Hostname: page-component-7c8c6479df-p566r Total loading time: 0 Render date: 2024-03-28T16:42:02.741Z Has data issue: false hasContentIssue false

Diabetic threesome (hyperglycaemia, renal function and nutrition) and advanced glycation end products: evidence for the multiple-hit agent?

Published online by Cambridge University Press:  30 January 2008

Kateřina Kaňková*
Affiliation:
Department of Pathophysiology, Faculty of Medicine, Masaryk University, Komenského nám 2, 662 43 Brno, Czech Republic
*
Corresponding author: Dr Kateřina Kaňková, fax +420 549 494 340, email kankov@med.muni.cz
Rights & Permissions [Opens in a new window]

Abstract

Complex chemical processes termed non-enzymic glycation that operate in vivo and similar chemical interactions between sugars and proteins that occur during thermal processing of food (known as the Maillard reaction) are one of the interesting examples of a potentially-harmful interaction between nutrition and disease. Non-enzymic glycation comprises a series of reactions between sugars, α-oxoaldehydes and other sugar derivatives and amino groups of amino acids, peptides and proteins leading to the formation of heterogeneous moieties collectively termed advanced glycation end products (AGE). AGE possess a wide range of chemical and biological properties and play a role in diabetes-related pathology as well as in several other diseases. Diabetes is, nevertheless, of particular interest for several reasons: (1) chronic hyperglycaemia provides the substrates for extracellular glycation as well as intracellular glycation; (2) hyperglycaemia-induced oxidative stress accelerates AGE formation in the process of glycoxidation; (3) AGE-modified proteins are subject to rapid intracellular proteolytic degradation releasing free AGE adducts into the circulation where they can bind to several pro-inflammatory receptors, especially receptor of AGE; (4) kidneys, which are principally involved in the excretion of free AGE adducts, might be damaged by diabetic nephropathy, which further enhances AGE toxicity because of diminished AGE clearance. Increased dietary intake of AGE in highly-processed foods may represent an additional exogenous metabolic burden in addition to AGE already present endogenously in subjects with diabetes. Finally, inter-individual genetic and functional variability in genes encoding enzymes and receptors involved in either the formation or the degradation of AGE could have important pathogenic, nutrigenomic and nutrigenetic consequences.

Type
Research Article
Copyright
Copyright © The Author 2008

Abbreviations:
AGE

advanced glycation end products

CML

Nε-carboxymethyl-lysine

CRP

C-reactive protein

FN3K

fructosamine-3-kinase

GAPDH

glyceraldehyde 3-phosphate dehydrogenase

RAGE

receptor for AGE

ROS

reactive oxygen species

T1DM

type 1 diabetes mellitus

T2DM

type 2 diabetes mellitus

Diabetes mellitus is the most common metabolic disease. Its prevalence is steadily rising (with few exceptions) globally, having already reached epidemic proportions in industrialised countries. Current WHO estimates predict that the number of individuals with diabetes will reach approximately 300 million by 2025(Reference King, Aubert and Herman1). The diabetes mellitus epidemic is paralleled by body-weight changes (overweight or obesity), and in both cases there is a clear-cut correlation; the faster the socio-economic progress (and subsequent lifestyle changes) in a given society the steeper the rise in the incidence and prevalence of diabetes mellitus and obesity. Both common types of diabetes mellitus, type 1 (T1DM) and type 2 (T2DM), are ethiopathogenetically ‘complex’ diseases with a certain extent of genetic predisposition and a substantial influence of environmental factors (particularly diet and physical (in)activity), although the phenotypic complexity and prevalence of T2DM is much higher. Since diabetes mellitus is characterised by profound abnormalities in energy-substrate partitioning as a result of deficient insulin action (either absolute or relative), nutrition plays a particularly important role in disease progression and its long-term outcomes.

Energy-substrate overload in diabetes (i.e. hyperglycaemia and high NEFA) leads to multiple alterations of intermediate metabolism, presumably as a result of hyperglycaemia- and NEFA-driven overproduction of reactive oxygen species (ROS) in the mitochondrial respiratory chain (since both glucose and NEFA provide the same electron donors NADH and FADH2)(Reference Brownlee2, Reference Du, Edelstein, Obici, Higham, Zou and Brownlee3). One of these alterations leads to the formation of heterogeneous moieties, collectively termed advanced glycation end products (AGE), by the processes of non-enzymic glycation and glycoxidation, which comprise a series of reactions between reducing sugars, α-oxoaldehydes and other sugar derivatives and amino groups of amino acids, peptides and proteins. A mounting body of evidence suggests that AGE possess a wide range of chemical and biological effects and play a role in diabetes-related pathology as well as in several other diseases(Reference Ahmed and Thornalley4Reference Munch, Gasic-Milenkovic and Arendt7). Similar chemical interactions between sugars and proteins occur during thermal processing of food and have been known to the food chemists for approximately 100 years as a result of the pioneering studies of Louis Camille Maillard(Reference Finot8). This chemical interaction is termed the Maillard reaction and the compounds formed are Maillard reaction products; chemically they are largely similar to the AGE formed in vivo. Diet is believed to be a potentially important source of exogenous AGE; however, assessment of the true impact of dietary AGE has recently been the subject of intensive research.

The current and in particular the future extent of the diabetes mellitus epidemic makes it crucial to understand its ethiopathogenesis (including the role of AGE) from the whole body to the molecular level. The following review summarises: (A) the contribution of the principal pathogenic features of diabetes (chronic hyperglycaemia) to the enhanced formation of endogenous AGE, and the biological effects of AGE; therefore establishing the ‘first’ hit of the multiple-hit hypothesis being put forward in the present paper for the role of AGE in diabetes; (B) the current knowledge of the metabolic fate of AGE, the renal excretion of AGE and its impairment by diabetic nephropathy (as well as other causes), which might thus constitute the ‘second’ hit in increasing the AGE burden; (C) the most controversial topic, and potentially the ‘third’ hit, i.e. whether increased dietary intake of AGE in highly-processed foods consumed by patients with diabetes might represent an important source of AGE for the organism in question and, therefore, pose the additional metabolic burden (the ‘third’ hit); (D) in addition, based on data available, the possibility that an individual's genetic make-up may influence the formation and processing of AGE (i.e. nutrigenetics) and that AGE, conversely, may influence the integrity and expression of the genome (i.e. nutrigenomics).

The ‘first’ hit: increased formation of endogenous advanced glycation end products

Metabolic derangements in diabetes and mechanisms of cell damage by hyperglycaemia

Absolute (T1DM) or relative (T2DM) insulin deficiency results in impaired postprandial glucose uptake by skeletal muscle, decreased glucose processing into glycogen, inadequate suppression of hepatic gluconeogenesis and lipolysis in adipose tissue (NEFA further interferes with glucose oxidation via the glucose–NEFA cycle). If the cell cannot efficiently down regulate insulin-independent glucose intake in hyperglycaemia it becomes overloaded, which has been convincingly shown for endothelial(Reference Gaudreault, Scriven and Moore9, Reference Kaiser, Sasson, Feener, Boukobza-Vardi, Higashi, Moller, Davidheiser, Przybylski and King10), mesangial(Reference Brosius and Heilig11) and renal tubular(Reference Linden, DeHaan, Zhang, Glowacka, Cox, Kelly and Rogers12) cells, all of which are targets for vascular damage by hyperglycaemia in diabetes. In the case of endothelial cells the expression of some types of facilitative glucose transporters stay stable or are moderately down regulated by glucose (GLUT1), while others (GLUT2) are up regulated(Reference Gaudreault, Scriven and Moore9, Reference Kaiser, Sasson, Feener, Boukobza-Vardi, Higashi, Moller, Davidheiser, Przybylski and King10). Intracellularly, glucose is metabolised via glycolysis and the Krebs cycle, and the electron donors (NADH and FADH2) generated then supply electrons for the electron transport chain in the mitochondria to be used to generate H+ membrane potential for the production of ATP via ATP synthase. Increased availability of glucose and glycolytic intermediates have been shown to provide substrates for several pathways that are believed to be largely responsible for the cell damage induced by hyperglycaemia(Reference Brownlee2): polyol and hexosamine pathways; dicarbonyl production (mainly methylglyoxal by degradation of triosephosphates) and non-enzymic glycation (AGE); de novo synthesis of diacylglycerol with subsequent activation of protein kinase C isoforms. The overall effects of these abnormalities are changes in gene expression by activation of certain transcription factors (mainly NF-κB), post-translational modification of intra- and extracellular proteins (glycation), alteration of cellular signalling (by activation of kinases), endoplasmic reticulum stress leading to unfolded protein response, impairment of antioxidant capacity etc. (for details, see Brownlee(Reference Brownlee2, Reference Brownlee13)).

Recently, a ‘unifying’ hypothesis was proposed that strengthens the role of mitochondrial ROS overproduction and subsequent DNA oxidative damage as a potent accelerator of these abnormalities (originally seen merely as a result of increased substrate availability because of allosteric inhibition and thus accumulation of glycolytic intermediates). Hyperglycaemia overloads the mitochondrial electron transport chain by increased availability of reducing equivalents (mainly on complex III) and a larger proportion of O2 is then partly reduced with single electrons to superoxide(Reference Brownlee13). Overexpression of manganese superoxide dismutase or uncoupling protein-1 abolishes hyperglycaemia-induced ROS generation in vitro (Reference Nishikawa, Edelstein and Du14). In fact, NADH:NAD+ rather than the absolute amount of NADH seems to be important for the intensity of mitochondrial superoxide generation(Reference Adam-Vizi and Chinopoulos15), and since hyperglycaemia-induced ROS generation directly contributes to an increase in the NADH:NAD+ (see later), diabetes represents a vicious circle in relation to oxidative stress. ROS can oxidatively damage most types of macromolecules; however, oxidative damage of DNA (strand breaks) is particularly important since it requires energy-consuming repair. Ribosylation of multiple proteins by poly-ADP-ribose polymerase, including auto-ribosylation of poly-ADP-ribose polymerase, itself initiates and facilitates DNA repair(Reference Soriano, Virag and Szabo16). NAD+ serves as a donor of ADP-ribose and its consumption by poly-ADP-ribose polymerase contributes to its decreased availability (an increase in NADH:NAD+), further ROS generation and a decrease in ATP formation. Pharmacological inhibition of poly-ADP-ribose polymerase in vitro can suppress all these abnormalities in streptozotocin-induced diabetes(Reference Soriano, Pacher, Mabley, Liaudet and Szabo17). Depletion of NAD+ also inhibits the key glycolytic enzyme glyceraldehyde 3-phosphate dehydrogenase (GAPDH) and causes the accumulation of upstream glycolytic intermediates and their processing in alternative metabolic pathways (see earlier discussion, p. 61). Experimentally, GAPDH antisense oligonucleotides applied to cells cultured in 5 mm-glucose completely mimics hyperglycaemia by activating pathways leading to vascular damage to the same extent as in 30 mm-glucose culture conditions(Reference Du, Matsumura, Edelstein, Rossetti, Zsengeller, Szabo and Brownlee18). GAPDH is generally viewed as a classical glycolytic enzyme; however, it is now clear that it is a multifunctional protein with diverse functions in numerous cellular processes including apoptosis and DNA repair(Reference Sirover19Reference Chuang, Hough and Senatorov21). GAPDH translocates to the nucleus under a variety of stressors, most of which are associated with oxidative stress. Interestingly, poly-ADP-ribosylation of GAPDH itself seems to be one of the likely events enabling its nuclear translocation (further slowing down the rate of glycolysis) and yet another factor responsible for the harmful effect of hyperglycaemia(Reference Du, Matsumura, Edelstein, Rossetti, Zsengeller, Szabo and Brownlee18).

Formation of advanced glycation end products in vivo by non-enzymic glycation

Non-enzymic glycation (also termed the Maillard reaction in some contexts) of proteins involves interplay of serial and parallel reactions between protein residues (mainly lysine and arginine) and sugars, α-oxoaldehydes (dicarbonyls) and other sugar derivatives (in vivo mainly triosephosphate derivatives) generating early (pre-AGE) and advanced glycation products (AGE)(Reference Thornalley, Battah, Ahmed, Karachalias, Agalou, Babaei-Jadidi and Dawnay22). Glycation is one of the most common types of protein modification and spontaneous damage of proteins, which affects approximately 0·1–0·2% of the arginine and lysine residues in vivo (Reference Thornalley, Battah, Ahmed, Karachalias, Agalou, Babaei-Jadidi and Dawnay22). Glycation of proteins proceeds with variable rate and extent during the lifespan of proteins in tissues and body fluids under physiological conditions, but it is more intensive in several disease conditions such as diabetes(Reference Vlassara and Palace23), atherosclerosis(Reference Baynes and Thorpe24), neurodegenerative diseases(Reference Munch, Deuther-Conrad and Gasic-Milenkovic25), osteoarthritis(Reference DeGroot26) and renal failure(Reference Thornalley27). The rate of reaction depends on variables such as temperature, pH and carbonyl:amine, and is thus topically and temporally variable. Generation of some AGE is substantially enhanced in the presence of ROS and some AGE are formed exclusively under oxidative conditions (hence the alternative term ‘glycoxidation’). Originally, glycation was considered to be a type of post-translational modification that occurs mostly in long-lived proteins (e.g. collagen, crystalline); however, it is now clear that AGE are formed also in short-lived proteins, including the cellular proteome(Reference Ahmed and Thornalley4, Reference Thornalley, Battah, Ahmed, Karachalias, Agalou, Babaei-Jadidi and Dawnay22).

Glycating agents in vivo include free sugars (glucose), glycolytic intermediates such as sugar phosphates (glucose- and fructose-6-phosphates) and dicarbonyls (methylglyoxal, glyoxal and 3-deoxyglucosone). Sugars are reactive towards lysine residues while dicarbonyls are mainly reactive towards arginine residues of proteins. Glycation by sugars proceeds through the early (Schiff's bases) and intermediate stages (Schiff's bases undergo the Amadori rearrangement to fructosamines) towards the formation of heterogeneous moieties collectively termed AGE. If the initial glycating agent is glucose the initial product is termed a fructosamine. Probably the best-studied fructosamine to date is HbA1c, which is glycated on the NH2-terminal valine of the β-globin chain, although Hb can also be glycated on several other lysine residues. Since its discovery in 1968(Reference Rahbar28) HbA1c has been established as a widely used and very useful marker of medium- to long-term compensation of diabetes with a certain predictive value(Reference Pickup, Pickup and Williams29). Quantitatively, however, fructoselysine is the most abundant early glycation product and fructosamine in vivo. Fructosamines are unstable products that slowly degrade to form AGE.

Fructosamines were traditionally considered to be a major source of AGE in vivo. Recently, however, as a result of advances in the methodology available for the precise chemical characterisation and quantification of pre-AGE and AGE (particularly liquid chromatography with tandem MS(Reference Ahmed and Thornalley30)), it has became apparent that dicarbonyl-derived AGE are the predominant class of AGE in vivo (Reference Thornalley, Battah, Ahmed, Karachalias, Agalou, Babaei-Jadidi and Dawnay22). Dicarbonyls are formed from the triosephosphates glyceraldehyde 3-phosphate and dihydroxyacetone phosphate (methylglyoxal), lipid peroxidation (glyoxal), fragmentation of early glycation products by the Namiki pathway(Reference Hayashi and Namiki31) (glyoxal, 3-deoxyglucosone), oxidative degradation of nucleotides (glyoxal) and ketone body metabolism (methylglyoxal) and by enzymic degradation of fructosamines(Reference Delpierre, Rider, Collard, Stroobant, Vanstapel, Santos and Van Schaftingen32, Reference Szwergold, Howell and Beisswenger33) (3-deoxyglucosone). In diabetes the triosephosphates are the most important source of dicarbonyls as a result of the previously described inhibitory effects of hyperglycaemia on GAPDH function. GAPDH activity has been shown to be the most important determinant of methylglyoxal concentration for a given glucose level (with an inverse correlation between GAPDH activity and methylglyoxal level)(Reference Beisswenger, Howell, Smith and Szwergold34). The products of sugar oxidation, methylglyoxal, glyoxal and 3-deoxyglucosone, react with mainly arginine residues of proteins and form AGE directly. Dicarbonyl-derived AGE on arginine residues, the hydroimidazolones (methylglyoxal-H, glyoxal-H, 3-deoxyglucosone-H), are the most abundant AGE in vivo in body fluids as well as in cellular protein (representing approximately 1% of the total arginine in some cell types)(Reference Thornalley, Battah, Ahmed, Karachalias, Agalou, Babaei-Jadidi and Dawnay22). The concentration of dicarbonyls as well as dicarbonyl-derived hydroimidazolones has been shown to be significantly increased in patients with diabetes (⩽3-fold for glycated plasma proteins, and ⩽10-fold for glycation-free adducts; see p. 65) compared with controls without diabetes(Reference Ahmed, Babaei-Jadidi, Howell, Beisswenger and Thornalley35, Reference Ahmed, Babaei-Jadidi, Howell, Thornalley and Beisswenger36). The level of dicarbonyls correlates with glycaemia, and variations in postprandial glucose contribute substantially to the rise of dicarbonyls (and thus dicarbonyl-derived AGE), while levels of HbA1c do not reflect the postprandial glucose fluctuations(Reference Beisswenger, Howell, O'Dell, Wood, Touchette and Szwergold37). Apart from the arginine-based dicarbonyl-derived AGE several other classes of AGE have been identified and chemically characterised (an overview of the current AGE classification based on their chemical structure and source is presented in Table 1).

Table 1. Current classification of advanced glycation end products (AGE)

3DG, 3-deoxyglucosone; FN, fructosamine; G, glyoxal; MG, methylglyoxal; AA, amino acid.

Enzymic defence against glycation

Although the formation of pre-AGE and AGE is a non-enzymic process, e.g. methylglyoxal arises by non-enzymic phosphate elimination from the two intermediates of glycolysis glyceraldehyde 3-phosphate and dihydroxyacetone phosphate (approximately 0·2–1% of the flux of triosephosphates under normoglycaemic conditions(Reference Phillips and Thornalley38)), organisms including man have developed several enzymic detoxification systems to deal with those products(Reference Thornalley39), since high levels of methylglyoxal, for example, are highly cytotoxic. Surprisingly, the physiological role of dicarbonyls including methylglyoxal is still rather obscure (methylglyoxal is known to target several proteins involved in the regulation of cell growth, differentiation and cell death). From an evolutionary point of view monosaccharides such as glucose have evolved as a universal source of energy and C for metabolism; however, because of their intrinsically-high reactivity with amines, non-enzymic glycation is an unavoidable ‘background’ reaction in all living systems(Reference Szwergold40). Since uncontrolled glycation is detrimental to the function and integrity of biological macromolecules (the production of ROS from early glycation products, impairment of enzyme functions, perturbations of peptide hormone signalling, activation of AGE-specific receptors, cross-linking of structural proteins, impairment of protein recycling, mutagenicity etc.; see later) the need for regulatory systems capable of limiting and reversing such protein modifications has evolved in parallel(Reference Szwergold40). Passive regulatory mechanisms to limit unwanted excessive glycation include preference for monosaccharides with the lowest reactivity towards amines (i.e. glucose) and the maintenance of the lowest glycaemia possible or necessary as a result of prompt disposal of excess glucose to storage forms under insulin regulation. However, exclusively active mechanisms (deglycating enzyme systems) can (at least partly) maintain homeostasis when there are physiological or pathophysiological variations in the levels of glycating substrates. One of these detoxification mechanisms is the glyoxalase system(Reference Thornalley41). Methylglyoxal reacts spontaneously with glutathione to form a hemithioacetal, which is converted into S-d-lactoyl-glutathione by glyoxalase I, and then further metabolised to d-lactate by glyoxalase II. The importance of methylglyoxal and its detoxification in hyperglycaemia has been confirmed experimentally; glyoxalase I overexpression in vitro completely protects against the hyperglycaemia-driven formation of methylglyoxal and AGE in bovine endothelial cells(Reference Shinohara, Thornalley, Giardino, Beisswenger, Thorpe, Onorato and Brownlee42). Similarly, fructoselysines may be enzymically degraded by fructosamine-3-kinase (FN3K)(Reference Delpierre, Rider, Collard, Stroobant, Vanstapel, Santos and Van Schaftingen32, Reference Szwergold, Howell and Beisswenger33). FN3K phosphorylates fructoselysine residues on glycated proteins yielding fructosamine 3-phosphates, which are unstable and spontaneously decompose to form 3-deoxyglucosone. Based on current knowledge of both these enzymic systems (and several others such as aldehyde reductases and dehydrogenases and amadoriase) it is clear that deglycation plays a regulatory role under the normal metabolic conditions and efficacy of the systems may, consequently, play an important role in hyperglycaemia(Reference Conner, Beisswenger and Szwergold43). Whether the genetic variability in the glyoxalase and FN3K loci influence their efficacy will be discussed later (p. 68).

Biological effects of advanced glycation end products

AGE exert multiple biologically-important effects in living organisms by direct modification of proteins (and their functions) and indirectly via AGE-binding receptors. The importance of direct glycation certainly applies to long-lived proteins such as extracellular matrix proteins (e.g. collagen); their cross-linking (lysine to lysine or lysine to arginine residues), rigidity and resistance to proteolysis contribute to the changes in physical properties observed during aging, with accelerated rates in tissues (e.g. skin collagen) from patients with diabetes(Reference Yu, Thorpe and Jenkins44). If glycation occurs at specific sites it can also affect the cell–collagen interaction, e.g. modification of arginine within the RGD or GFOGER motifs recognised by integrins α1β2 and α2β1, which reduces cell interactions during turnover and platelet interactions and can ultimately affect tissue repair and wound healing in patients with diabetes(Reference Avery and Bailey45). Arginine-derived hydroimidazolone modification of integrin-binding sites of vascular-basement-membrane type IV collagen has been shown to induce endothelial cell detachment, anoikis and inhibition of angiogenesis, therefore possibly contributing to vascular damage in diabetes(Reference Dobler, Ahmed, Song, Eboigbodin and Thornalley46). In proteins with a faster turnover abnormal protein–protein or enzyme–substrate interactions can be affected by glycation. The most-abundant plasma protein in man, human serum albumin, is an important target of glycation in vivo. Peptide mapping of human serum albumin has identified hot spots for glycation by dicarbonyls (methylglyoxal) located in the drug-binding site and the active site of albumin-associated esterase activity(Reference Ahmed and Thornalley47); hot spots for modification by glucose have also been identified (Reference Garlick and Mazer48). Similar hot spots for dicarbonyl-derived arginine-directed modification (in addition to glucose-derived lysine-directed modification) have been identified in Hb(Reference Gao and Wang49), both types of modifications have been shown to contribute to increased susceptibility to auto-oxidation, reduced α-helix content, increased thermolability and a weaker haem–globin linkage in glycated Hb(Reference Sen, Kar, Roy and Chakraborti50). These examples are just some of the glycated proteins studied; structural modification of many other proteins (such as apo, extracellular matrix proteins, crystalline etc.) also contributes to diabetic pathology. Another interesting effect of AGE has come from an in vitro study demonstrating that AGE dose-dependently impairs CD34+ progenitor cell function (i.e. incorporation of progenitor cells into the sprouting endothelium), which might be yet another pathophysiological factor (mediated by a hitherto unknown mechanism) of disturbed vascular function in diabetes(Reference Scheubel, Kahrstedt, Weber, Holtz, Friedrich, Borgermann, Silber and Simm51).

Although most of the research has focused on the effect of AGE on proteins, it has become increasingly obvious that DNA serves as another target for AGE modification. The genotoxicity of AGE has been documented in vitro in several cell lines(Reference Stopper, Schinzel, Sebekova and Heidland52, Reference Schupp, Schinzel, Heidland and Stopper53) and in vivo in patients with uraemia who are undergoing different types of haemodialysis(Reference Fragedaki, Nebel and Schupp54). DNA damage is induced by both the production of ROS via receptor-mediated signalling, probably involving the receptor for AGE (RAGE)/NAD(P)H oxidase pathway (estimated by 8-hydroxy-2-deoxyguanosine), and by the direct effect of AGE on DNA producing base modifications, stand breaks, photosensitisation and apurination or apyrimidination.

Circulating AGE-modified molecules also interact with specific cell-surface receptors on a range of cell types (in the circulation predominantly on monocytes and macrophages and endothelium). In some cases AGE undergo endocytosis and degradation, in others they activate intracellular signalling pathways and influence cellular phenotype. Several types of receptors for AGE-modified molecules have been identified: RAGE(Reference Neeper, Schmidt, Brett, Yan, Wang, Pan, Elliston, Stern and Shaw55, Reference Schmidt, Vianna, Gerlach, Brett, Ryan, Kao, Esposito, Hegarty, Hurley and Clauss56); galectin-3(Reference Vlassara, Li, Imani, Wojciechowicz, Yang, Liu and Cerami57); OST-48; 80K-H(Reference Li, Mitsuhashi, Wojciechowicz, Shimizu, Li, Stitt, He, Banerjee and Vlassara58); scavenger receptors class A (I and II)(Reference Araki, Higashi, Mori, Shibayama, Kawabe, Kodama, Takahashi, Shichiri and Horiuchi59) and B (CD36 and BI); lectin-like oxidised LDL receptor-1; FEEL-1 and -2; probably others as yet unidentified(Reference Horiuchi, Sakamoto and Sakai60). RAGE, isolated and characterised in 1992(Reference Neeper, Schmidt, Brett, Yan, Wang, Pan, Elliston, Stern and Shaw55, Reference Schmidt, Vianna, Gerlach, Brett, Ryan, Kao, Esposito, Hegarty, Hurley and Clauss61), is an Ig-type cell-surface receptor that recognises the tertiary structure of multiple ligands including S100/calgranulins(Reference Hofmann, Drury and Fu62) and high-mobility-group box-1 protein(Reference Hori, Brett and Slattery63), amyloid β-peptides(Reference Yan, Chen and Fu64) and AGE. RAGE has similarities with signalling pattern-recognition receptors such as Toll-like or mannose-receptors that are expressed by immune cells and recognise proteins associated with microbial pathogens or cellular stress and participate predominantly in the innate immune response. Accumulating experimental data on RAGE functions and actions have been extensively reviewed throughout the past decade, particularly in the most recent comprehensive reviews(Reference Yan, Barile, D'Agati, Du Yan, Ramasamy and Schmidt65Reference Bierhaus, Stern and Nawroth68). Binding of AGE to RAGE has been shown to activate several intracellular signal transduction pathways (probably tissue- and cell-type-specific), mainly mitogen-activated protein kinases including extracellular regulated kinase 1/2, p38 mitogen-activated protein kinase and stress-activated protein kinase/c-Jun N-terminal kinase; furthermore, other pathways identified are the janus kinase/signal transducers and activators of transcription pathway, 1-phosphatidylinositol 3-kinase and NAD(P)H oxidase. As AGE–RAGE signalling is probably cell-type and ligand specific, the array of cellular and tissue responses has to be quite broad. One of the well-described down-stream events following activation of upstream kinases by AGE ligation to RAGE is NF-κB activation. Unlike the physiological transient NF-κB activation promoting cell survival under a variety of stimuli, in diabetes (in vivo in mononuclear cells from patients with T1DM as well as in vitro) NF-κB activation has been shown to be sustained over long-term periods as a result of overproduction of the NF-κB p65 subunit, while at the same time degradation of the inhibitory subunit of NF-κB is increased(Reference Bierhaus, Schiekofer and Schwaninger69). Considerable data relating to the actions of RAGE have been generated using blockade of signalling by soluble RAGE, transfection of cells with a dominant negative RAGE plasmid or transgenic mice expressing RAGE tissue-specifically or RAGE-deficient mice. Interestingly, RAGE also exists in vivo in a soluble form similar to other secreted pattern-recognition receptors (such as pentraxins, mannan-binding lectin etc.). There have been several independent descriptions of the existence of endogenous RAGE variants derived by alternative splicing of the gene leading to C- or N-terminal truncation(Reference Malherbe, Richards, Gaillard, Thompson, Diener, Schuler and Huber70Reference Yonekura, Yamamoto and Sakurai72). A proportion of the RAGE isoforms produced is tissue-specific, regulated by as yet unknown mechanisms. Circulating C-terminally-truncated soluble RAGE of approximately 50 kDa produced by endothelial cells is regarded as a naturally-occurring competitive inhibitor of signalling pathways induced by the transmembrane RAGE. Following the discovery of soluble RAGE, decreased levels have been demonstrated in several diseases in association with enhanced formation of AGE, such as coronary disease(Reference Falcone, Emanuele, D'Angelo, Buzzi, Belvito, Cuccia and Geroldi73), carotid and femoral atherosclerosis(Reference Koyama, Shoji and Yokoyama74), essential hypertension(Reference Geroldi, Falcone, Emanuele, D'Angelo, Calcagnino, Buzzi, Scioli and Fogari75), diabetic retinopathy(Reference Hayaishi-Okano, Yamasaki and Kajimoto76, Reference Sakurai, Yamamoto and Tamei77), rheumatoid arthritis(Reference Pullerits, Bokarewa, Dahlberg and Tarkowski78) and Alzheimer's disease(Reference Emanuele, D'Angelo, Tomaino, Binetti, Ghidoni, Politi, Bernardi, Maletta, Bruni and Geroldi79).

The ‘second’ hit: decreased excretion of advanced glycation end products

Processing and metabolism of advanced glycation end products

The integrity of the cellular proteome is maintained by rapid proteolysis of misfolded or altered proteins by the proteasome(Reference Goldberg80). Proteasomes are large protein complexes that are located in the nucleus and cytoplasm of all eukaryotes. They can function as the 26S proteasome, which comprises one 20S core particle and two 19S regulatory caps that carry out ubiquitin- and ATP-dependent proteolysis, or as the 20S proteasome alone. Oxidised proteins are degraded directly by the 20S core particle without the involvement of the 19S regulatory cap and do not require ATP hydrolysis or tagging with ubiquitin(Reference Shringarpure, Grune, Mehlhase and Davies81). Cellular proteins modified by AGE are also subject to intracellular proteolytic degradation, releasing free AGE into the circulation (i.e. AGE-free adducts)(Reference Thornalley, Battah, Ahmed, Karachalias, Agalou, Babaei-Jadidi and Dawnay22), and are probably the most important source of AGE in circulation, even if cellular glycated proteins exhibit some extent of resistance to proteasome degradation(Reference Bulteau, Verbeke, Petropoulos, Chaffotte and Friguet82). In vitro glyoxal treatment of fibroblasts inhibits 20S proteasome peptidase activities without changing the proteasome content(Reference Bulteau, Verbeke, Petropoulos, Chaffotte and Friguet82). Additionally, a decrease in proteasome degradation can be, to some extent, a result of AGE modification of lysine residues that are then inaccessible for ubiquitinylation or glycation by the proteasome itself(Reference Carrard, Dieu, Raes, Toussaint and Friguet83, Reference Stolzing, Widmer, Jung, Voss and Grune84).

In the circulation the total pool of AGE consists of those bound to polypeptides (i.e. formed primarily on plasma proteins), peptides (probably released by extracellular proteolysis from tissue-immobilised AGE) and amino acids (formed by proteolytic degradation of intracellular proteins, direct modification of circulating amino acids and by exogenously-derived AGE, primarily those contained in food and absorbed from the intestine; see Fig. 1). The nomenclature of AGE in the context of their metabolism therefore distinguishes between (1) glycation adduct residues of proteins, (2) glycation adduct residues of peptides (0·5–12 kDa) and (3) glycation-free adducts (amino acids). The latter category represents the most abundant form of AGE in the plasma of healthy subjects as well as subjects with diabetes(Reference Ahmed, Babaei-Jadidi, Howell, Beisswenger and Thornalley35).

Fig. 1. Overview of advanced glycation end product (AGE) metabolism and factors influencing AGE turnover and the concentration of AGE in particular body compartments. MG, methylglyoxal; GLO I, glyoxalase I; FN3K, fructosamine 3 kinase; RAGE, receptor of AGE; ECM, extracellular matrix.

Role of the kidney in the excretion of advanced glycation end products

Renal clearance is the predominant means of excretion of AGE, particularly the low-molecular-weight fraction (glycation-free adducts)(Reference Ahmed, Babaei-Jadidi, Howell, Beisswenger and Thornalley35). The predominant plasma AGE (hydroimidazolones, Nε-carboxymethyl-lysine (CML) and Nε-carboxyethyl-lysine) have a high renal clearance(Reference Thornalley, Battah, Ahmed, Karachalias, Agalou, Babaei-Jadidi and Dawnay22) and their levels are inversely correlated with renal function(Reference Stam, Schalkwijk, van Guldener, ter Wee and Stehouwer85). AGE adducts and peptides are filtered into the glomeruli and a small proportion may be reabsorbed and degraded by proximal tubular cells(Reference Gugliucci and Bendayan86). Subjects with chronic renal failure have increased plasma levels of glycation-free adducts (fructoselysine, methylglyoxal-H, CML and Nε-carboxyethyl-lysine) compared with normal controls (⩽5-fold); however, in patients with end-stage renal disease and renal replacement therapy plasma glycation free adducts are increased ⩽18-fold on peritoneal dialysis and ⩽40-fold on haemodialysis(Reference Agalou, Ahmed, Babaei-Jadidi, Dawnay and Thornalley87). AGE accumulate in patients with uraemia to a much greater extent than in subjects with normal renal function(Reference Odetti, Cosso, Pronzato, Dapino and Gurreri88Reference Sebekova, Podracka, Blazicek, Syrova, Heidland and Schinzel91); however, AGE do not differ between patients with uraemia associated with diabetes and those with uraemia not associated with diabetes(Reference Miyata, Ueda, Shinzato, Iida, Tanaka, Kurokawa, van Ypersele de Strihou and Maeda92, Reference Miyata, Fu, Kurokawa, van Ypersele de Strihou, Thorpe and Baynes93). Renal replacement therapy itself also plays a role; the concentration of AGE in the peritoneal dialysate exceeds that of the plasma, suggesting that AGE might be formed de novo in the peritoneal cavity from the glucose and its derivatives formed during heat sterilisation of the dialysate(Reference Thornalley27). The origin of increased AGE in uraemia is complex and not entirely dependent on their decreased renal clearance; it is more likely to be a consequence of enhanced oxidative and carbonyl stress and inflammation induced by uraemic toxins and contact with dialysis membranes and fluids(Reference Kalousova, Zima, Tesar, Stipek and Sulkova94, Reference Miyata, van Ypersele de Strihou, Kurokawa and Baynes95).

The involvement of the liver in the removal of AGE has also been studied, as preliminary experimental data for animals injected with heavily-modified albumin indicate hepatic uptake by scavenger receptors and metabolism of AGE(Reference Smedsrod, Melkko, Araki, Sano and Horiuchi96, Reference Hansen, Svistounov, Olsen, Nagai, Horiuchi and Smedsrod97). Similarly, elevation of AGE in subjects with liver cirrhosis and subsequent normalisation after liver transplantation has been demonstrated(Reference Sebekova, Kupcova, Schinzel and Heidland98). However, extraction of AGE formed in vivo (minimally modified compared with the previous experiments) from the blood entering and leaving the liver of both healthy subjects and patients with liver cirrhosis does not suggest marked hepatic degradation(Reference Ahmed, Thornalley, Luthen, Haussinger, Sebekova, Schinzel, Voelker and Heidland99).

Aetiology of chronic renal failure and end-stage renal disease, diabetic nephropathy

Diabetic nephropathy is a serious long-term consequence of diabetes that affects approximately 30% of patients with T1DM and T2DM, and it is the major single cause of chronic kidney disease and end-stage renal disease in developed countries (approximately 28%). Other diseases leading to chronic kidney disease and end-stage renal disease include hypertension (approximately 24%), glomerulonephritis (21%) and several less-common diseases such as polycystic kidney disease, chronic pyelonephritis, nephrolithiasis and systemic lupus erythematosus.

The aetiology of diabetic nephropathy comprises both the metabolic and haemodynamic changes that accompany diabetes(Reference Forbes, Fukami and Cooper100). The principal metabolic changes associated with hyperglycaemia have been described earlier (pp. 61–62). Haemodynamic changes are represented initially by increased intraglomerular pressure, which is independent of systemic blood pressure and can be detected quite early after the onset of the disease. Blockade of the rennin–angiotensin system (by angiotensin-converting enzyme inhibitors and/or angiotensin II receptor blockers) has been convincingly shown to delay the onset and progression of diabetic nephropathy, not only by its effect on filtration pressure but also by its non-haemodynamic effects (suppression of ROS production and cytokine release)(Reference Forbes, Fukami and Cooper100Reference Jacobsen103). All parts of nephron (the vasculature, glomerular filtration barrier, tubuli and also the kidney interticium) undergo pathological changes during the course of diabetic nephropathy. Experimental data (mainly from animal studies using inhibitors of advanced glycation) suggest that the pathogenic mechanism in the diabetic kidney is mainly mediated by AGE(Reference Thomas, Forbes and Cooper104). Direct modification of renal proteins by AGE may produce changes in charge selectivity, solubility, conformation (e.g. type IV collagen of glomerular basement membrane and its interaction with podocytes via integrins) and cell turnover (mesangium). Moreover, AGE also interact with specific receptors and binding proteins to influence the renal expression of growth factors and cytokines implicated in the progression of diabetic renal disease. Immunohistochemical data obtained from human or rodent kidneys show that RAGE is expressed by podocytes (but not in mesangial cells or glomerular endothelium)(Reference Tanji, Markowitz, Fu, Kislinger, Taguchi, Pischetsrieder, Stern, Schmidt and D‘Agati105) and by proximal tubular cells(Reference Oldfield, Bach, Forbes, Nikolic-Paterson, McRobert, Thallas, Atkins, Osicka, Jerums and Cooper106). In rodent models activation of RAGE contributes to glomerular pathology (enhanced permeability, inflammation and glomerular basement membrane thickening), mesangial expansion and tubular changes(Reference Oldfield, Bach, Forbes, Nikolic-Paterson, McRobert, Thallas, Atkins, Osicka, Jerums and Cooper106Reference Morcos, Sayed and Bierhaus110). Engagement of RAGE by AGE (and probably also S100/calgranulins) in the diabetic kidney contributes through the cascade of signalling events using ROS as secondary messengers in the activation of transforming growth factor β, connective tissue growth factor and vascular endothelial growth factor axes directly responsible for renal remodelling. Experimental blockade of AGE formation by benfotiamine(Reference Babaei-Jadidi, Karachalias, Ahmed, Battah and Thornalley111), pyridoxamine(Reference Degenhardt, Alderson, Arrington, Beattie, Basgen, Steffes, Thorpe and Baynes112) and alagebrium(Reference Forbes, Thallas, Thomas, Founds, Burns, Jerums and Cooper113, Reference Thallas-Bonke, Lindschau, Rizkalla, Bach, Boner, Meier, Haller, Cooper and Forbes114), blockade of RAGE by neutralising antibodies(Reference Flyvbjerg, Denner, Schrijvers, Tilton, Mogensen, Paludan and Rasch115) or the absence of the latter in RAGE-knock-out mice(Reference Wendt, Tanji and Guo107) completely suppress structural and functional changes associated with diabetic nephropathy, thereby supporting the pathogenic role of AGE–RAGE interaction in diabetic nephropathy.

The third hit: increased intake of exogenous advanced glycation end products in the diet

Maillard reaction; dietary advanced glycation end products sources

In 1912 the French scientist Louis Camille Maillard described the ‘browning’ or Maillard reaction between reducing sugars and amino acids during cooking(Reference Maillard116). Products formed by this reaction (Maillard reaction products) contribute to the qualitative properties of foods such as colour, taste and aroma; moreover, it has become increasingly evident that they can also affect the nutritional and toxicological properties of food. The potential biological effect of endogenously-formed AGE, especially in situations like diabetes or uraemia, has also stimulated intensive interest in the potential contribution of dietary AGE to an individual's total AGE levels. Thermal processing of foods enhances their digestibility, sensory properties and shelf life, producing microbiologically-safe products with the desired nutritional quality that make a major contribution to the human diet. Although such treatment can lead to the formation of components with presumably health-promoting properties (such as antioxidants or melanoidins), it can also lead to the formation of potentially-harmful compounds (such as heterocyclic amines, acrylamide and Maillard reaction products)(Reference Somoza117). It is clear that food is a rich source of pre-AGE (mainly fructoselysine) and AGE (mainly methylglyoxal-H, CML, pyrraline, pentosidine). Most dietary AGE are derived from sugar- and protein-rich cooked foods such as bakery products, roasted meat, milk and some other drinks (coke drinks, for example, have a lower AGE content than untreated milk, and the AGE content of milk further increases during pasteurisation and sterilisation(Reference Ahmed, Mirshekar-Syahkal, Kennish, Karachalias, Babaei-Jadidi and Thornalley118)), although, the levels depend also on the preparation conditions, presence of metals and water content(Reference Goldberg, Cai, Peppa, Dardaine, Baliga, Uribarri and Vlassara119). Qualitative and quantitative assessment of AGE in food is currently an important area of food research(Reference Henle120).

Bioavailability, metabolic fate and excretion of dietary advanced glycation end products

It seems that highly-glycated proteins (containing fructoselysines and AGE) may not be digested efficiently (resistance to proteolysis) and that the majority of AGE present in foods are not absorbed from the gastrointestinal tract but excreted in the faeces. More precisely, approximately 10–30% of the ingested AGE are intestinally absorbed, mainly in the form of glycation-free adducts and peptides, and are subsequently found in the circulation(Reference Koschinsky, He, Mitsuhashi, Bucala, Liu, Buenting, Heitmann and Vlassara121Reference Faist, Wenzel and Randel123), where the peptides undergo rapid degradation to free adducts, as shown by comparing the concentration of glycated peptides in the portal and systemic venous circulation(Reference Ahmed, Thornalley, Luthen, Haussinger, Sebekova, Schinzel, Voelker and Heidland99). Absorbed AGE are excreted rapidly in the urine by subjects or experimental animals with normal renal function, as indicated by balance studies quantifying recovery in the urine of CML, pyrraline and pentosidine from different foods or after injection of labelled compounds(Reference Faist, Wenzel and Randel123Reference Bergmann, Helling, Heichert, Scheunemann, Mading, Wittrisch, Johannsen and Henle126). In these studies ⩽80% of AGE were recovered in the urine, predominantly in the glycation-free adduct form, although it may have been as low as 2% in the case of protein-bound AGE (e.g. pentosidine in the free adducts in brewed coffee v. that of pretzel sticks). In animal studies faecal excretion was reported to be ⩽26–29% for CML but only 1–3% for fructoselysine(Reference Faist and Erbersdobler122, Reference Faist, Wenzel and Randel123), which was considered to be an effect of the metabolisation of fructoselysine by colon microflora, e.g. Escherichia coli can degrade fructoselysine to glucose 6-phosphate and utilise it as a substrate(Reference Wiame, Delpierre, Collard and Van Schaftingen127). It is not clear whether AGE are similarly degraded by colon microflora(Reference Tuohy, Hinton, Davies, Crabbe, Gibson and Ames128).

It is difficult to interpret the findings of both animal and human studies of the bioactivity of AGE because of a series of confounding factors, experimental (study design, duration of administration, medical condition of subjects studied, species studied etc.) as well as methodological (different composition and preparation of high-AGE diet, analytical methods employed to quantify AGE etc.). While some studies have shown a significant effect of dietary AGE, other studies have not found such an effect (see following discussion). Nevertheless, such disparity certainly stimulates further carefully-designed experiments.

Bioactivity of dietary advanced glycation end products: animal studies

The biological effects of dietary AGE have been investigated in healthy rodents fed isoenergetic chow with different AGE contents (for study periods ranging from 6 weeks to 6 months). Rats on a high-AGE diet have been reported to show a higher increase in plasma AGE, urinary excretion of AGE, weight gain, renal protein excretion and renal expression of the pro-fibrotic cytokine transforming growth factor β(Reference Sebekova, Hofmann, Boor, Sebekova, Ulicna, Erbersdobler, Baynes, Thorpe, Heidland and Somoza129), while, increased fasting plasma insulin has been detected in mice(Reference Sandu, Song, Cai, Zheng, Uribarri and Vlassara130).

Similar findings have been obtained in studies of mice with diabetes (non-obese or db/db) fed a high-AGE diet (higher plasma AGE, weight gain and fasting plasma insulin)(Reference Hofmann, Dong, Li, Cai, Altomonte, Thung, Feng, Fisher and Vlassara131), as well as in another study in which there was also progressive development of diabetic nephropathy (higher urinary AGE excretion, renal protein excretion, glomerular hypertrophy, mesangial expansion and expression of transforming growth factor β) and shorter survival (non-obese NOD mice)(Reference Zheng, He, Cai, Hattori, Steffes and Vlassara132). A low-AGE diet has also been found to delay the onset of diabetes in diabetic (NOD) mice(Reference Peppa, He, Hattori, McEvoy, Zheng and Vlassara133), improve wound healing in db/db mice(Reference Peppa, Brem, Ehrlich, Zhang, Cai, Li, Croitoru, Thung and Vlassara134) and delay the progression of atherosclerosis(Reference Lin, Reis, Dore, Lu, Ghodsi, Fallon, Fisher and Vlassara135). In the latter study genetically-hypercholesterolaemic (apoE-deficient) mice fed a low-AGE diet for 4 weeks following experimental injury of the femoral artery were reported to develop a lesser extent of neointimal formation (scar tissue) than animals fed a high-AGE diet(Reference Lin, Reis, Dore, Lu, Ghodsi, Fallon, Fisher and Vlassara135). Moreover, in the same animal model rendered diabetic by treatment with streptozotocin, 2-month dietary AGE restriction was associated with lower serum AGE, reduced formation of atheromatous lesions, lower tissue expression of AGE receptors and infiltration with inflammatory cells without concomitant differences in plasma glucose, TAG or cholesterol(Reference Lin, Choudhury, Cai, Lu, Fallon, Fisher and Vlassara136).

The effect of dietary AGE on renal function has been studied using an experimental model of renal failure (subtotally-nephrectomised rats). Nephrectomised rats were found to exhibit higher weight gain, proteinuria, kidney hypertrophy and expression of transforming growth factor β than healthy rats(Reference Sebekova, Hofmann, Boor, Sebekova, Ulicna, Erbersdobler, Baynes, Thorpe, Heidland and Somoza129). Administration of alagebrium (AGE breaker) was shown to delay and/or reverse established diabetic nephropathy in db/db mice by reducing systemic AGE and facilitating their urinary excretion(Reference Peppa, Brem, Cai, Zhang, Basgen, Li, Vlassara and Uribarri137). Of particular interest is a study designed to elucidate the potential effect of AGE as a secondary renal insult in rats with a developmental nephron deficit. A low-protein diet fed to female rats during pregnancy and lactation was found to reduce nephron number in their offspring compared with that of the offspring of female rats fed a normal-protein diet; however, no obvious differences were found in systemic blood pressure or glomerular filtration rate. Administration of AGE for 4 weeks beginning at the 20th week of age was shown to induce greater expression of the genes encoding transforming growth factor β and procollagen III as well as renal accumulation of AGE when compared with rats with a normal nephron status(Reference Zimanyi, Denton, Forbes, Thallas-Bonke, Thomas, Poon and Black138).

Bioactivity of dietary advanced glycation end products: human studies

In a study of ninety healthy subjects AGE intake (assessed by 3 d food records) was shown to be correlated with circulating AGE, AGE-modified LDL and high-sensitivity C-reactive protein (CRP) concentrations(Reference Uribarri, Cai, Sandu, Peppa, Goldberg and Vlassara139). In a larger study of 172 healthy subjects of different ages it was found that AGE are directly influenced by dietary intake (independent of age or energy intake), and circulating AGE (CML and methylglyoxal derivatives) are elevated in older subjects and correlated with indicators of inflammation and oxidative stress across all ages(Reference Uribarri, Cai, Peppa, Goodman, Ferrucci, Striker and Vlassara140). A 6-week nutritional intervention study of subjects with diabetes has demonstrated that circulating AGE can be modulated by altering dietary AGE intake, and high AGE intake is associated with higher levels of inflammatory molecules (CRP, TNFα) and markers of endothelial dysfunction (vascular-cell adhesion molecule 1)(Reference Vlassara, Cai, Crandall, Goldberg, Oberstein, Dardaine, Peppa and Rayfield141). Furthermore, a comparison of the effects of low and high intakes of dietary AGE on the modification of plasma LDL in patients with diabetes who had equal glycaemic control and lipidaemia has shown that high AGE intake can transform circulating macromolecules to a much greater extent and render them more pro-atherogenic than those of subjects on a low AGE intake(Reference Cai, He, Zhu, Peppa, Lu, Uribarri and Vlassara142). Also, a high-AGE diet was found to induce more pronounced micro- and macrovascular dysfunction and endothelial dysfunction compared with a low-AGE diet in a cross-over study with twenty subjects with T2DM(Reference Negrean, Stirban and Stratmann143). The same effect (impaired flow-mediated macrovascular dilatation and endothelial dysfunction) was observed in a study in which forty-four subjects with T2DM and ten subjects without diabetes were administered a single oral AGE challenge(Reference Uribarri, Stirban, Sander, Cai, Negrean, Buenting, Koschinsky and Vlassara144).

When subjects with chronic renal insufficiency were assigned to either a high- or low-AGE diet the low intake was found to be associated with a decrease in AGE, CRP, plasminogen-activator inhibitor-1, vascular-cell adhesion molecule 1 and TNFα levels(Reference Peppa, Uribarri, Cai, Lu and Vlassara145). The importance of this finding is not clear, however, as a large cross-sectional study of >300 patients on haemodialysis that analysed the relationships between AGE and CRP and all-cause as well as cardiovascular mortality has shown better survival of subjects with high total serum fluorescent AGE and CML levels despite higher levels of inflammation markers (CRP)(Reference Schwedler, Metzger, Schinzel and Wanner146). The authors suggest that high serum AGE in patients on haemodialysis with better survival is either an epiphenomenon or reflects a better nutritional status. An inverse relationship between circulating AGE and glomerular filtration rate has also been reported for subjects with T1DM as well for subjects without diabetes(Reference Stam, Schalkwijk, van Guldener, ter Wee and Stehouwer85, Reference Lieuw, van Hinsbergh, Teerlink, Barto, Twisk, Stehouwer and Schalkwijk147); however, a correlation with inflammatory markers was only present in subjects with diabetes. An even larger study that followed 450 subjects with T2DM and diabetic nephropathy for 2·6 years has ruled out serum CML level as an independent risk factor for cardiovascular or renal outcomes(Reference Busch, Franke, Wolf, Brandstadt, Ott, Gerth, Hunsicker and Stein148).

Interesting findings have been reported from a study of plasma AGE in vegetarians compared with subjects on a typical Western diet. Vegetarians were found to have unexpectedly higher plasma AGE, although not in association with typical AGE toxicity such as induction of insulin resistance (glycaemia), nephrotoxicity (glomerular filtration rate) or inflammation (CRP)(Reference Sebekova, Krajcoviova-Kudlackova, Schinzel, Faist, Klvanova and Heidland149). On further investigation of the components of the metabolic syndrome in vegetarians and omnivores(Reference Sebekova, Boor, Valachovicova, Blazicek, Parrak, Babinska, Heidland and Krajcovicova-Kudlackova150), the authors have proposed that this finding might be the result of a higher intake of fructose (in fruit), which is a better glycating agent that glucose, and that the potential AGE toxicity may be counterbalanced by higher levels of antioxidants often detected in vegetarians.

Advanced glycation end products v. (nutri)genomics and (nutri)genetics

One of the greatest changes preceding and accompanying the current increase in complex diseases has been the change in the human diet, its quality as well as its quantity. Beyond the role of diet as an energy source, micronutrients (vitamins, Ca, Fe etc.) and macronutrients (NEFA, cholesterol, glucose) are potent environmental signals that influence cellular metabolic programming, homeostasis and gene expression. Nutrigenomics is a relatively new discipline that seeks to provide an understanding of how nutrients affect gene expression(Reference Ordovas and Mooser151), and studies of the effect of dietary AGE are no exception. Experiments performed in vitro with food-derived AGE have confirmed their potent effect, and the previously observed biological effects of AGE could theoretically also apply to food-derived AGE. AGE derived from common thermally-processed foods (animal products, vegetables, starches) have been studied in vitro before ingestion and have been shown to possess pro-oxidative (depletion of reduced glutathione, cross-link formation), pro-inflammatory (induction of TNFα) and signalling properties in endothelial cells (human umbilical vein endothelial cells)(Reference Cai, Gao, Zhu, Peppa, He and Vlassara152). Another study has investigated the effect of food compounds formed by heat treatment during processing of food on the expression of the RAGE and p44/42 mitogen-activated protein kinase activation(Reference Zill, Bek and Hofmann153). Dose-dependent activation of RAGE signal transduction pathways was found in response to food-derived AGE and other thermally-produced compounds, as well as inhibition of this activation by pre-incubation with anti-RAGE antibody or in cells expressing C-terminally-truncated RAGE. Nevertheless, because of the limited bioavailability of AGE the true effect of food-derived AGE is probably minor in healthy subjects, although definitely worthy of further investigation in vulnerable groups such as infants and patients with diabetes, uraemia and bowel disease.

Another important issue is to understand how genetic variability in relevant genes modulates the effect of dietary components on specific phenotypes, which is the working definition of nutrigenetics. The current human genome has evolved under environmental influences that were, until recently, predominantly harsh rather than hospitable. The resulting metabolic ‘thriftiness’ is therefore a logical outcome, but undesirable against a background of affluence. Some of the best examples of the nutrigenetic consequences come from lipid metabolism, i.e. how genetic polymorphisms in genes encoding proteins involved in lipoprotein metabolism modulate the effect of dietary lipid intake on plasma lipid levels or the response to lipid-lowering interventions. In the context of both endogenously-formed AGE and food-derived AGE genes with potential (nutri)genetic importance are (1) those encoding enzymes detoxifying AGE and their precursors (e.g. glyoxalase system or FN3K) and (2) those encoding AGE-binding scavenger and signalling receptors (e.g. RAGE, galectin-3 etc.). The former group, i.e. genetic variability in deglycating enzymes resulting in their functional variability, could contribute to the observed inter-individual variation in AGE levels that was found in a study of twins without diabetes(Reference Leslie, Beyan, Sawtell, Boehm, Spector and Snieder154). Approximately 74% heritability of AGE levels (assessed by serum CML) was found, which was independent of heritability of fasting glucose or HbA1c. The notion that genetic factors play an important role in determining AGE levels in the healthy state further emphasises their pathophysiological potential in diseases. However, in a study of the common polymorphism A111E in the glyoxalase I in diabetic nephropathy no significant association has been found(Reference Kankova, Stejskalova and Pacal155). The activity of another important deglycating enzyme, FN3K, has been shown to exhibit a wide inter-individual variability influenced by the two polymorphisms in the FN3K gene(Reference Delpierre, Veiga-da-Cunha, Vertommen, Buysschaert and Van Schaftingen156). Neither enzyme activity nor frequency of the two gene variants was found to differ between groups of subjects with T1DM and without diabetes, although the relationship with diabetes complications was not studied.

The author's group has been involved in the study of genetic variability in the RAGE gene since soon after its discovery, and together with others has contributed to the identification of several common polymorphisms in the RAGE(Reference Kankova, Zahejsky, Marova, Muzik, Kuhrova, Blazkova, Znojil, Beranek and Vacha157) and some of their potentially interesting phenotypic outcomes, e.g. circulating levels of selected non-enzymic antioxidants(Reference Kankova, Marova, Zahejsky, Muzik, Stejskalova, Znojil and Vacha158). Of special interest is the modulation by certain genetic variants in RAGE of the genetic risk of the development of diabetic nephropathy in patients with diabetes, since diabetic nephropathy increases the already-enhanced AGE formation by impairment of their excretion. Using relatively novel approaches to genetic epidemiology suitable for studying complex phenotypes (e.g. haplotype analysis and multi-locus association studies) it has been shown that risk variants in the RAGE gene increase the susceptibility to diabetic nephropathy and accelerate its onset(Reference Kankova, Stejskalova and Pacal155, Reference Kankova, Stejskalova, Hertlova and Znojil159). Other studies have also identified RAGE variants as risk factors for diabetic nephropathy(Reference Pettersson-Fernholm, Forsblom, Hudson, Perola, Grant and Groop160, Reference Matsunaga-Irie, Maruyama, Yamamoto, Motohashi, Hirose, Shimada, Murata and Saruta161).

Conclusions: the double-hit or, alternatively, triple-hit hypothesis in diabetes mellitus?

Glycation represents the most common type of post-translation modification of protein residues. Methylglyoxal-derived hydroimidazolones are the most abundant AGE both as protein residues and free adducts. AGE formation is increased as a consequence of hyperglycaemia in diabetes. Cells contain enzymic systems to detoxify precursors of AGE (dicarbonyls and fructosamines) and functional insufficiency of these systems (e.g. as a result of genetic factors) can be of critical importance. AGE-modified proteins are subject to proteasomal degradation, free adducts are released from cells into the plasma and, under normal circumstances, are rapidly excreted in the urine. A decline in renal function, e.g. as a result of diabetic nephropathy, leads to retention of AGE and aggravation of their toxicity, which is mediated by direct modification of macromolecules by AGE and also through their binding to the cell surface receptors (RAGE and others). The resulting increase in the expression of pro-inflammatory genes contributes to the development of diabetes complications including diabetic nephropathy. A vicious cycle may then ensue whereby AGE contribute to the development of renal impairment, and once developed renal impairment may then contribute to further accumulation of AGE. Although there have been few studies of the biological effects of dietary AGE, current knowledge indicates that the impact of dietary AGE is relatively low in metabolically-healthy subjects with preserved renal function, because limited bioavailability means that AGE intake represents only a minor contribution compared with endogenously-produced AGE. However, in selected groups such as patients with diabetes (especially those with diabetic nephropathy) and uraemia AGE-rich diet can further aggravate AGE-mediated pathology.

In conclusion, accumulation of AGE is intimately associated with the chronic course of diabetes and represents an example of a self-amplifying pathophysiological mechanism. The data presented support the proposal that AGE act as a multiple-hit agent; although further studies are needed to assess their impact more precisely.

Acknowledgements

The author thanks the Grant Agency of the Czech Academy of Sciences (KJB501620601) and the Ministry of Health of the Czech Republic (NR 9443–3/2007) for the support of her work.

References

1. King, H, Aubert, RE & Herman, WH (1998) Global burden of diabetes, 1995–2025. prevalence, numerical estimates, and projections. Diabetes Care 21, 14141431.CrossRefGoogle ScholarPubMed
2. Brownlee, M (2001) Biochemistry and molecular cell biology of diabetic complications. Nature 414, 813820.CrossRefGoogle ScholarPubMed
3. Du, X, Edelstein, D, Obici, S, Higham, N, Zou, MH & Brownlee, M (2006) Insulin resistance reduces arterial prostacyclin synthase and eNOS activities by increasing endothelial fatty acid oxidation. J Clin Invest 116, 10711080.CrossRefGoogle ScholarPubMed
4. Ahmed, N & Thornalley, PJ (2007) Advanced glycation endproducts: what is their relevance to diabetic complications? Diabetes Obes Metab 9, 233245.CrossRefGoogle ScholarPubMed
5. Ahmed, N (2005) Advanced glycation endproducts – role in pathology of diabetic complications. Diabetes Res Clin Pract 67, 321.CrossRefGoogle ScholarPubMed
6. Stitt, AW & Curtis, TM (2005) Advanced glycation and retinal pathology during diabetes. Pharmacol Rep 57, Suppl., 156168.Google ScholarPubMed
7. Munch, G, Gasic-Milenkovic, J & Arendt, T (2003) Effect of advanced glycation endproducts on cell cycle and their relevance for Alzheimer's disease. J Neural Transm 65, Suppl., 6371.Google Scholar
8. Finot, PA (2005) Historical perspective of the Maillard reaction in food science. Ann NY Acad Sci 1043, 18.CrossRefGoogle ScholarPubMed
9. Gaudreault, N, Scriven, DR & Moore, ED (2004) Characterisation of glucose transporters in the intact coronary artery endothelium in rats: GLUT-2 upregulated by long-term hyperglycaemia. Diabetologia 47, 20812092.CrossRefGoogle ScholarPubMed
10. Kaiser, N, Sasson, S, Feener, EP, Boukobza-Vardi, N, Higashi, S, Moller, DE, Davidheiser, S, Przybylski, RJ & King, GL (1993) Differential regulation of glucose transport and transporters by glucose in vascular endothelial and smooth muscle cells. Diabetes 42, 8089.CrossRefGoogle ScholarPubMed
11. Brosius, FC & Heilig, CW (2005) Glucose transporters in diabetic nephropathy. Pediatr Nephrol 20, 447451.CrossRefGoogle ScholarPubMed
12. Linden, KC, DeHaan, CL, Zhang, Y, Glowacka, S, Cox, AJ, Kelly, DJ & Rogers, S (2006) Renal expression and localization of the facilitative glucose transporters GLUT1 and GLUT12 in animal models of hypertension and diabetic nephropathy. Am J Physiol Renal Physiol 290, F205F213.CrossRefGoogle ScholarPubMed
13. Brownlee, M (2005) The pathobiology of diabetic complications: a unifying mechanism. Diabetes 54, 16151625.CrossRefGoogle ScholarPubMed
14. Nishikawa, T, Edelstein, D, Du, XL et al. (2000) Normalizing mitochondrial superoxide production blocks three pathways of hyperglycaemic damage. Nature 404, 787790.CrossRefGoogle ScholarPubMed
15. Adam-Vizi, V & Chinopoulos, C (2006) Bioenergetics and the formation of mitochondrial reactive oxygen species. Trends Pharmacol Sci 27, 639645.CrossRefGoogle ScholarPubMed
16. Soriano, FG, Virag, L & Szabo, C (2001) Diabetic endothelial dysfunction: role of reactive oxygen and nitrogen species production and poly(ADP-ribose) polymerase activation. J Mol Med 79, 437448.CrossRefGoogle ScholarPubMed
17. Soriano, FG, Pacher, P, Mabley, J, Liaudet, L & Szabo, C (2001) Rapid reversal of the diabetic endothelial dysfunction by pharmacological inhibition of poly(ADP-ribose) polymerase. Circ Res 89, 684691.CrossRefGoogle ScholarPubMed
18. Du, X, Matsumura, T, Edelstein, D, Rossetti, L, Zsengeller, Z, Szabo, C & Brownlee, M (2003) Inhibition of GAPDH activity by poly(ADP-ribose) polymerase activates three major pathways of hyperglycemic damage in endothelial cells. J Clin Invest 112, 10491057.CrossRefGoogle ScholarPubMed
19. Sirover, MA (2005) New nuclear functions of the glycolytic protein, glyceraldehyde-3-phosphate dehydrogenase, in mammalian cells. J Cell Biochem 95, 4552.CrossRefGoogle ScholarPubMed
20. Hara, MR, Cascio, MB & Sawa, A (2006) GAPDH as a sensor of NO stress. Biochim Biophys Acta 1762, 502509.CrossRefGoogle ScholarPubMed
21. Chuang, DM, Hough, C & Senatorov, VV (2005) Glyceraldehyde-3-phosphate dehydrogenase, apoptosis, and neurodegenerative diseases. Annu Rev Pharmacol Toxicol 45, 269290.CrossRefGoogle ScholarPubMed
22. Thornalley, PJ, Battah, S, Ahmed, N, Karachalias, N, Agalou, S, Babaei-Jadidi, R & Dawnay, A (2003) Quantitative screening of advanced glycation endproducts in cellular and extracellular proteins by tandem mass spectrometry. Biochem J 375, 581592.CrossRefGoogle ScholarPubMed
23. Vlassara, H & Palace, MR (2003) Glycoxidation: the menace of diabetes and aging. Mt Sinai J Med 70, 232241.Google Scholar
24. Baynes, JW & Thorpe, SR (2000) Glycoxidation and lipoxidation in atherogenesis. Free Radic Biol Med 28, 17081716.CrossRefGoogle ScholarPubMed
25. Munch, G, Deuther-Conrad, W & Gasic-Milenkovic, J (2002) Glycoxidative stress creates a vicious cycle of neurodegeneration in Alzheimer's disease – a target for neuroprotective treatment strategies? J Neural Transm 62, Suppl., 303307.Google Scholar
26. DeGroot, J (2004) The AGE of the matrix: chemistry, consequence and cure. Curr Opin Pharmacol 4, 301305.CrossRefGoogle ScholarPubMed
27. Thornalley, PJ (2006) Advanced glycation end products in renal failure. J Ren Nutr 16, 178184.CrossRefGoogle ScholarPubMed
28. Rahbar, S (1968) An abnormal hemoglobin in red cells of diabetics. Clin Chim Acta 22, 296298.CrossRefGoogle ScholarPubMed
29. Pickup, JC (2003) Diabetic control and its measurement. In Textbook of Diabetes, 3rd ed., vol. 1, 34.31–34.17 [Pickup, JC and Williams, Geditors]. Oxford: Blackwell Science.Google Scholar
30. Ahmed, N & Thornalley, PJ (2003) Quantitative screening of protein biomarkers of early glycation, advanced glycation, oxidation and nitrosation in cellular and extracellular proteins by tandem mass spectrometry multiple reaction monitoring. Biochem Soc Trans 31, 14171422.CrossRefGoogle ScholarPubMed
31. Hayashi, T & Namiki, M (1980) Formation of two-carbon sugar fragments at an early stage of the browning reaction of sugar and amine. Agric Biol Chem 44, 25752580.Google Scholar
32. Delpierre, G, Rider, MH, Collard, F, Stroobant, V, Vanstapel, F, Santos, HI & Van Schaftingen, E (2000) Identification, cloning, and heterologous expression of a mammalian fructosamine-3-kinase. Diabetes 49, 16271634.CrossRefGoogle ScholarPubMed
33. Szwergold, BS, Howell, S & Beisswenger, PJ (2001) Human fructosamine-3-kinase: purification, sequencing, substrate specificity, and evidence of activity in vivo. Diabetes 50, 21392147.CrossRefGoogle ScholarPubMed
34. Beisswenger, PJ, Howell, SK, Smith, K & Szwergold, BS (2003) Glyceraldehyde-3-phosphate dehydrogenase activity as an independent modifier of methylglyoxal levels in diabetes. Biochim Biophys Acta 1637, 98106.CrossRefGoogle ScholarPubMed
35. Ahmed, N, Babaei-Jadidi, R, Howell, SK, Beisswenger, PJ & Thornalley, PJ (2005) Degradation products of proteins damaged by glycation, oxidation and nitration in clinical type 1 diabetes. Diabetologia 48, 15901603.CrossRefGoogle ScholarPubMed
36. Ahmed, N, Babaei-Jadidi, R, Howell, SK, Thornalley, PJ & Beisswenger, PJ (2005) Glycated and oxidized protein degradation products are indicators of fasting and postprandial hyperglycemia in diabetes. Diabetes Care 28, 24652471.CrossRefGoogle ScholarPubMed
37. Beisswenger, PJ, Howell, SK, O'Dell, RM, Wood, ME, Touchette, AD & Szwergold, BS (2001) alpha-Dicarbonyls increase in the postprandial period and reflect the degree of hyperglycemia. Diabetes Care 24, 726732.CrossRefGoogle ScholarPubMed
38. Phillips, SA & Thornalley, PJ (1993) The formation of methylglyoxal from triose phosphates. Investigation using a specific assay for methylglyoxal. Eur J Biochem 212, 101105.CrossRefGoogle ScholarPubMed
39. Thornalley, PJ (2003) The enzymatic defence against glycation in health, disease and therapeutics: a symposium to examine the concept. Biochem Soc Trans 31, 13411342.CrossRefGoogle ScholarPubMed
40. Szwergold, BS (2005) Intrinsic toxicity of glucose, due to non-enzymatic glycation, is controlled in-vivo by deglycation systems including: FN3K-mediated deglycation of fructosamines and transglycation of aldosamines. Med Hypotheses 65, 337348.CrossRefGoogle ScholarPubMed
41. Thornalley, PJ (2003) Glyoxalase I – structure, function and a critical role in the enzymatic defence against glycation. Biochem Soc Trans 31, 13431348.CrossRefGoogle Scholar
42. Shinohara, M, Thornalley, PJ, Giardino, I, Beisswenger, P, Thorpe, SR, Onorato, J & Brownlee, M (1998) Overexpression of glyoxalase-I in bovine endothelial cells inhibits intracellular advanced glycation endproduct formation and prevents hyperglycemia-induced increases in macromolecular endocytosis. J Clin Invest 101, 11421147.CrossRefGoogle ScholarPubMed
43. Conner, JR, Beisswenger, PJ & Szwergold, BS (2005) Some clues as to the regulation, expression, function, and distribution of fructosamine-3-kinase and fructosamine-3-kinase-related protein. Ann NY Acad Sci 1043, 824836.CrossRefGoogle Scholar
44. Yu, Y, Thorpe, SR, Jenkins, AJ et al. (2006) Advanced glycation end-products and methionine sulphoxide in skin collagen of patients with type 1 diabetes. Diabetologia 49, 24882498.CrossRefGoogle ScholarPubMed
45. Avery, NC & Bailey, AJ (2006) The effects of the Maillard reaction on the physical properties and cell interactions of collagen. Pathol Biol (Paris) 54, 387395.CrossRefGoogle ScholarPubMed
46. Dobler, D, Ahmed, N, Song, L, Eboigbodin, KE & Thornalley, PJ (2006) Increased dicarbonyl metabolism in endothelial cells in hyperglycemia induces anoikis and impairs angiogenesis by RGD and GFOGER motif modification. Diabetes 55, 19611969.CrossRefGoogle ScholarPubMed
47. Ahmed, N & Thornalley, PJ (2005) Peptide mapping of human serum albumin modified minimally by methylglyoxal in vitro and in vivo. Ann NY Acad Sci 1043, 260266.CrossRefGoogle ScholarPubMed
48. Garlick, RL & Mazer, JS (1983) The principal site of nonenzymatic glycosylation of human serum albumin in vivo. J Biol Chem 258, 61426146.CrossRefGoogle ScholarPubMed
49. Gao, Y & Wang, Y (2006) Site-selective modifications of arginine residues in human hemoglobin induced by methylglyoxal. Biochemistry 45, 1565415660.CrossRefGoogle ScholarPubMed
50. Sen, S, Kar, M, Roy, A & Chakraborti, AS (2005) Effect of nonenzymatic glycation on functional and structural properties of hemoglobin. Biophys Chem 113, 289298.CrossRefGoogle ScholarPubMed
51. Scheubel, RJ, Kahrstedt, S, Weber, H, Holtz, J, Friedrich, I, Borgermann, J, Silber, RE & Simm, A (2006) Depression of progenitor cell function by advanced glycation endproducts (AGEs): potential relevance for impaired angiogenesis in advanced age and diabetes. Exp Gerontol 41, 540548.CrossRefGoogle ScholarPubMed
52. Stopper, H, Schinzel, R, Sebekova, K & Heidland, A (2003) Genotoxicity of advanced glycation end products in mammalian cells. Cancer Lett 190, 151156.CrossRefGoogle ScholarPubMed
53. Schupp, N, Schinzel, R, Heidland, A & Stopper, H (2005) Genotoxicity of advanced glycation end products: involvement of oxidative stress and of angiotensin II type 1 receptors. Ann NY Acad Sci 1043, 685695.CrossRefGoogle ScholarPubMed
54. Fragedaki, E, Nebel, M, Schupp, N et al. (2005) Genomic damage and circulating AGE levels in patients undergoing daily versus standard haemodialysis. Nephrol Dial Transplant 20, 19361943.CrossRefGoogle ScholarPubMed
55. Neeper, M, Schmidt, AM, Brett, J, Yan, SD, Wang, F, Pan, YC, Elliston, K, Stern, D & Shaw, A (1992) Cloning and expression of a cell surface receptor for advanced glycosylation end products of proteins. J Biol Chem 267, 1499815004.CrossRefGoogle ScholarPubMed
56. Schmidt, AM, Vianna, M, Gerlach, M, Brett, J, Ryan, J, Kao, J, Esposito, C, Hegarty, H, Hurley, W & Clauss, M (1992) Isolation and characterization of two binding proteins for advanced glycosylation end products from bovine lung which are present on the endothelial cell surface. J Biol Chem 267, 1498714997.CrossRefGoogle ScholarPubMed
57. Vlassara, H, Li, YM, Imani, F, Wojciechowicz, D, Yang, Z, Liu, FT & Cerami, A (1995) Identification of galectin-3 as a high-affinity binding protein for advanced glycation end products (AGE): a new member of the AGE-receptor complex. Mol Med 1, 634646.CrossRefGoogle ScholarPubMed
58. Li, YM, Mitsuhashi, T, Wojciechowicz, D, Shimizu, N, Li, J, Stitt, A, He, C, Banerjee, D & Vlassara, H (1996) Molecular identity and cellular distribution of advanced glycation endproduct receptors: relationship of p60 to OST-48 and p90 to 80K-H membrane proteins. Proc Natl Acad Sci USA 93, 1104711052.CrossRefGoogle ScholarPubMed
59. Araki, N, Higashi, T, Mori, T, Shibayama, R, Kawabe, Y, Kodama, T, Takahashi, K, Shichiri, M & Horiuchi, S (1995) Macrophage scavenger receptor mediates the endocytic uptake and degradation of advanced glycation end products of the Maillard reaction. Eur J Biochem 230, 408415.CrossRefGoogle ScholarPubMed
60. Horiuchi, S, Sakamoto, Y & Sakai, M (2003) Scavenger receptors for oxidized and glycated proteins. Amino Acids 25, 283292.CrossRefGoogle ScholarPubMed
61. Schmidt, AM, Vianna, M, Gerlach, M, Brett, J, Ryan, J, Kao, J, Esposito, C, Hegarty, H, Hurley, W & Clauss, M (1992) Isolation and characterization of two binding proteins for advanced glycosylation end products from bovine lung which are present on the endothelial cell surface. J Biol Chem 267, 1498714997.CrossRefGoogle ScholarPubMed
62. Hofmann, MA, Drury, S, Fu, C et al. (1999) RAGE mediates a novel proinflammatory axis: a central cell surface receptor for S100/calgranulin polypeptides. Cell 97, 889901.CrossRefGoogle Scholar
63. Hori, O, Brett, J, Slattery, T et al. (1995) The receptor for advanced glycation end products (RAGE) is a cellular binding site for amphoterin. Mediation of neurite outgrowth and co-expression of rage and amphoterin in the developing nervous system. J Biol Chem 270, 2575225761.CrossRefGoogle ScholarPubMed
64. Yan, SD, Chen, X, Fu, J et al. (1996) RAGE and amyloid-beta peptide neurotoxicity in Alzheimer's disease. Nature 382, 685691.CrossRefGoogle ScholarPubMed
65. Yan, SF, Barile, GR, D'Agati, V, Du Yan, S, Ramasamy, R & Schmidt, AM (2007) The biology of RAGE and its ligands: uncovering mechanisms at the heart of diabetes and its complications. Curr Diab Rep 7, 146153.CrossRefGoogle ScholarPubMed
66. Bierhaus, A, Humpert, PM, Morcos, M, Wendt, T, Chavakis, T, Arnold, B, Stern, DM & Nawroth, PP (2005) Understanding RAGE, the receptor for advanced glycation end products. J Mol Med 83, 876886.CrossRefGoogle ScholarPubMed
67. Hudson, BI, Wendt, T, Bucciarelli, LG, Rong, LL, Naka, Y, Yan, SF & Schmidt, AM (2005) Diabetic vascular disease: it's all the RAGE. Antioxid Redox Signal 7, 15881600.CrossRefGoogle ScholarPubMed
68. Bierhaus, A, Stern, DM & Nawroth, PP (2006) RAGE in inflammation: a new therapeutic target? Curr Opin Investig Drugs 7, 985991.Google ScholarPubMed
69. Bierhaus, A, Schiekofer, S, Schwaninger, M et al. (2001) Diabetes-associated sustained activation of the transcription factor nuclear factor-kappaB. Diabetes 50, 27922808.CrossRefGoogle ScholarPubMed
70. Malherbe, P, Richards, JG, Gaillard, H, Thompson, A, Diener, C, Schuler, A & Huber, G (1999) cDNA cloning of a novel secreted isoform of the human receptor for advanced glycation end products and characterization of cells co-expressing cell-surface scavenger receptors and Swedish mutant amyloid precursor protein. Brain Res Mol Brain Res 71, 159170.CrossRefGoogle ScholarPubMed
71. Schlueter, C, Hauke, S, Flohr, AM, Rogalla, P & Bullerdiek, J (2003) Tissue-specific expression patterns of the RAGE receptor and its soluble forms – a result of regulated alternative splicing? Biochim Biophys Acta 1630, 16.CrossRefGoogle ScholarPubMed
72. Yonekura, H, Yamamoto, Y, Sakurai, S et al. (2003) Novel splice variants of the receptor for advanced glycation end-products expressed in human vascular endothelial cells and pericytes, and their putative roles in diabetes-induced vascular injury. Biochem J 370, 10971109.CrossRefGoogle ScholarPubMed
73. Falcone, C, Emanuele, E, D'Angelo, A, Buzzi, MP, Belvito, C, Cuccia, M & Geroldi, D (2005) Plasma levels of soluble receptor for advanced glycation end products and coronary artery disease in nondiabetic men. Arterioscler Thromb Vasc Biol 25, 10321037.CrossRefGoogle ScholarPubMed
74. Koyama, H, Shoji, T, Yokoyama, H et al. (2005) Plasma level of endogenous secretory RAGE is associated with components of the metabolic syndrome and atherosclerosis. Arterioscler Thromb Vasc Biol 25, 25872593.CrossRefGoogle ScholarPubMed
75. Geroldi, D, Falcone, C, Emanuele, E, D'Angelo, A, Calcagnino, M, Buzzi, MP, Scioli, GA & Fogari, R (2005) Decreased plasma levels of soluble receptor for advanced glycation end-products in patients with essential hypertension. J Hypertens 23, 17251729.CrossRefGoogle ScholarPubMed
76. Hayaishi-Okano, R, Yamasaki, Y, Kajimoto, Y et al. (2003) Association of NAD(P)H oxidase p22 phox gene variation with advanced carotid atherosclerosis in Japanese type 2 diabetes. Diabetes Care 26, 458463.CrossRefGoogle ScholarPubMed
77. Sakurai, S, Yamamoto, Y, Tamei, H et al. (2006) Development of an ELISA for esRAGE and its application to type 1 diabetic patients. Diabetes Res Clin Pract 73, 158165.CrossRefGoogle ScholarPubMed
78. Pullerits, R, Bokarewa, M, Dahlberg, L & Tarkowski, A (2005) Decreased levels of soluble receptor for advanced glycation end products in patients with rheumatoid arthritis indicating deficient inflammatory control. Arthritis Res Ther 7, R817R824.CrossRefGoogle ScholarPubMed
79. Emanuele, E, D'Angelo, A, Tomaino, C, Binetti, G, Ghidoni, R, Politi, P, Bernardi, L, Maletta, R, Bruni, AC & Geroldi, D (2005) Circulating levels of soluble receptor for advanced glycation end products in Alzheimer disease and vascular dementia. Arch Neurol 62, 17341736.CrossRefGoogle ScholarPubMed
80. Goldberg, AL (2003) Protein degradation and protection against misfolded or damaged proteins. Nature 426, 895899.CrossRefGoogle ScholarPubMed
81. Shringarpure, R, Grune, T, Mehlhase, J & Davies, KJ (2003) Ubiquitin conjugation is not required for the degradation of oxidized proteins by proteasome. J Biol Chem 278, 311318.CrossRefGoogle Scholar
82. Bulteau, AL, Verbeke, P, Petropoulos, I, Chaffotte, AF & Friguet, B (2001) Proteasome inhibition in glyoxal-treated fibroblasts and resistance of glycated glucose-6-phosphate dehydrogenase to 20 S proteasome degradation in vitro. J Biol Chem 276, 4566245668.CrossRefGoogle ScholarPubMed
83. Carrard, G, Dieu, M, Raes, M, Toussaint, O & Friguet, B (2003) Impact of ageing on proteasome structure and function in human lymphocytes. Int J Biochem Cell Biol 35, 728739.CrossRefGoogle ScholarPubMed
84. Stolzing, A, Widmer, R, Jung, T, Voss, P & Grune, T (2006) Degradation of glycated bovine serum albumin in microglial cells. Free Radic Biol Med 40, 10171027.CrossRefGoogle ScholarPubMed
85. Stam, F, Schalkwijk, CG, van Guldener, C, ter Wee, PM & Stehouwer, CD (2006) Advanced glycation end-product peptides are associated with impaired renal function, but not with biochemical markers of endothelial dysfunction and inflammation, in non-diabetic individuals. Nephrol Dial Transplant 21, 677682.CrossRefGoogle Scholar
86. Gugliucci, A & Bendayan, M (1996) Renal fate of circulating advanced glycated end products (AGE): evidence for reabsorption and catabolism of AGE-peptides by renal proximal tubular cells. Diabetologia 39, 149160.CrossRefGoogle ScholarPubMed
87. Agalou, S, Ahmed, N, Babaei-Jadidi, R, Dawnay, A & Thornalley, PJ (2005) Profound mishandling of protein glycation degradation products in uremia and dialysis. J Am Soc Nephrol 16, 14711485.CrossRefGoogle ScholarPubMed
88. Odetti, P, Cosso, L, Pronzato, MA, Dapino, D & Gurreri, G (1995) Plasma advanced glycosylation end-products in maintenance haemodialysis patients. Nephrol Dial Transplant 10, 21102113.Google ScholarPubMed
89. Friedlander, MA, Wu, YC, Schulak, JA, Monnier, VM & Hricik, DE (1995) Influence of dialysis modality on plasma and tissue concentrations of pentosidine in patients with end-stage renal disease. Am J Kidney Dis 25, 445451.CrossRefGoogle ScholarPubMed
90. Miyata, T, Ueda, Y, Yoshida, A, Sugiyama, S, Iida, Y, Jadoul, M & Maeda, K, Kurokawa, K & van Ypersele de Strihou, C (1997) Clearance of pentosidine, an advanced glycation end product, by different modalities of renal replacement therapy. Kidney Int 51, 880887.CrossRefGoogle ScholarPubMed
91. Sebekova, K, Podracka, L, Blazicek, P, Syrova, D, Heidland, A & Schinzel, R (2001) Plasma levels of advanced glycation end products in children with renal disease. Pediatr Nephrol 16, 11051112.CrossRefGoogle ScholarPubMed
92. Miyata, T, Ueda, Y, Shinzato, T, Iida, Y, Tanaka, S, Kurokawa, K, van Ypersele de Strihou, C & Maeda, K (1996) Accumulation of albumin-linked and free-form pentosidine in the circulation of uremic patients with end-stage renal failure: renal implications in the pathophysiology of pentosidine. J Am Soc Nephrol 7, 11981206.CrossRefGoogle ScholarPubMed
93. Miyata, T, Fu, MX, Kurokawa, K, van Ypersele de Strihou, C, Thorpe, SR & Baynes, JW (1998) Autoxidation products of both carbohydrates and lipids are increased in uremic plasma: is there oxidative stress in uremia? Kidney Int 54, 12901295.CrossRefGoogle ScholarPubMed
94. Kalousova, M, Zima, T, Tesar, V, Stipek, S & Sulkova, S (2004) Advanced glycation end products in clinical nephrology. Kidney Blood Press Res 27, 1828.CrossRefGoogle ScholarPubMed
95. Miyata, T, van Ypersele de Strihou, C, Kurokawa, K & Baynes, JW (1999) Alterations in nonenzymatic biochemistry in uremia: origin and significance of ‘carbonyl stress’ in long-term uremic complications. Kidney Int 55, 389399.CrossRefGoogle ScholarPubMed
96. Smedsrod, B, Melkko, J, Araki, N, Sano, H & Horiuchi, S (1997) Advanced glycation end products are eliminated by scavenger-receptor-mediated endocytosis in hepatic sinusoidal Kupffer and endothelial cells. Biochem J 322, 567573.CrossRefGoogle ScholarPubMed
97. Hansen, B, Svistounov, D, Olsen, R, Nagai, R, Horiuchi, S & Smedsrod, B (2002) Advanced glycation end products impair the scavenger function of rat hepatic sinusoidal endothelial cells. Diabetologia 45, 13791388.Google ScholarPubMed
98. Sebekova, K, Kupcova, V, Schinzel, R & Heidland, A (2002) Markedly elevated levels of plasma advanced glycation end products in patients with liver cirrhosis – amelioration by liver transplantation. J Hepatol 36, 6671.CrossRefGoogle ScholarPubMed
99. Ahmed, N, Thornalley, PJ, Luthen, R, Haussinger, D, Sebekova, K, Schinzel, R, Voelker, W & Heidland, A (2004) Processing of protein glycation, oxidation and nitrosation adducts in the liver and the effect of cirrhosis. J Hepatol 41, 913919.CrossRefGoogle ScholarPubMed
100. Forbes, JM, Fukami, K & Cooper, ME (2007) Diabetic nephropathy: where hemodynamics meets metabolism. Exp Clin Endocrinol Diabetes 115, 6984.CrossRefGoogle ScholarPubMed
101. Gurley, SB & Coffman, TM (2007) The renin-angiotensin system and diabetic nephropathy. Semin Nephrol 27, 144152.CrossRefGoogle ScholarPubMed
102. Rossing, P, Parving, HH & de Zeeuw, D (2006) Renoprotection by blocking the RAAS in diabetic nephropathy – fact or fiction? Nephrol Dial Transplant 21, 23542357.CrossRefGoogle ScholarPubMed
103. Jacobsen, PK (2005) Preventing end-stage renal disease in diabetic patients – dual blockade of the renin-angiotensin system (Part II). J Renin Angiotensin Aldosterone Syst 6, 5568.CrossRefGoogle ScholarPubMed
104. Thomas, MC, Forbes, JM & Cooper, ME (2005) Advanced glycation end products and diabetic nephropathy. Am J Ther 12, 562572.CrossRefGoogle ScholarPubMed
105. Tanji, N, Markowitz, GS, Fu, C, Kislinger, T, Taguchi, A, Pischetsrieder, M, Stern, D, Schmidt, AM & D‘Agati, VD (2000) Expression of advanced glycation end products and their cellular receptor RAGE in diabetic nephropathy and nondiabetic renal disease. J Am Soc Nephrol 11, 16561666.CrossRefGoogle ScholarPubMed
106. Oldfield, MD, Bach, LA, Forbes, JM, Nikolic-Paterson, D, McRobert, A, Thallas, V, Atkins, RC, Osicka, T, Jerums, G & Cooper, ME (2001) Advanced glycation end products cause epithelial-myofibroblast transdifferentiation via the receptor for advanced glycation end products (RAGE). J Clin Invest 108, 18531863.CrossRefGoogle ScholarPubMed
107. Wendt, TM, Tanji, N, Guo, J et al. (2003) RAGE drives the development of glomerulosclerosis and implicates podocyte activation in the pathogenesis of diabetic nephropathy. Am J Pathol 162, 11231137.CrossRefGoogle ScholarPubMed
108. Bonnardel-Phu, E, Wautier, JL, Schmidt, AM, Avila, C & Vicaut, E (1999) Acute modulation of albumin microvascular leakage by advanced glycation end products in microcirculation of diabetic rats in vivo. Diabetes 48, 20522058.CrossRefGoogle ScholarPubMed
109. Yamamoto, Y, Kato, I, Doi, T et al. (2001) Development and prevention of advanced diabetic nephropathy in RAGE-overexpressing mice. J Clin Invest 108, 261268.CrossRefGoogle ScholarPubMed
110. Morcos, M, Sayed, AA, Bierhaus, A et al. (2002) Activation of tubular epithelial cells in diabetic nephropathy. Diabetes 51, 35323544.CrossRefGoogle ScholarPubMed
111. Babaei-Jadidi, R, Karachalias, N, Ahmed, N, Battah, S & Thornalley, PJ (2003) Prevention of incipient diabetic nephropathy by high-dose thiamine and benfotiamine. Diabetes 52, 21102120.CrossRefGoogle ScholarPubMed
112. Degenhardt, TP, Alderson, NL, Arrington, DD, Beattie, RJ, Basgen, JM, Steffes, MW, Thorpe, SR & Baynes, JW (2002) Pyridoxamine inhibits early renal disease and dyslipidemia in the streptozotocin-diabetic rat. Kidney Int 61, 939950.CrossRefGoogle ScholarPubMed
113. Forbes, JM, Thallas, V, Thomas, MC, Founds, HW, Burns, WC, Jerums, G & Cooper, ME (2003) The breakdown of preexisting advanced glycation end products is associated with reduced renal fibrosis in experimental diabetes. FASEB J 17, 17621764.CrossRefGoogle ScholarPubMed
114. Thallas-Bonke, V, Lindschau, C, Rizkalla, B, Bach, LA, Boner, G, Meier, M, Haller, H, Cooper, ME & Forbes, JM (2004) Attenuation of extracellular matrix accumulation in diabetic nephropathy by the advanced glycation end product cross-link breaker ALT-711 via a protein kinase C-alpha-dependent pathway. Diabetes 53, 29212930.CrossRefGoogle Scholar
115. Flyvbjerg, A, Denner, L, Schrijvers, BF, Tilton, RG, Mogensen, TH, Paludan, SR & Rasch, R (2004) Long-term renal effects of a neutralizing RAGE antibody in obese type 2 diabetic mice. Diabetes 53, 166172.CrossRefGoogle ScholarPubMed
116. Maillard, L (1912) Action des acides aminés sur les sucres: formation des mélanoidines par voie méthodique (Action of amino acids on sugars: pathway of formation of melanoidins). C R Acad Sci 154, 6668.Google Scholar
117. Somoza, V (2005) Five years of research on health risks and benefits of Maillard reaction products: an update. Mol Nutr Food Res 49, 663672.CrossRefGoogle ScholarPubMed
118. Ahmed, N, Mirshekar-Syahkal, B, Kennish, L, Karachalias, N, Babaei-Jadidi, R & Thornalley, PJ (2005) Assay of advanced glycation endproducts in selected beverages and food by liquid chromatography with tandem mass spectrometric detection. Mol Nutr Food Res 49, 691699.CrossRefGoogle ScholarPubMed
119. Goldberg, T, Cai, W, Peppa, M, Dardaine, V, Baliga, BS, Uribarri, J & Vlassara, H (2004) Advanced glycoxidation end products in commonly consumed foods. J Am Diet Assoc 104, 12871291.CrossRefGoogle ScholarPubMed
120. Henle, T (2005) Protein-bound advanced glycation endproducts (AGEs) as bioactive amino acid derivatives in foods. Amino Acids 29, 313322.CrossRefGoogle ScholarPubMed
121. Koschinsky, T, He, CJ, Mitsuhashi, T, Bucala, R, Liu, C, Buenting, C, Heitmann, K & Vlassara, H (1997) Orally absorbed reactive glycation products (glycotoxins): an environmental risk factor in diabetic nephropathy. Proc Natl Acad Sci USA 94, 64746479.CrossRefGoogle ScholarPubMed
122. Faist, V & Erbersdobler, HF (2001) Metabolic transit and in vivo effects of melanoidins and precursor compounds deriving from the Maillard reaction. Ann Nutr Metab 45, 112.CrossRefGoogle ScholarPubMed
123. Faist, V, Wenzel, E, Randel, G et al. (2000) In vitro and in vivo studies on the metabolic transit of N-carboxymethyllysine. Czech J Food Sci 18, 116119.Google Scholar
124. Foerster, A & Henle, T (2003) Glycation in food and metabolic transit of dietary AGEs (advanced glycation end-products): studies on the urinary excretion of pyrraline. Biochem Soc Trans 31, 13831385.CrossRefGoogle ScholarPubMed
125. Forster, A, Kuhne, Y & Henle, T (2005) Studies on absorption and elimination of dietary maillard reaction products. Ann NY Acad Sci 1043, 474481.CrossRefGoogle ScholarPubMed
126. Bergmann, R, Helling, R, Heichert, C, Scheunemann, M, Mading, P, Wittrisch, H, Johannsen, B & Henle, T (2001) Radio fluorination and positron emission tomography (PET) as a new approach to study the in vivo distribution and elimination of the advanced glycation endproducts N epsilon-carboxymethyllysine (CML) and N epsilon-carboxyethyllysine (CEL). Nahrung 45, 182188.3.0.CO;2-Q>CrossRefGoogle Scholar
127. Wiame, E, Delpierre, G, Collard, F & Van Schaftingen, E (2002) Identification of a pathway for the utilization of the Amadori product fructoselysine in Escherichia coli. J Biol Chem 277, 4252342529.CrossRefGoogle ScholarPubMed
128. Tuohy, KM, Hinton, DJ, Davies, SJ, Crabbe, MJ, Gibson, GR & Ames, JM (2006) Metabolism of Maillard reaction products by the human gut microbiota – implications for health. Mol Nutr Food Res 50, 847857.CrossRefGoogle ScholarPubMed
129. Sebekova, K, Hofmann, T, Boor, P, Sebekova, K Jr, Ulicna, O, Erbersdobler, HF, Baynes, JW, Thorpe, SR, Heidland, A & Somoza, V (2005) Renal effects of oral maillard reaction product load in the form of bread crusts in healthy and subtotally nephrectomized rats. Ann NY Acad Sci 1043, 482491.CrossRefGoogle ScholarPubMed
130. Sandu, O, Song, K, Cai, W, Zheng, F, Uribarri, J & Vlassara, H (2005) Insulin resistance and type 2 diabetes in high-fat-fed mice are linked to high glycotoxin intake. Diabetes 54, 23142319.CrossRefGoogle ScholarPubMed
131. Hofmann, SM, Dong, HJ, Li, Z, Cai, W, Altomonte, J, Thung, SN, Feng, Z, Fisher, E & Vlassara, H (2002) Improved insulin sensitivity is associated with restricted intake of dietary glycoxidation products in the db/db mouse. Diabetes 51, 20822089.CrossRefGoogle ScholarPubMed
132. Zheng, F, He, C, Cai, W, Hattori, M, Steffes, M & Vlassara, H (2002) Prevention of diabetic nephropathy in mice by a diet low in glycoxidation products. Diabetes Metab Res Rev 18, 224237.CrossRefGoogle ScholarPubMed
133. Peppa, M, He, C, Hattori, M, McEvoy, R, Zheng, F & Vlassara, H (2003) Fetal or neonatal low-glycotoxin environment prevents autoimmune diabetes in NOD mice. Diabetes 52, 14411448.CrossRefGoogle ScholarPubMed
134. Peppa, M, Brem, H, Ehrlich, P, Zhang, JG, Cai, W, Li, Z, Croitoru, A, Thung, S & Vlassara, H (2003) Adverse effects of dietary glycotoxins on wound healing in genetically diabetic mice. Diabetes 52, 28052813.CrossRefGoogle ScholarPubMed
135. Lin, RY, Reis, ED, Dore, AT, Lu, M, Ghodsi, N, Fallon, JT, Fisher, EA & Vlassara, H (2002) Lowering of dietary advanced glycation endproducts (AGE) reduces neointimal formation after arterial injury in genetically hypercholesterolemic mice. Atherosclerosis 163, 303311.CrossRefGoogle ScholarPubMed
136. Lin, RY, Choudhury, RP, Cai, W, Lu, M, Fallon, JT, Fisher, EA & Vlassara, H (2003) Dietary glycotoxins promote diabetic atherosclerosis in apolipoprotein E-deficient mice. Atherosclerosis 168, 213220.CrossRefGoogle ScholarPubMed
137. Peppa, M, Brem, H, Cai, W, Zhang, JG, Basgen, J, Li, Z, Vlassara, H & Uribarri, J (2006) Prevention and reversal of diabetic nephropathy in db/db mice treated with alagebrium (ALT-711). Am J Nephrol 26, 430436.CrossRefGoogle ScholarPubMed
138. Zimanyi, MA, Denton, KM, Forbes, JM, Thallas-Bonke, V, Thomas, MC, Poon, F & Black, MJ (2006) A developmental nephron deficit in rats is associated with increased susceptibility to a secondary renal injury due to advanced glycation end-products. Diabetologia 49, 801810.CrossRefGoogle ScholarPubMed
139. Uribarri, J, Cai, W, Sandu, O, Peppa, M, Goldberg, T & Vlassara, H (2005) Diet-derived advanced glycation end products are major contributors to the body's AGE pool and induce inflammation in healthy subjects. Ann NY Acad Sci 1043, 461466.CrossRefGoogle Scholar
140. Uribarri, J, Cai, W, Peppa, M, Goodman, S, Ferrucci, L, Striker, G & Vlassara, H (2007) Circulating glycotoxins and dietary advanced glycation endproducts: two links to inflammatory response, oxidative stress, and aging. J Gerontol A Biol Sci Med Sci 62, 427433.CrossRefGoogle ScholarPubMed
141. Vlassara, H, Cai, W, Crandall, J, Goldberg, T, Oberstein, R, Dardaine, V, Peppa, M & Rayfield, EJ (2002) Inflammatory mediators are induced by dietary glycotoxins, a major risk factor for diabetic angiopathy. Proc Natl Acad Sci USA 99, 1559615601.CrossRefGoogle Scholar
142. Cai, W, He, JC, Zhu, L, Peppa, M, Lu, C, Uribarri, J & Vlassara, H (2004) High levels of dietary advanced glycation end products transform low-density lipoprotein into a potent redox-sensitive mitogen-activated protein kinase stimulant in diabetic patients. Circulation 110, 285291.CrossRefGoogle ScholarPubMed
143. Negrean, M, Stirban, A, Stratmann, B et al. (2007) Effects of low- and high-advanced glycation endproduct meals on macro- and microvascular endothelial function and oxidative stress in patients with type 2 diabetes mellitus. Am J Clin Nutr 85, 12361243.CrossRefGoogle ScholarPubMed
144. Uribarri, J, Stirban, A, Sander, D, Cai, W, Negrean, M, Buenting, CE, Koschinsky, T & Vlassara, H (2007) Single oral challenge by advanced glycation end products acutely impairs endothelial function in diabetic and nondiabetic subjects. Diabetes Care 30, 25792582.CrossRefGoogle ScholarPubMed
145. Peppa, M, Uribarri, J, Cai, W, Lu, M & Vlassara, H (2004) Glycoxidation and inflammation in renal failure patients. Am J Kidney Dis 43, 690695.CrossRefGoogle ScholarPubMed
146. Schwedler, SB, Metzger, T, Schinzel, R & Wanner, C (2002) Advanced glycation end products and mortality in hemodialysis patients. Kidney Int 62, 301310.CrossRefGoogle ScholarPubMed
147. Lieuw, AFML, van Hinsbergh, VW, Teerlink, T, Barto, R, Twisk, J, Stehouwer, CD & Schalkwijk, CG (2004) Increased levels of N(epsilon)-(carboxymethyl)lysine and N(epsilon)-(carboxyethyl)lysine in type 1 diabetic patients with impaired renal function: correlation with markers of endothelial dysfunction. Nephrol Dial Transplant 19, 631636.CrossRefGoogle Scholar
148. Busch, M, Franke, S, Wolf, G, Brandstadt, A, Ott, U, Gerth, J, Hunsicker, LG, Stein, G & Collaborative Study Group (2006) The advanced glycation end product N(epsilon)-carboxymethyllysine is not a predictor of cardiovascular events and renal outcomes in patients with type 2 diabetic kidney disease and hypertension. Am J Kidney Dis 48, 571579.CrossRefGoogle ScholarPubMed
149. Sebekova, K, Krajcoviova-Kudlackova, M, Schinzel, R, Faist, V, Klvanova, J & Heidland, A (2001) Plasma levels of advanced glycation end products in healthy, long-term vegetarians and subjects on a western mixed diet. Eur J Nutr 40, 275281.Google ScholarPubMed
150. Sebekova, K, Boor, P, Valachovicova, M, Blazicek, P, Parrak, V, Babinska, K, Heidland, A & Krajcovicova-Kudlackova, M (2006) Association of metabolic syndrome risk factors with selected markers of oxidative status and microinflammation in healthy omnivores and vegetarians. Mol Nutr Food Res 50, 858868.CrossRefGoogle ScholarPubMed
151. Ordovas, JM & Mooser, V (2004) Nutrigenomics and nutrigenetics. Curr Opin Lipidol 15, 101108.CrossRefGoogle ScholarPubMed
152. Cai, W, Gao, QD, Zhu, L, Peppa, M, He, C & Vlassara, H (2002) Oxidative stress-inducing carbonyl compounds from common foods: novel mediators of cellular dysfunction. Mol Med 8, 337346.CrossRefGoogle ScholarPubMed
153. Zill, H, Bek, S, Hofmann, T et al. (2003) RAGE-mediated MAPK activation by food-derived AGE and non-AGE products. Biochem Biophys Res Commun 300, 311315.CrossRefGoogle ScholarPubMed
154. Leslie, RD, Beyan, H, Sawtell, P, Boehm, BO, Spector, TD & Snieder, H (2003) Level of an advanced glycated end product is genetically determined: a study of normal twins. Diabetes 52, 24412444.CrossRefGoogle ScholarPubMed
155. Kankova, K, Stejskalova, A, Pacal, L et al. (2007) Genetic risk factors for diabetic nephropathy on chromosomes 6p and 7q identified by the set-association approach. Diabetologia 50, 990999.CrossRefGoogle ScholarPubMed
156. Delpierre, G, Veiga-da-Cunha, M, Vertommen, D, Buysschaert, M & Van Schaftingen, E (2006) Variability in erythrocyte fructosamine 3-kinase activity in humans correlates with polymorphisms in the FN3K gene and impacts on haemoglobin glycation at specific sites. Diabetes Metab 32, 3139.CrossRefGoogle ScholarPubMed
157. Kankova, K, Zahejsky, J, Marova, I, Muzik, J, Kuhrova, V, Blazkova, M, Znojil, V, Beranek, M & Vacha, J (2001) Polymorphisms in the RAGE gene influence susceptibility to diabetes-associated microvascular dermatoses in NIDDM. J Diabetes Complications 15, 185192.CrossRefGoogle ScholarPubMed
158. Kankova, K, Marova, I, Zahejsky, J, Muzik, J, Stejskalova, A, Znojil, V & Vacha, J (2001) Polymorphisms 1704G/T and 2184A/G in the RAGE gene are associated with antioxidant status. Metabolism 50, 11521160.Google ScholarPubMed
159. Kankova, K, Stejskalova, A, Hertlova, M & Znojil, V (2005) Haplotype analysis of the RAGE gene: identification of a haplotype marker for diabetic nephropathy in type 2 diabetes mellitus. Nephrol Dial Transpl 20, 10931102.CrossRefGoogle ScholarPubMed
160. Pettersson-Fernholm, K, Forsblom, C, Hudson, BI, Perola, M, Grant, PJ & Groop, PH (2003) The functional-374 T/A RAGE gene polymorphism is associated with proteinuria and cardiovascular disease in type 1 diabetic patients. Diabetes 52, 891894.CrossRefGoogle ScholarPubMed
161. Matsunaga-Irie, S, Maruyama, T, Yamamoto, Y, Motohashi, Y, Hirose, H, Shimada, A, Murata, M & Saruta, T (2004) Relation between development of nephropathy and the p22phox C242T and receptor for advanced glycation end product G1704T gene polymorphisms in type 2 diabetic patients. Diabetes Care 27, 303307.CrossRefGoogle ScholarPubMed
Figure 0

Table 1. Current classification of advanced glycation end products (AGE)

Figure 1

Fig. 1. Overview of advanced glycation end product (AGE) metabolism and factors influencing AGE turnover and the concentration of AGE in particular body compartments. MG, methylglyoxal; GLO I, glyoxalase I; FN3K, fructosamine 3 kinase; RAGE, receptor of AGE; ECM, extracellular matrix.