American Journal of Transplantation 2014; 14: 764–778 Wiley Periodicals Inc.

 C

Copyright 2014 The American Society of Transplantation and the American Society of Transplant Surgeons doi: 10.1111/ajt.12653

Minireview

Transplantation Genetics: Current Status and Prospects B. Almoguera1,*, A. Shaked2,* and B. J. Keating1,2,3,*

Received 07 August 2013, revised 17 December 2013 and accepted for publication 31 December 2013

1

The Center for Applied Genomics, Abramson Research Center, The Children’s Hospital of Philadelphia, PA 2 Department of Surgery, Hospital of the University of Pennsylvania, Philadelphia, PA 3 Department of Pediatrics, School of Medicine University of Pennsylvania, Philadelphia, PA  Corresponding authors: Berta Almoguera, [email protected]; Abraham Shaked, [email protected]; and Brendan J. Keating, [email protected]

Over the last decade, advances in genetic technologies have accelerated our understanding of the genetic diversity across individuals and populations. Case– control and population-based studies have led to several thousand genetic associations across a range of phenotypes and traits being unveiled. Despite widespread and successful use of organ transplantation as a curative therapy for organ failure, genetic research has yet to make a major impact on transplantation practice aside from HLA matching. New studies indicate that non-HLA loci, termed minor histocompatibility antigens (mHAs), may play an important role in graft rejection. With several million common and rare polymorphisms observed between any two unrelated individuals, a number of these polymorphisms represent mHAs, and may underpin transplantation rejection. Genetic variation is also recognized as contributing to clinical outcomes including response to immunosuppressants, introducing the possibility of genotype-guided prescribing in the very near future. This review summarizes existing knowledge of the impact of genetics on transplantation outcomes and therapeutic responses, and highlights the translational potential that new genomic knowledge may bring to this field. Keywords: Allograft rejection, genomic variability, non-HLA-related rejection, pharmacogenomics Abbreviations: CNV, copy number variant; GWAS, genome-wide association studies; hdCNV, homozygous-deleted copy number variant; HSCT, hematopoietic stem cell transplantation; LoF, loss-of-function; mHAs, minor histocompatibility antigens; SNP, single nucleotide polymorphism; SNV, single nucleotide variant 764

Introduction Over the last two decades, more than 300 000 solid organ transplantations have been performed in the United States alone (1). However, despite improvements in surgical techniques and the development of more effective immunosuppressant therapies, allograft rejection still affects 60% of transplanted individuals and remains one of the major risk factors of graft loss (2,3). Up to 40% of graft recipients experience some form of rejection within the first postoperative year (4), with lung and heart recipients showing the highest rates of rejection, with 55% and 25% of patients, respectively (1,5,6), and kidney and liver the lowest, with 10% and 17% of patients experiencing rejection, respectively. Rejection can occur where genetic disparities exist between donors and recipients, which may lead to presentation of polymorphic peptides that the recipient’s immune system recognizes as nonself. Although key HLA loci have traditionally been considered to be the main contributor to the genetic variability of allograft rejection, some degree of rejection still occurs in HLAmatched sibling transplantations, which may be the result of noncompatible loci beyond HLA between donor and recipient (7). Indeed, new findings indicate that non-HLA polymorphisms can impact upon transplantation outcomes since they have the potential of generating histoincompatibilities (7–9) influencing allograft rejection (10), and impacting immunosuppressant responses (11). Approximately 3.5 million common and rare polymorphisms exist between two unrelated individuals of European ancestry and up to 10 million variants in individuals of African ancestry (12). However, investigations of non-HLA genetic determinants of clinical outcomes following organ transplantation have yet to be performed in any systematic fashion to date. Recent technological advances in genomics such as genome-wide association studies (GWAS) allow the characterization of hundreds of thousands to several million single nucleotide polymorphisms (SNPs) and copy number variants (CNVs) across the human genome rapidly and inexpensively (11,13). Figure 1 illustrates the concepts of structural variations, SNPs and insertion/deletion polymorphisms (InDels). Furthermore, whole exome and whole genome sequencing, which interrogates the coding regions and the entire human genome, respectively, are quickly

Transplantation Genetics

Figure 1: Illustration of common forms of human genomic variations: structural variation (copy number variants [CNVs]), singlenucleotide variants (SNVs) and insertion–deletion polymorphisms (InDels). (A) Structural variation. CNVs are the most common form of structural variation and include deletions and multiplications (typically duplications) >1000 bp. CNVs can affect whole genes or parts of them and may be present in heterozygosis (e.g. deletion of one copy of gene A in IND2 or duplication of gene B in IND1) or homozygosis (e.g. deletion of one copy of gene D in IND3). (B) SNVs and InDels. The presence of an A in IND1 instead of a C on the reference genome represents a graphic illustration of an SNV and the two red Ns in IND3 represent two nucleotides deletion.

becoming commonly used tools within the clinical diagnostic arena. These second-generation sequencing technologies have the ability to extensively characterize genomewide sources of histoincompatibility between donors and recipients, potentially unraveling specific genetic risk factors influencing rejection and immunosuppressant responses or severe adverse effects (14,15). In this review we aim to overview the current knowledge from existing genetics studies recently conducted for transplantation outcomes and therapeutic responses to immunosuppression therapies and we also discuss the translational components from this genetic knowledge that may be rapidly implemented in this field.

Genetics of Transplantation Outcomes Incompatibility across key HLA alleles has traditionally been considered the main factor influencing rejection in stem cell and solid organ transplants, and has therefore been the focus of extensive investigation. Because of high genetic variability in the HLA locus, a detailed characterization of American Journal of Transplantation 2014; 14: 764–778

HLA mismatches between donors and recipients is not routinely achieved. The effect of specific HLA mismatches in kidney transplantation have been characterized for sometime, whereas the importance of HLA matching on outcomes in organs such as the liver is still under debate (6,16–18). Even in HLA identically matched kidney transplantation, some degree of rejection is still evident. NonHLA or minor histocompatibility antigens (mHAs) resulting from a range of functional polymorphisms in the genome have been suggested to be capable of inducing strong cellular immune responses (reviewed in (7)). In contrast to the established role of HLA, Terasaki (2) estimated that over twice as many graft failures in HLA identical siblings at 10 years posttransplantation (38%) were due to immunological reactions to non-HLA factors compared to graft failures attributable to HLA (18%). Although our current knowledge of non-HLA antigens is still limited to a small number of loci (1,19,20), if one extrapolates the large number of genetic variants observed between two unrelated individuals within a donor–recipient pair, then the number of non-HLA discrepancies between any given donor and recipient would be expected to be very large. 765

Almoguera et al

The most rigorous way of investigating the association of mHAs with organ survival is within HLA-identical sibling donor–recipients, which is mostly restricted to hematopoietic stem cell transplantation (HSCT) and living kidney transplants. Therefore, the role of mHAs in other solid organ rejection may be more difficult to resolve due to incomplete or no matching. The first source of mHAs was identified in the Y chromosome (21) with at least six Y chromosome genes encoding various antigens presented by multiple MHC alleles recognized by donor T cells as foreign peptides after gender mismatched transplantation (19,22,23). Autosomal chromosomes also contain mHAs, with more than 40 described to date (reviewed in (20)). Such antigens arise as a consequence of common genetic variation, especially nonsynonymous SNPs in coding regions of the genome (24), leading to differences in the amino acid sequence proteins between donor and recipient. The number of identified autosomal mHAs identified to date is small considering the abundance of functional variants in the human genome, but it will undoubtedly grow with the recent genomic advances. Recently, Tennessen et al (25) applied whole exome sequencing to 15 585 human protein-coding genes in more than 2000 individuals of European and African ancestry and described more than 500 000 single nucleotide variants (SNVs) among the entire sample with an average of 13 595 variants per individual. The majority of these variants had not been previously described and were rare (minor allele frequency below 0.5%) and population-specific, with 2% predicted to impact the function of more than 300 genes per genome assessed (25), highlighting the need for a deeper genome-wide examination of the donor and recipient polymorphisms for potential mHAs. A distinct form of loss-of-function (LoF) genetic variants, homozygous-deleted CNVs (hdCNVs), is gaining traction as playing a role in transplant rejection. hdCNVs have been identified through large-scale GWAS and sequencing data sets (9,26). Possession of hdCNVs may manifest in a ‘‘normal’’ phenotype, with the term ‘‘disposable’’ or ‘‘dispensable,’’ genes often used to describe such variants. Cases where unaltered phenotypes are observed with carriage of hdCNVs suggest that there is either compensation for the hdCNV(s), or their function is no longer necessary. For example, hdCNVs of UGT2B17, a gene expressed in graft-versus-host disease-affected tissues, have been characterized as mHAs in HSCT (8,27). McCarroll et al (8) recently analyzed six common CNV deletions spanning genes in three HSCT cohorts (totaling 1345 HLAidentical sibling donor–recipient pairs). The authors found that risk of acute graft-versus-host disease was greater in recipients where UGT2B17 hdCNV was mismatched, that is zero gene copies present in the donor but one or two copies present in recipient (OR ¼ 2.5; 95% CI 1.4–4.6) (8). However, they only used markers for CNVs that were detectable on the initial GWAS arrays, which were limited in content, as many CNV regions known now were not 766

adequately probed or captured. Common hdCNVs may thus be important as they may play an active role in graft longevity. Recent whole exome and genome sequencing indicates that each individual carries numerous genetic variants predicted to cause LoF of protein-coding genes. MacArthur et al recently studied 2951 putative LoF variants obtained from whole genome sequencing from 185 human individuals from the 1000 Genome Project (derived from typical population samples), identifying and validating rare and likely deleterious LoF alleles, as well as common LoF variants in nonessential genes in this data set (26). They estimated that a typical human genome contains 100 genuine LoF variants with 20 genes inactivated in both copies, indicating unexpected redundancy in the human genome and suggesting that there are numerous mutations that are ‘‘private’’ to each personal genome. Since the immune system of an individual carrying LoF variants in both copies of a given gene may have had no previous exposure to protein (s) encoded by that gene(s), cellular or humoral immune recognition of that protein as an alloantigen in the grafted organ could very plausibly contribute to risk of rejection. LoF impacting the expression or translation of both copies of a given gene, especially in the transplanted organ of interest, is thus a plausible source of donor–recipient genomic incompatibilities, and may underpin rejection. Need for second-generation sequencing The emergence of large numbers of potential non-HLA incompatibilities with varying potential levels of immunogenicity highlights the clear need for deeper capture and assessment of SNVs, SNPs and CNVs in donors and recipients. Deep sequencing technologies will undoubtedly be key to unraveling common and rare LoFs and hdCNVs in population-scale diversity as well as capturing ‘‘private’’ individual-level polymorphisms to characterize the histocompatibility determinants involved in the biological mechanisms of alloimmunity. It should be noted that while there is low statistical power to detect such LoFs and hdCNVs at an association level, there is a lot of promise to follow up putative individual donor–recipient genomic incompatibility through the use of autoantibody testing in sera and/or tissues from the recipient postoperatively. Such approaches offer the promise of clinical utility through closer monitoring of a priori mHAs or potentially tolerizing recipients to a given gene product. Genetic risk factors for transplantation outcomes beyond HLA Over the last decade there have been unprecedented advances in the assessment of human genomic diversity across the major human populations through the development of high-throughput genotyping and deep sequencing technologies, as well as the development of population scale genomic maps such as the International HapMap Project (28). These tools have been applied to deeply American Journal of Transplantation 2014; 14: 764–778

Transplantation Genetics

characterize the genetic architecture of common and rare diseases and evoked conditions such as drug severe adverse events (29). Unfortunately, to date, the field of organ transplantation has not benefited from many of these advances. While over a thousand genetic association studies on organ rejection have been published so far, they are primarily candidate-gene based and suffer from many of the usual pitfalls of genetic association studies such as lack of adequate sample sizes, retrospective study designs, noninclusion of appropriate covariates such as ethnicity, lack of replication and proper statistical correction when multiple hypotheses are tested. Table 1 shows a review of appropriately designed published studies on organ transplantation outcomes. These studies were initially collated from PubMed and filtered for appropriate study size and design, defining appropriate as having at least a sample size of 200 individuals, including covariates, accounting for ethnicity and adjusting for multiple testing, when necessary. As observed in Table 1, genetic variation in up to 47 genes encoding cytokines, chemokines, celladhesion molecules, components of the renin–angiotensin–aldosterone pathway, coagulation and aggregation factors has been widely investigated and associated with outcomes such as delayed graft function, acute or chronic rejection or graft failure. However, results do not consistently replicate across studies. This lack of replication can be attributed to both the above-mentioned limitations of candidate-gene study designs and the complexity and diversity of clinical phenotypes. In addition to differences in protocols for immunosuppressive regimes, ascertainment of outcomes may differ by study—e.g. the criteria used to define acute rejection and chronic allograft dysfunction might vary among transplant centers: these all introduce further heterogeneity between studies. Specifically, in the case of rejection, studies may not always base the diagnosis on biopsy results and biopsy results themselves may be inaccurate due to interoperator differences. It is essential, therefore, to use objective and standard definitions for transplantation outcomes in order to obtain consistent results to homogenize studies. Thus future studies should aim to define phenotypes with precision and use a rigorous genetic approach, which preferably incorporates a hypothesis-free design such as GWAS to gain the most insight into genetic risk factors for organ transplantation.

Pharmacogenetics of immunosuppressant responses Genetic background is thought to account for as much as 95% of the variability in drug disposition and therapeutic effects (30). Regimes for immunosuppressants are typically characterized by wide pharmacokinetic interindividual variability and narrow therapeutic indexes, which often makes the ideal balance between sufficient immunosuppression and drug toxicity difficult to achieve. Thus, the American Journal of Transplantation 2014; 14: 764–778

discovery of genetic markers responsible for the interindividual variation in response to immunosuppressive therapy is an intensive area of ongoing research in transplantation. Ekberg et al evaluated the immunosuppressant drug exposures in 1645 renal transplant patients randomly assigned to four treatment groups: (1) standard-dose cyclosporine, (2) low-dose cyclosporine, (3) low-dose tacrolimus and (4) low-dose sirolimus (31), and observed that up to 90% of patients experienced at least one adverse event during treatment. Both biopsy-proven acute rejection and severe adverse events were similar in groups, ranging from 2.3% to 37% and 43% to 53%, respectively, depending on the drug, with the highest events observed in the low sirolimus dose group (31). In efforts to avoid such issues, therapeutic drug monitoring is being routinely performed, but it is currently assessed posttransplant and thus is not used for determining the optimal immunosuppressant starting dose, which is still established by an iterative postoperative approach. Alternative strategies incorporating pharmacogenetics hold great promise as complementary tools in drug monitoring to better guide individual therapies and doses. While the number of studies focusing on the pharmacogenetics of immunosuppressants has increased dramatically over the last few years, to date, many of these are underpowered that suffer from the typical candidate gene association study pitfalls. One clear exception is the robust association observed for rs776746 in CYP3A5, the primary enzyme involved in the metabolism of tacrolimus, the most frequently prescribed immunosuppressant drug worldwide. While nongenetic factors such as recipient gender, age, diabetes status and calcium channel blockers, such as some newer class of antifungals and grapefruit juice, influence tacrolimus levels, the rs776746 SNP is very important in tacrolimus clearance, with dosing requirements as well as time to therapeutic concentrations explaining up to 45% of the dose and 30% of clearance variability (13). The rs776746-A form is classically referred to as CYP3A5 1, while rs776746-G is referred to as CYP3A5 3. The latter is a noncoding variant that results in a cryptic splicing site, which causes 131 nucleotides of the intronic sequence to be inserted in the mRNA (32). The consequence is the introduction of a premature stop codon that truncatesCYP3A5 and leads to complete lack of CYP3A5 translation in  3 homozygotes (32). The prevalence of this CYP3A5 3 allele is as high as 90% in individuals of European ancestry while 90% of African Americans carry at least one of the common fully functional CYP3A5 1 alleles ( 1/ 1 or  1/ 3). These frequency differences in CYP3A5 3 make it one the most important genetic markers of interindividual and interethnic differences observed in CYP3A-dependent drug responses and clearance (33). There is also evidence of an additional rare variant, CYP3A5 6, which is associated with tacrolimus trough levels. Tables 2–4 review genetic association studies for tacrolimus, cyclosporine and mycophenolic acid regarding 767

768

Variant

rs4343

rs699

rs5186

rs1044240

rs2230199

C4L/S, CNV

rs1024611

109T/C, rs2107538, rs2280788 (a)

rs1799864

rs1799987 (b), rs333 (c)

rs2569190

rs20417

rs1801239, rs7918972

rs1801157

Gene

ACE

AGT

AGTR1

ALCAM

C3

C4

CCL2

CCL5

CCR2

CCR5

CD14

COX2

CUBN

CXCL12

12201365

22574174

19788502

19741468 17217435 17989610

19561149 15458467 12462338b 12201365c 17989610c

15458467 17989610 12201365 12239249

17989610 12462338a

17989610 12462338 12239249

21164027

22176838 19246358

20220571

AR, GS, PS

Proteinuria

GF

AR AR, BOS, GS CAN

AR AR, CAN AR, GS AR, GS, PS AR, SR

AR, CAN AR, CAN, SR AR, GS, PS GF

AR, CAN, SR AR, GS

AR, CAN, SR AR, GS GF

AR, GF, GL, GS

AR, DGF, PNF AR, GF, PS

GS

GF, serCr

GF, serCr

11923700 11923700

GF

GF GF, serCr RHC

Outcome

12548125

12548125 11923700 21659963

PMID

Table 1: Genetic association studies of organ transplantation outcomes

Liver

Kidney

1142R þ 1186D 207

Kidney

Kidney Lung Kidney

Kidney Kidney Liver Liver Kidney

Kidney Kidney Liver Kidney

Kidney Liver

Kidney Liver Kidney

Kidney

603

216D-R 226 436

243D-R 244D 209 207 436

244D 436 207 232

436 209

436 209 232

1969

Kidney Kidney

Kidney

954 þ 1002 1265D-R 1147

Kidney

Kidney

Kidney

Kidney Kidney Heart

Graft

224

224

210

210 224 532

Sample size

American Journal of Transplantation 2014; 14: 764–778

(Continued)

rs1801157 29.0% in dead patients compared to alive 17.1% and PS time (134 vs 98 months wt vs carriers) (p < 0.034)

ESRD: OR ¼ 1.39, elevated proteinuria 1y post (p < 0.015)

rs20417C and GL HR ¼ 2.43 [1.19-4.97], cohort 1; HR ¼ 1.72 [0.99-3.77], cohort 2 (p < 0.051)

NA Earlier AR, BOS and rs2569190TT: HR ¼ 1.65 [1.03–2.64], p ¼ 0.04 NA

# A alleles (D þ R) and AR, RR AR (p ¼ 0.039) NA NA NA

NA NA NA NA

NA NA

NA NA rs1024611GG vs A and GS: 67  14 vs 95  4 months; p ¼ 0.0052

NA

NA NA

NA

rs5186 and preserved long-term GF (p ¼ 0.037)

Increased Cr/t shorter t-to-sustained doubling of Cr, shorter t-to GL (p < 0.007) NA

NA NA RHC and rs4343: HR ¼ 0.58 [0.36-0.95]; p ¼ 0.031

Results

Almoguera et al

American Journal of Transplantation 2014; 14: 764–778

rs1126579, rs4674258

rs2228014

rs1799998

rs2814778

rs1800682

rs763110

rs17514136, rs17549193, rs3124952, rs3124953, rs7851696

CXCR2

CXCR4

CYP11B2

DARC

FAS

FASL

FCN2

(GT)n, rs2071746, rs2071747, rs2071748, rs2071749, rs2285112, rs5755720, rs6518952, rs8140669

rs2430561

rs1800871, rs1800872, rs1800896

HMOX1

IFNG

IL10

rs6025

rs2671222

CXCR1

FV

Variant

Gene

Table 1: Continued

15458467 15599305 20622753 15367225

15458467 17989610 15599305 20622753 21659963

18640487

AR, CAN AR, DGF GF, IIF GS

AR, CAN AR, CAN, SR AR, DGF GF, IIF RHC

DGF

AR, GF

BPAR, DGF, PNF, PS

22173060

11858477

RHC

RHC

AR, DGF

GF, serCr

BPAR, GS

AR

21659963

21659963

15327416

11923701

21304904

21452410

AR

BPAR, GS

21304904 21452410

Outcome

PMID

244D 291R þ 206D 218 2298 þ 1901

244D 436 291R þ 206D 218 532

965R

394

1272

532

532

222

224

335D-R

216

216

335D-R

Sample size

Kidney Kidney Kidney Kidney

Kidney Kidney Kidney Kidney Heart

Kidney

Kidney

Kidney

Heart

Heart

Kidney

Kidney

Kidney

Kidney

Kidney

Kidney

Graft

769

(Continued)

NA R ACCACC, ATAATA, GCCATA AR: OR ¼ 1.9 [1.1–3.1], p ¼ 0.016 rs1800871 TT high IIF grade (OR ¼ 3.27 [1.1–9.8]; p ¼ 0.035) NA

CAN (p < 0.008) NA NA NA NA

NA

GA no GF (RR ¼ 2.87 [1.01-8.26]), #dialysis until GF, risk for at least one AR episode (RR ¼ 3.83 [1.38-10.59]). Cr, protein excretion rate (p < 0.05).

NA

NA

RHC and rs1800682AA: HR ¼ 1.84 [1.25–2.69]; p ¼ 0.002

NA

rs1799998TT vs CC worse GF (p ¼ 0.002)

NA

NA

AR and D rs2671222A (OR ¼ 3.56 [1.37-9.27]; p ¼ 0.009)

D-rs1801157A and BPAR: OR ¼ 0.39, [0.20–0.76] and poor GS: HR ¼ 3.01; [1.19–7.60]; p < 0.020

Results

Transplantation Genetics

770

rs1801131, rs1801133

rs1805087

rs1799983

MTHFR

MTR

NOS3

rs12953, rs1131012, rs668

rs11003125, rs1800450, rs1800451, rs5030737, rs7095891, rs7096206

MBL2

PECAM1

rs72550870

rs1800795

IL6

MASP2

rs1801275

IL4R

rs5918

(CA)m(CT)n rs2069762

IL2

ITGB3

rs1143634, rs16944

IL1B

rs4073

rs187238

IL18

IL8

rs3212227

rs1554286, rs1878672, rs2222202, rs3021094, rs3024493, rs3024494, rs3024498

Variant

IL12B

Gene

Table 1: Continued

20220571

19349296

19349296

19349296

22173059 19104434

22173060

17928472

21452410

15458467 15599305 20622753 21659963

15367225

11981437 15458467

20622753

19077897

GS

GS

GS

GS

BPAR, DGF, PNF, PS BOS, GS

BPAR, DGF, PNF, PS

AR

AR

AR, CAN AR, DGF GF, IIF RHC

GS

AR AR, CAN

GF, IIF

AR

GF, IIF

RHC DGF

21659963 18640487

20622753

Outcome

PMID

Kidney Kidney

954 þ 1002

Kidney

Kidney

Kidney Lung

Kidney

Kidney

Kidney

262D-218R

262D-218R

262D-218R

1271 277 D-R

1272

445

216

Kidney Kidney Kidney Heart

Kidney

2298 þ 1901 244D 291R þ 206D 218 532

Heart Kidney

290 244D

Kidney

Kidney

226 þ 148C 218

Kidney

Heart Kidney

Graft

218

532 965R

Sample size

NA

GS: OR ¼ 9.192; [1.09-76.9]; p ¼ 0.0406

NA

NA

(Continued)

NA D-rs7096206 and GS and BOS free PS (p ¼ 0.007)

NA

rs5918TT and AR: OR ¼ 3.4; p ¼ 0.04

NA

NA NA NA NA

NA

Allele 135 and risk of AR (p < 0.03) NA



rs187238GG and AR: OR ¼ 3.653; p ¼ 0.015

NA

rs1800896 RHC: HR ¼ 0.49 [0.27–0.90]; p ¼ 0.020 NA

Results

Almoguera et al

American Journal of Transplantation 2014; 14: 764–778

American Journal of Transplantation 2014; 14: 764–778

rs5368

rs1131498, rs2205849, rs2229569

rs1800471, rs1800470 (d), rs1800472 (e)

rs3775290, rs3775291, rs3775296

rs10759932, rs1927914

rs1800629 (f), rs1800628 (g), rs3093662 (h) rs361525

rs1042522, rs12951053, rs1614984, rs1625895, rs17884306, rs4968187, rs9894946

rs1005230, rs2010963, rs3556934, rs699947

SELE

SELL

TGFB1

TLR3

TLR4

TNF

TP53

VEGF

h

21659963

RHC

DGF

AR, GS DGF GF, IIF RHC

15367225 18640487f, g, 20622753 21659963 18640487

AR AR AR, CAN, SR AR, DGF

AR

12682890 15458467 17989610 15599305

19741468

AR

DGF GF, IIF GS

18640487d,e 20622753 15367225 19741468

AR, CAN AR, CAN, SR AR, DGF

GS

20220571

15458467 17989610 15599305

AR

Outcome

21452410

PMID

532

Heart

Kidney

Kidney Kidney Kidney Heart

2298 þ 1901 965R 218 532 965R

Liver Kidney Kidney Kidney

Kidney

Kidney

Kidney Kidney Kidney

210D-R 244D 436 291R þ 206D

216D-R

216D-R

965R 218 2298 þ 1901

Kidney Kidney Kidney

Kidney

964 þ 1002

244D 436 291R þ 206D

Kidney

Graft

216

Sample size

NA

NA

NA NA NA R rs1800629AA and AR: OR ¼ 5.0 [3.0–8.3]; D rs1800629GA and DGF OR ¼ 1.6 [1.1–2.6], p < 0.040 GS rate among retransplants: 63.0% AA vs.79.5%G; p ¼ 0.0116 NA NA NA

R rs10759932C vs TT and AR (OR ¼ 0.25 [0.11-0.57]). rs10759932C higher rejection-free PS rates (p < 0.023)

NA

rs1800470 and AR (p ¼ 0.027) NA R rs1800471GG þD rs1800871T and AR: OR ¼ 1.8 [1.1–3.0], p ¼ 0.027 NA NA NA

NA

NA

Results

Summary of the most relevant studies performed to date reporting genes/variants associated with transplantation outcomes. Low case letters a to h in PMID column indicate that, amongst all the variants, these particular studies investigate exclusively those variants specified by the letters. For the rest of the studies, all variants are investigated. In alphabetical order, abbreviations for outcomes are: AR ¼ acute rejection; BOS ¼ bronchiolitis obliterans syndrome; BPAR ¼ biopsy proven acute rejection; CAN ¼ chronic allograft nephropathy; DGF ¼ delayed graft function; GF ¼ graft function; GL ¼ graft loss; GS ¼ graft survival; IIF ¼ infiltration of inflammatory cells; PNF ¼ primary non-function; PS ¼ patient survival; RHC ¼ rejection with hemodynamic compromise; serCr ¼ serum creatinine; SR ¼ subclinic rejection. In sample size and results columns, D ¼ donor and R ¼ recipient. For results, ESRD ¼ end-stage renal disease; NA ¼ No Association

Variant

Gene

Table 1: Continued

Transplantation Genetics

771

772

Variant

rs1045642 (a), rs1128503, rs2032582 (b)

rs3740066, 1446C > G, rs717620 (c)

rs2740574 (d) and rs28371759, rs35599367 (e), rs4987161, rs55785340, rs4986910

rs776746, rs10264272 (f), rs41303343 (g)

Gene

ABCB1

ABCC2

CYP3A4

CYP3A5

104 151 399D-R 301 185

832

18192894 15592326f 15454731 21806386f,g 21903774

19005401

CNI

185

21903774e

CsA

CNI

104

124 151 832

18192894

CsA

185 338

11375290d 15592326 19005401d

19890249c 19494809

MPA þ CNI

104

136 209D-R

18334918 21677300a 18192894

301 832 103 150D

124 151 294

11375290a 15592326a 20061922 21806386 19005401 16753004a,b 17885626

104

Sample size

18192894

PMID

CsA

Tac

CNI

CsA

Primary IST

Table 2: Pharmacogenetic studies on immunosuppressant therapy pharmacokinetics

Kidney

Kidney Kidney Kidney Kidney

Kidney

Kidney

Kidney Kidney and heart Kidney

Kidney

Kidney Kidney

Kidney

Kidney Kidney

Kidney Kidney Kidney Liver

Kidney and heart Kidney Heart

Kidney

Graft

NA NA rs776746AA Tac Co/D than G (p0.0001) Haplotype rs776746AA (CYP3A5)-rs35599367T or CC higher C than rs776746G (CYP3A5)-rs35599367CC (p < 0.03) rs2740574(CYP3A4)AA-rs776746GG and less time to achieve Tac C and higher C/D/Kg (p < 0.001)

NA

NA rs2740574G higher Cl than AA (p < 0.05) rs2740574AA-rs776746 GG (CYP3A5) less time to Tac C and higher C/D/Kg (p < 0.001) rs35599367T D lower than rs35599367CC (p ¼ 0.018). Haplotype rs776746AA (CYP3A5)-rs35599367T or CC higher C than rs776746G (CYP3A5)-rs35599367CC (p < 0.03)

NA

NA NA

NA

NA NA Higher Co/D-Co/weigh in rs1128503CC-rs2032582GGrs1045642CC than TT-TT-TT (p < 0.05) NA NA rs1045642 explains 3.7% of D variation (p ¼ 0.009) rs1128503T and rs2032582T hepatic C greater than CC and GG (p < 0.035) rs1045642TT lower D than C (p < 0.04) NA

Lower oral B: rs2032582GG, rs1128503TT, rs2229109Grs1128503C-rs2032582G-rs1045642C (p < 0.05)

Results

Almoguera et al

American Journal of Transplantation 2014; 14: 764–778

American Journal of Transplantation 2014; 14: 764–778

rs17868320, rs6714486, rs72551330, rs2741049

rs7438135, rs7439366, rs7668282

UGT1A9

UGT2B7

18641546 19494809

MPA MPA þ CNI

19890249

19494809

MPA þ CNI

338

332

185

338

185

180

12490779 19890249

178

150D

103 208 150D 209D-R 136

Sample size

15147425

MPA þ CNI

Tac

17885626

16753004 21566507 17885626f 21677300 18334918

Tac

Tac

PMID

Primary IST

Kidney

Kidney

Kidney

Kidney

Kidney

Liver

Kidney Kidney Liver Kidney Kidney

Graft

NA

NA

NA

rs1042597GG higher CsA AUC0–12; rs6714486A and/or rs17868320T lower CsA AUC0–12 (p < 0.05)

Tac patients rs4149117T decrease in MPAG/MPA; Cmax (p < 0.0003)

1 lower C, delay in achieving target C and  3/ 3 C above target (p < 0.003) CYP3AP1 1 vs  3 reduction in C (p < 0.05)



NA

rs776746A explains 35.3% of D variation (p < 0.001) NA rs776746G higher D than AA (p < 0.05) R rs776746G lower Co, Co/D (p < 0.05) rs776746G lower Co/D (p < 0.003)

Results

Relevant studies investigating genes/variants associated with calcineurin inhibitor and mycophenolic acid pharmacokinetics. Similarly to Table 1, letters a to g in PMID column indicate that particular studies investigate exclusively those variants specified by the letters. For Primary IST, CsA ¼ cyclosporine A; CNI ¼ CsA þ Tac; MPA ¼ mycophenolic acid; Tac ¼ tacrolimus. For sample size, D ¼ donor and R ¼ recipient. In results, AUC0–12 ¼ area under the curve from 0 to 12 hours; B ¼ bioavailability; C ¼ concentration; Cl ¼ clearance; Co ¼ trough concentration; D ¼ dose; MPAG/MPA ¼ MPA glucuronide to MPA ratio; NA ¼ No Asociation.

rs1042597, rs17863762

UGT1A8

1/ 3

1C

rs4149117, rs60140950





Variant

SLCO1B3

CYP3AP1

CYP3A7

Gene

Table 2: Continued

Transplantation Genetics

773

774

rs3740066, 1446C > G, rs717620

rs2740574

ABCC2

CYP3A4

PXR

rs11706052 (g), rs4974081 rs1523130, rs2276707,

IMPDH2

1/ 3

rs11770116, rs2228075, rs2288548, rs2288549, rs2288550, rs2288553, rs4731448 and rs2278293 (e), rs2278294 (f)



IMPDH1

CYP3AP1

rs776746

rs1045642 (a), rs1128503 (b), rs2032582 (c), rs717620 (d)

ABCB1

CYP3A5

Variant

Gene

CNI

MPA

MPA

Tac

CNI MPA þ Tac Tac

CsA

CNI MPA þ Tac

CsA

DGF

AR

AR

AR

GL GS AR CAD AR AR, DGF

AR AR, GL AR CAD

AR

CAD AR, DGF

MPA þ Tac Tac

191 456

191 456 178D-R

17851563g 20679962 22453193

178

227D-259R 399D-R 832 252 136 209D-R

124 227D-259R 832 252

338

17851563 20679962e,f

15147425

20505666 15454731 19005401 19762492 18334918 21677300

11375290 20505666 19005401 19762492

19494809

178D-R 252 209D-R

22453193d 19762492a-c 21677300a

147 832

18510642

DGF

227D-259R

294 170

124

Sample size

19005401

20505666

AR

MPA þ CNI

a

20061922 17198259a,c

11375290

PMID

AR, GL

AR

Primary outcome

CNI

CsA

Primary IST

Table 3: Pharmacogenetic studies of transplantation outcomes

Kidney Kidney Kidney

Kidney Kidney

Kidney

Kidney Kidney Kidney Kidney Kidney Kidney

Kidney Kidney Kidney Kidney

Kidney

Kidney Kidney Kidney

Kidney

Kidney

Kidney

Heart Heart

Kidney

Graft

1 earlier AR (p ¼ 0.005)

NA NA Donor 2276707TT higher risk for DGF (p < 0.05)

rs2278293A and rs2278294A lower risk (p < 0.05) rs2278294A lower risk (p ¼ 0.0075)



NA NA NA NA NA NA

NA NA NA NA

NA

NA rs1045642CC and haplotype rs2032582GG-rs1045642CC higher risk for >3A rejection (p ¼ 0.02) D haplotype rs1128503T-rs2032582T-rs1045642T higher risk for AR (p ¼ 0.018) and GL (p ¼ 0.0019) rs1045642T and rs2032582T higher risk for DGF (p < 0.034) Haplotype rs1128503C-rs2032582G-rs1045642T higher risk for AR (p ¼ 0.038) NA D and R rs104564TT higher IF/TA (p ¼ 0.001) NA

NA

Results

Almoguera et al

American Journal of Transplantation 2014; 14: 764–778

Relevant pharmacogenetic studies in transplantation outcomes. Letters a to g in PMID column indicate that particular studies investigate exclusively those variants specified by the letters. For Primary IST, CsA ¼ cyclosporine A; CNI ¼ CsA þ Tac; MPA ¼ mycophenolic acid and Tac ¼ tacrolimus. For sample size, D ¼ donor and R ¼ recipient. For primary outcome, AR ¼ acute rejection; CAD ¼ chronic allograft damage; DGF ¼ delayed graft function; GL ¼ graft loss; GS ¼ graft survival. For sample size, D ¼ donor and R ¼ recipient and in results, NA ¼ No Asociation.

rs7438135, rs7439366, rs7668282 UGT2B7

338 19494809 rs6714486, rs17868320 UGT1A9

AR rs3814055, rs1042597, rs7643645 rs17863762 UGT1A8

MPA þ CNI

Sample size Primary outcome Gene

Table 3: Continued

Variant

Primary IST

PMID

Graft

Kidney

Results

Tac patients with 6714486A and/or rs17868320T higher risk of AR (p < 0.05)

Transplantation Genetics

American Journal of Transplantation 2014; 14: 764–778

pharmacokinetics, transplant outcomes and adverse events. Due to the previously mentioned limitations of the candidate gene approach, we only focus on studies with over 100 individuals and appropriate quality control for phenotype and genotype data. Several studies have confirmed the major impact of CYP3A5 on tacrolimus dose requirements and renal clearance; however, similar to genetic association studies on graft outcomes, it has been challenging to robustly replicate additional findings. Apart from CYP3A5, to date, only the associations of ABCB1 rs1045642, rs1128503, rs2032582 and cyclosporine pharmacokinetics and CYP3A4 rs2740574 and rs3559936 with tacrolimus pharmacokinetics appear to be consistent. All of the rigorously designed studies to date conclude that there is a need to dose tacrolimus patients by the CYP3A5 rs776746 genotype. Thervet et al (15), during the first 6 days posttransplantation, studied the effect of guiding tacrolimus dosing based on the number of CYP3A5 functional alleles, on plasma drug concentration. Patients were randomly assigned to either a standard initial dose of tacrolimus (0.2 mg/kg/day) or a genotype-adjusted dose, who received 0.3 mg/kg/day if carrying allele  1 and 0.15 mg/kg/day if  3/  3. This randomized clinical trial found significantly higher number of patients achieving optimal dosing in the genotype-based dosing group, more rapidly and with fewer dose modifications (15). However, the trial was not designed to investigate hard clinical outcomes such as graft failure—the acid test of a pharmacogenetic test. Passey et al (34) established a dosing algorithm for tacrolimus including clinical, genetic and ethnic information. From a cohort of more than 600 kidney recipients, compared to those with G/G genotype, those with the rs776746 A/G genotype experienced a 69% increase in tacrolimus clearance and A/A genotype a 100% increase in clearance. The final dosing algorithm included CYP3A5 rs776746 genotype, days posttransplantation, age, steroid and calcium channel blocker use as independent predictors of the outcome (34). The incorporation of CYP3A5 genetic information into tacrolimus dosing algorithm will likely be a first major step toward precision genotype-guided dosing in the transplantation setting. Optimal use of pharmacogenetics will require a deeper knowledge of additional genetics factors governing drug disposition, efficacy and toxicity. The application of the genomic advances to this field has the real potential of optimizing dosing strategies for immunosuppressive drugs, avoiding serious adverse and improving patient management after organ transplantation. Ancestry as a genetic risk factor for transplantation outcomes It is also worth noting the large impact that donor and recipient ancestry has in transplantation. Many studies have reported greater risks of rejection and mortality in African Americans when their organs are used as donor 775

776 MPA

MPA

rs5186

rs2740574

rs776746

rs11770116, rs2228075, rs2288548, rs2288549, rs2288550, rs2288553, rs4731448 and rs2278293 (b), rs2278294 (c)

rs11706052, rs4974081

521T > C, rs2306283, rs4149015, rs4149056

rs4149117

98T > C, rs17868320, rs6714486

rs7438135, rs7439366

CYP3A4

CYP3A5

IMPDH1

IMPDH2

SLCO1B1

SLCO1B3

UGT1A9

UGT2B7

Blood pressure

Diarrhea,leukopenia, anemia, infections

456

332 218

18641546 (diarrhea, leukopenia) 21142914

218

20679962 21142914

191

218 456

191

399D-R

209D-R 304

304

233

17851563 (leukopenia)

21142914b,c 20679962 (leukopenia,infection) b,c

17851563 (leukopenia)

15454731

21677300 20526235

20526235

16477235

218

209D-R 304 218

21677300a 20526235 21142914a 21142914

294

Sample size

20061922

PMID

Kidney

Kidney

Kidney

Kidney

Kidney

Kidney Kidney

Kidney

Kidney

Kidney Kidney

Kidney

Liver

Kidney

Kidney Kidney Kidney

Heart

Graft

NA

NA

rs4149056C reduced risk of adverse events (p ¼ 0.002)

NA

NA

NA rs2278294 higher risk of leukopenia (p ¼ 0.0139)

NA

NA rs776746G higher risk of nephrotoxicity (p ¼ 0.01) NA

NA

rs4343DD higher risk of nephrotoxicity (p < 0.0001)

NA

NA NA NA

NA

Results

Pharmacogenetic studies on IST induced adverse events. Letters a to g in PMID column indicate that particular studies investigate exclusively those variants specified by the letters. For Primary IST, CsA ¼ cyclosporine A; CNI ¼ CsA þ Tac; MPA ¼ mycophenolic acid; Tac ¼ tacrolimus. For sample size, D ¼ donor and R ¼ recipient. For sample size, D ¼ donor and R ¼ recipient and in results, NA ¼ No Association.

MPA

Diarrhea,leukopenia, anemia, infections

Leukopenia, infections

Diarrhea,leukopenia, anemia, infections

CsA MPA

Nephrotoxicity

Nephrotoxicity

Nephrotoxicity

Diarrhea,leukopenia, anemia, infections

Diarrhea,leukopenia, anemia, infections

Nephrotoxicity

Adverse event

Tac

Tac

rs5051

AGTR1

CNI

MPA

AGT

rs2273697, rs3740066, 1446C > G, rs717620

ABCC2

MPA

Tac

CsA

rs4343

rs1045642 (a), rs1128503, rs2032582

ABCB1

Primary IST

ACE

Variant

Gene

Table 4: Pharmacogenetics of immunosuppressant therapy–induced adverse events

Almoguera et al

American Journal of Transplantation 2014; 14: 764–778

Transplantation Genetics

grafts or when they are a recipient of a graft (35–37). Callender et al also confirmed lower kidney survival rates among African Americans compared to all other ethnic groups and showed that this ethnic group also harbored the shortest half-life after kidney transplantation: only 5.3 years versus 12.2 years for Asians, 10.2 years for individuals of European descent and 9.0 years for Hispanics (36). In 2010, Li et al (38) published a pivotal study showing that Chinese and African human genomes differed by approximately 5 Mb of unique sequence and over 240 potential genes differing between these sequences. Such ‘‘pangenome’’ data sets illustrate the differences between populations and emphasize the need for further characterization of such genomic differences in the transplantation arena. In addition to the highly polymorphic differences in HLA loci already evident in the different ethnic groups, variants like LoFs or hdCNVs with the potential of acting as incompatibility antigens, or genetic risk factors for postoperative transplant outcomes as well as pharmacogenetic markers, exhibit substantial variability interethnically.

Future Perspectives The current pace of genomic discovery has revolutionized our understanding of the genetic architecture of common diseases (www.genome.gov/gwastudies) (29). In contrast, the pipeline of genomic research in transplantation has been relatively weak. This is surprising because in essence transplantation involves the direct exposure of a foreign donor genome into a recipient, creating unparalleled opportunities to investigate the implications this has on graft function. For example, recipient by donor genomic interactions may play a critical role in graft survival. However, analyses of these types will require large wellcharacterized data sets that have been harmonized for important clinical traits and genotyped using GWAS or preferably second-generation sequencing. While small candidate studies have started the collation of DNA and appropriately phenotyped recipients and donors, very few studies have sufficient numbers to be statistically powered to look at polygenic transplant-related intermediate phenotypes and outcomes. There is a clear need for wider collaborations among all appropriately harmonizable data sets using robust genomic platforms and statistical methods to bolster statistical power among existing studies. With such initiatives, we expect the genetic architecture of transplant rejection ranging from identification of LoF variants, which introduce new epitopes to donors, to genes involved in drug metabolism to change the way transplanted grafts are matched and/or recipients are monitored posttransplant. Furthermore, established genetic techniques such as hypothesis-free analyses and Mendelian randomization can be conducted in this setting to investigate the causal effect of modifiable risk factors on graft function and survival with the potential to discover new therapeutic targets. American Journal of Transplantation 2014; 14: 764–778

Large-scale NIH networks such as The electronic Medical Records and Genomics (eMERGE) network integrating large-scale hospital GWAS biobank resource with EMRs (www.genome.gov/27540473) will allow us a framework to consent, counsel and manage the return of results to physicians and graft recipients in a clinical environment, which will lead to greater patient monitoring and care. As waiting lists for kidney transplantation continue to grow (n ¼ 97 856, September 2013) and with solid organ transplants costing from $200 000 to $850 000 (USD) per patient, efforts such as eMERGE are currently attempting to integrate actionable pharmacogenomics data into patients’ EMRs, and tacrolimus-related dosing is being conducted in at least one eMERGE site thus far, which will likely result in a significant benefit to society in the very near future. Although beyond the scope of this review, a number of protein, RNA and microRNA biomarkers for rejection and graft damage injury are emerging that may aid in the diagnosis of rejection episodes. Specific biomarkers are also under investigation for immune quiescence and minimization of immune-suppressants (reviewed in (39)). We are now entering a period in genomics when tools are becoming available, which will dramatically change our understanding of the molecular mechanisms underpinning graft outcomes, and patient-specific treatments will begin to benefit transplant patient care and to define the future of transplantation.

Disclosure The authors of this manuscript have no conflicts of interest to disclose as described by the American Journal of Transplantation.

References

Electronic Resources Described A Catalog of Published Genome-Wide Association Studies: www.genome.gov/gwastudies/. The electronic Medical Records and Genomics (eMERGE) network: www.genome.gov/27540473. 1. Spierings E, Hendriks M, Absi L, et al. Phenotype frequencies of autosomal minor histocompatibility antigens display significant differences among populations. PLoS Genet 2007; 3: e103. 2. Terasaki PI. Deduction of the fraction of immunologic and nonimmunologic failure in cadaver donor transplants. Clin Transpl 2003; 449–452. 3. Burton CM, Iversen M, Carlsen J, et al. Acute cellular rejection is a risk factor for bronchiolitis obliterans syndrome independent of post-transplant baseline FEV1. J Heart Lung Transplant 2009; 28: 888–893.

777

Almoguera et al 4. Stehlik J, Edwards LB, Kucheryavaya AY, et al. The Registry of the International Society for Heart and Lung Transplantation: Twentyseventh official adult heart transplant report—2010. J Heart Lung Transplant 2010; 29: 1089–1103. 5. Trulock EP, Christie JD, Edwards LB, et al. Registry of the International Society for Heart and Lung Transplantation: Twentyfourth official adult lung and heart-lung transplantation report— 2007. J Heart Lung Transplant 2007; 26: 782–795. 6. Mangi AA, Mason DP, Nowicki ER, et al. Predictors of acute rejection after lung transplantation. Ann Thorac Surg 2011; 91: 1754–1762. 7. Gratwohl A, Dohler B, Stern M, Opelz G. H-Y as a minor histocompatibility antigen in kidney transplantation: A retrospective cohort study. Lancet 2008; 372: 49–53. 8. McCarroll SA, Bradner JE, Turpeinen H, et al. Donor–recipient mismatch for common gene deletion polymorphisms in graftversus-host disease. Nat Genet 2009; 41: 1341–1344. 9. Tan JC, Kim JP, Chertow GM, Grumet FC, Desai M. Donor– recipient sex mismatch in kidney transplantation. Gend Med 2012; 9: 335–347. 10. Courivaud C, Bamoulid J, Loupy A, et al. Influence of cyclooxygenase-2 (COX-2) gene promoter polymorphism-765 on graft loss after renal transplantation. Am J Transplant 2009; 9: 2752–2757. 11. Birdwell KA, Grady B, Choi L, et al. The use of a DNA biobank linked to electronic medical records to characterize pharmacogenomic predictors of tacrolimus dose requirement in kidney transplant recipients. Pharmacogenet Genomics 2012; 22: 32–42. 12. 1000 Genomes Project Consortium, Abecasis GR, Altshuler D, et al. A map of human genome variation from population-scale sequencing. Nature 2010; 467: 1061–1073. 13. Press RR, Ploeger BA, den Hartigh J, et al. Explaining variability in tacrolimus pharmacokinetics to optimize early exposure in adult kidney transplant recipients. Ther Drug Monit 2009; 31: 187–197. 14. Bandur S, Petrasek J, Hribova P, Novotna E, Brabcova I, Viklicky O. Haplotypic structure of ABCB1/MDR1 gene modifies the risk of the acute allograft rejection in renal transplant recipients. Transplantation 2008; 86: 1206–1213. 15. Thervet E, Loriot MA, Barbier S, et al. Optimization of initial tacrolimus dose using pharmacogenetic testing. Clin Pharmacol Ther 2010; 87: 721–726. 16. Roberts JP, Wolfe RA, Bragg-Gresham JL, et al. Effect of changing the priority for HLA matching on the rates and outcomes of kidney transplantation in minority groups. N Engl J Med 2004; 350: 545– 551. 17. Navarro V, Herrine S, Katopes C, Colombe B, Spain CV. The effect of HLA class I (A and B) and class II (DR) compatibility on liver transplantation outcomes: An analysis of the OPTN database. Liver Transpl 2006; 12: 652–658. 18. Clark B, Unsworth DJ. HLA and kidney transplantation. J Clin Pathol 2010; 63: 21–25. 19. Dierselhuis M, Goulmy E. The relevance of minor histocompatibility antigens in solid organ transplantation. Curr Opin Organ Transplant 2009; 14: 419–425. 20. Warren EH, Zhang XC, Li S, et al. Effect of MHC and non-MHC donor/recipient genetic disparity on the outcome of allogeneic HCT. Blood 2012; 120: 2796–2806.

778

21. Goulmy E, Termijtelen A, Bradley BA, van Rood JJ. Alloimmunity to human H-Y. Lancet 1976; 2: 1206. 22. Tan JC, Wadia PP, Coram M, et al. H-Y antibody development associates with acute rejection in female patients with male kidney transplants. Transplantation 2008; 86: 75–81. 23. Ofran Y, Kim HT, Brusic V, et al. Diverse patterns of T-cell response against multiple newly identified human Y chromosome-encoded minor histocompatibility epitopes. Clin Cancer Res 2010; 16: 1642–1651. 24. Roopenian D, Choi EY, Brown A. The immunogenomics of minor histocompatibility antigens. Immunol Rev 2002; 190: 86–94. 25. Tennessen JA, Bigham AW, O’Connor TD, et al. Evolution and functional impact of rare coding variation from deep sequencing of human exomes. Science 2012; 337: 64–69. 26. MacArthur DG, Balasubramanian S, Frankish A, et al. A systematic survey of loss-of-function variants in human protein-coding genes. Science 2012; 335: 823–828. 27. Murata M, Warren EH, Riddell SR. A human minor histocompatibility antigen resulting from differential expression due to a gene deletion. J Exp Med 2003; 197: 1279–1289. 28. Consortium TIHP. The International HapMap Project. Nature 2003; 426: 789–796. 29. Hindorff LA, Sethupathy P, Junkins HA, et al. Potential etiologic and functional implications of genome-wide association loci for human diseases and traits. Proc Natl Acad Sci USA 2009; 106: 9362–9367. 30. Evans WE, McLeod HL. Pharmacogenomics—Drug disposition, drug targets, and side effects. N Engl J Med 2003; 348: 538–549. 31. Ekberg H, Tedesco-Silva H, Demirbas A, et al. Reduced exposure to calcineurin inhibitors in renal transplantation. N Engl J Med 2007; 357: 2562–2575. 32. Kuehl P, Zhang J, Lin Y, et al. Sequence diversity in CYP3A promoters and characterization of the genetic basis of polymorphic CYP3A5 expression. Nat Genet 2001; 27: 383–391. 33. Xie HG, Wood AJ, Kim RB, Stein CM, Wilkinson GR. Genetic variability in CYP3A5 and its possible consequences. Pharmacogenomics 2004; 5: 243–272. 34. Passey C, Birnbaum AK, Brundage RC, Oetting WS, Israni AK, Jacobson PA. Dosing equation for tacrolimus using genetic variants and clinical factors. Br J Clin Pharmacol 2011; 72: 948– 957. 35. Christie JD, Kotloff RM, Pochettino A, et al. Clinical risk factors for primary graft failure following lung transplantation. Chest 2003; 124: 1232–1241. 36. Callender CO, Cherikh WS, Miles PV, et al. Blacks as donors for transplantation: Suboptimal outcomes overcome by transplantation into other minorities. Transplant Proc 2008; 40: 995–1000. 37. Nair S, Eustace J, Thuluvath PJ. Effect of race on outcome of orthotopic liver transplantation: A cohort study. Lancet 2002; 359: 287–293. 38. Li R, Li Y, Zheng H, et al. Building the sequence map of the human pan-genome. Nat Biotechnol 2010; 28: 57–63. 39. Sarwal MM, Benjamin J, Butte AJ, Davis MM, Wood K, Chapman J. Transplantomics and biomarkers in organ transplantation: A report from the first international conference. Transplantation 2011; 91: 379–382.

American Journal of Transplantation 2014; 14: 764–778

Transplantation genetics: current status and prospects.

Over the last decade, advances in genetic technologies have accelerated our understanding of the genetic diversity across individuals and populations...
245KB Sizes 3 Downloads 3 Views