids, and between BBM lipids and lipid uptake, must be made in enterocytes isolated from similar positions along the villus, rather than from homogenates of enterocytes isolated from all portions of the villus.

A fatty-acid-binding protein (FABPpm) has also been identified in BBM; it influences uptake of long-chain fatty acids and of cholesterol in isolated intestinal epithelial cells, with in vivo perfused segments of rat jejunum, with in vitro sheets of intestine and in BBM vesicles.[71,72] The availability of monoclonal antiserum and partial peptide sequence should facilitate the cloning of the FABPpm gene as well as the demonstration of the molecular basis for its control and potential for adaptation and therefore modification of lipid uptake.[73]

An amiloride-inhibitable uptake step for fatty acids across the BBM has been reported.[71] This step is presumably due to the presence of a BBM sodium­hydrogen exchanger (NHE).[74] Depending on the presence or absence of a combined sodium and hydrogen gradient across the BBM, one or another of these protein-mediated steps may be important in lipid uptake (Fig. 1). It is not yet known whether the activity of NHE or the abundance of FABPpm varies when lipid uptake is modified in models of intestinal adaptation. If this activity or abundance does vary, and if these proteins are heterogeneously distributed along the villus, then such findings as the increased lipid uptake in diabetes mellitus would need to be interpreted on the basis not only of known alterations in BBM lipid composition[75,76] but also of alterations in the activity or abundance of these proteins and their corresponding mRNA.

[Table of contents]

Importance of diet in modifying nutrient absorption

With increasing intake of energy-generating nutrients (such as glucose and nonessential amino acids), intestinal uptake generally increases. In contrast, there is a reciprocal relation between intake and uptake of essential nutrients or toxic substances. For example, when the intake of iron is low and iron deficiency develops, intestinal uptake of iron will increase.[77] Switching from a low- to a high-carbohydrate diet results in alterations in the crypt cells, and, after a lag period of 2 to 3 days (representing the enterocyte migration rate), glucose uptake will increase.[2,3,25] These carbohydrate-associated alterations in transport do not require metabolism of the sugar.

Diet plays an important role in the regulation of intestinal transport in early[78] as well as later life.[79,80] Although the intestine responds to alterations in the nutrients in the lumen, nutrient absorption is generally considered to be genetically programmed or "hard-wired," and early changes in the composition of the animal's diet may have lasting effects in later life.[76,79,80] Changing the lipids in the mother's diet modifies the absorption characteristics of the intestine of suckling rats.[81] Diet plays a role in determining total and specific maximal activity levels; it also plays a modulating role in the maturation of the longitudinal pattern of, for example, expression of the lactase gene.[82] However, there is less dietary regulation of intestinal nutrient absorption in aging animals than in younger animals.[83] The mechanism of impaired adaptation to diet in older animals is unknown.

In adult rats, intestinal nutrient transporters are adaptively regulated by the dietary level of their substrates.[6] Some studies have examined the differential responses of intestinal hexose transporters to levels of dietary sugars. For example, a high-glucose diet stimulates glucose transport activity and increases the levels of SGLT1 and GLUT2 expression.[84­86] By contrast, GLUT5 expression is increased only by fructose and galactose, and fructose also increases GLUT2 expression.[84,85] The increases in GLUT5 expression after eating fructose appear to be due primarily to post-transcriptional mechanisms.[86] In the case of amino acids, consumption of a high-protein diet increases the Vmax for the uptake of amino acids, with relatively greater changes in the Vmax for the uptake of essential than of nonessential amino acids. Aspartate, valine, lysine and glutamine induce their own transport.[86-89] However, amino acids do not necessarily regulate their own transport. For example, the dipolar amino acids glutamine and valine induce the transport of acidic amino acids;[86] the acidic amino acid aspartate is the best inducer of cationic amino-acid transport, and the cationic amino acid arginine (but not lysine) is a good inducer of acidic amino-acid transport.[88]

Changing the quality of dietary lipids by substituting one type of lipid for another has demonstrated that intestinal transport is modified by the ratio of saturated to polyunsaturated fatty acids, the ratio of n6 to n3 fatty acids and the presence of cholesterol in the diet.[76,77,90] Intake of isocaloric diets supplemented with saturated, rather than polyunsaturated, fatty acids up-regulates the intestinal absorption of sugars and lipids,[91] and the nature and extent of this effect is influenced by the type of lipid in the diet.[92,93] This phenomenon of the dietary lipid modulation of intestinal transport may be used to achieve a desired therapeutic aim, such as enhancing nutrient uptake in the remaining intestine after small-bowel resection, preventing malabsorption of nutrients caused by long-term ethanol ingestion or abdominal irradiation, or reducing the undesired hyperabsorption of nutrients in diabetes mellitus (and thereby diminishing the associated hyperlipidemia and normalizing the serum concentration of hemoglobin A1C).[94­96] Furthermore, dietary lipid changes modify the ontogeny of the intestine. In this regard, a "critical period" phenomenon and a late effect of early nutrition have been shown;[62,79,80,97,98] it is important to note that modifying the lipid content of a nursing dam's diet alters nutrient transport in her suckling offspring.[81]

Variations in the type of lipids in the diet influence the abundance of SGLT1 and its mRNA,[99] which implies an effect on genetic expression. Absorbed fatty acids also act as substrates for the intracellular metabolism of lipids, and the modification of lipid metabolic-enzyme activities through diet alters the composition of the EMMs.[90] These EMM lipids traffic to the BBM and to the BLM. It is believed that these lipids are modified between the times they are inserted into the EMM and into the BBM, since the diet-associated changes in EMM lipids are not reflected exactly by the changes in the BBM lipids (Monika Keelan and associates: unpublished observations, 1996). In other tissues, deacylation­reacylation enzymes modify the composition of membrane phospholipid fatty acids, but it is not known whether this is the case in the intestine. In addition, phosphatidylethanolamine methyl transferase is present in the BBM[69,92] and may influence postmicrosomal modification of BBM lipids.

Before the importance of the role of FABPpm and NHE in the BBM was appreciated,[71,72] it was assumed that any alteration in lipid uptake was due to a modification in the lipophilic properties in the BBM, reflected by variations in BBM lipids. However, an important portion of lipid uptake (about 40%, depending upon the experimental conditions) may be influenced by NHE or FABPpm,[71,72] and lipid uptake varies along the length of the villus.[22] Therefore, attempts to link alterations in the BBM lipids with lipid uptake must involve collection of enterocytes from along the villus and determination of the proportion of total lipid uptake that is passive and can therefore be correlated with BBM lipid content.

The dietary lipid-associated changes in sugar uptake are due to alterations in the Vmax of the sugar carriers[79,100­102] as well as to altered expression of the carriers and their respective mRNA.[99] In diabetes mellitus, the adaptation of sugar transport is achieved by both transcriptional and post-transcriptional means.[42] The signalling of changes in dietary carbohydrates is thought to be read in the intestinal crypt cells.[2,3,25] It is therefore important to establish whether the dietary lipid-associated alterations in nutrient uptake occur in just the crypt cells or also in enterocytes from along the length of the villus.

Most uptake of sugars and amino acids occurs from the upper quarter to the upper third of the villus.[21,22] Therefore, attempts to demonstrate concordance or disconcordance among transporter activity, expression of protein, and mRNA levels of abundance must correlate these components in enterocytes taken from different positions along the villus. Thus, the signals for genetic regulation and post-translational control of carrier activity must also be searched for in the corresponding enterocyte fraction. Furthermore, lipid uptake is greater from the upper than from the lower portions of the villus.[22] Hence, the same villus segment must be used in attempts to explain adaptation of lipid uptake achieved through a modification in the lipophilic properties of the BBM. Similarly, studies to explain BBM lipid changes as a function of alterations in the activities of lipid metabolic enzymes in the BBM itself must be performed in the appropriate villus segment. Although isocaloric changes in diet do not modify BBM cholesterol or phospholipids,[102] it is not known whether these changes alter the BBM phospholipid fatty acids or free fatty acids, and there have been no studies that have specifically examined the transporting enterocytes in the upper third of the villus.

[Table of contents]

Possible signals of diet-associated nutrient adaptation

The process or processes by which the stimulus delivered at the cell membrane by dietary changes results in gene transcription likely involve the phosphorylation and dephosphorylation of a variety of enzymes involved in intracellular signalling.[103,104] Phosphorylation of protein kinases occurs at serine, threonine or tyrosine residues. Protein kinase A (PKA) is activated by the catalytic subunit of adenylate cyclase. Signals that increase intracellular cyclic 3',5'-adenosine monophosphate (cAMP) activate cAMP.[105,106] Protein kinase C (PKC) is activated by calcium released from intracellular stores and by the phospholipid diacylglycerol.[107] Phospholipase C catalyzes the hydrolysis of phosphatidylinositol to diacylglycerol. PKC is activated by the increases in diacylglycerol generated during hydrolysis of phosphatidylinositol 4,5-biphosphate. The removal of phosphate groups by sequence-specific phosphatases is an additional mechanism by which the transcriptional activity of DNA-binding proteins may be altered.[104]

In addition to functioning as metabolic substrates and constituents of complex lipids, including membrane lipids, long-chain fatty acids have been recognized as mediators in signal-transduction pathways.[108] For example, arachidonic acid is a second messenger of certain G-protein-mediated cellular receptors, and palmitic and oleic acids (important dietary lipids) regulate gene expression at the transcriptional level.[109,110] The fatty-acid-induced effects are reversible upon removal of the effector and do not require metabolism of the fatty acid.[111] In addition, long-chain fatty acids act as signals in the induction of liver, fatty-acid-binding protein, peroxisomal b-oxidation and cytochrome P450 4A1.[112,113] This modulation of gene expression by fatty acids may be mediated by ligand-activated transcription factors.[114] Extracelluar and intracellular diet-induced regulation of gene expression in the pancreas has been reviewed.[115] It is unclear how these second messengers alter gene expression, but it is reasonable to speculate that dietary lipids influence the expression of transporter proteins and the enzymes responsible for lipid metabolism by altering some of the second messengers.

At the cellular level, both the direct and indirect mitogenic effects of nutrients appear to be mediated by the induction of ornithine decarboxylase (ODC) activity.[116] The importance of ODC and the polyamines is underscored by the finding that their levels increase with the cell proliferation that occurs in a variety of models of intestinal adaptation.[117] ODC is the key enzyme in the biosynthesis of the polyamines, which bind to nucleic acids, modify growth-regulating genes and stabilize membranes.[118] Polyamines regulate membrane-based enzymes, transport of ions and metabolites, calcium homeostasis, polyphosphoinositide metabolism, PKC, phospholipids and membrane fusion. As well, polyamines regulate growth-related genes such as c-myc, c-fos and H2A.[119,120]

Important growth factors in the gastrointestinal tract include members of the following families: epidermal growth factor, transforming growth factor b, insulin-like growth factor and fibroblast growth factor.[121,122] Additional families of peptides, generally classified on the basis of non-growth-related activities, including the cytokines and colony-stimulating factors, may also play a role in modifying the functional activities of the BBM. Many of these growth factors have been shown to stimulate ODC activity, and their mitogenic effects may be mediated through the actions of polyamines.[123]

Most receptors for peptide growth factors in the gastrointestinal tract possess intrinsic tyrosine kinase activity related to their intracellular domains.[122,124] Activation of this activity within the intracellular domain of the membrane receptor leads to phosphorylation of a wide variety of substrates, many of which are phosphokinases, including MAP and Raf kinases.[124] Growth factors modulate cellular proliferation and phenotype through the alteration of various genes.[125] Growth factor receptors may be coupled with a variety of secondary intracellular-messenger signalling systems, including adenyl cyclase-generated cAMP, G-coupled cyclic guanine monophosphate (cGMP), the products of the phospholipase C-phosphatidylinositol pathways (including diacylglycerol and phosphinositides) closely integrated with intracellular calcium, and the products of PKC.[124,126,127] Alteration in the expression of the nuclear transcriptional factors (e.g., fos and myc) appears to be involved in the induction or suppression of transcription of many genes required for cellular growth and proliferation.[124] This induction or suppression may play a central role in the adaptation of nutrient transport.

Transcriptional control elements have a critical function in modulating the expression of various genes, but the mechanisms of their action are not yet defined.[128] All genes are not transcribed concurrently, indicating that the cell has mechanisms for "silencing" some genes in response to extracelluar stimuli. The mechanisms for repressing genes may be general or sequence-specific[129] and may allow certain cellular functions to proceed unchecked. Thus, positive and negative regulators may interact, as they do during cellular proliferation and differentiation.[130] Many of the mechanisms in which transcription factors are transactivated involve protein dephosphorylation and phosphorylation by protein kinases.[104] The precise mechanisms of transcriptional activation are unknown. It appears that most cells respond to their environment by recruiting subsets of ubiquitous and promoter-specific transcription factors that synergistically produce the desired cellular phenotype.[128,131]

Post-translational processing of proteins, such as originally described with proinsulin, may occur.[132] After completion of protein synthesis in the endoplasmic reticulum, glycosylation, phosphorylation and sulfation may occur in the Golgi apparatus.[133] Transport vesicles have an outer coat of protein (coatamer). The fungal metabolite brefeldin A (BFA) blocks protein transport through the Golgi stack as well as binding of the coatamer complex to budding Golgi membranes.[134,135] As an example, BFA interferes with the post-translational processing of progastrin, which suggests that this pathway is involved in the sorting of prohormones as well as other soluble secretory proteins.[136] The primary amino-acid sequence of SGLT1 contains the crucial information required to target this transporter to the apical membrane domain.[118] GLUT2 cell-surface expression and intracellular transport have been examined in pancreatic b cells.[137] The data obtained from these studies demonstrate that, at least in this cell type, GLUT2 surface expression occurs via the constitutive pathway of protein secretion,[133] its transport can be blocked by BFA, and, once delivered to the cell surface, GLUT2 does not recycle in intracellular vesicles.[137] Whether BFA modifies the movement of transport proteins to the BBM or affects the trafficking of lipids destined for insertion into the BBM in the enterocyte remains to be defined.

Fatty acylation, usually with either myristic or palmitic acid, is a common mechanism for the post-translational modification of proteins[138] and may play a role in membrane targeting.[139] PKC is a candidate for the kinase-mediating PP1-linked receptor phosphorylation and desensitization. PKC is activated by the increases in diacylglycerol generated during the hydrolysis of phosphatidylinositol 4,5-biphosphate.[140] However, it is unknown whether the diet-induced changes in transcriptional and post-translational processing of transport proteins are mediated by that means.

[Table of contents]

A model of the mechanisms of dietary-lipid-associated intestinal adaptation

We propose a model of the control of intestinal adaptation in response to dietary lipid manipulations (Fig.  2). This model is also applicable to the testing of other examples of intestinal adaptation, such as that which occurs in diabetes mellitus or after bowel resection. In this model the following steps are proposed: (1) dietary lipids act as signals to release second messengers (such as ODC activity, intracellular calcium, phospholipase C, phosphyatidylinositol or PKC) that modify the genetic regulation of the expression of the mRNA for the BBM and BLM carrier proteins and modify the EMM lipid metabolic enzymes; (2) dietary lipids also act as substrate for the diet-associated changes in EMM lipid metabolic enzymes, which modify the EMM content of lipids available for trafficking to the BBM and the BLM; (3) the partial dissociation between diet-associated changes in EMM and in BBM and BLM lipids is due to postmicrosomal modification of EMM lipids by lipid metabolic enzymes in the BBM and by variations along the villus in the diet-associated changes in EMM and in BBM and BLM lipids; (4) the partial dissociation between diet-associated changes in lipid uptake and BBM lipids is due to the contribution of NHE and FABPpm to the total uptake of lipids and to the variable expression of those proteins along the villus; (5) there is post-transcriptional modulation of the sugar carriers in the Golgi apparatus as well as of the BBM and BLM composition of lipids; and (6) the diet-associated changes in sugar uptake necessarily modify the activity of BLM sodium­potassium ATPase, and the diet-associated alterations in lipid uptake modify the activity of NHE in the BBM, which further influences the required regulation of sodium­potassium ATPase.

[Table of contents]

Acknowledgements

The authors would like to thank Chandra Messier for typing the manuscript. The publication of this article was supported in part by a grant from the Fonds de la recherche en santé du Québec. We would like to thank Proctor & Gamble, Inc., for its sponsorship of the invited speakers' program.

[Table of contents]

References

  1. Suckling KE, Stange EF. Role of acyl-CoA: cholesterol acyltransferase in cellular cholesterol metabolism. J Lipid Res 1985; 26: 641-71.
  2. Ferraris RP, Villenas SA, Diamond J. Regulation of brush border enzyme activities and enterocyte migration rates in mouse small intestine. Am J Physiol 1992; 262: G1047-59.
  3. Ferraris RP, Villenas SA, Hirayama BA, Diamond J. Effect of diet on glucose transporter density along the intestinal crypt­villus axis. Am J Physiol 1992; 262: G1060-8.
  4. Thomson ABR, Keelan M. The development of the small intestine. Can J Physiol Pharmacol 1986; 64: 13-29.
  5. Thomson ABR, Keelan M. The aging gut. Can J Physiol Pharmacol 1986; 64: 30-8.
  6. Ferraris RP, Diamond JM. Specific regulation of intestinal nutrient transporters by their dietary substrates. Annu Rev Physiol 1989; 51: 125-41.
  7. Toloza EM, Diamond J. Ontogenic development of nutrient transporters in rat intestine. Am J Physiol 1992; 263: G593-604.
  8. Buddington RK, Chen JW, Diamond JM. Genetic and phenotypic adaptation of intestinal nutrient absorption to diet. Am J Physiol 1987; 393: 261-81.
  9. Buddington RK, Diamond JM. Ontogenetic development of intestinal nutrient transporters. Annu Rev Physiol 1989; 51: 601-19.
  10. Altmann GG. Influence of starvation and refeeding on mucosal size and epithelial renewal in the rat small intestine. Am J Anat 1972; 133: 391-400.
  11. Dowling RH, Booth CC. Structural and functional changes following small intestinal resection in the rat. Clin Sci 1967; 32: 132-49.
  12. Boyne R, Fell BJ, Robb I. The surface area of the intestinal mucosa in the lactating rat. J Physiol 1966; 183: 570-5.
  13. Miller DL, Hanson W, Schedl HP. Proliferation rate and transit time of mucosal cells in small intestine of the diabetic rat. Gastroenterology 1977; 73: 1326-32.
  14. Altmann GG. Influence of bile and pancreatic secretions on the size of the intestinal villi in the rat. Am J Anat 1971; 132: 167-77.
  15. Dowling RH. Small bowel adaptation and its regulation. Scand J Gastroenterol 1982; 74: 53-74.
  16. Westergaard H, Dietschy JM. Delineation of dimensions and permeability characteristics of the two major diffusion barriers to passive mucosal uptake in rabbit intestine. J Clin Invest 1974; 54: 718-32.
  17. Thomson ABR. Mechanisms of intestinal adaptation: unstirred layer resistance and membrane transport. Can J Physiol Pharmacol 1984; 62: 678-82.
  18. Hermiston ML, Simon TC, Crossman MW, Gordon JI. Model systems for studying cell fate specification and differentiation in the gut epithelium. In: Johnson LR, editor. Physiology of the gastrointestinal tract. 3rd ed. New York: Raven Press, 1994: 521-70.
  19. Rubin DC, Swietlicki E, Roth KA, Gordon JI. Use of fetal intestinal isografts from normal and transgenic mice to study the programming of positional information along the duodenal-to-colonic axis. J Biol Chem 1992; 267: 15122-33.
  20. Jenkins AP, Thompson RP. Mechanisms of small intestinal adaptation. Dig Dis 1994; 12: 15-27.
  21. Thomson ABR, Cheeseman CI, Keelan M, Fedorak R, Clandinin MT. Crypt cell production rate, enterocyte turnover time and appearance of transport along the jejunal villus of the rat. Biochim Biophys Acta 1994; 1191: 197-204.
  22. Fingerote RJ, Doring DA, Thomson ABR. Gradient for D-glucose and linoleic acid uptake along the crypt-villus axis of rabbit jejunal brush border membrane vesicles. Lipids 1994; 29: 117-27.
  23. Smith NW, Turvey A, Freeman TC. Appearance of phloridzin-sensitive glucose transport is not controlled at mRNA level in rat jejunal enterocytes. Exp Physiol 1992; 77: 525-8.
  24. Ferraris RP. Regulation of intestinal nutrient transport. In: Johnson LR, editor. Physiology of the gastrointestinal tract. 3rd ed. New York: Raven Press, 1994: 1821-44.
  25. Ferraris RP, Diamond J. Crypt-villus site of glucose transporter induction by dietary carbohydrate in mouse intestine. Am J Physiol 1992; 262: G1069-73.
  26. Fedorak RN, Chang EB, Madara JL, Field M. Intestinal adaptation to diabetes. J Clin Invest 1987; 79: 1571-8.
  27. Fedorak RN, Cheeseman CI, Thomson ABR, Porter VM. Altered glucose carrier expression: a mechanism of intestinal adaptation during streptozocin-induced diabetes in rats. Am J Physiol 1991; 261: G585-9.
  28. Madara JL, Trier JS. The functional morphology of the mucosa of the small intestine. In: Johnson LR, editor. Physiology of the gastrointestinal tract. 3rd ed. New York: Raven Press, 1994: 1577-622.
  29. Hopfer U. Membrane transport mechanisms for hexoses and amino acids in the small intestine. In: Johnson LR, editor. Physiology of the gastrointestinal tract. 2nd ed. New York: Raven Press, 1987: 1499-526.
  30. Carey MC, Small DM, Bliss CM. Lipid digestion and absorption. Annu Rev Physiol 1983; 45: 651-77.
  31. Pappenheimer JR. Paracellular intestinal absorption of glucose, creatinine, and nammitol in normal animals: relation to body size. Am J Phyiol 1990; 259: G290-9.
  32. Ferraris RP, Yasharpour S, Lloyd KCK, Mirzayan R, Diamond JM. Luminal glucose concentrations in the gut under normal conditions. Am J Physiol 1990;
    259: G822-37.
  33. Hediger MA, Rhoads DB. Molecular physiology of sodium-glucose cotransporters [review]. Physiol Rev 1994; 74: 993-1026.
  34. Hediger MA, Coady MJ, Ikeda TS, Wright EM. Expression cloning and cDNA sequencing of the Na+/glucose cotransporter. Nature 1987; 330: 379-81.
  35. Burant CF, Sivitz WI, Fukumoto H, Kayano T, Nagamatsu S, Seino S, et al. Mammalian glucose transporters: structure and molecular regulation [review]. Recent Prog Horm Res 1991; 47: 349-87; discussion 387-8.
  36. Burant CF, Takeda J, Brot-Laroche E, Bell GI, Davidson NO. Fructose transport in human spermatozoa and small intestine is GLUT5. J Biol Chem 1992; 267: 14523-6.
  37. Thorens B. Facilitated glucose transporters in epithelial cells. Annu Rev Physiol 1993; 55: 591-608.
  38. Horisberger JD, Lemas V, Kraehenbuhl JP, Rossier BC. Structure-function relationship of Na,K-ATPase. Annu Rev Physiol 1991; 53: 565-84.
  39. Freeman TC, Collins AJ, Heavens RP, Tivey DR. Genetic regulation of enterocyte function: a quantitative in situ hybridisation study of lactase-phlorizin hydrolase and Na(+)-glucose cotransporter mRNAs in rabbit small intestine. Pflugers Arch 1993; 422: 570-6.
  40. Freeman TC, Wood IS, Sirinathsinghji DJS, Beechey RB, Dyer J, Shirazi-Beechey SP. The expression of the Na+/glucose cotransporter (SGLT1) gene in lamb small intestine during postnatal development. Biochim Biophys Acta 1993; 1146: 203-12.
  41. Hopfer U. Membrane transport mechanisms for hexoses and amino acids in the small intestine. In: Johnson LR, editor. Physiology of the gastrointestinal tract. 2nd ed. New York: Raven Press, 1987: 1499-526.
  42. Burant CF, Flink S, Depaoli AM, Chen J, Lee WS, Hediger MA, et al. Small intestinal hexose transport in experimental diabetes: increased transporter mRNA and protein expression in enterocytes. J Clin Invest 1994; 93: 578-85.
  43. Hwang ES, Hirayama BA, Wright EM. Distribution of the SGLT1 Na/glucose cotransporter and mRNA along the crypt-villus axis of rabbit small intestine. Biochim Biophys Res Commun 1991; 181: 1208-17.
  44. Smith MW, Turvey A, Freeman TC. Appearance of phloridzin-sensitive glucose transport is not controlled at mRNA level in rabbit jejunal enterocytes. Exp Physiol 1992; 77: 525-8.
  45. Miyamoto K, Tatsumi S, Morimoto A, Minami H, Yamamoto H, Sone K, et al. Characterization of the rabbit intestinal fructose transporter (GLUT5). Biochem J 1994; 303: 877-83.
  46. Thorens B, Cheng ZQ, Brown D, Lodish HF. Liver glucose transporter: a basolateral protein in hepatocytes and intestine and kidney cells. Am J Physiol 1990; 259: C279-85.
  47. Shirazi-Beechey SP, Hirayama BA, Wang Y. Ontogenic development of the lamb intestinal sodium glucose cotransporter is regulated by diet. J Physiol (Lond) 1991; 437: 699-708.
  48. Lescale-Matys L, Dyer J, Scott D, Freeman TC, Wright EM, Shirazi-Beechey SP. Regulation of the ovine intestinal Na/glucose co-transporter (SGLT1) is dissociated from mRNA abundance. Biochem J 1993; 291: 435-40.
  49. Miyamoto K, Hase K, Taketani Y, Minami H, Oka T, Nakabou Y. Development changes in intestinal glucose transporter mRNA levels. Biochem Biophys Res Commun 1992; 183: 626-31.
  50. Castello A, Guma A, Sevilla L, Furriols M, Testar X, Palacin M, et al. Regulation of GLUT5 gene expression in rat intestinal mucosa -- regional distribution, circadian rhythm, perinatal development and effect of diabetes. Biochem J 1995; 309: 271-7.
  51. Leeper LL, Henning SJ. Development and tissue distribution of sucrase-isomaltase mRNA in rats. Am J Physiol 1990; 258: G52-8.
  52. Nanthakumar NN, Henning SJ. Ontogeny of sucrase-isomaltase gene expression in rat intestine: response to glucocorticoids. Am J Physiol 1993; 264: G306-11.
  53. Freund JN, Torp N, Duluc I, Foltzer-Jourdainne C, Daneilsen M, Raul F. Comparative expression of the mRNA for three intestinal hydrolases during postnatal development in the rat. Cell Mol Biol 1990; 36: 729-36.
  54. Sebastio G, Hunziker W, Ballabio A, Auricchio S, Semenza G. On the primary site of control in the spontaneous development of small-intestinal sucrase-isomaltase after birth. FEBS Lett 1986; 208: 460-4.
  55. Sebastio G, Hunziker W, O'Neill B, Malo C, Menard D, Auricchio S, et al. The biosynthesis of intestinal sucrase-isomaltase in human embryo is most likely controlled at the level of transcription. Biochem Biophys Res Commun 1987; 49: 830-9.
  56. Sebastio G, Villa M, Sartorio R, Guzzetta V, Poggi V, Auricchio S, et al. Control of lactase in human adult-type hypolactasia and in weaning rabbits and rats. Am J Hum Genet 1989; 45: 489-97.
  57. Villa M, Menard D, Semenza G, Mantei N. The expression of lactase enzymatic activity and mRNA in human fetal jejunum. FEBS Lett 1992; 301: 202-6.
  58. Maiuri L, Raia V, Potter J, Swallow D, Mosaic pattern of lactase expression by villus enterocytes in human adult-type hypolactasia. Gastroenterology 1991; 100; 359-69.
  59. Rings EH, De Boer PA, Moorman AF, Van Beers EH, Dekker J, Montgomery RK, et al. Lactase gene expression during early development of rat small intestine. Gastroenterology 1992; 103: 1154-61.
  60. Davidson NO. Cellular and molecular mechanisms of small intestinal lipid transport. In: Johnson LR, editor. Physiology of the gastrointestinal tract. 3rd ed. New York: Raven Press, 1994: 1909-34.
  61. Thomson ABR, Schoeller C, Keelan M, Smith L, Clandinin MT. Lipid absorption: passing through the unstirred layers, brush border membrane and beyond. Can J Physiol Pharmacol 1993; 71: 531-55.
  62. Tso P. Intestinal lipid absorption. In: Johnson LR, editor. Physiology of the gastrointestinal tract. 3rd ed. New York: Raven Press, 1994: 1867-907.
  63. Bloj B, Zilversmit DB. Heterogeneity of rabbit intestine brush border plasma membrane cholesterol. J Biol Chem 1982; 257: 7608-14.
  64. Spady DK, Turley SD, Dietschy JM. Receptor-independent low density lipoprotein transport in the rat in vivo. Quantitation, characterization and metabolic consequences. J Clin Invest 1985; 76: 1113-22.
  65. Thurnhoffer H, Schnabel J, Betz M, Lipka G, Pidgeon C, Hauser H. Cholesterol-transfer protein located in the intestinal brush border membrane. Partial purification and characterization. Biochim Biophys Acta 1991; 1064: 275-86.
  66. Huang Y, Hui DY. Metabolic fate of pancreas-derived cholesterol esterase in intestine: an in vitro study using Caco-2 cells. J Lipid Res 1990; 31: 2029-37.
  67. Purdy BH, Field FJ. Regulation of acylcoenzyme A cholesterol acyl transferase and 3-hydroxy-3-methylglutaryl coenzyme A reductase activity by lipoproteins in the intestine of parabiont rats. J Clin Invest 1984; 74: 351-7.
  68. Gallo LL, Myers S, Vahouny GV. Rat intestinal acyl coenzyme A: cholesterol acyl transferase properties and localization. Proc Soc Exp Biol Med 1984; 177: 188-96.
  69. Dudeja PK, Wali RK, Klitzke A, Brasitus RA. Intestinal D-glucose transport and membrane fluidity along crypt-villus axis of streptozocin-induced diabetic rats. Am J Physiol 1990; 259: G571-7.
  70. Meddings JB, Desouze D, Goel M, Thiesen S. Glucose transport and microvillus membrane physical properties along the crypt-villus axis of the rabbit. J Clin Invest 1990; 85: 1099-107.
  71. Schoeller C, Keelan M, Mulvey G, Stremmel W, Thomson ABR. Role of a brush border membrane fatty acid binding protein in oleic acid uptake into rat and rabbit jejunal brush border membrane. Clin Invest Med 1995; 18: 380-8.
  72. Schoeller C, Keelan M, Mulvey G, Stremmel W, Thomson ABR. Oleic acid uptake into rat and rabbit jejunal brush border membrane. Biochim Biophys Acta 1995; 1236: 51-64.
  73. Diede HE, Rodilla-Sala E, Gunawan J, Manns M, Stremmel W. Identification and characterization of a monoclonal antibody to the membrane fatty acid binding protein. Biochim Biophys Acta 1992; 1125: 13-20.
  74. Yun CH, Tse CM, Nath SK, Levine SA, Brant SR, Donowitz M. Mammalian Na/H exchanger gene family: structure and function studies. Am J Physiol 1995; 269: G1-11.
  75. Keelan M, Walker K, Thomson ABR: Intestinal brush border membrane marker enzymes, lipid composition and villus morphology: effect of fasting and diabetes mellitus in rats. Comp Biochem Physiol 1985; 82A: 83-9.
  76. Keelan M, Thomson ABR, Garg ML, Clandinin MT. Critical period programming of intestinal glucose transport via alterations in dietary fatty acid composition. Can J Physiol Pharmacol 1990; 68: 642-5.
  77. Thomson ABR, Valberg LS, Sinclair DG. Competitive nature of the intestinal transport mechanism for cobalt and iron in the rat. J Clin Invest 1971; 50: 2384-94.
  78. Hennings SJ, Rubin DC, Shulman RJ. Ontogeny of the intestinal mucosa. In: Johnson LR, editor. Physiology of the gastrointestinal tract. 3rd ed. New York: Raven Press, 1994: 571-610.
  79. Thomson ABR, Keelan M, Fedorak RN, Cheeseman CI, Garg ML, Sigalet D, et al. Enteroplasticity. In: Freeman HJ, editor. Inflammatory bowel disease: selected topics. Atlanta: CRC Press, 1989: 95-140.
  80. Thomson ABR, Keelan M, Garg ML, Clandinin MT. Intestinal aspects of lipid absorption: in review. Can J Physiol Pharmacol 1989; 67: 179-91.
  81. Perin N, Keelan M, Clandinin MT, Thomson ABR. Dietary lipids are modifiers of intestinal transport of fatty acids and cholesterol but not sugars in suckling and weanling rats. Gastroenterology 1995; 108: A745.
  82. Duluc I, Gassuser M, Raul F, Freund JN. Dietary control of the lactase mRNA distribution along the rat small intestine. Am J Physiol 1992; 262: G954-61.
  83. Ferraris RP, Vinnakota R. Regulation of intestinal nutrient transport is impaired in aged mice. J Nutr 1993; 123: 502-11.
  84. Miyamoto K, Hase K, Takagi T. Differential responses of glucose transporter mNRA transcripts to levels of dietary sugars. Biochem J 1993; 295: 211-5.
  85. Cheeseman CI. GLUT2 is the transporter for fructose across the rat intestinal basolateral membrane. Gastroenterology 1993; 105: 1050-6.
  86. Burant CF, Saxena M. Rapid reversible substrate regulation of fructose transporter expression in rat small intestine and kidney. Am J Physiol 1994; 267: G71-9.
  87. Bierhoff ML, Levine GM. Luminal and metabolic regulation of jejunal amino acid absorption in the rat. Gastroenterology 1988; 95: 63-8.
  88. Stein ED, Chang SD, Diamond JM. Comparison of different dietary amino acids as inducers of intestinal amino acid transport. Am J Physiol 1987; 252: G626-35.
  89. Salloum RM, Souba WW, Fernandez A, Stevens BR. Dietary modulation of small intestinal glutamine transport in intestinal brush border membrane vesicles of rats. J Surg Res 1990; 48: 635-8.
  90. Keelan M, Doring K, Tavernini M, Wierzbicki A, Clandinin MT, Thomson ABR: Dietary omega-3 fatty acids and cholesterol modify enterocyte microsomal membrane phospholipids, cholesterol content and phospholipid enzyme activities in diabetic rats. Lipids 1994; 29: 851-8.
  91. Thomson ABR, Rafotte R. Effect of dietary modification on the enhanced uptake of glucose, fatty acids, and alcohols in diabetic rats. Am J Clin Nutr 1983; 38: 398-403.
  92. Garg ML, Keelan M, Thomson ABR, Clandinin MT. Fatty acid desaturation in the intestinal mucosa. Biochim Biophys Acta 1988; 958: 139-41.
  93. Garg ML, Wierzbicki A, Keelan M, Thomson ABR, Clandinin MT. Fish oil prevents change in arachidonic acid and cholesterol content in rat caused by dietary cholesterol. Lipids 1989; 24: 266-70.
  94. Lakey JRT, Thomson ABR, Keelan M, Lopaschuk GL, Clandinin MT, Cheeseman CI, et al. Dietary lipid content influences the clinical and intestinal adaptive responses to islet transplantation in diabetic rats. Diabetes Res 1992; 19: 149-64.
  95. Churnratanakul S, Kirdeikis KL, Murphy GK, Wirzba BJ, Keelan M, Rajotte RV, et al. Dietary omega-3 fatty acids partially correct the enhanced in vivo uptake of glucose in diabetic rats. Diabetes Res 1990; 15: 117-23.
  96. Churnratanakul S, Wirzba BJ, Murphy GK, Kirdeikis KL, Keelan M, Clandinin MT, et al. The irradiation-associated decline in the in vivo uptake of glucose observed in rats fed fish oil is prevented by feeding a diet enriched in saturated fatty acids. J Lab Clin Med 1991; 118: 363-9.
  97. Thomson ABR, Keelan M, Tavernini M. Early feeding of a high-cholesterol diet enhances intestinal permeability to lipids in rabbits. Pediatr Res 1987; 21: 347-51.
  98. Thomson ABR, Keelan M, Cheng T, Clandinin MT. Delayed effects of early nutrition with cholesterol plus saturated or polyunsaturated fatty acids on intestinal morphology and transport function in the rat. Biochim Biophys Acta 1993; 1170: 80-91.
  99. Wild GE, Searles LE, Thompson JA, Turner R, Thomson ABR. Dietary modulation of small bowel Na-glucose c-transporter, Na/K ATPase and ornithine decarboxylase gene expression in a rat model of diabetes mellitus [abstract]. Clin Invest Med 1995; 18: A354.
  100. Karasov WH, Diamond JM. Adaptation of intestinal nutrient transport. In: Johnson LR, editor. Physiology of the gastrointestinal tract. 2nd ed. New York: Raven Press, 1987: 1489-97.
  101. Thomson ABR, Keelan M, Sigalet DL, Fedorak RN, Garg ML, Clandinin MT. Patterns, mechanisms and signals for intestinal adaptation. Dig Dis 1990; 8: 99-111.
  102. Thomson ABR, Keelan M, Clandinin MT, Walker K. Dietary fat selectively alters transport properties of rat jejunum. J Clin Invest 1986; 77: 279-88.
  103. Elson CO, Beagley KW. Cytokines and immune mediators. In: Johnson LR, editor. Physiology of the gastrointestinal tract. 3rd ed. New York: Raven Press, 1994: 243-65.
  104. Hunter T, Karin M. The regulation of transcription by phosphorylation. Cell 1992; 70: 375-87.
  105. Taylor SS. cAMP-dependent protein kinase. J Biol Chem 1989; 263: 8443-6.
  106. Roesler WJ, VandenBark GR, Hanson RW. Cyclic AMP and the induction of eukaryotic gene transcription. J Biol Chem 1988; 263: 9063-6.
  107. Nishizuka Y. Intracellular signalling by hydrolysis of phospholipids and activation of protein kinase C. Science 1992; 258: 607-14.
  108. Glatz JFC, Borchers T, Spener F, Van Der Vusse GJ. Fatty acids in cell signalling: modulation by lipid binding proteins. Prostaglandins Leukot Essent Fatty Acids 1995; 52: 121-7.
  109. Distel RJ, Robinson GS, Spiegelman BM. Fatty acid regulation of gene expression. Transcriptional and post-transcriptional mechanism. J Biol Chem 1992; 267: 5937-41.
  110. Clarke SD, Jump DB. Regulation of gene transcription by polyunsaturated fatty acids. Prog Lipid Res 1993; 32: 139-49.
  111. Amri EZ, Ailhaud G, Gimaldi PA. Fatty acids as signal transducing molecules: involvement in the differentiation of preadipose to adipose cells. J Lipid Res 1994; 35: 930-7.
  112. Kaikaus RM, Chan WK, Ortiz DE, Montellano PR, Bass NM. Mechanisms of regulation of liver fatty acid-binding protein. Mol Cell Biochem 1993; 123: 93-100.
  113. Kaikaus RM, Sui Z, Lysensko N, Wu NY, Ortiz De Montellano PR, Ockner RK, et al. Regulations of pathways of extramitochondrial fatty acid oxidation and liver fatty acid-binding protein by long-chain monocarboxylic fatty acids in hepatocytes. J Biol Chem 1993; 268: 26866-71.
  114. Gottlicher M, Widmark E, Li Q, Gustafsoon JA. Fatty acids activate a chimera of the clofibric acid-activated receptor and the glucocorticoid receptor. Proc Natl Acad Sci U S A 1992; 89: 4653-7.
  115. Scheele GA. Extracellular and intracellular messengers in diet-induced regulation of pancreatic gene expression. In: Johnson LR, editor. Physiology of the gastrointestinal tract. 3rd ed. New York: Raven Press, 1994: 1543-53.
  116. Wang J, McCormack SA, Viar MJ, Johnson LR. Stimulation of proximal small intestinal mucosal growth by luminal polyamines. Am J Physiol 1991; 261: G504-11.
  117. Luk GD, Baylin SB. Ornithine decarboxylase in intestinal maturation, recovery and adaptation. In: Robinson JWL, Dowling RH, Riecken EE, editors. Mechanisms of intestinal adaptation. Lancaster: MTP Press, 1982: 65-78.
  118. Schuber F. Influence of polyamines on membrane function. Biochem J 1989; 260: 1-10.
  119. Heikkila R, Schwab G, Wickstrom E, Loongloke S, Pluznik DH, Watt R, et al. A c-myc antisense oligodeoxynucleiotide inhibits entry into S phase but not progress from G0 to G1. Nature 1987; 328: 445-9.
  120. Celano P, Baylin SB, Casero RAJ. Polyamines differentially modulate the transcription of growth-association genes in human colon carcinoma cells. J Biol Chem 1988; 264: 8922-7.
  121. Podolsky DK. Peptide growth factors in the gastrointestinal tract. In: Johnson LR, editor. Physiology of the gastrointestinal tract. 3rd ed. New York: Raven Press, 1994: 129-67.
  122. Barnard JA, Beauchamp RD, Russel WE, Dubois RN, Coffey RJ. Epidermal growth factor-related peptides and their relevance to gastrointestinal pathology. Gastroenterology 1995; 108: 564-80.
  123. Luk GD, Yang P. Polyamines in intestinal and pancreatic adaptation. Gut 1987; 28: 95-101.
  124. Cantley LC, Auger KR, Carpenter C, Duckworth B, Graziani-Apeller R, Soltof S. Oncogenes and signal transduction. Cell 1991; 64: 281-302.
  125. Li L, Hu J, Olson EN. Different members of the jun proto-oncogene family exhibit distinct patterns of expression in response to type beta transforming growth factor. J Biol Chem 1990; 265: 1556-62.
  126. Assi A, Boscoboinik D, Hensey C. The protein kinase C family. Eur J Biochem 1992; 208: 547-57.
  127. Druker BJ, Mamon JH, Roberts TM. Oncogenes, growth factors, and signal transduction. N Engl J Med 1989; 321: 1383-91.
  128. Merchant JL, Dickinson CJ, Yamada T. Molecular biology of the gut: model of gastrointestinal hormones. In: Johnson LR, editor. Physiology of the gastrointestinal tract. 3rd ed. New York: Raven Press, 1994: 295-350.
  129. Levine M, Manley JL. Transcriptional repression of eukaryotic promoters. Cells 1989; 59: 405-8.
  130. Ponta H, Cato ACB, Herrlich P. Interference of pathway specific transcription factors. Biochim Biophys Acta 1992; 1129: 255-61.
  131. Diamond MI, Miner JN, Yoshinaga SK, Yamamoto KR. Transcription factor interactions: selectors of positive or negative regulation from a single DNA element. Science 1990; 249: 1266-72.
  132. Steiner DF, Oyer PE. The biosynthesis of insulin and a probable precursor of insulin by a human islet cell adenoma. Proc Natl Acad Sci U S A 1967; 57: 473-80.
  133. Rothman JE, Orci L. Molecular dissection of the secretory pathway. Nature 1992; 355: 409-15.
  134. Pelham HR. Multiple targets for brefeldin A. Cell 1991; 67: 449-51.
  135. Klausen RD, Donaldson JG, Lippincott-Schwartz J. Brefeldin A: Insights into the control of membrane traffic and organelle structure. J Cell Biol 1991; 116: 1071-80.
  136. Varro A, Dockray GJ. Post-translational processing of progastrin: inhibition of cleavage and phosphorylation by brefeldin A. Regul Pept 1992; 40: 268.
  137. Thorens B, Gerard N, Deriaz N. GLUT2 surface expression and intracellular transport via the constitutive pathway in pancreatic beta cells and insulinoma. J Cell Biol 1993; 123: 1687-94.
  138. James G, Olson EN. Fatty acylated proteins as components of intracellular signalling pathways. Biochemistry 1990; 29: 2623-34.
  139. Molenaar CM, Prange R, Gallwitz D. A carboxyl-terminal cysteine residue is required for palmitic acid binding and biological activity of the ras-related yeast YPTq protein. EMBO J 1988; 7: 971-6.
  140. Kopp R, Mayer P, Pfeiffer A: Agonist-induced desensitization of cholinergically stimulated phosphoinositide breakdown is independent of endogenously activated protein kinase C in HT-29 human colon carcinoma cells. Biochem J 1990; 269: 73-8.

[Table of contents]


| CIM: October 1996 / MCE: octobre 1996 |
CMA Webspinners / >