touchENDOCRINOLOGY touchENDOCRINOLOGY
Diabetes
Read Time: 12 mins

Type 1 Diabetes—Pathogenesis, Prediction, and Prevention

Copy Link
Published Online: Jun 6th 2011 US Endocrinology, 2009;5(1):79-84 DOI: http://doi.org/10.17925/USE.2009.05.1.79
Authors: George S Eisenbarth, Peter R Baker II
Quick Links:
Abstract
Article
Article Information
Abstract:
Overview

Abstract
Type 1 diabetes affects over 1.4 million people in the US, with a rising incidence in many western nations. It is clear that there is a strong hereditary component and that autoimmunity plays a large role in disease pathogenesis. In the last two decades novel technologies have been developed to study the genetics, biochemistry, and molecular pathology of type 1 diabetes. These, in turn, have allowed for early recognition of disease as well as the potential for prevention trials and early insulin treatment. This article highlights the prediction of type 1 diabetes risk and developing immunotherapeutic concepts.

Keywords
Autoimmune polyendocrine syndrome (APS), autoantibodies, autoimmunity, environmental risk, genome-wide association study, human leukocyte antigen (HLA), insulin, major histocompatability complex (MHC), pancreatic beta cells, trimolecular complex

Disclosure: The authors have no conflicts of interest to declare.
Acknowledgments: This work was supported by the National Institutes of Health (DK32083, DK057538, UO1DK062418, DK32493), Autoimmunity Prevention Center (AI050864), Diabetes Endocrine Research Center (P30 DK57516), Clinical Research Centers (MO1 RR00069, MO1 RR00051), the Immune Tolerance Network (AI15416), the American Diabetes Association, the Juvenile Diabetes Research Foundation, the Children’s Diabetes Foundation, and the Brehm Coalition. Peter R Baker II, MD, is a Fellow of the Pediatric Scientist Development Program. The project described was supported by Award K12-HD000850 from the Eunice Kennedy Shriver National Institute of Child Health and Human Development.
Received: October 12, 2009 Accepted: November 23, 2009
Correspondence: George S Eisenbarth, MD, PhD, Barbara Davis Center for Childhood Diabetes, 1775 Ursula St, Room 4201E, Building M20, PO Box 6511, Box B140, Aurora, CO 80045-6511. E: George.Eisenbarth@ucdenver.edu

Article:

Type 1 diabetes has become one of the most studied polygenic disorders. It effects over 1.4 million people in the US, with a rising incidence in many western nations.1,2 It is clear that there is a strong hereditary component in the development of disease, with siblings at higher risk than offspring and both at higher risk than the general population. It is also clear that autoimmunity plays a large role in disease pathogenesis.

Type 1 diabetes has become one of the most studied polygenic disorders. It effects over 1.4 million people in the US, with a rising incidence in many western nations.1,2 It is clear that there is a strong hereditary component in the development of disease, with siblings at higher risk than offspring and both at higher risk than the general population. It is also clear that autoimmunity plays a large role in disease pathogenesis. Development is insidious and chronic, but the initial presentation is often acute (hyperglycemia, ketoacidosis, and cerebral edema) and can be deadly as it is often unexpected.3

In the last two decades novel technologies have been developed to study the genetics, biochemistry, and molecular pathology of type 1 diabetes. These, in turn, have allowed for early recognition of disease, as well as the potential for prevention trials and early insulin treatment. In this article we will highlight the prediction of type 1 diabetes risk and developing immunotherapeutic concepts.

Immunology and Pathophysiology of Type 1 Diabetes
An overwhelming amount of evidence in the last several decades points to type 1 diabetes being an autoimmune, specifically T-cell-mediated, disease. Through study of multiple animal models, including the nonobese diabetic (NOD) mouse (which has many similarities to human type 1 diabetes), the pathogenesis of autoimmune beta-cell destruction is becoming clearer. Central to T-cell response, including autoimmune responses, are components of the trimolecular complex. For CD4- positive T cells this complex consists of the T-cell receptor (TCR), an antigenic peptide, and a human leukocyte antigen (HLA) molecule on antigen-presenting cells (APCs) (e.g. the NOD mouse I-Ag7, homologous to HLA class II DQ of humans).
Figure 1 illustrates the trimolecular complex, which can be likened to a hotdog (the peptide), bun (class II or class I molecules of the major histocompatability complex or MHC), and barbecue instruments (the TCR). The peptide sequence being presented to the TCR sits in the
groove of class II or class I molecules on the surface of an APC. The TCR is then able to recognize it, bind to it (with varying affinity dependent on molecular shape and charge), and mount an immune response. The TCR is crucial for T-cell selection in the thymus as well as immune targeting of peptides by mature T-lymphocytes. The receptor structure includes variable alpha and beta chains, both with germline-encoded V and J sequences. These segment sequences are ‘randomly’ combined to form literally billions of different TCRs that recognize specific antigenic sequences.4,5 When an antigenic peptide is presented by thymic epithelium (via major histocompatibility [MHC] class I and II molecules), the TCR binds to it. In the absence of any TCR engagement in the thymus, T cells will die by ‘neglect.’ If recognition in the developing thymus is modest (due to weak binding related to variations in the presenting molecule, the antigenic peptide, and the TCR binding
sequence), T-cells fail to be ‘deleted.’ They then leave the thymus and enter the peripheral circulation. A subset of autoreactive T-cells fail to be deleted in the thymus and can react with self-antigens in the periphery.
Self-antigen reactivity can occur by several mechanisms, including modification of self-molecules in the periphery but not the thymus (e.g. citrinylated peptides), failure of the thymus to express certain peripheral antigens in concentrations that are sufficient to delete all self-reactive T cells, and innate immune activation of self-reactive T-cells in the periphery. These peripherally activated, autoreactive T cells can then trigger a cascade of events leading to a large-scale immune response that ultimately ends in tissue destruction.
Insulin peptide sequences are now thought to be central to the development of autoimmunity in the NOD mouse.6–14 The insulin B:9–23 peptide sequence may be of particular importance in loss of tolerance leading to diabetes.9,14 In the mouse there are two insulin genes (insulin 1 and 2) that form nearly identical preproinsulin molecules. Insulin 1 and 2 are both expressed in pancreatic islet beta cells, but only insulin 2 is expressed in the thymus. Ideally, insulin-reactive T cells would be deleted in the thymus. Knocking out insulin 2 accelerates the development of type 1 diabetes, while knocking out insulin 1 prevents the majority of type 1 diabetes development.15 Thus, it is likely that attenuated expression of insulin 2 in the thymus of knockout mice enhances autoimmunity by decreasing negative selection in the thymus, while eliminating Insulin 1 in the periphery may remove an important islet target peptide. Of note, knocking out both insulin genes and providing mice with a single mutated insulin gene (replacing beta-chain 16 tyrosine with alanine) prevents all diabetes of NOD mice.13,16–18

A specific V alpha segment of the TCR, TRAV5D-4*04, appears to play a unique role in targeting the B:9–23 sequence of the insulin molecule. Experiments involving variation of this specific alpha chain sequence but conservation of other elements in the beta and alpha construct of the TCR have led to the hypothesis that this V-alpha segment is important in enhancing diabetes susceptibility.19–21 The final component in the trimolecular complex in the NOD mouse is I-Ag7, homologous to the DQ8 HLA class II molecule in humans. HLA class II molecules play a major role in the development of autoimmunity (see below). Human DR3–DQ2 and DR4–DQ8 haplotypes, which are closely associated with type 1 diabetes risk, have similar polymorphisms to I-Ag7.22,23 These polymorphisms alter the peptides bound and presented to TCRs, and thus alter self-antigen recognition as described above.
Although in the NOD mouse model there are convincing data supporting the hypothesis that insulin is the primary autoantigen, studies in humans are not definitive. In particular, although insulin autoimmunity is prominent and polymorphisms of the insulin gene influence diabetes risk, there are multiple islet autoantigens targeted in humans. Autoantibodies to IA-2, glutamic acid decarboxylase (GAD), and the newly discovered autoantigen ZnT8 (discussed in more detail below) are important markers of disease risk. Furthermore, loss of tolerance and development of autoimmunity clearly depend on more than the trimolecular complex recognition of insulin. Environmental factors and polymorphisms of non-MHC genes involving maintenance of tolerance can play a distinct role in disease development (see below). Inability to maintain tolerance is a key aspect of the NOD mouse model and humans.
Disease Prediction in Type 1 Diabetes
Genetic Markers
Approximately one in 300 individuals in the general population in the US develop type 1 diabetes, while approximately one in 20 first-degree relatives of patients with type 1 diabetes (offspring or sibling) develop diabetes.24,25 More than 60% of monozygotic twins with a twin-mate having type 1 diabetes will develop diabetes and more than 70% develop anti-islet antibodies.26 Dizygotic twin risk is much lower and similar to that of siblings (again, one in 20 or 5%).27
Environmental factors play a role in the development of type 1 diabetes. This is evident by the lack of 100% concordance in twin studies, the increasing incidence of type 1 diabetes worldwide (at a rate too fast to be explained by genetic changes alone), potential disease links to medications, temporal associations with environmental factors (e.g. diet and viral infections), and variability of disease penetrance in mouse models with different environmental exposures.28–34 One recent study followed monozygotic and dizygotic twins for 10 years and reported that 88% of phenotypic variance was due to genetics, while 12% could be attributed to the environment.35 More research is warranted in this field, and studies such as The Environmental Determinates of Diabetes in the Young (TEDDY) are under way to help explore these issues.
Regarding genetic susceptibility to diabetes, there are well-known single-gene causes of autoimmune diabetes. They include autoimmune polyendocrine syndrome type 1 (APS1) caused by mutations in the AIRE gene and immunodysregulation, polyendocrinopathy, enteropathy, X-linked (IPEX) syndrome caused by mutations in the FOXP3 gene. Both of these syndromes are well studied and have contributed to current understanding of diabetes pathophysiology. The FOXP3 gene in particular is essential for the development of regulatory T cells, and approximately 80% of children with FOXP3 mutations develop type 1 diabetes with onset as early as the first days of life.36 The AIRE gene controls expression of peripheral antigens such as insulin in the thymus and it is hypothesized that lack of multiple peripheral antigens in the thymus of individuals with mutations of the AIRE gene contribute to their widespread autoimmunity.37 Other diseases contributing to the knowledge of diabetes pathogenesis include autoimmune diseases known to be associated with type 1 diabetes, e.g. Addison’s disease, celiac disease, pernicious anemia, and thyroiditis.38 These polygenic conditions reflect the immunogenetics behind common forms of autoimmune diabetes. Approximately one-third of children with newonset type 1 diabetes have associated organ-specific autoimmunity (e.g. thyroid peroxidase, transglutaminase, 21-hydroxylase autoantibodies).
The MHC class II region on chromosome 6 has long been linked to diabetes susceptibility. Effects of high-risk alleles in this region are consistent across different ethnicities despite large differences in allele frequencies.39 There is also significant homology between species, with defined genes and regions of risk in man, NOD mice, and susceptible rat strains.23,40–43
The extremely high-risk genotype DR3/4–DQ2/DQ8 (DR3–DQA1* 501–DQB1*201; DR4–DQA1*301–DQB1*302) occurs in 2.4% of Denver newborns, and 30–40% of all type 1 diabetes patients carry this heterozygous genotype. Children with this genotype have an absolute risk of one in 15 versus one in 300 in the general population.24,25 In longterm follow-up studies of 30,000 newborns (selected for either highrisk HLA or a first-degree relative with type 1 diabetes), we find that 41% of DR3/4 siblings, as well as 16% of offspring of type 1 diabetes patients, expressed islet autoantibodies by seven years of age.

Furthermore, in siblings identical by descent for both DR3/4 haplotypes, 63% had positive autoantibodies by seven years of age and 85% were positive by 15 years of age (see Figure 2). This is in contrast to only 20% developing autoantibodies in DR3/4 siblings sharing no or one haplotype identical by descent.44 Within the general population, the DR3/4 genotype combined with analysis of DP alleles (absence of protective DPB1*0402) and DR4 (absence of DRB1*0403) confers a 20% risk of developing islet autoimmunity.45
While 2.4% of the population of Denver carry the DR3/4 heterozygous genotype, over 30% of patients with diabetes have this genotype.27 Interestingly, either DR3 or DR4 haplotypes in homozygous form (DR3/DR3 or DR4/DR4) are lower-risk compared with the DR3/4 genotype. The mechanism for the increased heterozygote risk is not completely delineated, but it has been hypothesized that the DQA1*0501 allele of DR3 haplotype and the DQB1*0302 allele of DR4 haplotype combine, creating a ‘chimeric’ molecule (DQA1*0501, DQB1*0302) for antigen presentation that increases the risk of diabetes.40 There are also MHC class II alleles that are protective (see Table 1): DQB1*0602 is present in 20% of the population but only 1% of children with type 1 diabetes.
Other protective alleles include DRB1*0403 (even when DQB1*0302 is present) and DRB1*1401.40,46 Even within DRB1*04 alleles there are variations conferring greater and lesser risk.47,48
MHC class I loci (HLA-A, B, and C) play a lesser role in diabetes susceptibility. A24 is associated with a younger age at presentation and A30 is associated with higher risk; A1 is lower-risk than other HLA-A alleles when associated with the DR3–B8 haplotype.49 HLA-B18, B39, B44, and B8 are all associated as well, with HLA-B39 conferring higher risk in three different populations and HLA-B8 being lower-risk when linked with the DR3 allele in the well-known extended haplotype with DR3, HLA-B8, and HLA-A1.50,51 HLA-C3, C8, and C16 have been reported to increase susceptibility.51 Extensive long-range linkage dysequilibrium between alleles of genes of the MHC make it difficult to pinpoint specific genes contributing to risk. One of the most common extended haplotypes consists of DR3–B8–A1 alleles, termed the 8.1 haplotype (containing DRB1*0303–DQA1*0501–DQB1*0201–HLA-B8–HLA-A1). This is the most common extended haplotype in the Caucasian population, with over 99% identity across the MHC by single nucleotide polymorphism (SNP) analysis. Interestingly, it is increased in type 1 diabetes individuals (18%, versus 9% of Caucasian controls)52 due presumably to the DR3 and DQ2 alleles and not as much to the HLA-B8 and HLA-A1 portion of the haplotype. The DR3–B8–A1 haplotype confers less risk than other DR3 haplotypes. Higher risk was found in the less common DR3–B18–A30 haplotype (Basque haplotype) as well as other non-B8 DR3-positive individuals.53,54 This would point to susceptibility loci telomeric to class II alleles. For several years, interest has turned to regions outside the MHC region for susceptibility to type 1 diabetes. Because there is a known correlation between development of diabetes and anti-insulin antibodies (see below), the insulin gene has been of particular interest. There is a variable number tandem repeat sequence (VNTR) at the 5´ end of the insulin gene that has been known for over two decades to be associated with risk in type 1 diabetes. Longer repeats are protective and are associated with increased insulin expression in the thymus.55,56
Differences in expression of the insulin 2 versus insulin 1 gene in NOD mouse thymus presumably relate to the same mechanism (see above). PTPN22 is located on 1p13 and encodes for protein tyrosine phosphatase non-receptor type 22 (PTPN22)/lymphoid phosphatase (LyP). Position 1858 contains a non-synonymous SNP that changes arginine to tryptophan at position 620. This polymorphism results in a gain of function that increases inhibition of TCR signaling. Many groups have confirmed its presence in type 1 diabetes patients in many different populations, with an odds ratio of 3.4 in its homozygous form.57–59 It is hypothesized that this SNP decreases T-cell signaling, thereby decreasing negative selection in the thymus. This risk allele is therefore associated with many autoimmune diseases.60 Genomewide association studies have also been performed on type 1 diabetes patients in hopes of finding other loci of interest. High-density SNP analysis (over 300,000 per individual) and follow-up meta-analyses have added to the list of regions involved in type 1 diabetes. Several signals of interest include confirmation of the MHC, PTPN22, cytotoxic T-lymphocyte antigen 4 (CTLA4), and insulin gene loci. Other genes of interest include those encoding CD25/interleukin-2 receptor alpha (IL2RA) and interferon-induced helicase C domain-containing protein 1 (IFIH1); the strongest signal from the latter three genes was in CD25/IL2RA, with more than one SNP associated with risk. CTLA4, despite an odds ratio of 1.1–1.2, has been implicated in multiple studies and is known to be strongly involved in T-cell signaling.61–67 More than 40 loci are now firmly associated with risk of type 1 diabetes, with the strongest signals by far associated with the MHC (see Figure 3).68

Serological Markers
Autoantibody development and insulitis are the end result of loss of self-tolerance and a component of the heightened immune response that results in destruction of beta cells in the pancreatic islets. Numerous antibodies are generated from a very early age in type 1 diabetes-susceptible patients. The most important autoantibodies include (in typical order of appearance chronologically) anti-insulin, anti- GAD65, anti-IA-2, and anti-ZnT8 antibodies.69–71 Presence of multiple islet autoantibodies is the most important predictor of progression to disease in type 1 diabetes.72 Development of autoantibodies can begin as early as four to 12 months of age, with earlier development correlating with greater risk of progression to overt disease (type 1 diabetes).73 For some patients with pre-diabetes, autoantibodies do not appear before 50 years of age. Insulin autoantibodies, often the first to appear, can be a predictor of severity as their levels are inversely related to age at disease onset.74 Furthermore, if two or more of the above antibodies are elevated (with each assay set at positivity ≥99 percentile of normal populations), both relatives and individuals in the general population will ‘inevitably’ develop overt disease (>90%) versus individuals with only one antibody (20%).44,75,76
In the DAISY Study, the development of autoantibodies in DR3/4 general population individuals is influenced by DP alleles (see Figure 4). The rate of progression to diabetes increases in direct proportion with autoantibody positivity, with autoantibody development often preceding disease onset.77 Individuals can express autoantibodies for decades prior to hyperglycemia. Despite a percentage of false or transiently positive individuals, people deemed ‘high-risk’ via genotyping can be followed with a reasonable prediction of disease progression such that prevention trials are under way.78 These individuals, found in studies such as DAISY, can also then be followed with glucose tolerance testing and glycated hemoglobin (HbA1c) to diagnose hyperglycemia early, and often they can be started on insulin therapy without hospitalization or the development of ketoacidosis.79,80
Disease Modification in Type 1 Diabetes
For the last few decades standards of management for type 1 diabetes have centered on glucose monitoring and insulin replacement therapy. While the technology has greatly improved with regard to insulin pumps and continuous glucose monitoring to simulate as best as possible physiological pancreatic beta-cell function, it is by no means a perfect solution to the disease. Currently, efforts are under way to prevent betacell loss or replace lost cells via beta-cell regeneration or transplantation of islets. Subjects who undergo islet cell transplant often initially develop insulin independence and markedly improved glucose levels. However, these benefits are short-lived with a significant number losing insulin independence over long-term (two- and five-year) follow-up. There remains some benefit with improved blood glucose control and prevention of severe hypoglycemia, however.81 Toxicity of immunosuppressive regimens and failure of islet grafts with time suggest that for most patients complications of therapy outweigh the benefits, and for now islet transplantation is still in development.82
It was reported more than 20 years ago that horse anti-thymocyte globulin or cyclosporine therapy prolongs the honeymoon phase in newonset diabetic patients.83,84 Since the initial publications, many therapies have been studied to modulate the immune system both generally as well as through targeting specific antigens. General immunosupression is a poor option for treatment and prevention of diabetes, given the very high cost–benefit ratio. Trials using oral insulin (through the National Institutes of Health [NIH] Diabetes Prevention Trial) to stimulate self-tolerance in the subgroup of individuals with high levels of antiinsulin antibodies showed some promise only in individuals with high levels of insulin autoantibodies by delaying the onset of diabetes. In several well-powered studies, however, the onset of disease could not be prevented overall.69,85–89 Other promising therapies include vaccination with GAD65 (a known target of anti islet antibodies) and monoclonal antibody therapy with anti-CD3 and anti-CD20.89–91 Again, long-term arrest of disease progression has not been found with these therapies. While antigen-specific therapies are safer than broad immunomodulation, there is less evidence for efficacy. Still, there is optimism and ongoing research devoted to this area. Several phase III trials are either under way or planned with goals to delay loss of beta cells after onset of hyperglycemia or in autoantibody-positive high-risk individuals. In North America, individuals can be screened for islet autoantibodies and considered for participation in NIH-sponsored trials (either new onset or pre-diabetic) by contacting Trialnet (1-800-HALT-DM1).
Conclusion
There has been great progress in understanding type 1 diabetes in the last two decades. We can predict disease through genetic testing of alleles of genes in the MHC region combined with analysis of islet autoantibodies and metabolic function. The realization that type 1 diabetes is an autoimmune disorder associated with a series of additional autoimmune diseases, many with shared genetic loci (e.g. celiac disease, Addison’s disease, thyroid autoimmunity), has led many centers to screen for these associated disorders. Current knowledge has also led to progress in trials of preventive therapies through manipulation of the immune response. It is hoped that in the years to come diabetes will be a preventable disease and the results of current phase III clinical trials will hopefully inform clinical care.

References

  1. Gillespie KM, Bain SC, Barnett AH, et al., The rising incidence of childhood type 1 diabetes and reduced contribution of high-risk HLA haplotypes, Lancet, 2004;364:1699–1700.
  2. U.S. Department of Health and Human Services NIoH 2008 National Institute of Diabetes and Digestive and Kidney Diseases: National Diabetes Statistics, 2007 fact sheet.
  3. Dunger DB, Sperling MA, Acerini CL, et al., European Society for Paediatric Endocrinology/Lawson Wilkins Pediatric Endocrine Society consensus statement on diabetic ketoacidosis in children and adolescents, Pediatrics, 2004;113:e133–40.
  4. Eisenbarth SC, Homann D, Primer immunology and autoimmunity. In: Eisenbarth GS (ed.), Type I Diabetes: Molecular, Cellular and Clinical Immunology, 2006. Online edition Version 3.0, updated August 2008. Available at: www.uchsc.edu/misc/diabetes/books.html
  5. Simone EA, Yu L, Wegmann DR, Eisenbarth GS, T cell receptor gene polymorphisms associated with anti-insulin, autoimmune T cells in diabetes-prone NOD mice, J Autoimmun, 1997;10:317–21.
  6. Homann D, Eisenbarth GS, An immunologic homunculus for type 1 diabetes, J Clin Invest, 2006;116:1212–15.
  7. Jaeckel E, Lipes MA, von Boehmer H, Recessive tolerance to preproinsulin 2 reduces but does not abolish type 1 diabetes, Nat Immunol, 2004;5:1028–35.
  8. Krishnamurthy B, Dudek NL, McKenzie MD, et al., Responses against islet antigens in NOD mice are prevented by tolerance to proinsulin but not IGRP, J Clin Invest, 2006;116:3258–65.
  9. Nakayama M, Abiru N, Moriyama H, et al., Prime role for an insulin epitope in the development of type 1 diabetes in NOD mice, Nature, 2005;435:220–23.
  10. Peng JT, Wong FS, Du W, et al., Insulin Reactive T Regulatory Cell TCR Transgenic NOD Mouse, Diabetes, 2005;54(Suppl. 1): A93.
  11. Solvason N, Lou YP, Peters W, et al., Improved efficacy of a tolerizing DNA vaccine for reversal of hyperglycemia through enhancement of gene expression and localization to intracellular sites, J Immunol, 2008;181:8298–8307.
  12. Suri A, Levisetti MG, Unanue ER, Do the peptide-binding properties of diabetogenic class II molecules explain autoreactivity? Curr Opin Immunol, 2008;20:105–10.
  13. Thebault-Baumont K, Dubois-LaForgue D, Krief P, Acceleration of type 1 diabetes mellitus in proinsulin 2- deficient NOD mice, J Clin Invest, 2003;111:851–7.
  14. Fukushima K, Abiru N, Nagayama Y, Combined insulin B:9- 23 self-peptide and polyinosinic-polycytidylic acid accelerate insulitis but inhibit development of diabetes by increasing the proportion of CD4+Foxp3+ regulatory T cells in the islets in non-obese diabetic mice, Biochem Biophys Res Commun, 2008;367:719–24.
  15. Moriyama H, Abiru N, Paronen J, et al., Evidence for a primary islet autoantigen (preproinsulin 1) for insulitis and diabetes in the nonobese diabetic mouse, Proc Natl Acad Sci U S A, 2003;100:10376–81.
  16. Jasinski JM, Yu L, Nakayama M, et al., Transgenic insulin (B:9-23) T-cell receptor mice develop autoimmune diabetes dependent upon RAG genotype, H-2g7 homozygosity, and insulin 2 gene knockout, Diabetes, 2006;55:1978–84.
  17. Mathews CE, Pietropaolo SL, Pietropaolo M, Reduced thymic expression of islet antigen contributes to loss of self-tolerance, Ann N Y Acad Sci, 2003;1005:412–17.
  18. Nakayama M, Babaya N, Miao D, Thymic expression of mutated B16:A preproinsulin messenger RNA does not reverse acceleration of NOD diabetes associated with insulin 2 (thymic expressed insulin) knockout, J Autoimmun, 2005;25:193–8.
  19. Hernandez Prada JA, Ferreira AJ, Katovich MJ, Structurebased identification of small-molecule angiotensinconverting enzyme 2 activators as novel antihypertensive agents, Hypertension, 2008;51:1312–17.
  20. Kobayashi M, Jasinski J, Liu E, et al., Conserved T cell receptor alpha-chain induces insulin autoantibodies, Proc Natl Acad Sci U S A, 2008;105:10090–94.
  21. Burton AR, Vincent E, Arnold PY, et al., On the pathogenicity of autoantigen-specific T-cell receptors, Diabetes, 2008;57:1321–30.
  22. Corper AL, Stratmann T, Apostolopoulos V, et al., A structural framework for deciphering the link between IAg7 and autoimmune diabetes, Science, 2000;288:505–11.
  23. Todd JA, Acha-Orbea H, Bell JI, et al., 1988 A molecular basis for MHC class II associated autoimmunity, Science, 1988;240:1003–9.
  24. Rewers M, Bugawan TL, Norris JM, Newborn screening for HLA markers associated with IDDM: diabetes autoimmunity study in the young (DAISY), Diabetol, 1996;39:807–12.
  25. She J-X, Susceptibility to type I diabetes: HLA-DQ and DR revisited, Immunol Today, 1996;17:323–9.
  26. Redondo MJ, Jeffrey J, Fain PR, Eisenbarth GS, Orban T, Concordance for islet autoimmunity among monozygotic twins, N Engl J Med, 2008;359:2849–50.
  27. Redondo MJ, Fain PR, Krischer JP, et al., Expression of betacell autoimmunity does not differ between potential dizygotic twins and siblings of patients with type 1 diabetes, J Autoimmun, 2004;23:275–9.
  28. Furukawa N, Miyamura N, Nishida K, et al., Possible relevance of alpha lipoic acid contained in a health supplement in a case of insulin autoimmune syndrome, Diabetes Res Clin Pract, 2007;75:366–7.
  29. Graves PM, Rotbart HA, Nix WA, et al., Prospective study of enteroviral infections and development of beta-cell autoimmunity. Diabetes AutoImmunity Study in the Young (DAISY), Diabetes Res Clin Pract, 2003;59:51–61.
  30. Norris JM, Yin X, Lamb MM, et al., Omega-3 polyunsaturated fatty acid intake and islet autoimmunity in children at increased risk for type 1 diabetes, JAMA, 2007;298:1420–28.
  31. Poole JA, Barriga K, Leung DY, et al., Timing of initial exposure to cereal grains and the risk of wheat allergy, Pediatrics, 2006;117:2175–82.
  32. Salminen KK, Vuorinen T, Oikarinen S, et al., Isolation of enterovirus strains from children with preclinical Type 1 diabetes, Diabet Med, 2004;21:156–64.
  33. Stene LC, Joner G, Use of cod liver oil during the first year of life is associated with lower risk of childhood-onset type 1 diabetes: a large, population-based, case-control study, Am J Clin Nutr, 2003;78:1128–34.
  34. Ziegler AG, Schmid S, Huber D, et al., Early infant feeding and risk of developing type 1 diabetes-associated autoantibodies, JAMA, 2003;290:1721–8.
  35. Hyttinen V, Kaprio J, Kinnunen L, et al., Genetic liability of type 1 diabetes and the onset age among 22,650 young Finnish twin pairs, Diabetes, 2003;52:1052–5.
  36. Eisenbarth GS, Gottlieb PA, Autoimmune polyendocrine syndromes, N Engl J Med, 2004;350:2068–79.
  37. Perheentupa J, Autoimmune polyendocrinopathy– candidiasis–ectodermal dystrophy, J Clin Endocrinol Metab, 2006;91:2843–50.
  38. Barker JM, Clinical review: type 1 diabetes-associated autoimmunity: natural history, genetic associations, and screening, J Clin Endocrinol Metab, 2006;91:1210–17.
  39. Park Y, She JX, Wang CY, et al., Common susceptibility and transmission pattern of human leukocyte antigen DRB1- DQB1 haplotypes to Korean and Caucasian patients with type 1 diabetes, JCEM, 2000;85:4538–42.
  40. Erlich H, Valdes AM, Noble J, et al., HLA DR-DQ haplotypes and genotypes and type 1 diabetes risk: analysis of the type 1 diabetes genetics consortium families, Diabetes, 2008;57:1084–92.
  41. Hattori M, Buse JB, Jackson RA, et al., The NOD mouse: recessive diabetogenic gene within the major histocompatibility complex, Science, 1986;231:733–5.
  42. Kloting I, Schmidt S, Kovacs P, Mapping of novel genes predisposing or protecting diabetes development in the BB/OK rat, Biochem Biophys Res Commun, 1998;245:483 6.
  43. Wallis RH, Wang K, Marandi L, et al., Type 1 diabetes in the BB rat: a polygenic disease, Diabetes, 2009;58:1007–17.
  44. Aly TA, Ide A, Jahromi MM, et al., Extreme genetic risk for type 1A diabetes, Proc Natl Acad Sci U S A, 2006;103:14074–9.
  45. Bonifacio E, Hummel M, Walter M, Schmid S, Ziegler AG, IDDM1 and multiple family history of type 1 diabetes combine to identify neonates at high risk for type 1 diabetes, Diabetes Care, 2004;27:2695–2700.
  46. Wucherpfennig KW, Eisenbarth GS, Type 1 Diabetes, Nat Immunol, 2001;2:1–3.
  47. Aly TA, Baschal EE, Jahromi MM, Analysis of single nucleotide polymorphisms identifies major type 1A diabetes locus telomeric of the major histocompatibility complex, Diabetes, 2008;57:770–76.
  48. Baschal EE, Eisenbarth GS, Extreme genetic risk for type 1A diabetes in the post-genome era, J Autoimmun, 2008;31:1–6.
  49. Noble JA, Valdes AM, Bugawan TL, et al., The HLA class I A locus affects susceptibility to type 1 diabetes, Human Immunol, 2002;63:657–64.
  50. Nejentsev S, Howson JM, Walker NM, et al., Localization of type 1 diabetes susceptibility to the MHC class I genes HLA-B and HLA-A, Nature, 2007;450:887–92.
  51. Valdes AM, Erlich HA, Noble JA, Human leukocyte antigen class I B and C loci contribute to Type 1 Diabetes (T1D) susceptibility and age at T1D onset, Hum Immunol, 2005;66:301–13.
  52. Alper CA, Larsen CE, Dubey DP, The haplotype structure of the human major histocompatibility complex, Hum Immunol, 2006;67:73–84.
  53. Bilbao JR, Calvo B, Aransay AM, et al., Conserved extended haplotypes discriminate HLA-DR3-homozygous Basque patients with type 1 diabetes mellitus and celiac disease, Genes Immun, 2006;7:550–54.
  54. Baschal EE, Aly TA, Jasinski JM, et al., The frequent and conserved DR3-B8-A1 extended haplotype confers less diabetes risk than other DR3 haplotypes, Diabetes Obes Metab, 2009;11(Suppl. 1):25–30.
  55. Hanahan D, Peripheral-antigen-expressing cells in thymic medulla: factors in self-tolerance and autoimmunity, Curr Opin Immunol, 1998;10:656–62.
  56. Chentoufi AA, Polychronakos C, Insulin expression levels in the thymus modulate insulin-specific autoreactive T-cell tolerance, Diabetes, 2002;51:1383–90.
  57. Bottini N, Musumeci L, Alonso A, et al., A functional variant of lymphoid tyrosine phosphatase is associated with type I diabetes, Nat Genet, 2004;36:337–8.
  58. Zoledziewska M, Perra C, Orru V, et al., Further evidence of a primary, causal association of the PTPN22 620W variant with type 1 diabetes, Diabetes, 2008;57:229–34.
  59. Zheng W, She JX, Genetic association between a lymphoid tyrosine phosphatase (PTPN22) and type 1 diabetes, Diabetes, 2005;54:906–8.
  60. Lee YH, Rho YH, Choi SJ, et al., The PTPN22 C1858T functional polymorphism and autoimmune diseases—a meta-analysis, Rheumatology (Oxford), 2007;46:49–56.
  61. Genome-wide association study of 14,000 cases of seven common diseases and 3,000 shared controls, Nature, 2007;447:661–78.
  62. Concannon P, Erlich HA, Julier C, et al., Type 1 diabetes: evidence for susceptibility loci from four genome-wide linkage scans in 1,435 multiplex families, Diabetes, 2005;54:2995–3001.
  63. Ueda H, Howson JM, Esposito L, et al., Association of the T-cell regulatory gene CTLA4 with susceptibility to autoimmune disease, Nature, 2003;423:506–11.
  64. Burton PR, Clayton DG, Cardon LR, et al., Association scan of 14,500 nonsynonymous SNPs in four diseases identifies autoimmunity variants, Nat Genet, 2007;39:1329–37.
  65. Lowe CE, Cooper JD, Brusko T, et al., Large-scale genetic fine mapping and genotype-phenotype associations implicate polymorphism in the IL2RA region in type 1 diabetes, Nat Genet, 2007;39:1074–82.
  66. Todd JA, Walker NM, Cooper JD, et al., Robust associations of four new chromosome regions from genome-wide analyses of type 1 diabetes, Nat Genet, 2007;39:857–64.
  67. Smyth DJ, Cooper JD, Bailey R, et al., A genome-wide association study of nonsynonymous SNPs identifies a type 1 diabetes locus in the interferon-induced helicase (IFIH1) region, Nat Genet, 2006;38:617–19.
  68. Concannon P, Rich SS, Nepom GT, Genetics of type 1A diabetes, N Engl J Med, 2009;360:1646–54.
  69. Achenbach P, Bonifacio E, Williams AJ, et al., Autoantibodies to IA-2beta improve diabetes risk assessment in high-risk relatives, Diabetologia, 2008;51:488–92.
  70. Achenbach P, Ziegler AG, Diabetes-related antibodies in euglycemic subjects, Best Pract Res Clin Endocrinol Metab, 2005;19:101–17.
  71. Achenbach P, Koczwara K, Knopff A, et al., Mature highaffinity immune responses to (pro)insulin anticipate the autoimmune cascade that leads to type 1 diabetes, J Clin Invest, 2004;114:589–97.
  72. Yu L, Eisenbarth GS, Humoral autoimmunity in type 1 diabetes: cellular, molecular and clinical immunology. In: Eisenbarth GS (ed.), Type I Diabetes: Molecular, Cellular and Clinical Immunology, 2009. Available at: www.uchsc.edu/ misc/diabetes/books.html
  73. Yu L, Robles DT, Abiru N, et al., Early expression of antiinsulin autoantibodies of humans and the NOD mouse: evidence for early determination of subsequent diabetes, Proc Natl Acad Sci U S A, 2000;97:1701–6.
  74. Vardi P, Ziegler AG, Matthews JH, et al., Concentration of insulin autoantibodies at onset of type I diabetes. Inverse log-linear correlation with age, Diab Care, 1998;11:736–9.
  75. Verge CF, Gianani R, Kawasaki E, et al., Prediction of type I diabetes in first-degree relatives using a combination of insulin, GAD, and ICA512bdc/IA-2 autoantibodies, Diabetes, 1996;45:926–33.
  76. Bingley PJ, Christie MR, Bonifacio E, et al., Combined analysis of autoantibodies improves prediction of IDDM in islet cell antibody-positive relatives, Diabetes, 1994;43:1304–10.
  77. Baschal EE, Aly TA, Babu SR, et al., HLA-DPB1*0402 Protects Against Type 1A Diabetic Autoimmunity in the Highest Risk DR3-DQB1*0201/DR4-DQB1*0302 DAISY Population, Diabetes, 2007;56:2405–9.
  78. Barker JM, Barriga K, Yu L, et al, Prediction of autoantibody positivity and progression to type 1 diabetes: Diabetes Autoimmunity Study in the Young (DAISY), J Clin Endocrinol Metab, 2004;89:3896–3902.
  79. Barker JM, Goehrig SH, Barriga K, et al., Clinical characteristics of children diagnosed with type 1 diabetes through intensive screening and follow-up, Diabetes Care, 2004;27:1399–1404.
  80. Stene LC, Barriga K, Hoffman M, et al., Normal but increasing hemoglobin A1c levels predict progression from islet autoimmunity to overt type 1 diabetes: Diabetes Autoimmunity Study in the Young (DAISY), Pediatr Diabetes, 2006;7:247–53.
  81. Ryan EA, Paty BW, Senior PA, et al., Five-year follow-up after clinical islet transplantation, Diabetes, 2005;54:2060–69.
  82. Alejandro R, Barton FB, Hering BJ, Wease S, 2008 update from the Collaborative Islet Transplant Registry, Transplantation, 2008;86:1783–8.
  83. Eisenbarth GS, Srikanta S, Jackson R, et al., Anti-thymocyte globulin and prednisone immunotherapy of recent onset type I diabetes mellitus, Diabetes Res, 1985;2:271–6.
  84. Assan R, Feutren G, Debray-Sachs M, et al., Metabolic and immunological effects of cyclosporine in recently Type I diabetes mellitus, Lancet, 1985;1(8420):67–71.
  85. Sherr J, Sosenko J, Skyler JS, Herold KC, Prevention of type 1 diabetes: the time has come, Nat Clin Pract Endocrinol Metab, 2008;4:334–43.
  86. Zhang ZJ, Davidson L, Eisenbarth G, Weiner HL, Suppression of diabetes in nonobese diabetic mice by oral administration of porcine insulin, Proc Natl Acad Sci U S A, 1991;88:10252–6.
  87. Bingley PJ, Mahon JL, Gale EA, Insulin resistance and progression to type 1 diabetes in the European Nicotinamide Diabetes Intervention Trial (ENDIT), Diabetes Care, 2008;31:146–50.
  88. Bingley PJ, Gale EA, Progression to type 1 diabetes in islet cell antibody-positive relatives in the European Nicotinamide Diabetes Intervention Trial, Diabetologia, 2006;49:881–90.
  89. Ludvigsson J, Faresjo M, Hjorth M, et al., GAD treatment and insulin secretion in recent-onset type 1 diabetes, N Engl J Med, 2008;359:1909–20.
  90. Herold KC, Hagopian W, Auger JA, et al., Anti-CD3 monoclonal antibody in new-onset type 1 diabetes mellitus, N Engl J Med, 2002;346:1692–8.
  91. Keymeulen B, Vandemeulebroucke E, Ziegler AG, et al., Insulin needs after CD3-antibody therapy in new-onset type 1 diabetes, N Engl J Med, 2005;352:2598–2608.

Further Resources

Share this Article
Related Content In Diabetes
  • Copied to clipboard!
    accredited arrow-down-editablearrow-downarrow_leftarrow-right-bluearrow-right-dark-bluearrow-right-greenarrow-right-greyarrow-right-orangearrow-right-whitearrow-right-bluearrow-up-orangeavatarcalendarchevron-down consultant-pathologist-nurseconsultant-pathologistcrosscrossdownloademailexclaimationfeedbackfiltergraph-arrowinterviewslinkmdt_iconmenumore_dots nurse-consultantpadlock patient-advocate-pathologistpatient-consultantpatientperson pharmacist-nurseplay_buttonplay-colour-tmcplay-colourAsset 1podcastprinter scenerysearch share single-doctor social_facebooksocial_googleplussocial_instagramsocial_linkedin_altsocial_linkedin_altsocial_pinterestlogo-twitter-glyph-32social_youtubeshape-star (1)tick-bluetick-orangetick-red tick-whiteticktimetranscriptup-arrowwebinar Sponsored Department Location NEW TMM Corporate Services Icons-07NEW TMM Corporate Services Icons-08NEW TMM Corporate Services Icons-09NEW TMM Corporate Services Icons-10NEW TMM Corporate Services Icons-11NEW TMM Corporate Services Icons-12Salary £ TMM-Corp-Site-Icons-01TMM-Corp-Site-Icons-02TMM-Corp-Site-Icons-03TMM-Corp-Site-Icons-04TMM-Corp-Site-Icons-05TMM-Corp-Site-Icons-06TMM-Corp-Site-Icons-07TMM-Corp-Site-Icons-08TMM-Corp-Site-Icons-09TMM-Corp-Site-Icons-10TMM-Corp-Site-Icons-11TMM-Corp-Site-Icons-12TMM-Corp-Site-Icons-13TMM-Corp-Site-Icons-14TMM-Corp-Site-Icons-15TMM-Corp-Site-Icons-16TMM-Corp-Site-Icons-17TMM-Corp-Site-Icons-18TMM-Corp-Site-Icons-19TMM-Corp-Site-Icons-20TMM-Corp-Site-Icons-21TMM-Corp-Site-Icons-22TMM-Corp-Site-Icons-23TMM-Corp-Site-Icons-24TMM-Corp-Site-Icons-25TMM-Corp-Site-Icons-26TMM-Corp-Site-Icons-27TMM-Corp-Site-Icons-28TMM-Corp-Site-Icons-29TMM-Corp-Site-Icons-30TMM-Corp-Site-Icons-31TMM-Corp-Site-Icons-32TMM-Corp-Site-Icons-33TMM-Corp-Site-Icons-34TMM-Corp-Site-Icons-35TMM-Corp-Site-Icons-36TMM-Corp-Site-Icons-37TMM-Corp-Site-Icons-38TMM-Corp-Site-Icons-39TMM-Corp-Site-Icons-40TMM-Corp-Site-Icons-41TMM-Corp-Site-Icons-42TMM-Corp-Site-Icons-43TMM-Corp-Site-Icons-44TMM-Corp-Site-Icons-45TMM-Corp-Site-Icons-46TMM-Corp-Site-Icons-47TMM-Corp-Site-Icons-48TMM-Corp-Site-Icons-49TMM-Corp-Site-Icons-50TMM-Corp-Site-Icons-51TMM-Corp-Site-Icons-52TMM-Corp-Site-Icons-53TMM-Corp-Site-Icons-54TMM-Corp-Site-Icons-55TMM-Corp-Site-Icons-56TMM-Corp-Site-Icons-57TMM-Corp-Site-Icons-58TMM-Corp-Site-Icons-59TMM-Corp-Site-Icons-60TMM-Corp-Site-Icons-61TMM-Corp-Site-Icons-62TMM-Corp-Site-Icons-63TMM-Corp-Site-Icons-64TMM-Corp-Site-Icons-65TMM-Corp-Site-Icons-66TMM-Corp-Site-Icons-67TMM-Corp-Site-Icons-68TMM-Corp-Site-Icons-69TMM-Corp-Site-Icons-70TMM-Corp-Site-Icons-71TMM-Corp-Site-Icons-72