Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2010 Feb 3.
Published in final edited form as: J Am Coll Cardiol. 2009 Feb 3;53(5 Suppl):S35–S42. doi: 10.1016/j.jacc.2008.09.055

The Continuum of Diabetic Vascular Disease: From Macro- to Micro-

Gabriela Orasanu 1, Jorge Plutzky 1
PMCID: PMC2663393  NIHMSID: NIHMS92383  PMID: 19179216

Abstract

Hyperglycemia can promote vascular complications by multiple mechanisms, with formation of advanced glycation endproducts (AGEs) and increased oxidative stress proposed to contribute to both macrovascular and microvascular complications. Many of the earliest pathologic responses to hyperglycemia are manifest in the vascular cells that directly encounter elevated blood glucose levels. In the macrovasculature, these include endothelial cells (ECs) and vascular smooth muscle cells (VSMCs). In the microvasculature, these include ECs, pericytes (in retinopathy), and podocytes (in renal disease). Additionally, neovascularization arising from the vasa vasorum may promote atherosclerotic plaque progression and contribute to plaque rupture, thereby interconnecting macro- and microangiopathy.

Keywords: glucose, atherosclerosis, advanced glycation endproducts, oxidative stress, inflammation

Introduction

Type 2 diabetes mellitus (T2DM) is diagnosed, and hence largely defined, by hyperglycemia. Although this definition has framed the perspective on T2DM, the pathologic imprint of this disease often involves the vasculature, with the hyperglycemia promoting both microvascular and macrovascular complications. Not surprisingly, given complications such as stroke and acute coronary syndromes, much attention has focused on the diabetic macrovascular disease. However, the morbidity associated with diabetic microvasculature disease, including retinopathy, neuropathy, nephropathy, and limb ischemia, is staggering. Given the impact of diabetic vascular disease, prodigious effort has been directed towards improving vascular outcomes in T2DM. Improving macrovascular outcomes through glucose-lowering interventions has remained a difficult, complicated, and to date, largely unsuccessful enterprise. In contrast, tighter glucose control does limit microvascular disease. These seemingly paradoxical trends force re-examination of the diabetic vascular disease spectrum.

Diabetic Macrovascular Complications

Hyperglycemia can promote vascular complications by multiple postulated mechanisms (Table 1). Increased glucose concentrations can activate nuclear factor-κB (1), a key mediator that regulates multiple pro-inflammatory and pro-atherosclerotic target genes in endothelial cells (ECs), vascular smooth muscle cells (VSMCs), and macrophages. Elevated glucose can foster glycation of proteins, promoting formation of advanced glycation endproducts (AGEs) protein cross-linking and reactive oxygen species formation. Hyperglycemia itself can stimulate oxidative stress, which has been strongly implicated as a driving force in atherosclerosis.

Table 1.

Examples of Mechanisms Implicated in Diabetic Macrovascular Disease

Risk factors Targets Mechanisms

Glucose Endothelium NF-κB activation
AGE Decreased NO production
Increased reactive oxygen species
Increased harmful metabolites (peroxynitrite, nitrotyrosine)
Increased lipid peroxidation products
Impaired endothelial-dependent relaxation

Monocyte-derived macrophages Increased IL1β, IL6, CD36, MCP-1
Induction of protein kinase C

Vascular smooth muscle cells Increased proliferation
Increased migration into intima
Altered matrix components (chondroitin, dermatan sulphate proteoglycans)
Increased matrix degradation (elastin)
Increased non-enzymatic collagen glycation

AGE = advanced glycation endproducts; IL1β = interleukin 1β; IL6 = interleukin 6, MCP-1 = monocyte chemoattractant protein-1; NF-κB = nuclear factor-κB; NO = nitric oxide.

Not surprisingly, many early pathologic responses to glucose are manifest in the vascular cells that directly encounter hyperglycemia. The loss of the non-adhesive property of the endothelium, with monocyte adhesion to ECs, is an early atherogenic step. Hyperglycemia increases monocyte adhesion to cultured ECs (2). Hyperglycemia and AGEs can also stimulate EC production of superoxide (1,3), suggesting links between hyperglycemia, AGEs, and oxidative stress. Glucose may also activate matrix-degrading metalloproteinases, enzymes implicated in plaque rupture and arterial remodeling, inducing similar responses in VSMC. Glucose may also stimulate VSMC proliferation, migration, and altered reactivity, e.g., through renin-angiotensin activation.

Inflammation has been strongly implicated in both atherosclerosis and type 2 diabetes mellitus (T2DM) (46). Despite this, no single mechanism yet explains why this pattern is found in diabetic patients. Monocytes grown in the presence of high glucose concentrations or isolated from individuals with poorly controlled diabetes appear activated (7), with induction of many inflammatory mediators such as protein kinase C and nuclear factor-κB. These targets, as well as others, may promote oxidative stress (8). In vitro studies suggest similar pro-atherogenic effects of hyperglycemia on T lymphocytes, inflammatory cells also involved in atherosclerosis.

Hyperglycemia vs. dyslipidemia in the pathogenesis of atherosclerosis

Attempts to improve cardiovascular (CV) outcomes through glucose control contrast strikingly with the benefits seen with statins in patients with diabetes in most trials. Such data challenge the focus on glucose as the prime determinant of pathologic, or at least vascular, outcomes among patients with diabetes. The relative effects of hyperglycemia vs. dyslipidemia in atherogenesis have been difficult to separate. For example, dyslipidemia can be exacerbated by hyperglycemia. At the same time, some data suggest possible independent effects of hyperglycemia on atherosclerosis (9,10). For example, fat-fed diabetic pigs reportedly developed atherosclerosis more rapidly than similar dyslipidemic fat-fed pigs without diabetes (9). In low-density lipoprotein receptor-deficient mice with a novel form of diabetes induced by a T-cell-directed viral antigen, consumption of a cholesterol-free diet resulted in hyperglycemia without changes in lipids or lipoproteins (10). Adding increasing amounts of dietary cholesterol led to dyslipidemia, which was the major factor in atherosclerosis progression independent of hyperglycemia (10).

Endoplasmic reticulum (ER) stress may contribute to possible glucose-independent promotion of atherosclerosis. All secretory and membrane proteins, many pathogens, and diverse nutrients, including glucose, pass through the ER. Hyperglycemia alone can induce ER stress in multiple tissues, including the liver and fat, activating pathways involved in oxidation and inflammation (11). ER stress, which can also be stimulated by hypoxia, elevated free fatty acids, and other non-glucose pathways may promote both diabetes and atherosclerosis (12).

It is of interest to overlay these various postulated mechanisms that promote inflammation and macrovascular disease onto microvascular disease. As noted, in contrast to macrovascular disease, the impact of better glycemic control on microvascular disease is unequivocal. This divergence in clinical experience raises fundamental questions about the nature of microvascular disease, how hyperglycemia modifies the microvasculature, and the implications of differing glucose effects on arterial disease, based on vessel size.

Diabetic Microvascular Complications

Pathological changes in the diabetic microvasculature alter organ perfusion, including organs heavily dependent on their microvasculature supply, namely the retina, kidneys, and peripheral nervous system. The clinical problems associated with these changes–retinopathy, nephropathy, and neuropathy–drive a large burden of T2DM morbidity. Microvascular disease also contributes to peripheral vascular disease, reduced myocardium vascularization, and poor wound healing (13). To some extent, diabetic microvascular disease has been overlooked in terms of its clinical impact and research attention. Further consideration of microvascular disease should begin with an overview of the anatomic nature of the microvasculature.

Structural and Functional Differences: Micro vs. Macro

Microvessels–the smallest functional unit of the CV system–consist of arterioles, capillaries, and venules. These vessels differ significantly from macrovessels with respect to architecture and cellular components. In contrast to larger vessels providing blood to organs, microvessels have specific roles regulating blood pressure and offering nutrient delivery. The microcirculation also has regulatory systems such as vasomotion, permeability, and myogenic responses that can adapt flow to local metabolic needs (14,15). Disturbances in microvascular function may arise before overt hyperglycemia and vascular pathologic changes (14,15). This timeline underscores the importance of understanding the distinct role of the microvasculature in the natural history of T2DM.

The most consistent structural diabetic microvascular modification is a thickening of the capillary basement membrane, including arterioles in the glomeruli, retina, myocardium, skin, and muscle, resulting in the classic diabetic microangiopathy. This thickening alters vessel function, directly promoting clinical problems like hypertension, reduced wound healing, and tissue hypoxia. Ultimately, in later stages, a frank loss of microvessels occurs, with microvessel drop-out and pruning classically associated with T2DM. The possibility that microvascular pathology contributes to systemic diabetic complications, including macrovascular atherosclerosis, remains an intriguing hypothesis worthy of further exploration.

Mechanisms of Diabetic Microvascular Disease

Glucose and the microvasculature

A linear relationship exists between hyperglycemia and microvascular complications. The impact of improved glucose control in preventing or limiting progression of microvascular disease strongly implicates hyperglycemia in these complications (16). Indeed, the current fasting plasma glucose parameters used to diagnose diabetes derive largely from diabetes-specific microvascular complication data, especially retinopathy (17). More recently, cross-sectional data from the Blue Mountains Eye Study, the Australian Diabetes, Obesity and Lifestyle Study, and the Multi-Ethnic Study of Atherosclerosis showed no uniform glycemic threshold for retinopathy across different populations. These data suggest that microvascular complications do not occur at an arbitrary glycemic threshold, a notion also raised for macrovascular disease (18). Ultimately, the interaction between glucose levels and microvascular disease requires a molecular explanation. Interestingly, hyperglycemia-induced diabetic complications are especially evident in insulin-insensitive cells that are thus unable to regulate glucose handling. Capillary ECs in the retina, mesangial cells in the renal glomerulus, and neurons and Schwann cells in peripheral nerves can all be categorized in this way (1921).

Various mechanisms have been proposed for diabetic microvascular complications (Table 2). Of note, the endothelium is often implicated in diabetic microvascular disease pathways, not unlike its important role in macrovascular disease. Together, these data place ECs in the pathologic center of T2DM. ECs display remarkable heterogeneity in structure and function. The endothelium arising from vessels of different sizes and from different anatomical compartments can express different phenotypic properties in normal and diseased states (22,23). Recently, DNA microarrays examined differences among gene expression profiles of ECs from large and smaller vessels (24). The differentially expressed genes uncovered in this study have diverse, well-established roles in endothelial biology, including extracellular matrix formation, neuronal signaling and migration, angiogenesis, and lipid metabolism (24).

Table 2.

Examples of Mechanisms Implicated in Diabetic Microvascular Disease

Risk factors: Hyperglycemia
Targets: Endothelium, Podocytes, Pericytes, VSMCs
Increased aldose
reductase pathway
Protein kinase
activation
Increased
oxidative
stress
Protein
glycation
Increased
hexosamine
pathway

↑Sorbitol: Osmotic cellular damage
↓(Na+ & K+) ATPase activity
↑NADH/NAD+
↓NADPH
↑VEGF
↑ROS
NF-κB activation
Inhibition of eNOS activity
↑Endothelin-1
↑ROS ↑AGE
Apoptotic death
NFκB activation
↑ROS
↑PAI-1
Inhibition of eNOS activity

VSMCs = vascular smooth muscle cells; eNOS = endothelial nitric oxide synthase; NAD = nicotine adenine dinucleotide; NADH = nicotine adenine dinucleotide reduced; NADPH = nicotine adenine dinucleotide phosphate reduced; PAI-1 = plasminogen activator inhibitor-1; ROS = reactive oxygen species; VEGF = vascular endothelial growth factor; other abbreviations as in Table 1.

Other specific cell types may also play defined roles in microvascular disease in certain tissue beds. An early, specific retinal change induced by hyperglycemia is the death of microvascular contractile cells known as pericytes. Pericytes provide vessel stability, control of endothelial proliferation, and regulate angiogenesis (25). In the kidney, podocytes and ECs form the glomerular capillary, which, together with the basement membrane, constitutes the glomerular filtration barrier. Podocyte injury and loss, which also involves apoptosis, are cardinal features of diabetic nephropathy (26). Many of the mechanisms implicated in microvascular injury are common to ECs, pericytes, and podocytes. Recent clinical evidence reveals that cerebral microangiopathy may play a role in promoting vascular dementia, ventricular hypertrophy, lacunar infarcts, hemorrhage, and may be a predisposing factor for Alzheimer’s disease in patients with diabetes (27). It is also worth noting that the heart itself contains a large and extensive microvasculature. Indeed, coronary microangiopathy has been raised as a major complication in diabetics. Compared to controls, hyperemic myocardial flow was decreased 28% in patients with T2DM and macrovascular coronary disease but even further reduced (57%) among patients with T2DM who also had evidence for coronary microangiopathy (28,29). As suggested in other tissue beds, glycemic control may influence the degree of coronary microangiopathy (28).

AGEs, strongly implicated in diabetic vascular injury, have been localized to retinal blood vessels in patients with T2DM and correlated with the degree of retinopathy (30). Retinal vascular ECs exposed to AGEs show abnormal endothelial nitric oxide synthase expression (31) and induction of vascular endothelial growth factor (VEGF) expression (32). AGEs reportedly signal via the receptor for AGE. In neuronal-associated vessels, the AGE-receptor has been localized with its putative ligand N-epsilon-carboxymethyl lysine and nuclear factor-κB, and interleukin-6 (33). The blockade of AGE formation by aminoguanidine improved neural signal transmission in diabetic rats, suggesting this as a therapeutic strategy in diabetic vascular complications (34).

Oxidative stress has been implicated in micro- and macrovascular disease. Hyperglycemia promotes formation of reactive oxygen species, which can interact with both DNA and proteins, causing damage. Mitochondrial DNA may be an especially relevant target (35). Interestingly, reactive oxygen species-mediated cellular damage may be a form of pathologic “memory” in the microvasculature that persists even after glucose normalization, as suggested in human retinal vascular ECs (35). The microvasculature may be more sensitive to such changes simply on the basis of mass. Oxidative stress may also link hyperglycemia with other pathways implicated in diabetic vascular complications, including AGE formation, protein kinase C activation, increased polyol flux, and hexosamine formation (36,37). For example, oxidative stress in response to AGE formation may promote diabetic neurovascular dysfunction (38,39).

Thrombospondin-1, a potent anti-angiogenic and pro-atherogenic protein, has received some attention as a potential mediator linking hyperglycemia to both micro- and macrovascular disease (40). Glucose alters both cell- and tissue-specific thrombospondin-1 expression and its post-transcriptional regulation in ECs, VSMCs, and fibroblasts (41). In contrast, thrombospondin-1 levels are dramatically decreased by high glucose in microvascular ECs and retinal pigment epithelial cells, making this protein an example of differences between macro- and microvascular disease.

Recently, thrombomodulin-dependent formation of activated protein C (APC) was identified as a potential mechanism for hyperglycemia-induced changes in mesangial ECs and podocytes (42). The endothelial thrombomodulin-protein C system is impaired in T2DM, as evident in the increased soluble thrombomodulin and decreased APC levels among such patients (43,44). APC may protect glomerular ECs against apoptosis and has potent antithrombotic and other cytoprotective, fibrinolytic, and anti-inflammatory properties. APC modulates the mitochondrial apoptosis pathway via the protease-activated receptor-1 and the endothelial protein C receptor in glucose-stressed cells. Loss of thrombomodulin-dependent APC formation interrupts cross-talk between the vascular compartment and podocytes, causing glomerular apoptosis and diabetic nephropathy (42).

Microalbuminuria in T2DM reflects a generalized disturbance of microvascular function related to endothelium-dependent mechanisms. Microalbuminuria may be a marker for the risk of retinopathy, nephropathy, and neuropathy. Interestingly, microalbuminuria may also predict CV disease (45).

The vasa vasorum and neovascularization in diabetes

The vasa vasorum is a network of small vessels normally found only in the adventitia and outer medial layer of larger arteries and the aorta (46). Neovascularization arising from the vasa vasorum may promote atherosclerosis and predict plaque rupture (47). Neovascularization in atherosclerotic arteries occurs by growth from both the adventitial layer outward and the arterial lumen inward, towards the intima (46). The vasa vasorum also provides the arterial wall with a vast absorptive endothelial surface that influences lipid metabolism and delivery and removal of neurohumoral factors (46).

In T2DM, angiogenesis is increased and associated with plaque rupture (48). Neovasculature microangiopathy may accelerate diabetic atherosclerosis (Fig. 1). The initial angiogenic response in the adventitial vasa vasorum appears stimulated by hypoxia and ischemia, perhaps through increased hypoxia induced factor-1 and VEGF action (49). VEGF also increases vascular permeability to macromolecules, monocyte chemotaxis, and tissue factor production, possible contributors to microvascular complications (50,51). VEGF is also associated with diabetic nephropathy (52). Increased vascular permeability instigates an inflammatory response, with recruited monocyte-macrophages serving as a source of VEGF (5356). Conversely, VEGF treatment may limit diabetic neuropathy by restoring microcirculation in the vasa nervorum, as suggested by rodent VEGF gene transfer experiments (57).

Fig. 1. Intersection of Microangiopathy and Macroangiopathy in Accelerated Atherosclerosis.

Fig. 1

In type 2 diabetes, both angiogenesis and microangiopathy are increased and may contribute to accelerated atherosclerosis and the development of vulnerable plaque. Hyperglycemia is a driving force in both large and small vessel disease. Indeed, the two disorders may be interconnected, with microvascular disease promoting atherosclerosis through processes such as hypoxia and changes in the vasa vasorum.

In the eye, pigment epithelium-derived factor (PEDF) may offset VEGF action, providing another example of tissue-specific settings in microvascular disease. PEDF is a neurotrophic factor and a potent angiogenic inhibitor (58). In proliferative diabetic retinopathy, VEGF levels are increased while PEDF levels are decreased (59). Decreased PEDF levels may also contribute to diabetic nephropathy. Other growth factors may foster proliferative retinopathy, including insulin-like growth factor 1, basic fibroblast growth factor, and hepatocyte growth factor (60,61).

Additional contributors to diabetic microvascular disease

Dyslipidemia is strongly associated with the development and progression of microvascular disease. Increased levels of dense low-density lipoprotein, as well as low-density lipoprotein modified by glycation and oxidation, may foster retinopathy, neuropathy, and nephropathy while attenuated levels and function of high-density lipoprotein have also been linked to retinopathy (6265). T2DM is often characterized by abnormal very-low-density lipoprotein and triglyceride levels. Elevated very-low-density lipoprotein and triglyceride levels appear involved with retinopathy and albuminuria. This may be due to changes in the function of lipoprotein lipase, a key enzyme in triglyceride hydrolysis active in the microvasculature. Interestingly, lipoprotein lipase action can generate natural peroxisome proliferator-activated receptor (PPAR) ligands as we and others have found (66,67). Perhaps lipoprotein lipase dysfunction in T2DM promotes microvascular disease through a loss of endogenous PPAR agonists. Potentially consistent with this, in the Fenofibrate Intervention and Event Lowering in Diabetes study, the PPAR-α agonist fenofibrate reduced the need for laser treatment for diabetic retinopathy (68). Of note, this effect did not appear to be mediated by improved lipid, glycemic, or blood pressure profiles, leaving the mechanism involved unclear. PPAR-α agonists inhibit the VEGF pathway, which may promote angiogenesis, inflammation, and cell migration (69,70); regulate retinal EC survival; limit apoptotic cell death (71); and improve vascular reactivity (70). In the Diabetes Atherosclerosis Intervention Study trial, the improved lipid profiles with fenofibrate in patients with T2DM were associated with reduced progression to microalbuminuria (72).

In animal studies, thiazolidinediones reportedly decreased proteinuria or delayed progression to nephropathy; results were independent of insulin-sensitization, glycemic control, and lipid metabolism, but were associated with reductions in blood pressure (73). In six randomized, active-controlled trials up to 12 months in duration with either pioglitazone or rosiglitazone, significant 10%–30% reductions in albumin-to-creatinine ratio were demonstrated (74,75).

Adiposity may promote microvascular disease. Increased fat mass and resistance to insulin-mediated inhibition of lipolysis increase elevated free fatty acid levels, which can directly impair microvascular function and increase diabetic microangiopathy (76). Increased visceral fat is a source of inflammatory mediators such as tumor necrosis factor-α, interleukins, and the pro-coagulant plasminogen activator inhibitor-1. In addition to increasing C-reactive protein levels and oxidative stress, these mediators also stimulate endothelial degradation and leukocyte adhesion, with the possible obstruction of microvessels (77,78).

Inflammation also promotes diabetic retinopathy, nephropathy, and neuropathy (79,80). Leukocyte adherence and accumulation within the retinal vasculature is an early change in experimental diabetes (81,82). Overexpression of adhesion molecules (83) and certain chemokines (84) may promote diabetic nephropathy. In diabetic retinopathy, this may result from elevated tumor necrosis factor-α levels (85). In addition to promoting leukocyte infiltration and activation, tumor necrosis factor-α may also enhance microvascular cell apoptosis (86). In keeping with these observations, tumor necrosis factor-α inhibition decreases leukostasis (87). Consistent with their opposing roles discussed earlier, VEGF can also trigger inflammation (87) while PEDF appears to limit it (88).

The Intersection of Diabetic Microvascular and Macrovascular Disease

T2DM may be best characterized by its complexity. Arising over decades, T2DM involves multiple pathologic forces and resulting clinical issues. This complexity is evident in the problems of diabetic macrovascular and microvascular disease. Diabetic subjects with microvascular complications, approximately 25%–30% of all those with diabetes, appear particularly prone to accelerated atherosclerosis and premature death. Neovascularization arising from the vasa vasorum may interconnect macro- and microangiopathy (Fig. 1).

Why the benefits of glycemic control are readily apparent in microvascular but not macrovascular outcomes remains obscure. The similarity and differences between diabetic microvascular and macrovascular disease, including differing responses to therapeutic interventions, remain important, unresolved issues in the field of T2DM; clarity in these areas could lead to treatments that improve outcomes in patients with diabetes.

The challenge to better understand how all forms of vascular disease occur in T2DM and how to intervene allows us to re-focus on perhaps the most obvious clinical issue at hand–implementing treatments known to improve T2DM outcomes. Microvascular disease is dramatically improved by tighter glycemic control, which should be implemented as early as is safely possible and maintained for as long as possible. Such interventions can significantly reduce a large burden of disease. In terms of macrovascular disease, diabetic control involves appropriate control of blood pressure as well as lipids. These steps should be taken early in the natural history of T2DM. While we wait for greater insight and better therapeutic options in T2DM, these simple steps would have a major impact on improving outcomes for millions of people.

Acknowledgments

The authors thank Ruzena Tupy for excellent editorial assistance.

Abbreviations and Acronyms

AGEs

advanced glycation endproducts

APC

activated protein C

CV

cardiovascular

ECs

endothelial cells

ER

endoplasmic reticulum

PEDF

pigment epithelium-derived factor

PPAR

peroxisome proliferator-activated receptor

T2DM

type 2 diabetes mellitus

VEGF

vascular endothelial growth factor

VSMCs

vascular smooth muscle cells

Footnotes

Disclosure statement: Dr. Orasanu has no conflict of interest to report. Dr. Plutzky has received grant/research support from sanofi-aventis and consulting fees or honoraria from Dainippon Sumitomo Pharma America, Inc.; GlaxoSmithKline; Novo Nordisk; Ono Pharmaceutical Co., Ltd.; sanofi-aventis; Takeda Pharmaceuticals North America; and Merck & Co., Inc.

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Piga R, Naito Y, Kokura S, Handa O, Yoshikawa T. Short-term high glucose exposure induces monocyte-endothelial cells adhesion and transmigration by increasing VCAM-1 and MCP-1 expression in human aortic endothelial cells. Atherosclerosis. 2007;193:328–34. doi: 10.1016/j.atherosclerosis.2006.09.016. [DOI] [PubMed] [Google Scholar]
  • 2.Otsuka A, Azuma K, Lesaki T, et al. Temporary hyperglycaemia provokes monocyte adhesion to endothelial cells in rat thoracic aorta. Diabetologia. 2005;48:2667–74. doi: 10.1007/s00125-005-0005-6. [DOI] [PubMed] [Google Scholar]
  • 3.Quagliaro L, Piconi L, Assaloni RG, et al. Intermittent high glucose enhances ICAM-1, VCAM-1 and E-selectin expression in human umbilical vein endothelial cells in culture: the distinct role of protein kinase C and mitochondrial superoxide production. Atherosclerosis. 2005;183:259–67. doi: 10.1016/j.atherosclerosis.2005.03.015. [DOI] [PubMed] [Google Scholar]
  • 4.Hansson GK. Inflammation, atherosclerosis, and coronary artery disease. N Engl J Med. 2005;352:1685–95. doi: 10.1056/NEJMra043430. [DOI] [PubMed] [Google Scholar]
  • 5.Biddinger SB, Kahn CR. From mice to men: insights into the insulin resistance syndromes. Ann Rev Physiol. 2006;68:123–58. doi: 10.1146/annurev.physiol.68.040104.124723. [DOI] [PubMed] [Google Scholar]
  • 6.Shoelson SE, Lee J, Goldfine AB. Inflammation and insulin resistance. J Clin Invest. 2006;116:1793–801. doi: 10.1172/JCI29069. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Dasu MR, Devaraj S, Jialal I. High glucose induces IL-1 beta expression in human monocytes: mechanistic insights. Am J Physiol Endocrinol Metab. 2007;293:E337–46. doi: 10.1152/ajpendo.00718.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Venugopal SK, Devaraj S, Yang T, Jialal I. Alpha-tocopherol decreases superoxide anion release in human monocytes under hyperglycemic conditions via inhibition of protein kinase C-alpha. Diabetes. 2002;51:3049–54. doi: 10.2337/diabetes.51.10.3049. [DOI] [PubMed] [Google Scholar]
  • 9.Gerrity RG, Natarajan R, Nadler JL, Kimsey T. Diabetes-induced accelerated atherosclerosis in swine. Diabetes. 2001;50:1654–65. doi: 10.2337/diabetes.50.7.1654. [DOI] [PubMed] [Google Scholar]
  • 10.Renard CB, Kramer F, Johansson F, et al. Diabetes and diabetes-associated lipid abnormalities have distinct effects on initiation and progression of atherosclerotic lesions. J Clin Invest. 2004;114:659–68. doi: 10.1172/JCI17867. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Gregor MF, Hotamisligil GS. Adipocyte stress: the endoplasmic reticulum and metabolic disease. J Lipid Res. 2007;48:1905–14. doi: 10.1194/jlr.R700007-JLR200. [DOI] [PubMed] [Google Scholar]
  • 12.Gargalovic PS, Gharavi NM, Clark MJ. The unfolded protein response is an important regulator of inflammatory genes in endothelial cells. Arterioscler Thromb Vasc Biol. 2006;26:2490–6. doi: 10.1161/01.ATV.0000242903.41158.a1. [DOI] [PubMed] [Google Scholar]
  • 13.He Z, Rask-Madsen C, King GL. Pathogenesis of diabetic microvascular complications. In: De Fronzo RA, Ferrannini E, Keen H, Zimmet P, editors. International Textbook of Diabetes Mellitus. Vol. 2. Hoboken, NJ: John Wiley & Sons; 2004. pp. 1135–59. [Google Scholar]
  • 14.Wiernsperger NF. In defense of microvascular constriction in diabetes. Clin Hemorheol and Microcirc. 2001;25:55–62. [PubMed] [Google Scholar]
  • 15.Sheetz MJ, King GL. Molecular understanding of hyperglycemia’s adverse effects for diabetic complications. JAMA. 2002;288:2579–88. doi: 10.1001/jama.288.20.2579. [DOI] [PubMed] [Google Scholar]
  • 16.The Diabetes Control and Complications Trial Research Group. The effect of intensive treatment of diabetes on the development and progression of long-term complications in insulin-dependent diabetes mellitus. N Engl J Med. 1993;329:977–86. doi: 10.1056/NEJM199309303291401. [DOI] [PubMed] [Google Scholar]
  • 17.World Health Organization. [Accessed August 26, 2008];Definition and diagnosis of diabetes mellitus and intermediate hyperglycaemia [online] Available at: http://www.who.int/diabetes/publications/Definition%20and%20diagnosis%20of%20diabetes_new.pdf.
  • 18.Wong TY, Liew G, Tapp RJ, et al. Relation between fasting glucose and retinopathy for diagnosis of diabetes: three population-based cross-sectional studies. Lancet. 2008;371:736–43. doi: 10.1016/S0140-6736(08)60343-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Brownlee M. The pathobiology of diabetic complications: a unifying mechanism. Diabetes. 2005;54:1615–25. doi: 10.2337/diabetes.54.6.1615. [DOI] [PubMed] [Google Scholar]
  • 20.Kaiser N, Sasson S, Feener EP, et al. Differential regulation of glucose transport and transporters by glucose in vascular endothelial and smooth muscle cells. Diabetes. 1993;42:80–9. doi: 10.2337/diab.42.1.80. [DOI] [PubMed] [Google Scholar]
  • 21.Heilig CW, Concepcion LA, Riser BL, Freytag SO, Zhu M, Cortes P. Overexpression of glucose transporters in rat mesangial cells cultured in a normal glucose milieu mimics the diabetic phenotype. J Clin Invest. 1995;96:1802–14. doi: 10.1172/JCI118226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Aird WC. Phenotypic heterogeneity of the endothelium: II. Representative vascular beds. Circ Res. 2007;100:174–90. doi: 10.1161/01.RES.0000255690.03436.ae. [DOI] [PubMed] [Google Scholar]
  • 23.Aird WC. Phenotypic heterogeneity of the endothelium: I. Structure, function, and mechanisms. Circ Res. 2007;100:158–73. doi: 10.1161/01.RES.0000255691.76142.4a. [DOI] [PubMed] [Google Scholar]
  • 24.Chi J-T, Chang HY, Haraldsen G, et al. Endothelial cell diversity revealed by global expression profiling. Proc Natl Acad Sci USA. 2003;100:10623–8. doi: 10.1073/pnas.1434429100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Ejaz S, Chekarova I, Ejaz A, Sohail A, Lim CW. Importance of pericytes and mechanisms of pericyte loss during diabetes retinopathy. Diabetes Obes Metab. 2008;10:53–63. doi: 10.1111/j.1463-1326.2007.00795.x. [DOI] [PubMed] [Google Scholar]
  • 26.Wolf G, Chen S, Ziyadeh FN. From the periphery of the glomerular capillary wall toward the center of disease: podocyte injury comes of age in diabetic nephropathy. Diabetes. 2005;54:1626–34. doi: 10.2337/diabetes.54.6.1626. [DOI] [PubMed] [Google Scholar]
  • 27.Huber JD. Diabetes, cognitive function, and the blood-brain barrier. Curr Pharm Des. 2008;14:1594–600. doi: 10.2174/138161208784705441. [DOI] [PubMed] [Google Scholar]
  • 28.Yokoyama I, Yonekura K, Ohtake T, et al. Coronary microangiopathy in type 2 diabetic patients: relation to glycemic control, sex, and microvascular angina rather than to coronary artery disease. J Nucl Med. 2000;41:978–85. [PubMed] [Google Scholar]
  • 29.Schindler TH, Zhang XL, Vincenti G, Mhiri L, Lerch R, Schelbert HR. Role of PET in the evaluation and understanding of coronary physiology. J Nucl Cardiol. 2007;14:589–603. doi: 10.1016/j.nuclcard.2007.05.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Murata T, Nagai R, Ishibashi T, Inomuta H, Ikeda K, Horiuchi S. The relationship between accumulation of advanced glycation end products and expression of vascular endothelial growth factor in human diabetic retinas. Diabetologia. 1997;40:764–9. doi: 10.1007/s001250050747. [DOI] [PubMed] [Google Scholar]
  • 31.Chakravarthy U, Hayes RG, Stitt AW, McAuley E, Archer DB. Constitutive nitric oxide synthase expression in retinal vascular endothelial cells is suppressed by high glucose and advanced glycation end products. Diabetes. 1998;47:945–52. doi: 10.2337/diabetes.47.6.945. [DOI] [PubMed] [Google Scholar]
  • 32.Lu M, Kuroki M, Amano S, et al. Advanced glycation end products increase retinal vascular endothelial growth factor expression. J Clin Invest. 1998;101:1219–24. doi: 10.1172/JCI1277. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Kislinger T, Fu C, Huber B, et al. N(epsilon)-(carboxymethyl)lysine adducts of proteins are ligands for receptor for advanced glycation end products that activate cell signaling pathways and modulate gene expression. J Biol Chem. 1999;274:31740–9. doi: 10.1074/jbc.274.44.31740. [DOI] [PubMed] [Google Scholar]
  • 34.Kihara M, Schmelzer JD, Poduslo JF, Curran GL, Nickander KK, Low PA. Aminoguanidine effects on nerve blood flow, vascular permeability, electrophysiology, and oxygen free radicals. Proc Natl Acad Sci USA. 1991;88:6107–11. doi: 10.1073/pnas.88.14.6107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Xie L, Zhu X, Hu Y, et al. Mitochondrial DNA oxidative damage triggers mitochondrial dysfunction and apoptosis in high glucose-induced human retinal vascular endothelial cells. Investigative Ophthalmology and Visual Science. 2008;49:125–32. doi: 10.1167/iovs.07-1364. [DOI] [PubMed] [Google Scholar]
  • 36.Nishikawa T. Normalizing mitochondrial superoxide production blocks three pathways of hyperglycaemic damage. Nature. 2000;404:787–90. doi: 10.1038/35008121. [DOI] [PubMed] [Google Scholar]
  • 37.Brownlee M. Biochemistry and molecular biology of diabetic complications. Nature. 2001;414:813–20. doi: 10.1038/414813a. [DOI] [PubMed] [Google Scholar]
  • 38.Cameron NE, Eaton SE, Cotter MA, Tesfaye S. Vascular factors and metabolic interactions in the pathogenesis of diabetic neuropathy. Diabetologia. 2001;44:1973–88. doi: 10.1007/s001250100001. [DOI] [PubMed] [Google Scholar]
  • 39.Gruden G, Bruno G, Chaturvedi N, et al. for the EURODIAB Prospective Complications Study Group. Serum heat shock protein 27 and diabetes complications in the EURODIAB prospective complications study: a novel circulating marker for diabetic neuropathy. Diabetes. 2008;57:1966–70. doi: 10.2337/db08-0009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Stenina OI, Krukovets I, Wang K, et al. Increased expression of thrombospondin-1 in vessel wall of diabetic Zucker rat. Circulation. 2003;107:3209–15. doi: 10.1161/01.CIR.0000074223.56882.97. [DOI] [PubMed] [Google Scholar]
  • 41.Bhattacharyya S, Marinic TE, Krukovets I, Hoppe G, Stenina OI. Cell type-specific post-transcriptional regulation of production of the potent antiangiogenic and proatherogenic protein thrombospondin-1 by high glucose. J Biol Chem. 2008;283:5699–707. doi: 10.1074/jbc.M706435200. [DOI] [PubMed] [Google Scholar]
  • 42.Isermann B, Vinnikov IA, Madhusudhan T, et al. Activated protein C protects against diabetic nephropathy by inhibiting endothelial and podocyte apoptosis. Nat Med. 2007;13:1349–58. doi: 10.1038/nm1667. [DOI] [PubMed] [Google Scholar]
  • 43.Borcea V, Morcos M, Isermann B, et al. Influence of ramipril on the course of plasma thrombomodulin in patients with diabetes mellitus. Vasa. 1999;28:172–80. doi: 10.1024/0301-1526.28.3.172. [DOI] [PubMed] [Google Scholar]
  • 44.Fujiwara Y, Tagami S, Kawakami Y. Circulating thrombomodulin and hematological alterations in type 2 diabetic patients with retinopathy. J Atheroscler Thromb. 1998;5:21–8. doi: 10.5551/jat1994.5.21. [DOI] [PubMed] [Google Scholar]
  • 45.Nathan DM, Cleary PA, Backlund JY, et al. for the Diabetes Control and Complications Trial/Epidemiology of Diabetes Interventions and Complications (DCCT/EDIC) Study Research Group. Intensive diabetes treatment and cardiovascular disease in patients with type 1 diabetes. N Engl J Med. 2005;353:2643–53. doi: 10.1056/NEJMoa052187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Hayden MR, Tyagi SC. Vasa vasorum in plaque angiogenesis, metabolic syndrome, type 2 diabetes mellitus, and atheroscleropathy: a malignant transformation. Cardiovasc Diabetol. 2004;3:1. doi: 10.1186/1475-2840-3-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Langheinrich AC, Kampschulte M, Buch T, Bohle RM. Vasa vasorum and atherosclerosis—Quid novi? Thromb Haemost. 2007;97:873–9. [PubMed] [Google Scholar]
  • 48.Moreno PR, Fuster V. New aspects in the pathogenesis of diabetic atherothrombosis. J Am Coll Cardiol. 2004;44:2293–300. doi: 10.1016/j.jacc.2004.07.060. [DOI] [PubMed] [Google Scholar]
  • 49.Pugh CW, Ratcliffe PJ. Regulation of angiogenesis by hypoxia: role of the HIF system. Nat Med. 2003;9:677–84. doi: 10.1038/nm0603-677. [DOI] [PubMed] [Google Scholar]
  • 50.Dvorak HF, Brown LF, Detmar M, Dvorak AM. Vascular permeability factor/vascular endothelial growth factor, microvascular hyperpermeability, and angiogenesis. Am J Pathol. 1995;146:1029–39. [PMC free article] [PubMed] [Google Scholar]
  • 51.Miyamoto K, Khosrof S, Bursell SE, et al. Vascular endothelial growth factor (VEGF)-induced retinal vascular permeability is mediated by intercellular adhesion molecule-1 (ICAM-1) Am J Pathol. 2000;156:1733–9. doi: 10.1016/S0002-9440(10)65044-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Khamaisi M, Schrijvers BF, De Vriese AS, Raz I, Flyvbjerg A. The emerging role of VEGF in diabetic kidney disease. Nephrol Dial Transplant. 2003;18:1427–30. doi: 10.1093/ndt/gfg242. [DOI] [PubMed] [Google Scholar]
  • 53.Hayden MR, Tyagi SC. Atherosclerosis: Implications of angiotensin II and the AT-1 receptor. In: Dhalla NS, Zahradka P, Dixon IMC, Beamish RE, editors. Angiotensin II Blockade: Physiological and Clinical Implications. Boston, MA: Kluwer Academic Publishers; 1998. pp. 233–43. [Google Scholar]
  • 54.Hayden MR, Tyagi SC. Arterial vascular remodeling: the endothelial cell’s central role. Mo Med. 1998;95:213–7. [PubMed] [Google Scholar]
  • 55.Williamson JR, Kilo C. Hyperglycemia “pseudohypoxia” and diabetic complications. Diabetes. 1993;42:801–13. doi: 10.2337/diab.42.6.801. [DOI] [PubMed] [Google Scholar]
  • 56.Zhang Y, Cliff WJ, Schoefl GI, Higgins G. Immunohistochemical study of intimal microvessels in coronary atherosclerosis. Am J Pathol. 1993;143:164–72. [PMC free article] [PubMed] [Google Scholar]
  • 57.Schratzberger P, Walter DH, Rittig K, et al. Reversal of experimental diabetic neuropathy by VEGF gene transfer. J Clin Invest. 2001;107:1083–92. doi: 10.1172/JCI12188. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Campochiaro PA. Retinal and choroidal neovascularization. J Cell Physiol. 2000;184:301–10. doi: 10.1002/1097-4652(200009)184:3<301::AID-JCP3>3.0.CO;2-H. [DOI] [PubMed] [Google Scholar]
  • 59.Gao G, Li Y, Zhang D, Gee S, Crosson C, Ma J. Unbalanced expression of VEGF and PEDF in ischemia-induced retinal neovascularization. FEBS Lett. 2001;489:270–6. doi: 10.1016/s0014-5793(01)02110-x. [DOI] [PubMed] [Google Scholar]
  • 60.Miller JW, Adamis AP, Aiello LP. Vascular endothelial growth factor in ocular neovascularization and proliferative diabetic retinopathy. Diabetes Metab Rev. 1997;13:37–50. doi: 10.1002/(sici)1099-0895(199703)13:1<37::aid-dmr174>3.0.co;2-k. [DOI] [PubMed] [Google Scholar]
  • 61.Shinoda K, Ishida S, Kawashima S, et al. Comparison of the levels of hepatocyte growth factor and vascular endothelial growth factor in aqueous fluid and serum with grades of retinopathy in patients with diabetes mellitus. Br J Ophthalmol. 1999;83:834–7. doi: 10.1136/bjo.83.7.834. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Lyons TJ, Jenkins AJ, Zheng D, et al. Diabetic retinopathy and serum lipoprotein subclasses in the DCCT/EDIC cohort. Invest Ophthalmol Vis Sci. 2004;45:910–8. doi: 10.1167/iovs.02-0648. [DOI] [PubMed] [Google Scholar]
  • 63.Jenkins AJ, Lyons TJ, Zheng D, et al. for the DCCT/EDIC Research Group. Lipoproteins in the DCCT/EDIC cohort: associations with diabetic nephropathy. Kidney Int. 2003;64:817–28. doi: 10.1046/j.1523-1755.2003.00164.x. [DOI] [PubMed] [Google Scholar]
  • 64.Song W, Barth JL, Yu Y, et al. Effects of oxidized and glycated LDL on gene expression in human retinal capillary pericytes. Invest Ophthalmol Vis Sci. 2005;46:2974–82. doi: 10.1167/iovs.04-1505. [DOI] [PubMed] [Google Scholar]
  • 65.Lyons TJ, Li W, Wells-Knecht MC, Jokl R. Toxicity of mildly modified low-density lipoproteins to cultured retinal capillary endothelial cells and pericytes. Diabetes. 1994;43:1090–5. doi: 10.2337/diab.43.9.1090. [DOI] [PubMed] [Google Scholar]
  • 66.Ziouzenkova O, Perrey S, Asatryan L, et al. Lipolysis of triglyceride-rich lipoproteins generates PPAR ligands: evidence for an antiinflammatory role for lipoprotein lipase. Proc Natl Acad Sci USA. 2003;100:2730–5. doi: 10.1073/pnas.0538015100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Chawla A, Lee CH, Barak Y, et al. PPARdelta is a very low-density lipoprotein sensor in macrophages. Proc Natl Acad Sci USA. 2003;100:1268–73. doi: 10.1073/pnas.0337331100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Keech AC, Mitchell P, Summanen PA, et al. for the FIELD study investigators. Effect of fenofibrate on the need for laser treatment for diabetic retinopathy (FIELD study): a randomised controlled trial. Lancet. 2007;370:1687–97. doi: 10.1016/S0140-6736(07)61607-9. [DOI] [PubMed] [Google Scholar]
  • 69.Ferrara N. Role of vascular endothelial growth factor in regulation of physiological angiogenesis. Am J Physiol Cell Physiol. 2001;280:C1358–66. doi: 10.1152/ajpcell.2001.280.6.C1358. [DOI] [PubMed] [Google Scholar]
  • 70.Malik J, Melenovsky V, Wichterle D, et al. Both fenofibrate and atorvastatin improve vascular reactivity in combined hyperlipidaemia (fenofibrate versus atorvastatin trial—FAT) Cardiovasc Res. 2001;52:290–8. doi: 10.1016/s0008-6363(01)00382-0. [DOI] [PubMed] [Google Scholar]
  • 71.Kim J, Ahn JH, Kim JH, et al. Fenofibrate regulates retinal endothelial cell survival through the AMPK signal transduction pathway. Exp Eye Res. 2007;84:886–93. doi: 10.1016/j.exer.2007.01.009. [DOI] [PubMed] [Google Scholar]
  • 72.Ansquer JC, Foucher C, Rattier S, Taskinen MR, Steiner G for the DAIS Investigators. Fenofibrate reduces progression to microalbuminuria over 3 years in a placebo-controlled study in type 2 diabetes: results from the Diabetes Atherosclerosis Intervention Study (DAIS) Am J Kidney Dis. 2005;45:485–93. doi: 10.1053/j.ajkd.2004.11.004. [DOI] [PubMed] [Google Scholar]
  • 73.Sarafidis PA, Bakris GL. Protection of the kidney by thiazolidinediones: an assessment from bench to bedside. Kidney Int. 2006;70:1223–33. doi: 10.1038/sj.ki.5001620. [DOI] [PubMed] [Google Scholar]
  • 74.Rohatgi A, McGuire DK. Effects of the thiazolidinedione medications on micro- and macrovascular complications in patients with diabetes-update 2008. Cardiovasc Drugs Ther. 2008;22:233–40. doi: 10.1007/s10557-008-6093-z. [DOI] [PubMed] [Google Scholar]
  • 75.Kanda T, Wakino S, Hayashi K, Plutzky J. Cardiovascular disease, chronic kidney disease, and type 2 diabetes mellitus: proceeding with caution at a dangerous intersection. J Am Soc Nephrol. 2008;19:4–7. doi: 10.1681/ASN.2007111182. [DOI] [PubMed] [Google Scholar]
  • 76.de Jongh RT, Serné EH, Ijzerman RG, de Vries G, Stehouwer CD. Free fatty acid levels modulate microvascular function: relevance for obesity-associated insulin resistance, hypertension, and microangiopathy. Diabetes. 2004;53:2873–82. doi: 10.2337/diabetes.53.11.2873. [DOI] [PubMed] [Google Scholar]
  • 77.Park HS, Park JY, Yu R. Relationship of obesity and visceral adiposity with serum concentrations of CRP, TNF-alpha and IL-6. Diabetes Res Clin Pract. 2005;69:29–35. doi: 10.1016/j.diabres.2004.11.007. [DOI] [PubMed] [Google Scholar]
  • 78.Wiernsperger NF, Bouskela E. Microcirculation in insulin resistance and diabetes: more than just a complication. Diabetes Metab. 2003;29:6S77–87. doi: 10.1016/s1262-3636(03)72791-8. [DOI] [PubMed] [Google Scholar]
  • 79.Crane IJ, Wallace CA, McKillop Smith S, Forrester JV. Control of chemokine production at the blood-retina barrier. Immunology. 2000;101:426–33. doi: 10.1046/j.1365-2567.2000.00105.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Meleth AD, Agron E, Chan C, et al. Serum inflammatory markers in diabetic retinopathy. Invest Ophthalmol Vis Sci. 2005;46:4295–301. doi: 10.1167/iovs.04-1057. [DOI] [PubMed] [Google Scholar]
  • 81.Miyamoto K, Khosrof S, Bursell SE, et al. Prevention of leukostasis and vascular leakage in streptozotocin-induced diabetic retinopathy via intercellular adhesion molecule-1 inhibition. Proc Natl Acad Sci USA. 1999;96:10836–41. doi: 10.1073/pnas.96.19.10836. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Canas Barouch F, Miyamoto K, Allport JR, et al. Integrin-mediated neutrophil adhesion and retinal leukostasis in diabetes. Invest Ophthalmol Vis Sci. 2000;41:1153–8. [PubMed] [Google Scholar]
  • 83.Sugimoto H, Shikata K, Hirata K, et al. Increased expression of intercellular adhesion molecule-1 (ICAM-1) in diabetic rat glomeruli: glomerular hyperfiltration is a potential mechanism of ICAM-1 upregulation. Diabetes. 1997;46:2075–81. doi: 10.2337/diab.46.12.2075. [DOI] [PubMed] [Google Scholar]
  • 84.Shikata K, Makino H. Role of macrophages in the pathogenesis of diabetic nephropathy. Contrib Nephrol. 2001;134:46–54. doi: 10.1159/000060147. [DOI] [PubMed] [Google Scholar]
  • 85.Limb GA, Chignell AH, Green W, LeRoy F, Dumonde DC. Distribution of TNF-alpha and its reactive vascular adhesion molecules in fibrovascular membranes of proliferative diabetic retinopathy. Br J Ophthalmol. 1996;80:168–73. doi: 10.1136/bjo.80.2.168. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Behl Y, Krothapalli P, Desta T, DiPiazza A, Roy S, Graves DT. Diabetes-enhanced tumor necrosis factor-alpha production promotes apoptosis and the loss of retinal microvascular cells in type 1 and type 2 models of diabetic retinopathy. Am J Pathol. 2008;172:1411–8. doi: 10.2353/ajpath.2008.071070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Lu M, Perez VL, Ma N, et al. VEGF increases retinal vascular ICAM-1 expression in vivo. Invest Ophthalmol Vis Sci. 1999;40:1808–12. [PubMed] [Google Scholar]
  • 88.Zhang SX, Wang JJ, Gao G, Shao C, Mott R, Ma JX. Pigment epithelium-derived factor (PEDF) is an endogenous antiinflammatory factor. FASEB J. 2006;20:323–5. doi: 10.1096/fj.05-4313fje. [DOI] [PubMed] [Google Scholar]

RESOURCES