Вы находитесь на странице: 1из 98

Volume 16 • Number 4 • November 2017

NephSAP
®

Nephrology Self-Assessment Program

Transplantation
Co-Editors:
John P. Vella, MD
Alexander C. Wiseman, MD

Co-Directors:
Gerald Hladik, MD
Jerry Yee, MD
Preface
CO-DIRECTOR,
NephSAP
Gerald A. Hladik, MD
University of North Carolina at Chapel Hill NephSAPÒ is one of the premiere educational activities of the American Society of
Chapel Hill, NC
Nephrology (ASN). Its primary goals are self-assessment, education, and the provision of
CO-DIRECTOR, Continuing Medical Education (CME) credits and Maintenance of Certification (MOC) points
NephSAP for individuals certified by the American Board of Internal Medicine. Members of the ASN
Jerry Yee, MD, FASN receive NephSAP electronically through the ASN website by clicking on the NephSAP link
Henry Ford Hospital
Detroit, MI under “Education and Meetings” tab.
MANAGING EDITOR EDUCATION: Medical and nephrologic information continually accrues at a rapid pace.
Gisela Deuter, BSN, MSA Bombarded from all sides with demands on their time, busy practitioners, academicians, and
Washington, DC trainees at all levels are increasingly challenged to review and understand new and evolving
ASSOCIATE EDITORS evidence. Each bimonthly issue of NephSAP is dedicated to a specific theme, i.e., to a specific
Debbie L. Cohen, MD area of clinical nephrology, hypertension, dialysis, and transplantation, and consists of an
University of Pennsylvania School of Medicine editorial, a syllabus, and self-assessment questions, to serve as a self-study device. Over the
Philadelphia, PA
course of 24 months, all clinically relevant and key elements of nephrology will be reviewed
Richard J. Glassock, MD
Professor Emeritus, The David Geffen School of
and updated. The authors of each issue digest, assimilate, and interpret key studies published
Medicine at the University of California since the release of the previous issues and integrate this new material with the body of
Los Angeles, CA existing information. Occasionally a special edition is produced to cover an area not ordinarily
Stanley Goldfarb, MD addressed by core issues of NephSAP.
University of Pennsylvania School of Medicine
Philadelphia, PA
SELF-ASSESSMENT: Thirty, single-best-answer questions will follow the 80 to 100 pages of
Karen A. Griffin, MD, FASN
Loyola University Medical Center
syllabus text. The examination is available online with immediate feedback. Those answering
Maywood, IL 75% correctly will receive MOC and CME credit, and receive the answers to all the questions
Jay L. Koyner, MD along with brief discussions and an updated bibliography. Members will find a new area
University of Chicago
Chicago, IL
reviewed every 2 months, and they will be able to test their understanding with our quiz. This
Holly J. Kramer, MD format will help readers stay up to date in developing areas of clinical nephrology,
Loyola University Medical Center hypertension, dialysis, and transplantation, and the review and update will support those
Maywood, IL
taking certification and recertification examinations.
Ruediger W. Lehrich, MD
Duke University CONTINUING MEDICAL EDUCATION: Most state and local medical agencies as well as
Durham, NC
Kevin J. Martin, MBBCh
hospitals are demanding documentation of requisite CME credits for licensure and for staff
St. Louis University School of Medicine appointments. A maximum of 50 credits annually can be obtained by successfully completing
St. Louis, MO the NephSAP examinations. In addition, individuals enrolled in Maintenance of Certification
John P. Middleton, MD
Duke University
(MOC) through the American Board of Internal Medicine may obtain points toward MOC by
Durham, NC successfully completing the self-assessment examination of NephSAP.
Sankar D. Navaneethan, MD, MPH
Baylor College of Medicine This paper meets the requirements of ANSI/NISO Z39.48-1921 (Permanence of Paper),
Houston, TX effective with July 2002, Vol. 1, No. 1.
Asghar Rastegar, MD
Yale School of Medicine
New Haven, CT
Brad H. Rovin, MD
Ohio State University Medical Center
Columbus, OH
Manoocher Soleimani, MD
University of Cincinnati
Cincinnati, OH
Charuhas V. Thakar, MD
University of Cincinnati
Cincinnati, OH
John P. Vella, MD
Maine Medical Center
Portland, ME
Alexander C. Wiseman, MD
University of Colorado at Denver
Denver, CO

FOUNDING EDITORS
Richard J. Glassock, MD
Editor-in-Chief Emeritus NephSAPÒ
Robert G. Narins, MD Ó2017 by The American Society of Nephrology
Volume 16, Number 4, November 2017

Editorial 295 Recipient Factors

295 Delayed Graft Function and Rejection Risk


273 Modern Hepatitis C Virus Therapy and Its Effect on
Transplantation 296 Immunosuppression
Andrew F. Malone, Daniel C. Brennan
296 Induction Therapy
Syllabus 298 Maintenance Therapy

280 NephSAP, Volume 16, Number 4, November 2017— 298 Tacrolimus


Transplantation Pharmacokinetics
300
John P. Vella, Alexander C. Wiseman
301 Extended Release Tacrolimus
280 Learning Objectives
301 Antimetabolite Therapy
280 Access to Transplantation and Outcomes
302 Belatacept
280 Patient and Graft Survival
302 Steroid Therapy
282 Ethnicity
303 Mammalian Target of Rapamycin Inhibitor Therapy
283 Physical Frailty
305 Rejection
283 Cardiovascular Outcomes
307 Temporal Dynamics
284 Obesity
307 Transfusion
284 Deceased Donor Kidney Allocation
307 Novel Rejection Mediators
285 Discarded Kidneys
308 Adherence
286 Deceased Donor Hypothermia
308 Gene Activation
286 Use of Hepatitis C Virus–Seropositive Kidneys
308 Subclinical Rejection
286 Prior Living Donors in Need of Transplantation
309 Antibody-Mediated Rejection
287 Public Health Service Increased Risk Donors
310 Immunoglobulin Subtypes
288 Living Kidney Donation
311 Histoincompatibility Strategies
288 Kidney Paired Donation
312 ABO-Incompatible Transplantation
289 Living Donor Kidney Function Assessment
314 HLA Antibody Desensitization
289 Medically Complex Living Donors
314 Novel Desensitization Strategies
290 Surgery Complications

290 Pregnancy Risk 315 Clinical Tolerance

291 Kidney Failure Risk 316 Clinical Tolerance Induction

291 Prior Living Donors and Kidney Transplantation 317 Biomarkers

292 APOL1 319 Glomerular Injury Post-Transplant


292 Other Complications 319 Proteinuria in the Post-Transplant Setting
293 Delayed Graft Function 320 Glomerular Disease: Post-Transplant Outcomes
293 Premortem AKI 320 Focal and Segmental Glomerulosclerosis
294 Predonation Interventions 321 Membranous Nephropathy
294 Procedure Times 321 APOL1 and Kidney Transplantation
Volume 16, Number 4, November 2017

321 Complement-Mediated Glomerular Disease 340 Interventions and Outcomes

323 Infection after Transplant 342 Bone and Mineral Disorders


324 Cytomegalovirus 342 Hyperparathyroidism

324 BK Virus 343 Osteoporosis

325 HIV 345 Multiorgan Transplant


327 Hepatitis C Virus 345 Simultaneous Pancreas-Kidney Transplantation
328 Hepatitis B Virus 346 Simultaneous Liver-Kidney Transplantation
328 Epstein Barr Virus 347 Multiorgan Transplantation: Utility Considerations
329 Malignancy and Kidney Transplantation 347 Pregnancy
330 Risk Factors for Malignancy in the Transplant 349 Acid-Base and Electrolyte Disorders after Transplant
Recipient
333 Transplant Candidate Cancer Screening CME Self-Assessment Questions
334 Recipient Screening for Malignancy 350 NephSAP, Volume 16, Number 4, November 2017—
335 Cancer Management in the Transplant Recipient Transplantation

336 Cardiovascular Disease Upcoming Issues


336 Pretransplant Screening Hypertension
Debbie L. Cohen, MD and Karen A. Griffin, MD
337 Hypertension
March 2018
338 Post-Transplant Diabetes Mellitus
Primary and Secondary Glomerular Diseases
338 Hyperlipidemia Richard J. Glassock, MD and Brad H. Rovin, MD
Obesity June 2018
339
339 Nontraditional and Novel Cardiovascular Risk Interventional Nephrology and Dialysis Access
Factors under Investigation Anil Agarwal, MD, Lalathaksha Kumbar, MD and Vandana Dua Niyyar, MD
July 2018
339 Structural Parameters
Disorders of Divalent Ions, Renal Bone Disease, and Nephrolithiasis
339 Biochemical Markers
Stanley Goldfarb, MD and Kevin J. Martin, MBBCh
340 Clinical Features September 2018
Volume 16, Number 4, November 2017

The Editorial Board of NephSAP and KSAP extends its sincere appreciation to the following reviewers. Their efforts and insights help improve
the quality of these postgraduate education offerings.
NephSAP Review Panel
Mustafa Ahmad, MD, FASN Chokchai Chareandee, MD, FASN Pedram Fatehi, MD
King Fahad Medical City University of Minnesota Stanford Medicine
Riyadh, Saudi Arabia Minneapolis, MN Palo Alto, CA
Nasimul Ahsan, MD, FASN Joline L. Chen, MD William H. Fissell, MD
University of Florida and Long Beach Veteran Affairs Vanderbilt University Medical Center
Oscar G. Johnson Veteran Affairs Healthcare System Nashville, TN
Medical Center Orange, CA
Iron Mountain, MI D. Kevin Flood, MD
Karen Ching, MD Mike O’Callaghan Federal Medical Center
Jafar Al-Said, MD, FASN Hawaii Permanente Medical Group Nellis AFB, NV
Bahrain Specialist Hospital Honolulu. HI
Manama, Bahrain Lynda A. Frassetto, MD, FASN
W. James Chon, MD University of California at
Carmichael Angeles, MD, FASN University of Arkansas for Medical Sciences San Francisco
John T. Mather Memorial Hospital Little Rock, AR San Francisco, CA
Stony Brook, NY Jason Cobb, MD Tibor Fulop, MD
Kisra Anis, MBBS Emory University School of Medicine University of Mississippi Medical Center
Jacobi Medical Center/ Atlanta, GA Jackson, MS
Albert Einstein College of Medicine Armando Coca, MD, PhD
Bronx, NY Maurizio Gallieni, MD, FASN
Hospital Clínico Universitario
University of Milano
Valladolid, Spain
Naheed Ansari, MD, FASN Milano, Italy
Jacobi Medical Center/Albert Einstein Scott D. Cohen, MD, FASN
College of Medicine Duvuru Geetha, MD, FASN
George Washington University
Bronx, NY Johns Hopkins University
Washington, DC
Baltimore, MD
Nabeel Aslam, MD, FASN Beatrice Concepcion, MD
Mayo Clinic Florida Vanderbilt University Medical Center Ilya Glezerman, MD
Jacksonville, FL Nashville, TN Memorial Sloan Kettering Cancer Center
New York, NY
Nisha Bansal, MD Gabriel Contreras, MD
University of Washington University of Miami Carl S. Goldstein, MD, FASN
Seattle, WA Miami, FL Rutgers University
New Brunswick, NJ
Krishna M. Baradhi, MD Patrick Cunningham, MD
University of Oklahoma University of Chicago Basu Gopal, MBBS, FASN
Tulsa, OK Chicago, IL Royal Adelaide Hospital
Adelaide, Australia
Emmy Bell, MD, MSPH Kevin A. Curran, MD
University of Alabama at Birmingham Kevin A. Curran, MD, PA Steven Gorbatkin, MD, PhD
Birmingham, AL Canton, TX Emory University and
Atlanta Veteran Affairs Medical Center
Bruce E. Berger, MD Rajiv Dhamija, MD
Decatur, GA
Case Western Reserve University Rancho Los Amigos National
Cleveland, OH Rehabilitation Center Aditi Gupta, MD
Downey, CA University of Kansas Medical Center
Mona B. Brake, MD
Robert J. Dole Veteran Affairs Alejandro Diez, MD Kansas City, KS
Medical Center The Ohio State University Susan Hedayati, MD
Wichita, KS Columbus, OH University of Texas Southwestern
Pooja Budhiraja, MBBS John J. Doran, MD Dallas, TX
University of Kansas Medical Center Emory School of Medicine Marie C. Hogan, MBBCh, PhD
Kansas City, KS Atlanta, GA Mayo Clinic
Ruth C. Campbell, MD Randa A. El Husseini, MD, FASN Rochester, MN
Medical University of South Carolina HealthPartners Medical Group
Susie Hu, MD
Charleston, SC St. Paul, MN
Warren Alpert Medical School of Brown
Chia-Ter Chao, MD Mahmoud El-Khatib, MD University,
National Taiwan University Hospital University of Cincinnati Rhode Island Hospital
Taipei, Taiwan Cincinnati, OH Providence, RI
Edmund Huang, MD Tingting Li, MD Todd Pesavento, MD
University of California at Washington University in St. Louis Ohio State University
Los Angeles School of Medicine St. Louis, MO Columbus, OH
Los Angeles, CA
Orfeas Liangos, MD, FASN Phuong-Thu Pham, MD
Ekambaram Ilamathi, MD, FASN Klinikum Coburg David Geffen School of Medicine at UCLA
Northwell Health, Southside Hospital Coburg, Bayern, Germany Los Angeles, CA
Bayshore, NY
Michael Lioudis, MD Pairach Pintavorn, MD, FASN
Joshua M. Kaplan, MD Cleveland Clinic Nephrology East Georgia Kidney and Hypertension
Rutgers New Jersey Medical School Cleveland, OH Augusta, GA
Newark, NJ
Ajit Mahapatra, MD
Roberto Pisoni, MD
Amir Kazory, MD The Permanente Medical Group
University of Florida Medical University of South Carolina
Santa Clara, CA
Gainesville, FL Charleston, SC
A. Bilal Malik, MBBS
Quresh T. Khairullah, MD University of Washington James M. Pritsiolas, MD, FASN
St. Clair Nephrology Seattle, WA CarePoint Health - Bayonne Medical Center
Roseville, MI Bayonne, NJ
Jolanta Malyszko, MD, PhD
Apurv Khanna, MD Medical University Paul H. Pronovost, MD, FASN
State University of New York Bialystok, Poland Yale University School of Medicine
Upstate Medical University Waterbury, CT
Syracuse, NY Ernest Mandel, MD
Brigham and Women’s Hospital Mohammad A. Quasem, MD, FASN
Yong-Lim Kim, MD, PhD Boston, MA State University of New York Medical
Kyungpook National University Hospital University
Daegu, South Korea Naveed N. Masani, MD Binghamton, NY
Winthrop University Hospital
Nitin V. Kolhe, MD, FASN Mineola, NY
Derby Teaching Hospital NHS Trust Wajeh Y. Qunibi, MD
Derby, Derbyshire, UK University of Texas
Teri Jo Mauch, MD, PhD
Health Science Center
University of Nebraska College of Medicine
Farrukh M. Koraishy, MD, PhD San Antonio, TX
Omaha, NE
St. Louis University
St. Louis, MO Hanna W. Mawad, MD, FASN Pawan K. Rao, MD, FASN
University of Kentucky St. Joseph’s Hospital and Health Center
Eugene C. Kovalik, MD
Lexington, KY Syracuse, NY
Duke University Medical Center
Durham, NC Ellen T. McCarthy, MD Hernan Rincon-Choles, MD
University of Kansas Medical Center, Cleveland Clinic Foundation
Steven Kraft, MD Cleveland, OH
Kidney Institute
Western Nephrology
Kansas City, KS
Lafayette, CO Dario Roccatello, MD
Vineeta Kumar, MD Kirtida Mistry, MBBCh San Giovanni Hospital and
University of Alabama at Birmingham Children’s National Medical Center University of Torino
Birmingham, AL Washington, DC Torino, Italy

Sarat Kuppachi, MD Lawrence S. Moffatt, MD Helbert Rondon-Berrios, MD, FASN


University of Iowa Carolinas Medical Center University of Pittsburgh
Iowa City, IA Charlotte, NC School of Medicine
Pittsburgh, PA
Norbert H. Lameire, MD, PhD David B. Mount, MD
University Hospital Brigham and Women’s Hospital, Ehab R. Saad, MD, FASN
Gent, East Flanders, Belgium Harvard Medical School Medical College of Wisconsin
Boston, MA Milwaukee, WI
Sheron Latcha, MD
Memorial Sloan Kettering Cancer Center Thangamani Muthukumar, MD Mark C. Saddler, MBChB
New York, NY Weill Cornell Medicine Mercy Regional Medical Center
New York, NY Durango, CO
Vincent Weng Seng Lee, MBBS, PhD
Westmead Hospital Mohanram Narayanan, MD Neil Sanghani, MD
Sydney, NSW Baylor Scott & White Health Vanderbilt University
Australia Temple, TX Nashville, TN

Paolo Lentini, MD, PhD Macaulay A. Onuigbo, MD Mohammad N. Saqib, MD


St. Bortolo Hospital Mayo Clinic Lehigh Valley Hospital
Bassano del Grappa, Italy Rochester, MN Allentown, PA
Oliver Lenz, MD Rosemary Ouseph, MD Hitesh H. Shah, MD
University of Miami Health System St. Louis University Hofstra Northwell School of Medicine
Miami, FL Webster Groves, MO Great Neck, NY
Michiko Shimada, MD, PhD Hung-Bin Tsai, MD Nand K. Wadhwa, MD
Hirosaki University National Taiwan University Stony Brook University
Hirosaki, Japan Hospital Taipei, Taiwan Stony Brook, NY
Shayan Shirazian, MD Katherine Twombley, MD Connie Wang, MD
Winthrop-University Hospital Medical University of South Carolina Hennepin County Medical Center
State University of New York at Charleston, SC Minneapolis, MN
Stony Brook Kausik Umanath, MD Maura A. Watson, DO
Mineola, NY Henry Ford Hospital Walter Reed National Military Medical Center
Arif Showkat, MD, FASN Detroit, MI Bethesda, MD
University of Tennessee Puchimada M. Uthappa, MBBS, FASN Dawn Wolfgram, MD
Memphis, TN Columbia Asia Hospital Medical College of Wisconsin
Mysore, Karnataka, India Milwaukee, WI
Stephen M. Sozio, MD
Johns Hopkins University School of Medicine Anthony M. Valeri, MD Sri Yarlagadda, MBBS
Baltimore, MD Columbia University Medical Center University of Kansas Medical Center
New York, NY Kansas City, KS
Ignatius Yun-Sang Tang, MD, FASN Allen W. Vander, MD, FASN Brian Young, MD
University of Illinois at Chicago Kidney Center of South Santa Clara Valley Medical Center
Chicago, IL Louisiana San Jose, CA
Ahmad R. Tarakji, MD Thibodaux, LA Mario Javier Zarama, MD
King Saud University, Jon R. Von Visger, MD, PhD Kidney Specialists of
King Khalid University Hospital The Ohio State University Minnesota, PA
Riyadh, Saudi Arabia Columbus, OH Saint Paul, MN
Volume 16, Number 4, November 2017

Program Mission and Objectives


The Nephrology Self-Assessment Program (NephSAP) provides a learning vehicle for clinical nephrologists to renew and
refresh their clinical knowledge, diagnostic, and therapeutic skills. This enduring material provides nephrologists challenging,
clinically oriented questions based on case vignettes, a detailed syllabus that reviews recent publications, and an editorial on an
important and evolving topic. This combination of materials enables clinicians to rigorously assess their strengths and
weaknesses in the broad domain of nephrology.

Accreditation Statement
The American Society of Nephrology (ASN) is accredited by the Accreditation Council for Continuing Medical Education to
provide continuing medical education for physicians.

AMA Credit Designation Statement


The ASN designates this enduring material for a maximum of 10 AMA PRA Category 1 Credits™. Physicians should claim
only the credit commensurate with the extent of their participation in the activity.

Original Release Date


November 2017

CME Credit Termination Date


October 31, 2019

Examination Available Online


On or before Wednesday, November 15, 2017

Estimated Time for Completion


10 hours

Answers with Explanations


•Provided with a passing score after the first and/or after the second attempt
•November 2019: posted on the ASN website when the issue is archived.

Target Audience
• Nephrology certification and recertification candidates
• Practicing nephrologists
• Internists

Method of Participation
•Read the syllabus that is supplemented by original articles in the reference lists.
•Complete the online self-assessment examination.
•Each participant is allowed two attempts to pass the examination (.75% correct) for CME credit.
•Upon completion, review your score and incorrect answers and print your certificate.
•Answers and explanations are provided with a passing score or after the second attempt.
Volume 16, Number 4, November 2017

Activity Evaluation and CME Credit Instructions


• Go to www.asn-online.org/cme, and enter your ASN login on the right.
• Click the ASN CME Center.
• Locate the activity name and click the corresponding ENTER ACTIVITY button.
• Read all front matter information.
• On the left-hand side, click and complete the Demographics & General Evaluations.
• Complete and pass the examination for CME credit.
• Upon completion, click Claim Your CME Credits, check the Attestation Statement box, and enter the number of
CME credits commensurate with the extent of your participation in the activity.
• If you need a certificate, Print Your Certificate on the left.

For your complete ASN transcript, click the ASN CME Center banner, and click View/Print Transcript on the left.

Instructions to obtain American Board of Internal Medicine (ABIM) Maintenance of Certification


(MOC) Points
Each issue of NephSAP provides 10 MOC points. Respondents must meet the following criteria:
• Be certified by ABIM in internal medicine and/or nephrology and enrolled in the ABIM–MOC program
• Enroll for MOC via the ABIM website (www.abim.org).
• Enter your (ABIM) Candidate Number and Date of Birth prior to completing the examination.
• Take the self-assessment examination within the timeframe specified in this issue of NephSAP.
• Upon completion, click Claim Your MOC points, the MOC points submitted will match your CME credits claimed,
check the Attestation Statement box and submit.
• ABIM will notify you when MOC points have been added to your record.

Maintenance of Certification Statement


Successful completion of this CME activity, which includes participation in the evaluation component, enables the participant
to earn up to 10 MOC points in the American Board of Internal Medicine’s (ABIM) Maintenance of Certification (MOC)
program. Participants will earn MOC points equivalent to the amount of CME credits claimed for the activity. It is the CME activity
provider’s responsibility to submit participant completion information to ACCME for the purpose of granting ABIM MOC credit.
MOC points will be applied to only those ABIM candidates who have enrolled in the MOC program. It is your responsibility to
complete the ABIM MOC enrollment process.

System Requirements
Compatible Browser and Software
The ASN website (asn-online.org) has been formatted for cross-browser functionality, and should display correctly in all
modern web browsers. To view the interactive version of NephSAP, your browser must have Adobe Flash Player installed or
have HTML5 capabilities. NephSAP is also available in Portable Document Format (PDF), which requires Adobe Reader or
comparable PDF viewing software.

Monitor Settings
The ASN website was designed to be viewed in a 1024 · 768 or higher resolution.

Medium or Combination of Media Used


The media used include an electronic syllabus and online evaluation and examination.

Technical Support
If you have difficulty viewing any of the pages, please refer to the ASN technical support page for possible solutions. If you
continue having problems, contact ASN at email@asn-online.org.
Volume 16, Number 4, November 2017

Disclosure Information
The ASN is responsible for identifying and resolving all conflicts of interest prior to presenting any educational activity to learners to ensure that
ASN CME activities promote quality and safety, are effective in improving medical practice, are based on valid content, and are independent of the
control from commercial interests and free of bias. All faculty are instructed to provide balanced, scientifically rigorous and evidence-based
presentations. In accordance with the disclosure policies of the Accreditation Council for Continuing Medical Education (ACCME), individuals who are
in a position to control the content of an educational activity are required to disclose relationships with a commercial interest if (a) the relation is financial
and occurred within the past 12 months; and (b) the individual had the opportunity to affect the content of continuing medical education with regard to that
commercial interest. For this purpose, ASN consider the relationships of the person involved in the CME activity to include financial relationships of a spouse
or partner. Peer reviewers are asked to abstain from reviewing topics if they have a conflict of interest. Disclosure information is made available to learners
prior to the start of any ASN educational activity.

EDITORIAL BOARD for this Issue:


Gerald A. Hladik, MD, FASN—Current Employer: University of North Carolina at Chapel Hill; Honoraria: Renal Research Institute;
Scientific Advisor/Membership: ASN Co-Director, NephSAP; Board Member, Renal Research Institute
Ruediger W. Lehrich, MD—Current Employer: Duke University Medical Center
Manoocher Soleimani, MD—Current Employer: University of Cincinnati; Scientific Advisor/Membership: Editorial Board: American Journal of
Kidney Disease
John P. Vella, MD, FASN—Current Employer: Maine Nephrology Associates PA; Research Funding: Bristol-Myers Squibb, Astellas;
Scientific Advisor/Membership: UpToDate
Alexander C. Wiseman, MD—Current Employer: University of Colorado at Denver and Health Sciences Center; Research Funding:
Novartis, Alexion, Bristol-Myers Squibb, Oxford Immunotec, Quark, Astellas; Scientific Advisor/Membership: American Society of
Transplantation, American Journal of Transplantation
Jerry Yee, MD, FASN—Current Employer: Henry Ford Hospital; Consultancy: OptumHealth, Merck, Vasc-Alert, Akebia, Mallinckrodt,
CV-RX, Keryx, DynaMed/Ebsco, Abbvie; Ownership Interest: Merck, Gilead, Vasc-Alert; Honoraria: Washington University of
St. Louis, National Kidney Foundation; Patents/Inventions: Vasc-Alert, Henry Ford Hospital; Scientific Advisor/Membership: NKF:
Editor-in-Chief of Advances in CKD (journal); Editorial Board: American Journal of Nephrology, American Journal of Hypertension,
CJASN, Heart Failure Journal; Section Editor: EBSCO DynaMed, ASN Co-Director, NephSAP; Other Interests: Elsevier and NKF,
Editor-in-Chief, Advances in CKD

EDITORIAL AUTHORS:
Daniel C. Brennan, MD—Current Employer: Washington University in St. Louis; Consultancy: Alexion, CareDx, Immucor, Sanofi,
Veloxis; Research Funding: Alexion, Astellas, Bristol-Myers Squib, CareDx, Oxford Immunotec, Shire, Veloxis; Honoraria:
Alexion, Sanofi, Veloxis; Scientific Advisor/Membership: Editorial Board: Transplantation, UpToDate; Speakers Bureau:
Alexion, Astellas, Novartis, Sanofi, Veloxis
Andrew F. Malone, MBChB—Current Employer: Washington University in St. Louis; Ownership Interests: Gilead

ASN STAFF:
Gisela A. Deuter, BSN, MSA—Nothing to disclose

Commercial Support
There is no commercial support for this issue.
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

Editorial
Modern Hepatitis C Virus Therapy and Its Effect on Transplantation
Andrew F. Malone, MB, BCh, MRCPI, and Daniel C. Brennan, MD, FACP
Division of Nephrology, Department of Medicine, Washington University School of Medicine,
St. Louis, Missouri

The hepatitis C virus (HCV) is a small, enveloped, anti–HCV-positive serology at the time of referral for
single-stranded RNA virus from the family Flavivir- transplant (compared with anti–HCV-negative serol-
idae. This virus is transmitted by contact with small ogy patients) had a relative risk of death from all causes
amounts of blood and can also be transmitted sexually of 1.41 (95% confidence interval [95% CI], 1.01 to
and from mother to baby; it is estimated that 71 million 1.97) and had a relative risk of death due to liver
persons worldwide are infected with HCV, and it is most disease or infection of 2.39 (95% CI, 1.28 to 4.48). For
prevalent in eastern European regions. Between 15% and those HCV-positive patients who went on to get trans-
45% of infected persons spontaneously clear the virus planted (compared with those who remained on di-
within 6 months of infection without any treatment. The alysis), the relative risks of death from all causes were
remaining 55%–85% will develop chronic HCV infection. 4.75 (95% CI, 2.76 to 8.17) between 0 and 3 months,
The risk of cirrhosis of the liver in those with chronic 1.76 (95% CI, 0.75 to 4.13) between 3 and 6 months,
HCV infection is between 15% and 30% within 20 years 0.31 (95% CI, 0.18 to 0.54) between 3 months and 4
(1). Infection is commonly occult, and nearly 80% of years, and 0.84 (95% CI, 0.51 to 1.37) after 4 years (3).
those who become infected do not have any symptoms. Other more recent studies (but still 10 years old) have
This observation renders screening for HCV important confirmed this increased risk in dialysis patients with
among at-risk groups and in areas of high prevalence. HCV. A significant proportion of this risk was due to an
excess of cardiovascular-associated mortality and liver
HCV and ESRD disease–related risk (4–6).
HCV is a significant issue among patients with
ESRD. Patients on hemodialysis are at risk of infection HCV Infection in the Post-Transplant Patient
due to exposure at hemodialysis units, and there is There are complications in the post-transplant
currently no vaccination for HCV in contrast to period related to HCV infection. There is an increase
hepatitis B virus. The prevalence of HCV likely varies in secondary infections in patients with HCV post-
between different dialysis units due to patient behav- transplant, and this is the second most common cause of
iors and intravenous drug use in the surrounding death after cardiovascular disease among these patients
community. Seroconversion rates also vary between (7–11). Extrahepatic malignancies, such as post-transplant
dialysis units and likely reflect facility protocols and lymphoproliferative disorder and myeloma, have been
practices. The Dialysis Outcomes and Practice Patterns associated with HCV infection post-transplantation
Study examined HCV infection in dialysis units by (12–14). Furthermore, there is a direct effect of HCV on
country (2). The unadjusted prevalence of HCV in lymphoid carcinogenesis, and regression of Hodgkin lym-
United States centers was estimated at over 14%, with phoma with treatment of HCV has been shown (15,16).
a seroconversion rate of three per 100 patient-years Patients with HCV are at increased risk of post-transplant
(between 1997 and 2001). Prevalence also increased diabetes mellitus, and post-transplant diabetes mellitus is an
with time on dialysis, rising to over 20% after 10 years. independent risk factor for death and graft loss (17).
Perhaps the increased mortality and morbidity associ- From the renal aspect, recurrence or de novo glomerular
ated with HCV infection are more important. A nearly disease is also a concern, and post-transplant HCV
20-year-old study using New England Organ Bank infection, membranoproliferative GN (cryoglobulinemic
samples and data showed that waitlisted patients with and noncryoglobulinemic), membranous glomerulopathy,
273
274 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

anticardiolipin-associated glomerulopathy, and acute and (among HCV-positive recipients) and when we should
chronic transplant glomerulopathy have all been de- use HCV-positive kidneys. Kucirka et al. (24) explored
scribed. Whether there is an increased risk of allograft the use of HCV kidneys in HCV recipients and found
rejection in patients with HCV infection is debated, but in that African Americans (RR, 1.56; 95% CI, 1.39 to 1.75),
general, HCV infection likely confers an increased degree patients with diabetes (RR, 1.29; 95% CI, 1.18 to 1.40),
of immunosuppression due to reduced number and and those at a center with long wait times (RR, 1.19 per
responsiveness of T-helper lymphocytes (18). Finally, quartile of waiting time; 95% CI, 1.06 to 1.33, P50.002)
HCV infection increases the risk of hepatocellular carci- were more likely to receive an HCV-positive kidney.
noma and liver cirrhosis (19). However, those who received HCV-positive kidneys
Despite the concerns for these complications, waited 310 days less on the waitlist, and the authors
a clear survival advantage exists with kidney trans- suggested that this likely offsets the higher patient and
plantation compared with remaining on dialysis with graft loss associated with HCV-positive kidneys (24).
HCV. Ingsathit et al. (20) performed a meta-analysis of
data reporting outcomes of HCV-positive patients who Modern Era of HCV Treatment: DAAs
were transplanted compared with those who were not A paradigm shift in the approach to HCV man-
transplanted. This study used a random effects model agement due to the recent introduction of DAAs is likely
and reported a pooled risk ratio (RR) for death at 5 years to make such issues outlined above irrelevant soon. In
of 2.19 (95% CI, 1.50 to 3.20) for those remaining on the 2011, the first generation DAAs for treatment of HCV
waiting list compared with those transplanted. were approved by the US Food and Drug Administra-
These results and the advantage of transplantation tion: telaprevir and boceprevir. Efficacy of these oral
existed before the modern era of direct-acting antiviral protease inhibitors was tested in two trials: A New
(DAA) therapy treatment of HCV infection post- Direction in HCV Care: A Study of Treatment-Naive
transplantation. Before the advent of DAA drugs, Hepatitis C Patients with Telaprevir and Serine Protease
treatment was rarely successful, and the use of IFN alfa is Inhibitor Therapy 2 Study (25,26). When added to dual
associated with increased risk of rejection and graft loss. therapy of pegylated IFN (peg-IFN) and ribavirin for the
Thus, treatment options were limited to ribavirin mono- treatment of HCV genotype 1 infections, these agents
therapy; however, this strategy was controversial: im- enhanced sustained virologic response (SVR) rates to
provement in liver histology was variable, and therapy about 70% but at the expense of increased toxicities.
was limited by a dose-dependent hemolytic anemia (21). Since 2013, there has been explosive prolifera-
tion of new oral DAAs in the clinical arena with better
Using HCV-Positive Donor Kidneys in the Pre-DAA tolerability, side effect profiles, and SVR rates (27–
Era 39); also, treatment regimens are typically only 12
There is no doubt that, even in the pre-DAA weeks in duration. DAAs for HCV are NS5A inhib-
therapy era, ESRD patients with HCV infection benefited itors, NS5B inhibitors, and NS3/4A protease inhibitors.
from kidney transplantation. However, whether this Current therapies are combinations of two or more of
benefit remained when kidneys from donors with HCV these drug classes. These viral proteins are HCV-
were transplanted into HCV-positive recipients has encoded nonstructural proteins that constitute the
always been controversial, and this practice varies HCV RNA replication machinery. NS5B is an RNA-
widely. Decisions as to whether to use a kidney from dependent RNA polymerase, NS5A is responsible for
an HCV-positive donor are center specific. It is known the assembly of the membrane-bound replication com-
that transmission of HCV infection from donor to plex, and NS3/4A is a serine protease, an enzyme
recipient is almost 100%. Therefore, only transplantation involved in post-translational processing and replica-
of HCV-positive donors to known HCV-positive recip- tion of HCV (40). Sofosbuvir is an NS5B inhibitor
ients has occurred in routine practice to date (22). HCV- and currently, the most commonly prescribed DAA in
positive patients receiving an HCV-positive kidney combination therapy. This utilization pattern is largely
transplant have worse outcomes compared with HCV- attributable to the success of the Sofosbuvir/Velpatasvir
positive patients receiving an HCV-negative kidney Fixed Dose Combination for 12 Weeks in Adults With
(23). These issues led to the questions of for whom Chronic HCV Infection (ASTRAL-1) Trial, and the
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 275

American Association for the Study of Liver Diseases

Listed are the most common DAA combinations available currently. Columns 3–13 represent HCV genotypes 1–6 and subgenotypes (1a and 1b). Each genotype is further divided into indications for C2 and C1 patients. DAA
1a, C1 1b, C2 1b, C1 2, C2 2, C1 3, C2 3, C1 4, C2 4, C1 5/6, C2/1
guidelines state that sofosbuvir/velpatasvir is the

x
only combination DAA therapy currently recom-
mended for use in all genotypes, regardless of the
presence of compensated cirrhosis (Table 1) (34,35).

x
x

x
DAAs and CKD
From the renal point of view, sofosbuvir’s
hepatic metabolite is renally eliminated by 80%;

x
x

x
however, renal elimination is not significant for all
other DAAs. All currently licensed DAA regimens

combinations with activity are marked x. Column 2 refers to FDA-recommended GFR limitations of use. FDA, Food and Drug Administration; C2, noncirrhotic; C1, cirrhotic.
can be used in patients with a GFR.30 ml/min per

x
x
1.73 m2. Two studies have examined the use of DAA
regimens in CKD patients with lower GFRs. In the

x
x
RUBY-1 Trial, 12 weeks of paritaprevir/ritonavir/
ombitasvir/dasabuvir combination therapy were used
in 20 noncirrhotic patients with drug-naı̈ve genotype

x
1 and GFR,30 ml/min per 1.73 m2 or on dialysis
(41). Genotype 1a patients also received IFN. The 12-
week post-treatment SVR was reported in 90% of

x
patients. In the C-SURFER Trial, 116 patients with
CKD stage 4/5 or ESRD on maintenance dialysis
were treated for 12 weeks with elbasvir/grazaprevir.
x
x
x

x
These patients had genotype 1 HCV, and most were
treatment naı̈ve and noncirrhotic. With this combination,
99% experienced an SVR at 12 weeks post-treatment
x
x
x

x
x
x
(42). DAAs were well tolerated in both trials, except for
ribavirin-treated patients, in whom anemia was common.
The HCV-Therapeutic Registry and Research Network
x
x

x
Table 1. Spectrum of activity of direct-acting antiviral therapy

Trial is an observational trial following patients with


renal disease on sofosbuvir-based regimens. Some reg-
imens included ribavirin, and 12.3% of regimens
1a, C2

included peg-IFN. SVR among the 73 patients with


x
x
x

x
x
x

GFR,45 ml/min per 1.73 m2 was 83%. Anemia,


worsening renal function, and serious side effects
were more common in those with reduced GFR. The
DAA Combination Recommendation

Reduce dose; ESRD


FDA Label GFR

efficacy of lower doses of sofosbuvir in patients with


Over 30 ml/min

Over 30 ml/min
Over 30 ml/min
Over 30 ml/min
Any; ESRD not

low GFR was assessed by Bhamidimarri et al. (43).


not studied

This group used 200 mg daily or 400 mg on alternate


studied

days with a standard dose of simeprevir in genotype 1


patients with GFRs,30 ml/min per 1.73 m2 or on
Any

dialysis (most had cirrhosis as well). SVRs were 75%–


ombitasvir/dasabuvir

Sofosbuvir/velpatasvir
Daclatasvir/sofosbuvir
Simeprevir/sofosbuvir

100%, with best responses in noncirrhotic genotype 1a


ombitasvir/ribavirin
Ledipasvir/sofosbuvir
Paritaprevir/ritonavir/

Paritaprevir/ritonavir/
Elbasvir/grazaprevir

patients who were not on dialysis and were treated


with sofosbuvir at 200 mg daily (43).

DAAs and HCV Treatment Post-Transplantation


HCV has been successfully treated with modern
DAA regimens in the postkidney transplant population.
276 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

This was first reported in 2016 in the United States and isoenzymes 3A4/5 or the drug transporter P-glycopro-
European cohorts. In the United States study, 20 tein. Therefore, no clinically significant interactions
patients, primarily of genotype 1 (88%; genotype 2 with common transplant immunosuppression drugs are
was also treated) and most without cirrhosis (n516 of expected. All DAAs in use today as single agents or in
20), received DAA drug therapy at a median of 888 days combination, except for sofosbuvir alone, have drug-
post-transplant. Maintenance immunosuppression was drug interactions that need to be monitored (Table 2).
calcineurin inhibitors, antimetabolite agents, and pred- In two previous trials of post-transplant HCV treat-
nisone. Serum creatinine levels at the initiation of ment, tacrolimus trough levels were decreased post-
treatment ranged from 0.86 to 2.18 mg/dl. All received treatment by approximately 45%, despite no dose
a sofosbuvir-based regime (primarily treatment with reductions. This decrease was postulated to result from
sofosbuvir/simeprevir or sofosbuvir/ledipasvir), and improved hepatic enzyme function post-HCV viral
three patients also received ribavirin. All patients cleared clearance.
the virus and had an SVR at 12 weeks post-treatment The optimal treatment regime in post-transplant
(44). The French group treated 25 patients for post- patients is dependent on GFR and genotype, with
transplant HCV infection. Genotypes 1a, 1b, 2, 3, and 4 close monitoring of immunosuppression drug levels
were all represented, but genotype 1b was most common required. The ideal regimen has an SVR of 100%, is
(n515). The median time from transplant to the start of not cleared by the kidneys, treats all genotypes, and is
DAA treatment was 146 months. All received sofosbuvir- easily available from a cost and regulation stand-
based regimens (simeprevir/sofosbuvir or ledipasvir/ point. No such combination currently exists. How-
sofosbuvir), with two patients receiving a ribavirin-based ever, treatment efficacy in the post-transplant patient
regimen and one receiving sofosbuvir/ribavirin/peg-IFN. with available drugs is excellent. Given the huge
All drugs except ribavirin were not dose reduced for disparity between the waitlist population and num-
GFR. All patients had GFRs.30 ml/min per 1.73 m2 at ber of patients transplanted each year, new ways to
the start of therapy. The SVR at 12 weeks post-treatment increase the donor pool are always welcome. Accord-
was 100% (45). In both studies, the average tacrolimus ing to the Centers for Disease Control and Prevention,
trough levels were lower post-therapy compared with the 3.4 million people in the United States are living with
levels at the start of and during therapy. In both studies, HCV infection (46). Using Organ Procurement Trans-
drugs were well tolerated, with no significant side effects plant Network donor registration data from 2012,
noted, and no rejection episodes occurred. 2.4% of deceased donors used were HCV serology
It is important to understand the pharmacokinet- positive (47). Approximately 50% of HCV-positive
ics of DAAs in the setting of transplantation so that kidneys are discarded, and about 500 of these are high-
appropriate monitoring and dose adjustments can be quality kidneys (48,49). Among nonextended criteria
made during and after therapy. Daclatasvir and donor organs, HCV infection is the most likely reason
sofosbuvir are not substrates of the cytochrome P450 for an organ to be discarded (24).

Table 2. Direct-acting antiviral therapy and immunosuppression drug interactions


Paritaprevira
/Ritonavir/
Immunosuppressive Sofosbuvir/ Ledipasvir/ Elbasvir/ Ombitasvir/
Agent Velpatasvir Sofosbuvir Grazoprevira Dasabuvir Simeprevira Sofosbuvir Daclatasvir
Tac A A C C C A A
Ciclo A A D C D A A
Myf A A A C A A A
Aza A A A A A A A
Everol C C C C C A C
Sirol A A C C C A A
According to www.hep-druginteractions.org (University of Liverpool). Tac, tacrolimus; A, no interactions; C, potential interactions; Ciclo, ciclosporin; D, do not administer; Myf,
myfortic/cellcept; Aza, azathioprine; Everol, everolimus; Sirol, sirolimus.
a
NS3/4A protease inhibitors (most Cyp 3A inhibition).
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 277

Can We and Should We Use HCV-Positive Donor been studied in the immediate post-transplant period.
Kidneys in the HCV-Negative Recipient? This may represent a significant limitation in certain
Recent studies have examined the feasibility of regions, and 50%–70% of donors in United States
transplanting HCV-positive kidneys into HCV-negative cohorts (the THINKER Trial, National Health and
donors followed by immediate treatment with DAA Nutrition Examination Survey, and Kaiser Permanente)
regimens. The Zepatier for Treatment of Hepatitis C– are genotype 1a or 1b (54). Many insurance compa-
Negative Patients Who Receive Kidney Transplants nies and some Medicaid programs will only approve
from Hepatitis C–Positive Donors (THINKER) Trial is the use of DAA therapy for patients with advanced
a single-center, open-label trial of the use of HCV- biopsy-proven liver cirrhosis (Metavir fibrosis stages
positive kidneys in HCV-negative recipients (50). Early 2–4). This is likely due to the current high cost of
results have been reported, and the trial is still ongoing. DAAs and the opinions expressed by some profes-
This trial determined the safety and efficacy of using sional societies. Namely, those with the greatest need
kidneys from HCV genotype 1 viremic donors fol- should be treated first (55,56). Treatment of HCV
lowed by elbasvir/grazoprevir treatment. Recipients post-transplantation will continue to be a challenge
with any liver disease, HIV, hepatitis B virus, or HCV until payers make DAAs available to all HCV-infected
infection were excluded. HCV viremia post-transplant patients.
and genotype were confirmed in study subjects post- In summary, DAA agents are highly effective, and
transplantation and before initiation of DAA therapy. their efficacy is established in patients with reduced
Initiation of therapy on day 3 post-transplant was GFR, with minimal side effects. DAAs can effectively
established because of the delay in confirming viremia treat HCV in the post-transplantation period, including
and HCV genotype. Elbasvir/grazoprevir is only licensed immediately post-transplant. Drug-drug interactions with
for use in genotype 1 and 4 patients. Therefore, the DAA are mild for most classes. With careful monitoring,
investigators limited the THINKER Trial to genotype 1 graft function remains stable, with rejection events un-
patients. Ten patients were studied, and all patients likely. The transplantation of HCV-positive donor kidneys
cleared HCV and had an SVR at 12 weeks post-treatment. into HCV-negative recipients has the potential to increase
Median 6-month creatinine levels and GFRs were 1.1 mg/ the donor pool, while reducing waiting list time and
dl and 62.8 ml/min per 1.73 m2, respectively. One patient dialysis-related morbidity.
had delayed graft function, two patients had transient
transaminitis, and one patient developed transient class I Disclosures
de novo donor-specific antibodies. None.
Before using HCV-positive kidneys in HCV-
negative recipients, informed consent is required. References
However, most recipients who are HCV negative have 1. WHO: Hepatitis C Fact Sheet, Geneva, Switzerland, World Health
Organization, 2017
little knowledge of their illness. One study determined 2. Fissell RB, Bragg-Gresham JL, Woods JD, Jadoul M, Gillespie B,
how willing an HCV-negative recipient would be to Hedderwick SA, Rayner HC, Greenwood RN, Akiba T, Young EW:
Patterns of hepatitis C prevalence and seroconversion in hemodialysis
accept an HCV-positive kidney. This study questioned units from three continents: The DOPPS. Kidney Int 65: 2335–2342,
185 patients and found that 29% of patients would 2004 PubMed
accept under any conditions and that over 50% would 3. Pereira BJ, Natov SN, Bouthot BA, Murthy BV, Ruthazer R,
Schmid CH, Levey AS; The New England Organ Bank Hepatitis
accept under certain conditions (51). Use of HCV- C Study Group: Effects of hepatitis C infection and renal trans-
positive kidneys in HCV-negative patients is also cost plantation on survival in end-stage renal disease. Kidney Int 53: 1374–
effective on the basis of the assumption that cure rates 1381, 1998 PubMed
4. Butt AA, Skanderson M, McGinnis KA, Ahuja T, Bryce CL, Barnato
remain .97% (52). The cost benefit of using HCV- AE, Chang CC: Impact of hepatitis C virus infection and other
positive kidneys is also suggested by a United King- comorbidities on survival in patients on dialysis. J Viral Hepat 14:
688–696, 2007 PubMed
dom study (53). The elbasvir/grazoprevir combination 5. Fabrizi F, Takkouche B, Lunghi G, Dixit V, Messa P, Martin P: The impact
was likely chosen in the THINKER Trial for its of hepatitis C virus infection on survival in dialysis patients: Meta-analysis
nonrenal elimination. Notably, this drug combination of observational studies. J Viral Hepat 14: 697–703, 2007 PubMed
6. Kalantar-Zadeh K, Daar ES, Eysselein VE, Miller LG: Hepatitis C
is not active against genotypes 2, 3, 5, and 6. Pange- infection in dialysis patients: A link to poor clinical outcome? Int Urol
notypic treatments that include sofosbuvir have not Nephrol 39: 247–259, 2007 PubMed
278 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

7. Pereira BJ, Levey AS: Hepatitis C virus infection in dialysis and renal Rizzetto M, Shouval D, Sola R, Terg RA, Yoshida EM, Adda N,
transplantation. Kidney Int 51: 981–999, 1997 PubMed Bengtsson L, Sankoh AJ, Kieffer TL, George S, Kauffman RS,
8. Domínguez-Gil B, Morales JM: Transplantation in the patient with Zeuzem S; ADVANCE Study Team: Telaprevir for previously un-
hepatitis C. Transpl Int 22: 1117–1131, 2009 PubMed treated chronic hepatitis C virus infection. N Engl J Med 364: 2405–
9. Mitwalli AH, Alam A, Al-Wakeel J, Al Suwaida K, Tarif N, Schaar TA, 2416, 2011 PubMed
Al Adbha B, Hammad D: Effect of chronic viral hepatitis on graft 26. Poordad F, McCone J Jr., Bacon BR, Bruno S, Manns MP, Sulkowski
survival in Saudi renal transplant patients. Nephron Clin Pract 102: MS, Jacobson IM, Reddy KR, Goodman ZD, Boparai N, DiNubile MJ,
c72–c80, 2006 PubMed Sniukiene V, Brass CA, Albrecht JK, Bronowicki JP; SPRINT-2
10. Roth D, Gaynor JJ, Reddy KR, Ciancio G, Sageshima J, Kupin W, Investigators: Boceprevir for untreated chronic HCV genotype 1
Guerra G, Chen L, Burke GW 3rd: Effect of kidney transplantation on infection. N Engl J Med 364: 1195–1206, 2011 PubMed
outcomes among patients with hepatitis C. J Am Soc Nephrol 22: 1152– 27. Curry MP, O’Leary JG, Bzowej N, Muir AJ, Korenblat KM, Fenkel
1160, 2011 PubMed JM, Reddy KR, Lawitz E, Flamm SL, Schiano T, Teperman L,
11. Morales JM, Campistol JM, Domínguez-Gil B, Andrés A, Esforzado N, Fontana R, Schiff E, Fried M, Doehle B, An D, McNally J, Osinusi A,
Oppenheimer F, Castellano G, Fuertes A, Bruguera M, Praga M: Long- Brainard DM, McHutchison JG, Brown RS Jr., Charlton M;
term experience with kidney transplantation from hepatitis C-positive ASTRAL-4 Investigators: Sofosbuvir and velpatasvir for HCV in
donors into hepatitis C-positive recipients. Am J Transplant 10: 2453– patients with decompensated cirrhosis. N Engl J Med 373: 2618–
2462, 2010 PubMed 2628, 2015 PubMed
12. Burra P, Buda A, Livi U, Rigotti P, Zanus G, Calabrese F, Caforio A, 28. Zeuzem S, Dusheiko GM, Salupere R, Mangia A, Flisiak R, Hyland
Menin C, Canova D, Farinati F, Luciana Aversa SM: Occurrence of RH, Illeperuma A, Svarovskaia E, Brainard DM, Symonds WT,
post-transplant lymphoproliferative disorders among over thousand Subramanian GM, McHutchison JG, Weiland O, Reesink HW, Ferenci
adult recipients: Any role for hepatitis C infection? Eur J Gastroenterol P, Hézode C, Esteban R; VALENCE Investigators: Sofosbuvir and
Hepatol 18: 1065–1070, 2006 PubMed ribavirin in HCV genotypes 2 and 3. N Engl J Med 370: 1993–2001,
13. Caillard S, Lelong C, Pessione F, Moulin B; French PTLD Working 2014 PubMed
Group: Post-transplant lymphoproliferative disorders occurring after 29. Sulkowski MS, Gardiner DF, Rodriguez-Torres M, Reddy KR, Hassa-
renal transplantation in adults: Report of 230 cases from the French nein T, Jacobson I, Lawitz E, Lok AS, Hinestrosa F, Thuluvath PJ,
Registry. Am J Transplant 6: 2735–2742, 2006 PubMed Schwartz H, Nelson DR, Everson GT, Eley T, Wind-Rotolo M, Huang
14. Caillard S, Agodoa LY, Bohen EM, Abbott KC: Myeloma, Hodgkin SP, Gao M, Hernandez D, McPhee F, Sherman D, Hindes R, Symonds
disease, and lymphoid leukemia after renal transplantation: Char- W, Pasquinelli C, Grasela DM; AI444040 Study Group: Daclatasvir
acteristics, risk factors and prognosis. Transplantation 81: 888–895, plus sofosbuvir for previously treated or untreated chronic HCV
2006 PubMed infection. N Engl J Med 370: 211–221, 2014 PubMed
15. Kondo Y, Shimosegawa T: Direct effects of hepatitis C virus on the 30. Kowdley KV, Gordon SC, Reddy KR, Rossaro L, Bernstein DE, Lawitz
lymphoid cells. World J Gastroenterol 19: 7889–7895, 2013 PubMed E, Shiffman ML, Schiff E, Ghalib R, Ryan M, Rustgi V, Chojkier M,
16. Takahashi K, Nishida N, Kawabata H, Haga H, Chiba T: Regression of Herring R, Di Bisceglie AM, Pockros PJ, Subramanian GM, An D,
Hodgkin lymphoma in response to antiviral therapy for hepatitis C virus Svarovskaia E, Hyland RH, Pang PS, Symonds WT, McHutchison JG,
infection. Intern Med 51: 2745–2747, 2012 PubMed Muir AJ, Pound D, Fried MW; ION-3 Investigators: Ledipasvir and
17. Cole EH, Johnston O, Rose CL, Gill JS: Impact of acute rejection and sofosbuvir for 8 or 12 weeks for chronic HCV without cirrhosis. N Engl
new-onset diabetes on long-term transplant graft and patient survival. J Med 370: 1879–1888, 2014 PubMed
Clin J Am Soc Nephrol 3: 814–821, 2008 PubMed 31. Afdhal N, Zeuzem S, Kwo P, Chojkier M, Gitlin N, Puoti M, Romero-
18. Corell A, Morales JM, Mandroño A, Muñoz MA, Andrés A, Fuertes A, Gomez M, Zarski JP, Agarwal K, Buggisch P, Foster GR, Bräu N, Buti
Arnaiz-Villena A: Immunosuppression induced by hepatitis C virus M, Jacobson IM, Subramanian GM, Ding X, Mo H, Yang JC, Pang PS,
infection reduces acute renal-transplant rejection. Lancet 346: 1497– Symonds WT, McHutchison JG, Muir AJ, Mangia A, Marcellin P;
1498, 1995 PubMed ION-1 Investigators: Ledipasvir and sofosbuvir for untreated HCV
19. Fabrizi F, Martin P, Dixit V, Bunnapradist S, Dulai G: Hepatitis C virus genotype 1 infection. N Engl J Med 370: 1889–1898, 2014 PubMed
antibody status and survival after renal transplantation: Meta-analysis of 32. Lawitz E, Mangia A, Wyles D, Rodriguez-Torres M, Hassanein T,
observational studies. Am J Transplant 5: 1452–1461, 2005 PubMed Gordon SC, Schultz M, Davis MN, Kayali Z, Reddy KR, Jacobson
20. Ingsathit A, Kamanamool N, Thakkinstian A, Sumethkul V: Survival IM, Kowdley KV, Nyberg L, Subramanian GM, Hyland RH, Arter-
advantage of kidney transplantation over dialysis in patients with burn S, Jiang D, McNally J, Brainard D, Symonds WT, McHutchison
hepatitis C: A systematic review and meta-analysis. Transplantation JG, Sheikh AM, Younossi Z, Gane EJ: Sofosbuvir for previously
95: 943–948, 2013 PubMed untreated chronic hepatitis C infection. N Engl J Med 368: 1878–1887,
21. Kamar N, Ribes D, Izopet J, Rostaing L: Treatment of hepatitis C virus 2013 PubMed
infection (HCV) after renal transplantation: Implications for HCV- 33. Jacobson IM, Gordon SC, Kowdley KV, Yoshida EM, Rodriguez-
positive dialysis patients awaiting a kidney transplant. Transplantation Torres M, Sulkowski MS, Shiffman ML, Lawitz E, Everson G, Bennett
82: 853–856, 2006 PubMed M, Schiff E, Al-Assi MT, Subramanian GM, An D, Lin M, McNally J,
22. Pereira BJ, Milford EL, Kirkman RL, Levey AS: Transmission of Brainard D, Symonds WT, McHutchison JG, Patel K, Feld J, Pianko S,
hepatitis C virus by organ transplantation. N Engl J Med 325: 454–460, Nelson DR; POSITRON Study; FUSION Study: Sofosbuvir for
1991 PubMed hepatitis C genotype 2 or 3 in patients without treatment options. N
23. Bucci JR, Matsumoto CS, Swanson SJ, Agodoa LY, Holtzmuller KC, Engl J Med 368: 1867–1877, 2013 PubMed
Peters TG, Abbott KC: Donor hepatitis C seropositivity: Clinical 34. Foster GR, Afdhal N, Roberts SK, Bräu N, Gane EJ, Pianko S, Lawitz
correlates and effect on early graft and patient survival in adult E, Thompson A, Shiffman ML, Cooper C, Towner WJ, Conway B,
cadaveric kidney transplantation. J Am Soc Nephrol 13: 2974–2982, Ruane P, Bourlière M, Asselah T, Berg T, Zeuzem S, Rosenberg W,
2002 PubMed Agarwal K, Stedman CA, Mo H, Dvory-Sobol H, Han L, Wang J,
24. Kucirka LM, Singer AL, Ros RL, Montgomery RA, Dagher NN, Segev McNally J, Osinusi A, Brainard DM, McHutchison JG, Mazzotta F,
DL: Underutilization of hepatitis C-positive kidneys for hepatitis C- Tran TT, Gordon SC, Patel K, Reau N, Mangia A, Sulkowski M;
positive recipients. Am J Transplant 10: 1238–1246, 2010 PubMed ASTRAL-2 Investigators; ASTRAL-3 Investigators: Sofosbuvir and
25. Jacobson IM, McHutchison JG, Dusheiko G, Di Bisceglie AM, Reddy velpatasvir for HCV genotype 2 and 3 infection. N Engl J Med 373:
KR, Bzowej NH, Marcellin P, Muir AJ, Ferenci P, Flisiak R, George J, 2608–2617, 2015 PubMed
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 279

35. Feld JJ, Jacobson IM, Hézode C, Asselah T, Ruane PJ, Gruener N, transplant recipients with direct-acting antiviral agents. Am J Transplant
Abergel A, Mangia A, Lai CL, Chan HL, Mazzotta F, Moreno C, 16: 1588–1595, 2016 PubMed
Yoshida E, Shafran SD, Towner WJ, Tran TT, McNally J, Osinusi A, 45. Kamar N, Marion O, Rostaing L, Cointault O, Ribes D, Lavayssière L,
Svarovskaia E, Zhu Y, Brainard DM, McHutchison JG, Agarwal K, Esposito L, Del Bello A, Métivier S, Barange K, Izopet J, Alric L:
Zeuzem S; ASTRAL-1 Investigators: Sofosbuvir and velpatasvir for Efficacy and safety of sofosbuvir-based antiviral therapy to treat
HCV genotype 1, 2, 4, 5, and 6 infection. N Engl J Med 373: 2599– hepatitis C virus infection after kidney transplantation. Am J Transplant
2607, 2015 PubMed 16: 1474–1479, 2016 PubMed
36. Pianko S, Flamm SL, Shiffman ML, Kumar S, Strasser SI, Dore GJ, 46. Edlin BR, Eckhardt BJ, Shu MA, Holmberg SD, Swan T: Toward
McNally J, Brainard DM, Han L, Doehle B, Mogalian E, McHutch- a more accurate estimate of the prevalence of hepatitis C in the United
ison JG, Rabinovitz M, Towner WJ, Gane EJ, Stedman CA, Reddy States. Hepatology 62: 1353–1363, 2015 PubMed
KR, Roberts SK: Sofosbuvir plus velpatasvir combination therapy 47. Hart A, Smith JM, Skeans MA, Gustafson SK, Stewart DE, Cherikh
for treatment-experienced patients with genotype 1 or 3 hepatitis C WS, Wainright JL, Kucheryavaya A, Woodbury M, Snyder JJ, Kasiske
virus infection: A randomized trial. Ann Intern Med 163: 809–817, BL, Israni AK: OPTN/SRTR 2015 Annual Data Report: Kidney. Am J
2015 PubMed Transplant 17[Suppl 1]: 21–116, 2017 PubMed
37. Zeuzem S, Ghalib R, Reddy KR, Pockros PJ, Ben Ari Z, Zhao Y, Brown 48. Reese PP, Abt PL, Blumberg EA, Goldberg DS: Transplanting hepatitis
DD, Wan S, DiNubile MJ, Nguyen BY, Robertson MN, Wahl J, Barr E, C-positive kidneys. N Engl J Med 373: 303–305, 2015 PubMed
Butterton JR: Grazoprevir-elbasvir combination therapy for treatment- 49. Goldberg DS, Blumberg E, McCauley M, Abt P, Levine M: improving
naive cirrhotic and noncirrhotic patients with chronic hepatitis C virus organ utilization to help overcome the tragedies of the opioid epidemic.
genotype 1, 4, or 6 infection: A randomized trial. Ann Intern Med 163: Am J Transplant 16: 2836–2841, 2016 PubMed
1–13, 2015 PubMed 50. Goldberg DS, Abt PL, Blumberg EA, Van Deerlin VM, Levine M,
38. Sulkowski M, Hezode C, Gerstoft J, Vierling JM, Mallolas J, Pol S, Reddy KR, Bloom RD, Nazarian SM, Sawinski D, Porrett P, Naji A,
Kugelmas M, Murillo A, Weis N, Nahass R, Shibolet O, Serfaty L, Hasz R, Suplee L, Trofe-Clark J, Sicilia A, McCauley M, Farooqi M,
Bourliere M, DeJesus E, Zuckerman E, Dutko F, Shaughnessy M, Gentile C, Smith J, Reese PP: Trial of transplantation of HCV-
Hwang P, Howe AY, Wahl J, Robertson M, Barr E, Haber B: Efficacy infected kidneys into uninfected recipients. N Engl J Med 376: 2394–
and safety of 8 weeks versus 12 weeks of treatment with grazoprevir 2395, 2017 PubMed
(MK-5172) and elbasvir (MK-8742) with or without ribavirin in 51. Reese PP, Goldberg D, Mussell A, McCauley M, Sawinski D, Molina
patients with hepatitis C virus genotype 1 mono-infection and HIV/ N, Tomlin R, Doshi S, Abt P, Blumberg E, Thiessen C, Kulkarni S,
hepatitis C virus co-infection (C-WORTHY): A randomised, open-label Esnaola G: Willingness of end-stage renal disease patients (ESRD)
phase 2 trial. Lancet 385: 1087–1097, 2015 PubMed without Hepatitis C (HCV) to accept a HCV1 kidney [Abstract]. Am J
39. Sperl J, Horvath G, Halota W, Ruiz-Tapiador JA, Streinu-Cercel A, Transplant 17[Suppl 3], 2017. Available at: http://atcmeetingabstracts.
Jancoriene L, Werling K, Kileng H, Koklu S, Gerstoft J, Urbanek P, com/abstract/willingness-of-end-stage-renal-disease-patients-esrd-without-
Flisiak R, Leiva R, Kazenaite E, Prinzing R, Patel S, Qiu J, Asante-Appiah hepatitis-c-hcv-to-accept-a-hcv-kidney/. Accessed June 6, 2017.
E, Wahl J, Nguyen BY, Barr E, Platt HL: Efficacy and safety of elbasvir/ 52. Kiberd B, Doucette K, Tennankore K: Use of hepatitis c infected organs
grazoprevir and sofosbuvir/pegylated interferon/ribavirin: A phase III for kidney transplantation: A cost-effective analysis [Abstract]. Am J
randomized controlled trial. J Hepatol 65: 1112–1119, 2016 PubMed Transplant 17[Suppl 3], 2017. Available at: http://atcmeetingabstracts.
40. Liang TJ: Current progress in development of hepatitis C virus vaccines. com/abstract/use-of-hepatitis-c-infected-organs-for-kidney-transplantation-
Nat Med 19: 869–878, 2013 PubMed a-cost-effective-analysis/ Accessed June 6, 2017.
41. Pockros PJ, Reddy KR, Mantry PS, Cohen E, Bennett M, Sulkowski 53. Trotter P, Robb M, Ushiro-Lumb I, Powell J, Watson C, Bradley J,
MS, Bernstein DE, Cohen DE, Shulman NS, Wang D, Khatri A, Neuberger J: Transplantation of organs from donors with hepatitis C:
Abunimeh M, Podsadecki T, Lawitz E: Efficacy of direct-acting The potential to substantially increase transplant activity. [Abstract]. Am
antiviral combination for patients with hepatitis c virus genotype 1 J Transplant 17[Suppl 3], 2017. Available at: http://atcmeetingabstracts.
infection and severe renal impairment or end-stage renal disease. com/abstract/transplantation-of-organs-from-donors-with-hepatitis-c-the-
Gastroenterology 150: 1590–1598, 2016 PubMed potential-to-substantially-increase-transplant-activity/ Accessed June 6, 2017.
42. Roth D, Nelson DR, Bruchfeld A, Liapakis A, Silva M, Monsour H Jr., 54. Goldberg D, Van V, Farooqi M, Sicilia A, Hasz R, Bloom R, Blumberg
Martin P, Pol S, Londoño MC, Hassanein T, Zamor PJ, Zuckerman E, E, Gentile C, Abt P, Reese P: Hepatitis C genotypes among deceased
Wan S, Jackson B, Nguyen BY, Robertson M, Barr E, Wahl J, Greaves W: organ donors in the United States [Abstract]. Am J Transplant 17[Suppl 3],
Grazoprevir plus elbasvir in treatment-naive and treatment-experienced 2017. Available at: http://atcmeetingabstracts.com/abstract/hepatitis-c-
patients with hepatitis C virus genotype 1 infection and stage 4-5 chronic genotypes-among-deceased-organ-donors-in-the-united-states/ Accessed
kidney disease (the C-SURFER study): A combination phase 3 study. Lancet June 6, 2017.
386: 1537–1545, 2015 PubMed 55. Reau N, Fried MW, Nelson DR, Brown RS Jr., Everson GT, Gordon
43. Bhamidimarri KR, Czul F, Peyton A, Levy C, Hernandez M, Jeffers L, SC, Jacobson IM, Lim JK, Pockros PJ, Reddy KR, Sherman KE: HCV
Roth D, Schiff E, O’Brien C, Martin P: Safety, efficacy and tolera- Council–critical appraisal of data: Recommendations for clinical
bility of half-dose sofosbuvir plus simeprevir in treatment of Hepatitis practice in a rapidly evolving therapeutic landscape. Liver Int 36:
C in patients with end stage renal disease. J Hepatol 63: 763–765, 488–502, 2016 PubMed
2015 PubMed 56. Ooka K, Connolly JJ, Lim JK: Medicaid reimbursement for oral direct
44. Sawinski D, Kaur N, Ajeti A, Trofe-Clark J, Lim M, Bleicher M, Goral antiviral agents for the treatment of chronic hepatitis C. Am J Gastro-
S, Forde KA, Bloom RD: Successful treatment of hepatitis C in renal enterol 112: 828–832, 2017 PubMed
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

Syllabus
NephSAP, Volume 16, Number 4, November 2017—Transplantation

John P. Vella, MD, FASN, FAST* and Alexander C. Wiseman, MD, FASTy
*Division of Nephrology and Transplantation, Maine Medical Center and Tufts University School
of Medicine, Portland, Maine; and yDivision of Renal Diseases and Hypertension, Transplant
Center, University of Colorado Denver, Aurora, Colorado
(1) Implementation of the new Kidney Allocation
Learning Objectives System on December 4, 2014 led to an increase
1. Describe current transplant outcomes, includ- in deceased donor kidney transplants among
ing delayed graft function, rejection, graft, and black candidates and those with calculated panel
patient survival, and living donor outcomes reactive antibodies 98%–100% but a decrease
2. Discuss early outcomes associated with the among candidates ages 65 years old or older.
2014 changes to the Kidney Allocation System (2) There are positive trends in graft and patient
3. Examine the spectrum of complications that survival for deceased and living donor kidney
occur after transplantation, including infec- transplants, but the challenge of limited kidney
tion, malignancy, and cardiovascular disease, supply persists.
and accompanying individual diagnostic and (3) The total number of patients on the waiting list
therapeutic strategies decreased for the first time in a decade due to
4. Review the status of immunosuppression, in- fewer waiting list additions, an increased number
compatible transplantation, and efforts to achieve of removals due to medical deterioration, and an
immunologic tolerance increase in the total number of transplants.
(4) Waiting list deaths remain stable, presumably
from the removal of inactive candidates too sick
for transplantation.
This syllabus reviews the major clinical studies
published between January of 2015 and January of Smith et al. (2) reported that the new allocation
2017 in pertinent journals with the highest impact policy is associated with increased quality-associated
factors. The authors have focused on randomized trials, life-years at lower cost compared with the old policy,
case series, and registry data whenever possible. In the thereby yielding net benefits to the US Medicare system
interest of brevity, case reports, basic science studies, totaling $271 million United States dollars in the first
preclinical studies, and data published only in abstract year and $55 million United States dollars yearly after.
form have been, in general, excluded.
Patient and Graft Survival
The total number of kidney transplant recipients
Access to Transplantation and Outcomes
alive with a functioning graft exceeded 200,000 within
The Scientific Registry of Transplant Recipients the United States as of June of 2015, more than doubling
(SRTR) evaluates the clinical and scientific status of since 2000 (1). Although allograft function and survival
organ transplantation in the United States in close col- remained better for living than for deceased donor
laboration with the Organ Procurement and Trans- transplants, long-term outcomes continued to improve
plantation Network, which is administered by the for both. All-cause and death-censored graft failure at
United Network for Organ Sharing under contract with 1, 3, 5, and 10 years continued to decline for living and
the Health Resources and Services Administration. deceased donor transplants.
Each year, the SRTR reports on the state of trans- For deceased donor transplants, 10-year graft failure
plantation in the United States (1). Important findings for transplants in 2005 was 52.8%, which decreased from
include the following. 59.2% in 1995 (Figures 1 and 2). Similarly, 10-year graft
280
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 281

Figure 1. Adult deceased donor transplant recipient death- Figure 3. Adult live donor transplant recipient death with
censored graft failure rates are improving over time. Death- function rates are stable or marginally improving over
censored graft failure among adult deceased donor kidney time. Graft failure among adult living donor kidney
transplant recipients. Estimates are unadjusted and computed transplant recipients. Estimates are unadjusted and com-
using Kaplan–Meier competing risk methods. Recipients are puted using Kaplan–Meier competing risk methods. Re-
followed to the earliest of kidney graft failure; kidney retrans- cipients are followed to the earliest of kidney graft failure;
plant; return to dialysis; death; or 6 months or 1, 3, 5, or 10 kidney retransplant; return to dialysis; death; or 6 months
years post-transplant. Death-censored graft failure is defined as or 1, 3, 5, or 10 years post-transplant. All-cause graft
a return to dialysis, reported graft failure, or kidney retransplant. failure is defined as any of the prior outcomes before 6
OPTN, Organ Procurement and Transplantation Network. months or 1, 3, 5, or 10 years. OPTN, Organ Procurement
Reprinted with permission from Hart A, Smith JM, Skeans and Transplantation Network. Reprinted with permission
MA, Gustafson SK, Stewart DE, Cherikh WS, Wainright JL, from Hart A, Smith JM, Skeans MA, Gustafson SK, Stewart
Kucheryavaya A, Woodbury M, Snyder JJ, Kasiske BL, Israni DE, Cherikh WS, Wainright JL, Kucheryavaya A, Woodbury
AK: OPTN/SRTR 2015 Annual Data Report: Kidney. Am J M, Snyder JJ, Kasiske BL, Israni AK: OPTN/SRTR 2015
Transplant 17[Suppl 1]: 21–116, 2017. Annual Data Report: Kidney. Am J Transplant 17[Suppl 1]:
21–116, 2017.
failure for living donor transplants in 2005 was 37.3%,
down from 44.8% in 1995 (Figures 3 and 4). Death with donor graft survival was lowest for patients with diabetes
a functioning graft remained essentially constant for liv- or hypertension as a cause of kidney failure: 70.4% and
ing and deceased donor transplants. Five-year deceased 71.8%, respectively. As previously reported, the overall
outcomes of donation after brain and cardiac death trans-
plants continue to be identical, which is reassuring (Figure
5). Such observations should not lead to complacency.
Lim et al. (3), studying donation after cardiac death
outcomes, examined the association between delayed
graft function (DGF) and graft and patient outcomes. A
greater proportion of recipients with DGF had experi-
enced overall graft loss and death-censored graft loss at
3 years compared with those without DGF (14% versus
4%; P50.04 and 11% versus 0%, P,0.01, respec-
Figure 2. Adult deceased donor transplant recipient death with
function rates are stable over time. Death with function among tively). Compared with recipients without DGF, the
adult deceased donor kidney transplant recipients. Estimates are adjusted hazard ratio (HR) for overall graft loss at 3
unadjusted and computed using Kaplan–Meier competing risk years for recipients with DGF was 4.31 (95% confi-
methods. Recipients are followed to the earliest of kidney graft dence interval [95% CI], 1.13 to 16.44).
failure; kidney retransplant; return to dialysis; death; or 6 The kidney donor profile index (KDPI) is a numeric
months or 1, 3, 5, or 10 years post-transplant. Death with score that has been developed to assess deceased donor
function is defined as death without prior graft failure, return to kidney quality, and it is an integral part of the Kidney
dialysis, or retransplant. OPTN, Organ Procurement and Trans-
Allocation System. Factors include donor age, height,
plantation Network. Reprinted with permission from Hart A,
Smith JM, Skeans MA, Gustafson SK, Stewart DE, Cherikh weight, creatinine, diabetes, hypertension, cause of death,
WS, Wainright JL, Kucheryavaya A, Woodbury M, Snyder JJ, hepatitis C status, and donation after cardiac death, and they
Kasiske BL, Israni AK: OPTN/SRTR 2015 Annual Data are incorporated into a calculator (https://optn.transplant.
Report: Kidney. Am J Transplant 17[Suppl 1]: 21–116, 2017. hrsa.gov/resources/allocation-calculators/kdpi-calculator/).
282 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

Ethnicity
Taber et al. (4) examined the outcomes of black
kidney transplant recipients, a group known to experi-
ence higher rates of graft loss for the last 40 years. They
studied 3306 non-Hispanic whites (67%) compared
with 1612 non-Hispanic blacks (33%) with 6 years of
follow-up. In the unadjusted analysis, black recipients
were significantly more likely to have overall graft loss
(HR, 1.19; 95% CI, 1.07 to 1.33) and death-censored
Figure 4. Adult live donor transplant recipient death censored graft
failure rates are improving improving over time. Death-censored graft loss (HR, 1.67; 95% CI, 1.45 to 1.92) but had
graft failure among adult living donor kidney transplant recipients. lower mortality (HR, 0.83; 95% CI, 0.72 to 0.96). In
Estimates are unadjusted and computed using Kaplan–Meier fully adjusted models, only death-censored graft loss
competing risk methods. Recipients are followed to the earliest of remained significant (HR, 1.38; 95% CI, 1.12 to 1.71).
kidney graft failure; kidney retransplant; return to dialysis; death; or The authors concluded that non-Hispanic black kidney
6 months or 1, 3, 5, or 10 years post-transplant. Death-censored graft transplant recipients experience a substantial disparity
failure is defined as a return to dialysis, reported graft failure, or
kidney retransplant. OPTN, Organ Procurement and Transplantation
in graft loss but not mortality. Recognition that sus-
Network. Reprinted with permission from Hart A, Smith JM, ceptibility to kidney disease has a genetic component
Skeans MA, Gustafson SK, Stewart DE, Cherikh WS, Wainright has been evaluated by candidate gene studies identify-
JL, Kucheryavaya A, Woodbury M, Snyder JJ, Kasiske BL, Israni ing alterations in a variety of genes as reviewed in
AK: OPTN/SRTR 2015 Annual Data Report: Kidney. Am J a prior Nephrology Self-Assessment Program on trans-
Transplant 17[Suppl 1]: 21–116, 2017. plantation (5). Chief among them are two risk variants
Lower scores are better. Not surprisingly, there continues to of the apolipoprotein L1 gene (APOL1) in association
be a marked graft survival outcome discrepancy between with nondiabetic nephropathy in blacks. Freedman
the highest KDPI kidneys (.85%) compared with lower et al. (6) evaluated the impact of APOL1 positivity in
KDPI kidneys (,85%) (Figure 5). black kidney donors with data from 478 newly ana-
lyzed deceased donor recipients reported to the SRTR.
Shorter renal allograft survival in recipients of APOL1
kidneys was confirmed (HR, 2.00; 95% confidence
interval [95% CI], 1.39 to 3.02, P50.03). Combined
analysis of 1153 deceased donor kidney transplants
from black donors revealed that donor APOL1 high-
risk genotype (HR, 2.05), older donor age (HR, 1.18;
95% CI, 1.00 to 1.40), and younger recipient age (HR,
0.70) significantly impacted allograft survival adversely.
Although prolonged allograft survival was seen in many
Figure 5. Graft survival among adult deceased donor kidney recipients of APOL1 kidneys, follow-up serum creati-
transplant recipients associates with graft quality as measured by nine concentrations were higher than those in recipients
the Kidney Donor Profile Index (KDPI) (lower is better). Those
who receive kidneys with a KDPI .85% have the worse long
without APOL1 renal risk variant kidneys. A competing
term graft survival rates. Graft survival among adult deceased risk analysis revealed that APOL1 impacted renal
donor kidney transplant recipients in 2010 by KDPI. Graft allograft survival but not recipient survival.
survival was estimated using unadjusted Kaplan–Meier meth- Arce et al. (7) reported that risks for graft failure
ods. The reference population for the KDRI to KDPI conversion and mortality (HR, 0.69; 95% CI, 0.65 to 0.73) and all-
is all deceased donor kidneys recovered for transplant in the cause graft failure (HR, 0.79; 95% CI, 0.75 to 0.83) are
United States in 2015. OPTN, Organ Procurement and Trans- lower in Hispanics compared with non-Hispanic trans-
plantation Network. Reprinted with permission from Hart A,
Smith JM, Skeans MA, Gustafson SK, Stewart DE, Cherikh
plant recipients. Furthermore, the association between
WS, Wainright JL, Kucheryavaya A, Woodbury M, Snyder JJ, Hispanic ethnicity and graft failure excluding death was
Kasiske BL, Israni AK: OPTN/SRTR 2015 Annual Data Report: significantly modified by age. Compared with non-Hispanic
Kidney. Am J Transplant 17[Suppl 1]: 21–116, 2017. whites, graft failure, excluding death with a functioning
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 283

graft, did not differ in Hispanics ages 18–39 years old (HR, conducted a retrospective study using health care data-
0.96; 95% CI, 0.89 to 1.05) or 40–59 years old (HR, 1.08; bases in Ontario, Canada to determine whether the
95% CI, 1.00 to1.16) but was 13% lower in those ages $60 incidence of cardiovascular events after kidney trans-
years old (HR, 0.87; 95% CI, 0.78 to 0.98). plantation has changed over time. Recipients (n54954)
Adler et al. (8) analyzed socioeconomic status were older and had more baseline comorbidity in recent
(SES) and racial/ethnic gradients between donors and years. A total of 445 recipients (9.0%) died or experi-
recipients. In a retrospective cohort study, traditional enced a major cardiovascular event within 3 years of
demographic and socioeconomic factors as well as an transplantation. There was no significant change in the
SES index were compared in 56,697 deceased kidney incidence of the composite outcome or death-censored
donor and recipient pairs transplanted between 2007 and cardiovascular events over time (P50.41 and P50.92,
2012. Kidneys were more likely to be transplanted in respectively). After adjusting for age, sex, and comor-
recipients of the same racial/ethnic group as the donor bidities, the risk of death or major cardiovascular event
(P,0.001). Kidneys tended to go to recipients of lower steadily declined across the years of transplant (2006–
SES index (50.5% of the time; P,0.001), a relationship 2009 adjusted HR, 0.70; P,0.01; referent, 1994–1997).
that remained after adjusting for other available markers When recipients were matched for age, sex, and date of
of donor organ quality and SES (P,0.001). cohort entry to members of the general population and to
the CKD population, the risk was lowest in the general
Physical Frailty population and highest in the CKD population.
Frailty is associated with inferior survival and Shin et al. (11) analyzed 2684 primary kidney
increased resource requirements among kidney transplant transplant recipients who had a functioning graft at 6
candidates. Lynch et al. (9) examined US Renal Data months after transplantation to assess the association of
System data from 51,111 adult ESRD patients waitlisted renin-angiotensin system (RAS) blockade with patient
for transplant from 2000 to 2011. Frequently admitted and graft survival. RAS blockade was associated with
patients had higher subsequent resource requirements, an adjusted HR of 0.63 (95% CI, 0.53 to 0.75) for total
increased waitlist mortality, and decreased likelihood of graft loss, an adjusted HR of 0.69 (95% CI, 0.55 to
transplant (death after listing: 1–7 days: HR, 1.24; 95% 0.86) for death, and an adjusted HR of 0.62 (95% CI,
CI, 1.20 to 1.28; 8–14 days: HR, 1.49; 95% CI, 1.42 to 0.49 to 0.78) for death-censored graft failure. The
1.56; $15 days: HR, 2.07, 95% CI; 1.99 to 2.15 versus associations of RAS blockade with a lower risk of total
0 days). Graft and recipient survival was inferior, with graft loss and mortality were stronger with more severe
higher admissions, although survival benefit was pre- proteinuria. The RAS blockade was associated with
served. A model, including waitlist admissions alone, a twofold higher risk of hyperkalemia. By comparison,
performed better in predicting postlisting mortality than Dad et al. (12) analyzed the Folic Acid for Vascular
estimated post-transplant survival. Similarly, McAdams- Outcome Reduction in Transplantation Trial database
DeMarco et al. (10) studied the association between and reported that aspirin use did not associate with
prospectively assessed physical frailty and postkidney reduced risk for incident cardiovascular disease, all-
transplantation mortality and found that frailty was cause mortality, or kidney failure in stable kidney
independently associated with a 2.17-fold higher risk transplant recipients without a history of cardiovas-
of death (95% CI, 1.01 to 4.65; P50.05). cular disease. Pihlstrøm et al. (13) investigated the
association between baseline parathyroid hormone
(PTH) levels and major cardiovascular events, renal
graft loss, and all-cause mortality. Significant asso-
Physical frailty is associated with decreased
ciations between PTH and all three outcomes were
likelihood of transplantation and inferior sur-
found in univariate analyses. When adjusting for
vival after kidney transplantation.
a range of plausible confounders, including measures
of renal function and serum mineral levels, PTH
remained significantly associated with all-cause mor-
Cardiovascular Outcomes tality (4% increased risk per 10 U; P50.004) and
Cardiovascular death remains the leading cause of graft loss (6% increased risk per 10 U; P,0.001) but
mortality in kidney transplant recipients. Lim et al. (3) not major cardiovascular events.
284 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

Obesity renal transplant recipients: The Wisconsin Allograft Recipient Database


(WisARD). Transplantation 100: 1541–1549, 2016 PubMed
The impact of pretransplant body mass index 12. Dad T, Tighiouart H, Joseph A, Bostom A, Carpenter M, Hunsicker L,
(BMI) on long-term allograft outcomes after kidney Kusek JW, Pfeffer M, Levey AS, Weiner DE: Aspirin use and incident
transplantation remains controversial. Naik et al. cardiovascular disease, kidney failure, and death in stable kidney
transplant recipients: A post hoc analysis of the Folic Acid for Vascular
(14) studied 108,654 transplant recipients and de- Outcome Reduction in Transplantation (FAVORIT) Trial. Am J Kidney
termined that increasing BMI level was associated Dis 68: 277–286, 2016 PubMed
with increased risk of long-term allograft failure. In 13. Pihlstrøm H, Dahle DO, Mjøen G, Pilz S, März W, Abedini S, Holme I,
Fellström B, Jardine AG, Holdaas H: Increased risk of all-cause
an adjusted model with BMI of 18.5 to ,25 kg/m2 as mortality and renal graft loss in stable renal transplant recipients with
referent, the subhazard ratios (SHRs) for BMI were hyperparathyroidism. Transplantation 99: 351–359, 2015 PubMed
,18.5 kg/m2: SHR, 0.96; P50.41; 25 to ,30 kg/m2: 14. Naik AS, Sakhuja A, Cibrik DM, Ojo AO, Samaniego-Picota MD,
Lentine KL: The impact of obesity on allograft failure after kidney
SHR, 1.05; P50.01; 30 to ,35 kg/m2: SHR, 1.15; transplantation: A competing risks analysis. Transplantation 100:
P#0.001; 35 to ,40 kg/m2: SHR, 1.21; P,0.001; and 1963–1969, 2016 PubMed
.40 kg/m2: SHR, 1.13; P50.002. Mechanistically, it 15. Wu D, Dawson NA, Levings MK: Obesity-associated adipose tis-
sue inflammation and transplantation. Am J Transplant 16: 743–750,
has been suggested that obesity is often associated with 2016 PubMed
the development of an adiposity-induced inflammatory
state that induces metabolic dysfunction that increases
Deceased Donor Kidney Allocation
the risk for developing type 2 diabetes (15).
The Kidney Allocation System (KAS), a major
References change to deceased donor kidney allocation implemented
1. Hart A, Smith JM, Skeans MA, Gustafson SK, Stewart DE, Cherikh
WS, Wainright JL, Kucheryavaya A, Woodbury M, Snyder JJ, Kasiske by the United Network for Organ Sharing (UNOS) in
BL, Israni AK: OPTN/SRTR 2015 Annual Data Report: Kidney. Am J December of 2014, incorporated specific goals, which
Transplant 17[Suppl 1]: 21–116, 2017 PubMed
2. Smith JM, Schnitzler MA, Gustafson SK, Salkowski NJ, Snyder JJ,
included directing the highest quality organs to
Kasiske BL, Israni AK: Cost implications of new national allocation younger, healthier recipients; increasing access to
policy for deceased donor kidneys in the United States. Transplantation deceased donor kidney transplantation for highly
100: 879–885, 2016 PubMed
3. Lim WH, McDonald SP, Russ GR, Chapman JR, Ma MK, Pleass H,
sensitized patients, racial/ethnic minorities, and those
Jaques B, Wong G: Association between delayed graft function and with an extended duration on dialysis while mitigat-
graft loss in donation after cardiac death kidney transplants - a paired ing regional variability.
kidney registry analysis [published online ahead of print July 29, 2016].
TransplantationPubMed To assess the impact of this major change in
4. Taber DJ, Gebregziabher M, Payne EH, Srinivas T, Baliga PK, Egede allocation policy, Stewart et al.(1) compared Organ
LE: Overall graft loss versus death-censored graft loss: Unmasking the Procurement and Transplantation Network data 1 year
magnitude of racial disparities in outcomes among US kidney transplant
recipients. Transplantation 101: 402–410, 2017 PubMed before and 1 year after implementation of the KAS.
5. Vella JP, Wiseman A: Nephrol Self Assess Program 14: 375–439, 2015 Transplants in which the donor and recipient ages differed
6. Freedman BI, Pastan SO, Israni AK, Schladt D, Julian BA, Gautreaux MD, by .30 years declined by 23%. Initial sharp increases in
Hauptfeld V, Bray RA, Gebel HM, Kirk AD, Gaston RS, Rogers J, Farney
AC, Orlando G, Stratta RJ, Mohan S, Ma L, Langefeld CD, Bowden DW, transplants were observed for calculated panel reactive
Hicks PJ, Palmer ND, Palanisamy A, Reeves-Daniel AM, Brown WM, Divers antibody 99%–100% recipients and recipients with at
J: APOL1 genotype and kidney transplantation outcomes from deceased
least 10 years on dialysis, with a subsequent tapering of
African American donors. Transplantation 100: 194–202, 2016 PubMed
7. Arce CM, Lenihan CR, Montez-Rath ME, Winkelmayer WC: Com- transplants to these groups, suggesting a bolus effect. It
parison of longer-term outcomes after kidney transplantation between was noted that kidneys were more often being shipped
Hispanic and non-Hispanic whites in the United States. Am J Transplant
15: 499–507, 2015 PubMed
over long distances, leading to longer cold ischemic
8. Adler JT, Hyder JA, Elias N, Nguyen LL, Markmann JF, Delmonico times and higher delayed graft function (DGF) rates.
FL, Yeh H: Socioeconomic status and ethnicity of deceased donor Importantly, 6-month graft survival rates had not changed
kidney recipients compared to their donors. Am J Transplant 15: 1061–
1067, 2015 PubMed
significantly. Building on the prior observations, Massie
9. Lynch RJ, Zhang R, Patzer RE, Larsen CP, Adams AB: First-year et al.(2) compared kidney distribution, transplant rates for
waitlist hospitalization and subsequent waitlist and transplant outcome. waitlist registrants, and recipients for the 2 years pre-
Am J Transplant 17: 1031–1041, 2017 PubMed
10. McAdams-DeMarco MA, Law A, King E, Orandi B, Salter M, Gupta N,
and post-KAS. Some of their key findings included an
Chow E, Alachkar N, Desai N, Varadhan R, Walston J, Segev DL: increase in regional imports from 8.8% pre-KAS to
Frailty and mortality in kidney transplant recipients. Am J Transplant 12.5% post-KAS and an increase in national imports
15: 149–154, 2015 PubMed
11. Shin JI, Palta M, Djamali A, Kaufman DB, Astor BC: The association from 12.7% pre-KAS to 19.1% post-KAS (P,0.001).
between renin-angiotensin system blockade and long-term outcomes in The proportion of recipients .30 years older than their
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 285

donor decreased from 19.4% to 15.0% (P,0.001). The entire population (adjusted odds ratio [aOR], 0.97; 95% CI,
proportion of recipients with calculated panel reac- 1.04 to 1.10; P=0.30) or in any KDPI stratum. However,
tive antibody 100% increased from 1.0% to 10.3% among kidneys in which ECD and KDPI indicators were
(P,0.001). This latter observation is truly remarkable, discordant, high-risk standard criteria donor (SCD) kidneys
because such patients previously had low transplant (with KDPI.0.85) were at increased risk of discard in the
rates with markedly prolonged waiting times. The KDPI era (aOR, 1.07; 95% CI, 1.42 to 1.89; P=0.02).
overall deceased donor transplant rate did not change. However, recipients of these kidneys were at much lower
However, rates increased for blacks (incidence ratio risk of death (adjusted risk ratio, 0.56; 95% CI, 0.77 to 0.94
[IR], 1.19; 95% confidence interval [95% CI], 1.13 to at 2 years post-transplant) compared with those remaining
1.25), Hispanics (IR, 1.13; 95% CI, 1.05 to 1.20), and on dialysis and waiting for low-KDPI kidneys. The authors
candidates ages 18–40 years old (IR, 1.47; 95% CI, suggested that there might be an unexpected harmful
1.38 to 1.57) but declined for candidates ages .50 labeling effect of reporting a high KDPI for SCD kidneys
years old (IR, 0.93; 95% CI, 0.87 to 0.98), those 51–60 without the expected advantage of providing a more gran-
years old (IR, 0.90; 95% CI, 0.85 to 0.96), and those ular risk index. Such concerns warrant ongoing vigilance.
ages .70 years old. Of some concern, the DGF rate Stewart et al.(4) also evaluated trends in donor
increased from 24.8% pre-KAS to 29.9% post-KAS characteristics with specific reference to discard kidneys.
(P,0.001). In summary, KAS has been associated They reported that most of the kidney discard rate rise
with improved access to transplantation for minorities, was explained by the broadening donor pool. However,
younger candidates, and highly sensitized patients but the presence of an unexplained residual increase sug-
declines for older candidates as projected by preimple- gests that behavioral factors, such as programmatic risk
mentation modeling. DGF increased substantially. This aversion or allocation inefficiencies, may also play a role.
is of concern due to the association of DGF with in- Snyder et al. (5) examined whether accepting high-risk
ferior long-term outcomes (discussed below). kidneys (KDPI$0.85) is associated with transplant pro-
grams receiving inferior evaluations. Despite a clear
relationship between KDPI, graft failure, and mortality,
Changes in United States national kidney there was no relationship between a program’s use of
allocation implemented in December of 2014 high-KDPI kidneys and subsequently receiving a poor
have improved transplantation access for performance evaluation after risk adjustment. Excluding
minorities, younger candidates, and those high-KDPI donor transplants did not alter the programs
who are highly sensitized. This is at the ex- identified as underperforming, because in every case,
pense of reduced access for older patients underperforming programs also had worse than ex-
and higher delayed graft function rates. pected outcomes among lower-risk donor transplants.
The authors concluded that there is no evidence that
programs that accept higher-KDPI kidneys are at greater
Discarded Kidneys risk for poor performance evaluations, and risk aversion
One of the central features of the new system was may limit access to transplant for candidates while
the development of the kidney donor profile index providing no measurable benefit to program evaluations.
(KDPI), a numeric score on the basis of ten clinical criteria Therefore, do we have predictors of kidney discard?
that describe organ quality implemented pre-KAS in 2012 Marrero et al.(6) performed an analysis of the Organ
(lower score is better; calculator is available online:https:// Procurement and Transplantation Network database
optn.transplant.hrsa.gov/resources/allocation-calculators/ from 2000 to 2012 of all solid organ donors. Perhaps
kdpi-calculator/). It is known that allograft prognosis is as expected, predictors of increased discard included
inferior when the KDPI exceeds 85%, which includes age older than 50 years old, performance of a kidney
a group of donor kidneys that approximates the previous biopsy, cytomegalovirus seropositive status, donation
designation of expanded criteria donors (ECDs). Bae et al. after cardiac death, hepatitis B and C seropositive status,
(3) studied the discard rates of the kidneys recovered for cigarette use, diabetes, hypertension, terminal creatinine
transplantation from the ECD era with those of the KDPI .1.5 mg/dl, and blood type AB.
era. There was no significant change in discard rate from Does timing of kidney procurement impact the
the ECD era (18.1%) to the KDPI era (18.3%) among the discard rate? Indeed, the answer is yes given a recent
286 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

report by Mohan et al.(7) that studied the impact of the group (28%) and 112 recipients of kidneys from donors
day of the week of the procurement and subsequent in the normothermia group (39%; odds ratio, 0.62; 95%
utilization or discard of deceased donor kidneys in the CI, 0.43 to 0.92; P=0.02). The authors concluded that
United States. Compared with weekday kidneys, or- mild hypothermia compared with normothermia in
gans procured on weekends were significantly more organ donors after declaration of death per neurologic
likely to be discarded than transplanted (OR, 1.16; 95% criteria significantly reduced the rate of DGF among
CI, 1.13 to 1.19), even after adjusting for organ quality recipients. It is now up to organ procurement organi-
(aOR, 1.13; 95% CI, 1.10 to 1.17). Weekend discards zations to either repeat the study or implement the
were of a significantly higher quality than weekday protocols for deceased donors.
discards, although the effect was small (KDPI: 76.5%
versus 77.3%). Considerable geographic variation was Use of Hepatitis C Virus–Seropositive Kidneys
also noted in the proportion of transplants that occurred The impact of transplanting hepatitis C virus
over the weekend (Figure 6). (HCV)–seropositive deceased donor kidneys was stud-
ied by Scalea et al.(9), who analyzed 1679 adult
Deceased Donor Hypothermia allografts between 2000 and 2012. Of these, 195 hep-
DGF, which is reported in up to 50% of kidney atitis C virus–seropositive recipients (R+) received
transplant recipients, is associated with increased costs kidneys from hepatitis C virus–seropositive donors
and inferior prognosis (see below). Targeting mild (D+) in contrast to 1418 HCV-negative recipients who
hypothermia in organ donors before organ recovery received grafts from hepatitis C virus–negative donors
may impact DGF risk. Niemann et al.(8) enrolled (D–) and 66 R+ patients who received D– kidneys.
organ donors (after declaration of death by neurologic Death-censored graft survival in the D+/R+ population
criteria) from two large donation service areas and was better than graft survival for D–/R+ patients,
randomly assigned them to one of two targeted tem- despite D+/R+ patients having higher rates of hyper-
perature ranges: 34C to 35C (hypothermia) or 36.5C tension and black ethnicity. Waitlist times for patients
to 37.5C (normothermia). The study was terminated accepting HCV-seropositive grafts were 318 days (for
early on the recommendation of an independent data D+/R+ patients) versus 613 days (D–/HCV-negative
and safety monitoring board after the interim analysis recipients) or 570 days (D–/R+). The authors reason-
showed efficacy of hypothermia. DGF developed in 79 ably concluded that HCV D+/R+ patients spent less
recipients of kidneys from donors in the hypothermia time on the transplant waitlist, which contributed to
improved death-censored graft survival compared with
D–/R+ patients.

Prior Living Donors in Need of Transplantation


The incidence of ESRD in prior living kidney
donors (PLDs) has been estimated at 30 in 10,000,
higher than four in 10,000 in the general population
(reviewed in prior issues of the Nephrology Self-
Assessment Program on transplantation). Wainright
et al.(10) investigated the early effects of the KAS on
the access of PLDs to deceased donor kidney trans-
plants. Using the Organ Procurement and Transplanta-
Figure 6. There is marked geographic variation in the tion Network data, this group compared prevalent and
proportion of deceased donor transplants that are performed incident cohorts of PLDs in the 1-year periods before
over the weekend. Geographic variation in the proportion of and after KAS implementation. Importantly, transplant
deceased donor kidney transplants performed over the weekend
rates were not statistically different before or after KAS
by state, 2000–2013. Reprinted with permission from Mohan S,
Foley K, Chiles MC, Dube GK, Patzer RE, Pastan SO, Crew implementation for either prevalent (2.37 versus 2.29;
RJ, Cohen DJ, Ratner LE: The weekend effect alters the relative risk [RR], 0.96; 95% CI, 0.62 to 1.49) or
procurement and discard rates of deceased donor kidneys in the incident (4.76 versus 4.36; RR, 0.92; 95% CI, 0.53 to
United States. Kidney Int90: 157–163, 2016. 1.60) candidates. Median waiting time to deceased
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 287

donor kidney transplant for prevalent PLDs in the post- (10) Deceased donors whose medical/behavioral
KAS cohort was 102.6 days compared with 82.3 days history cannot be determined.
in the pre-KAS cohort (P=0.98). The median KDPI for (11) Deceased donors whose blood specimen has
PLD recipients was 31% with KAS versus 23% before been hemodiluted.
KAS (P=0.02). Despite a sharp decrease in the waiting
time for highly prioritized candidates with calculated The UNOS regulations specify that such donors
panel reactive antibodies of 98%–100% (from .7000 be screened for hepatitis B virus, HCV, and HIV by
to 1164 days), PLDs still had much shorter waiting both serology and nucleic acid testing (NAT). These
times (mean of 102.6 days). kidneys can only be offered as Public Health Service
Increased Risk Donor if hepatitis B virus, HCV, and
HIV testing is negative. It should be understood that
The incidence of ESRD in prior living donors there is still risk of a false negative result should such
is 30 in 10,000, which is higher than that in a donor remain within the window period after infection
nondonors (four in 10,000). This risk is but before viral nucleic acid has sufficiently amplified to
mitigated to an extent in that prior donors yield a positive result. Such kidneys are only offered to
have ready access to transplantation. recipients who have been informed of the risks and
consented to receive such a kidney. Kidneys that test
positive for HCV or hepatitis B virus can be offered to
recipients who are HCV seropositive or hepatitis B
Public Health Service Increased Risk Donors
virus positive, respectively. Kidneys from HIV-positive
The US Public Health Service redefined donors
donors can now be transplanted into HIV-positive recip-
previously classified by the Centers for Disease
ients under experimental protocols since the HIV Organ
Control as being at high risk for transmission of
Policy Equity Act was enacted in 2013.
hepatitis B, HCV, and HIV. These categories include
Importantly, the number of such deceased donors
the following.
has increased markedly over the last decade driven by
(1) People who have had sex with a person known or the epidemic of opiate addiction within the United States.
suspected to have HIV, hepatitis B, or HCV Kucirka et al.(11) quantified the impact of the new
infection within the previous year. guidelines and found that 19.5% of donors were labeled
(2) Men who have had sex with men within the increased risk donors (IRD) compared with 10.4%,
previous year. 12.2%, and 12.3% in the 3 most recent years under the
(3) Women who have had sex with a man with a old guidelines (incidence rate ratio, 1.45; P,0.001).
history of having sex with men within the pre- Increases were consistent across organ procurement
vious year. organizations: 44 of 59 had an increase in the percent-
(4) People who have had sex for money or drugs age of donors labeled as being at increased risk, and 14
within the previous year. organ procurement organizations labeled 25% of their
(5) People who have had sex with a person who had donors as being at increased risk under the new guidelines
sex in exchange for money or drugs within the (versus five organ procurement organizations under the
previous year. old guidelines). Blacks were 52% more likely to be labeled
(6) People using nonmedical injection of drugs IRD under the new guidelines (RR, 1.52; P=0.01).
within the previous year. The risk of transmission of disease using the prior
(7) People who have sex with a person who used Centers for Disease Control and Prevention guidelines
nonmedical injection of drugs within the previous was previously reviewed by Kucirka et al.(12). They
year. found that increased risk organs are discarded at higher
(8) People who have been in lockup, jail, prison, or rates than SCD kidneys. Their estimated risk of un-
juvenile correctional facility for .72 consecutive detected window period HIV infection varies by IRD
hours. behavior category (range, 0.035–4.9 per 10,000 donors
(9) Persons who have been newly diagnosed with when NAT is used), and HCV risk is higher (range,
or treated for syphilis, gonorrhea, Chlamydia, or 0.027–32.4 per 10,000). Recipients who receive such
genital ulcers in the preceding 12 months. kidneys are tested at time 0 and again at defined time
288 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

points until the end of the first post-transplant year by access to deceased donor kidney transplants: An early look.Am J
Transplant 17:1103–1111,2017 PubMed
NAT. Information about recipients who test positive 11. Kucirka LM, Bowring MG, Massie AB, Luo X, Nicholas LH, Segev
needs to be shared with the UNOS Disease Transmission DL:Landscape of deceased donors labeled increased risk for disease
Advisory Committee. Updated information on actual transmission under new guidelines.Am J Transplant 15:3215–3223,
2015 PubMed
disease transmission on the basis of the new criteria is 12. Kucirka LM, Singer AL, Segev DL:High infectious risk donors: what
eagerly awaited. are the risks and when are they too high?Curr Opin Organ Transplant
16:256–261,2011 PubMed

Deceased donor kidneys procured from


individuals who meet the US Public Health Living Kidney Donation
Service definitions of increased risk now
make up approximately 20% of the entire In 2015, 5626 living donor transplants were per-
donor pool. These donors are tested for viral formed, an increase from 5539 in 2014 but a decrease
infections by both serology and NAT. Organs from the peak of 6647 in 2004 (Figure 7) (1). This de-
from these donors can be offered for trans- cline is due to a decrease in related kidney donations,
plantation when the testing is negative. because unrelated donations have been stable. Although
Point estimates of risk for transmission of the number of paired donations has increased, this is not
HIV, HCV, and hepatitis B virus are all <1%. enough to offset the overall decline in related donations.
These kidneys tend to be from young donors Women made up 63.5% of living donors, an increase
with low KDPIs and may be associated with from 59.2% in 2005. The proportion of black donors has
reduced waiting time for transplantation. continued to decline over 10 years from 13.4% to 9.6%.
Laparoscopic nephrectomy made up .97% of procedures
in 2015. Readmissions after living donation, although
References infrequent, are not rare, with rates of 2.5% at 6 weeks,
1. Stewart DE, Kucheryavaya AY, Klassen DK, Turgeon NA, Formica RN,
Aeder MI:Changes in deceased donor kidney transplantation one year 3.9% at 6 months, and 5.2% at 1 year. Overall, compli-
after KAS implementation.Am J Transplant 16:1834–1847,2016 PubMed cations postnephrectomy occured in 5.3% of living donors
2. Massie AB, Luo X, Lonze BE, Desai NM, Bingaman AW, Cooper M,
Segev DL:Early changes in kidney distribution under the new allocation
at 6 weeks, 7.2% of living donors at 6 months, and 8.8%
system.J Am Soc Nephrol 27:2495–2501,2016 PubMed of living donors at 12 months. The distribution of body
3. Bae S, Massie AB, Luo X, Anjum S, Desai NM, Segev DL:Changes in mass index (BMI) remained relatively stable over 10
discard rate after the introduction of the kidney donor profile index
(KDPI).Am J Transplant 16:2202–2207,2016 PubMed years, with a slight decrease in the proportion of donors
4. Stewart DE, Garcia VC, Rosendale JD, Klassen DK, Carrico BJ: with BMI.35 kg/m2 (from 4.6% to 2.6%) and an
Diagnosing the decades-long rise in the deceased donor kidney discard increase in the proportion with BMI530–35 kg/m2 (from
rate in the United States.Transplantation 101:575–587,2017 PubMed
5. Snyder JJ, Salkowski N, Wey A, Israni AK, Schold JD, Segev DL, 16.1% to 19.5%). From 2011 to 2015, 17 deaths within 1
Kasiske BL:Effects of high-risk kidneys on Scientific Registry of Trans- year of donation were reported to the Organ Procurement
plant Recipients Program Quality Reports.Am J Transplant 16:2646– and Transplantation Network (OPTN). The most common
2653,2016 PubMed
6. Marrero WJ, Naik AS, Friedewald JJ, Xu Y, Hutton DW, Lavieri MS, causes of death were medical (including donation related)
Parikh ND:Predictors of deceased donor kidney discard in the United in seven and accident/homicide in five living donors.
States.Transplantation 101:1690–1697,2017 PubMed
7. Mohan S, Foley K, Chiles MC, Dube GK, Patzer RE, Pastan SO, Crew
RJ, Cohen DJ, Ratner LE:The weekend effect alters the procurement
and discard rates of deceased donor kidneys in the United States.Kidney
Int 90:157–163,2016 PubMed
The number of living donors has declined
8. Niemann CU, Feiner J, Swain S, Bunting S, Friedman M, Crutchfield in the United States, despite increasing kid-
M, Broglio K, Hirose R, Roberts JP, Malinoski D:Therapeutic hypo- ney paired donation (KPD).
thermia in deceased organ donors and kidney-graft function.N Engl J
Med 373:405–414,2015 PubMed
9. Scalea JR, Redfield RR, Arpali E, Leverson GE, Bennett RJ, Anderson
ME, Kaufman DB, Fernandez LA, D’Alessandro AM, Foley DP,
Mezrich JD:Does DCD donor time-to-death affect recipient outcomes? Kidney Paired Donation
Implications of time-to-death at a high-volume center in the United Approximately 30% of living donors are incom-
States.Am J Transplant 17:191–200,2017 PubMed
10. Wainright JL, Kucheryavaya AY, Klassen DK, Stewart DE:The impact patible with their intended recipients due to either pre-
of the new kidney allocation system on prior living kidney donors’ formed ABO or donor-specific antibodies. KPD is the
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 289

facilitated 47 transplants, and thus far, eight of their


paired recipients have received a kidney within a mean
of 178 (range, 10–562) days. Fumo et al. (3) studied
potential swaps that failed and found three primary
reasons that could be prevented by changes in protocol
or software: positive laboratory crossmatch (28%),
transplant center declined donor (17%), and pair trans-
planted outside the APD (14%). Making such changes
allowed their group to improve the success rate and the
number of transplants performed annually.

Living Donor Kidney Function Assessment


Per UNOS regulations, all living kidney donor
candidates undergo evaluation of GFR. Guidelines
recommend measured GFR (mGFR) via an endoge-
nous filtration marker or creatinine clearance rather
Figure 7. The relationship between living donors and their than eGFR by any methodology. However, measure-
intended recipients has evolved over time with an overall
ment methods are challenging, time consuming, and
decline in those that are biologically related. Kidney paired
donation is a growing trend however still makes up a small costly, and they can yield variable results. Huang et al.
proportion of live donor transplants. Kidney transplants from (4) investigated whether GFR estimated from serum
living donors by donor relation. Numbers of living donor creatinine with or without sequential cystatin C was
donations and characteristics recorded on the OPTN living sufficiently accurate to identify donor candidates with
donor registration form. Reprinted with permission from Hart A, high probability that mGFR is above or below thresh-
Smith JM, Skeans MA, Gustafson SK, Stewart DE, Cherikh olds for clinical decision making. They combined the
WS, Wainright JL, Boyle G, Snyder JJ, Kasiske BL, Israni AK: pretest probability for mGFR thresholds ,60, ,70,
Kidney. Am J Transplant 16[Suppl 2]: 11–46, 2016.
$80, and $90 ml/min per 1.73 m2 on the basis of
most exciting novel approach gaining traction that demographic characteristics from the National Health
addresses incompatibility, while avoiding the financial and Nutrition Examination Survey with test perfor-
costs, adverse events, and inferior outcomes associated mance of eGFR (Chronic Kidney Disease Epidemiol-
with desensitization. Numerous programs have been ogy Collaboration) to compute post-test probabilities.
created to promote KPD, including those managed by the Using data from the Scientific Registry of Transplant
United Network for Organ Sharing (UNOS), the National Recipients, 53% of recent living donors had predonation
Kidney Registry, the Alliance for Paired Donation, and GFR estimated from serum creatinine levels sufficiently
other regional or single-center initiatives. At core, these high to ensure $95% probability that predonation mGFR
programs promote the search for compatible living was $90 ml/min per 1.73 m2, suggesting that mGFR may
donors when candidate recipients have living donors who not be necessary in a large proportion of donor candi-
are incompatible. The concept of an open chain, often dates. They developed a web-based application (http://
triggered by an altruistic donor, frequently garners interest ckdepi.org/equations/donor-candidate-gfr-calculator/) to
by the popular press. The practice of shipping live donor compute the probability on the basis of eGFR that mGFR
kidneys is now firmly established, eliminating some of for a donor candidate is above or below a range of
the logistic barriers previously encountered. thresholds useful in living donor evaluation and selection.
Flechner et al. (2) described an interesting twist Although intriguing, such methodology does not cur-
on the general concept of KPD in the description of rently meet UNOS regulatory requirements.
their Advanced Donation Program (ADP) that permits
a donor who desires to donate by a specific date but Medically Complex Living Donors
does not yet have a matched donor to effectively “pay it Living donation has been deemed the chief
forward.” After obtaining informed consent from donor weapon in the fight to redistribute the gap between
and paired recipient, ten KPD chains were constructed supply and demand for kidneys given the continued
using an ADP donor. These ten ADP donors have expansion of the national waiting list. Ahmadi et al. (5)
290 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

reviewed current guidelines pertaining to the selection donors who did not develop ESRD. This association
of medically complex donors with regard to older age, was similar among related, spousal, and unrelated non-
obesity, hypertension, vascular anomalies/multiplicity, spousal donors.
nulliparous women, and minors as donors. Their summary
conclusions included the following statements: live kid- Surgery Complications
ney donation in older donors (up to 70 years of age) seems Lentine et al. (8) studied perioperative complications
to be safe when outcome is compared with that of younger associated with living kidney donation. Nephrectomies
donors; obese donors have comparable outcomes to lean were predominantly laparoscopic (94%). Overall, 16.8%
donors, at least in short- and midterm follow-up; given the of donors experienced a perioperative complication,
paucity of data pertaining to hypertensive donor safety, most commonly gastrointestinal (4.4%), bleeding (3.0%),
caution is advised; vascular multiplicity poses no direct respiratory (2.5%), surgical/anesthesia-related injuries
danger to the donor; women of childbearing age can be (2.4%), and other complications (6.6%). Major Clavien
safely included as donors; and minors should only be Classification of Surgical Complications grade IV
considered as kidney donors if no other options exist. (life threatening, organ dysfunction) or grade V (death)
To facilitate decision making when a transplant affected 2.5% of donors. After adjustment for demo-
candidate has multiple living donors, Massie et al. (6) graphic, clinical (including comorbidities), procedure,
developed a living kidney donor profile index (LKDPI) and center factors, blacks had increased risk of any
using the same scale as the deceased donor kidney complication (adjusted odds ratio [aOR], 1.26; P50.001)
donor profile index. Multivariate analysis indicated that as well as Clavien grade II (requiring pharmacologic
donor age over 50 years old (hazard ratio [HR] per 10 treatment or transfusion) or higher (aOR, 1.39; P,0.001),
years, 1.15; 95% confidence interval [95% CI], 1.24 to grade III (requiring anesthesia/surgery) or higher (aOR,
1.33), elevated BMI (HR per 10 U, 1.01; 95% CI, 1.09 1.56, P,0.001), and grade IV or higher (aOR, 1.56;
to 1.16), black race (HR, 1.15; 95% CI, 1.25 to 1.37), P50.004) events. Other significant correlates of Clavien
and cigarette use (HR, 1.09; 95% CI, 1.16 to 1.23) grade IV or higher events included obesity (aOR, 1.55;
as well as ABO incompatibility (HR, 1.03; 95% CI, 1.27 P,0.001), predonation hematologic (such as thrombo-
to 1.58), HLA B mismatches (HR, 1.03; 95% CI, 1.08 to cytopenia; aOR, 2.78; P,0.001) and psychiatric
1.14), and DR mismatches (HR, 1.04; 95% CI, 1.09 to (aOR, 1.45; P50.04) conditions, and robotic nephrectomy
1.15) were associated with greater risk of graft loss after (aOR, 2.07; P50.002), whereas an annual center volume
living donor transplantation (all P,0.05). The median .50 transplants (aOR, 0.55; P,0.001) was associated
LKDPI score was 13 (1–27); 24.2% of donors had with lower risk. The same group found that predonation
LKDPI,0 (less risk than any deceased donor kidney), narcotic analgesia use was independently associated with
and 4.4% of donors had LKDPI.50 (more risk than the readmission after donor nephrectomy (9).
median deceased donor kidney). An online calculator is
available http://www.transplantmodels.com/lkdpi/.
Muzaale et al. (7) questioned whether it was possible
that living kidney donors had undetected, subclinical, Postnephrectomy surgical complications
preexisting kidney disease that subsequently contrib- occur in approximately 17% of donors and
uted to risk of ESRD. To test this theory, the authors are more common in blacks and those
compared a cohort of 257 recipients whose donors with obesity or hematologic or psychiatric
subsequently developed ESRD with a matched cohort conditions. The risk of more serious compli-
whose donors remained ESRD free. The median cations (Clavien grade IV or higher) is 2.5%.
follow-up time from transplantation was 12.5 years
(interquartile range, 7.4–17.9; maximum 520 years).
Recipients of allografts from donors who developed Pregnancy Risk
ESRD had increased death-censored graft loss (74% Garg et al. (10) recently reported that gestational
versus 56% at 20 years; adjusted HR, 1.7; 95% CI, 1.5 hypertension or preeclampsia was more common
to 2.0; P,0.001) and mortality (61% versus 46% at 20 among living kidney donors than among nondonors
years; adjusted HR, 1.5; 95% CI, 1.2 to 1.8; P,0.001) (occurring in 15 of 131 pregnancies [11%] versus 38 of
compared with matched recipients of allografts from 788 pregnancies [5%]). The odds ratio (OR) for donors
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 291

was 2.4 (95% CI, 1.2 to 5.0; P50.01). Each component person with health characteristics that were like those of
of the primary outcome was also more common among age-matched kidney donors, the 15-year projections of
donors (OR, 2.5 for gestational hypertension and OR, the risk of ESRD in the absence of donation varied ac-
2.4 for preeclampsia). On a positive note, there were no cording to race and sex; the risk was 0.24% among black
significant differences between donors and nondonors men, 0.15% among black women, 0.06% among white
with respect to rates of preterm birth (8% and 7%, men, and 0.04% among white women. Risk projections
respectively) or low birth weight (6% and 4%, re- were higher in the presence of lower eGFR, higher
spectively). There were no reports of maternal death, albuminuria, hypertension, current or former smoking,
stillbirth, or neonatal death among donors. Most women diabetes, and obesity. The risk of ESRD was highest
had uncomplicated pregnancies after donation. among persons in the youngest age group, particularly
among young blacks. The 15-year observed risks after
donation among kidney donors in the United States were
Gestational hypertension and preeclamp- 3.5–5.3 times as high as the projected risks in the
sia are more common in living donors than absence of donation. Looking at this question another
nondonors. way, Ross (13) reported that .325 living kidney donors
have developed ESRD and have been listed on the
OPTN/UNOS deceased donor kidney waitlist. Unfortu-
nately, because of lack of standardized data collection,
Kidney Failure Risk the denominator is unknown.
There have been few prospective, controlled studies Anjum et al. (14) studied the causes of ESRD
of kidney donors. Kasiske et al. (11) posited that among PLDs. Overall, 125,427 donors were observed
understanding the pathophysiologic effects of kidney for a median of 11.0 years. The cumulative incidence of
donation may be important for judging donor safety ESRD increased from ten events per 10,000 at 10 years
and improving our understanding of the consequences postdonation to 85 events per 10,000 at 25 years
of reduced kidney function in CKD. In their pro- postdonation (late versus early ESRD adjusted for age,
spective, controlled, observational cohort study, this race, and sex: incidence rate ratio, 1.3; 95% CI, 1.7 to
group reported that GFR measured by plasma iohexol 2.3). Early postdonation ESRD was predominantly
clearance declined 0.3667.55 ml/min per year in 194 reported as GN; however, late postdonation ESRD was
controls but increased 1.4765.02 ml/min per year in more frequently reported as diabetes and hypertension.
198 donors (P,0.01) between 6 and 36 months. BP
was not different between donors and controls at any
visit. Urinary protein-to-creatinine and albumin-to- Living kidney donors have a small but
creatinine ratios were not increased in donors com- significantly increased risk of ESRD com-
pared with controls. From 6 to 36 months postdonation, pared with nondonors. This is driven early
serum parathyroid hormone, uric acid, homocysteine, postdonation by GN and late postdonation
and potassium levels were higher, whereas hemoglobin by diabetes and hypertension.
levels were lower in donors compared with controls.
The authors concluded that kidney donors manifest several
of the findings of mild CKD. However, at 36 months after
donation, kidney function continues to improve in donors Prior Living Donors and Kidney Transplantation
(presumably because of hyperfiltration), whereas controls The impact of the new Kidney Allocation System
have expected age-related declines in function. (KAS) on the access of PLDs to deceased donor
Previous reports on the risk of ESRD in prior living kidney transplants was studied by Wainright et al. (15).
kidney donors (PLDs) have been variously criticized due Using data from the OPTN, the numbers of PLDs
to small sample size, inappropriate comparators, short registered for transplantation before and after KAS
follow-up time, and incomplete data acquisition. Grams were examined (pre-KAS group: December 4, 2013 to
et al. (12) studied data from almost 5 million nondonors December 3, 2014 [n550, newly listed PLDs]; post-
studied for up to 16 years compared with 52,998 living KAS group: December 4, 2014 to December 3, 2015
kidney donors in the United States. For a 40-year-old [n539]). Transplant rates were similar before and after
292 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

KAS implementation for both prevalent (2.37 versus multiple additional analyses. From a cohort of 1333
2.29; relative risk, 0.96; 95% CI, 0.62 to 1.49) and living donors with 5332 matched controls, Lin et al. (19)
incident (4.76 versus 4.36; relative risk, 0.92; 95% CI, reported that the overall incidence of peptic ulcer disease
0.53 to 1.60) candidates. Median waiting time to was 1.74-fold higher in the living donor cohort than in
deceased donor kidney transplantation for prevalent the nonliving donor cohort (2.14 versus 1.48 per 1000
PLDs in the post-KAS cohort was 102.6 days com- person-years). This finding persisted after adjustment for
pared with 82.3 days in the pre-KAS cohort (P50.98). age, sex, monthly income, and comorbidities.
The median kidney donor profile index for PLD
recipients was 31% with KAS versus 23% before References
KAS (P50.02). Despite a sharp decrease in the waiting 1. Hart A, Smith JM, Skeans MA, Gustafson SK, Stewart DE, Cherikh
WS, Wainright JL, Boyle G, Snyder JJ, Kasiske BL, Israni AK: Kidney.
times for highly prioritized candidates with calculated Am J Transplant 16[Suppl 2]: 11–46, 2016 PubMed
panel reactive antibodies of 98%–100% (from .7000 2. Flechner SM, Leeser D, Pelletier R, Morgievich M, Miller K,
to 1164 days), PLDs still had much shorter waiting Thompson L, McGuire S, Sinacore J, Hil G: The incorporation of an
advanced donation program into kidney paired exchange: Initial
times (102.6 days). In conclusion, the new system experience of the national kidney registry. Am J Transplant 15:
continues to provide PLDs with quick access to high- 2712–2717, 2015 PubMed
3. Fumo DE, Kapoor V, Reece LJ, Stepkowski SM, Kopke JE, Rees SE,
quality kidneys for transplantation.
Smith C, Roth AE, Leichtman AB, Rees MA: Historical matching
strategies in kidney paired donation: The 7-year evolution of a web-based
APOL1 virtual matching system. Am J Transplant 15: 2646–2654, 2015 PubMed
4. Huang N, Foster MC, Lentine KL, Garg AX, Poggio ED, Kasiske BL,
Individuals of black ancestry who express two Inker LA, Levey AS: Estimated GFR for living kidney donor evalua-
variant copies of the gene encoding apo1 (APOL1), an tion. Am J Transplant 16: 171–180, 2016 PubMed
HDL that binds to apoA-1, make up 13% of the black 5. Ahmadi AR, Lafranca JA, Claessens LA, Imamdi RM, IJzermans JN,
Betjes MG, Dor FJ: Shifting paradigms in eligibility criteria for live kidney
population and are at increased risk of ESRD (16). donation: a systematic review. Kidney Int 87: 31–45, 2015 PubMed
Limited studies suggest that the survival of allografts 6. Massie AB, Leanza J, Fahmy LM, Chow EK, Desai NM, Luo X, King
from donors expressing two APOL1 risk alleles is EA, Bowring MG, Segev DL: A risk index for living donor kidney
transplantation. Am J Transplant 16: 2077–2084, 2016 PubMed
inferior to that of allografts with zero or one risk allele. 7. Muzaale AD, Massie AB, Anjum S, Liao C, Garg AX, Lentine KL,
In living kidney donation, two case reports describe Segev DL: Recipient outcomes following transplantation of allografts
donors expressing two APOL1 risk alleles who de- from live kidney donors who subsequently developed end-stage renal
disease. Am J Transplant 16: 3532–3539, 2016 PubMed
veloped ESRD (17). Given the potential impact of 8. Lentine KL, Lam NN, Axelrod D, Schnitzler MA, Garg AX, Xiao H,
APOL1 variants on the utility and safety of kidney Dzebisashvili N, Schold JD, Brennan DC, Randall H, King EA, Segev
transplantation and living kidney donation, the Amer- DL: Perioperative complications after living kidney donation: A
national study. Am J Transplant 16: 1848–1857, 2016 PubMed
ican Society of Transplantation convened a meeting 9. Lentine KL, Lam NN, Schnitzler MA, Hess GP, Kasiske BL, Xiao H,
with the goals of summarizing the current state of Axelrod D, Garg AX, Schold JD, Randall H, Dzebisashvili N, Brennan
knowledge with respect to transplantation and APOL1, DC, Segev DL: Predonation prescription opioid use: A novel risk factor
for readmission after living kidney donation. Am J Transplant 17: 744–
identifying knowledge gaps and studies to address 753, 2017 PubMed
these gaps, and considering approaches to integrating 10. Garg AX, Nevis IF, McArthur E, Sontrop JM, Koval JJ, Lam NN,
Hildebrand AM, Reese PP, Storsley L, Gill JS, Segev DL, Habbous S,
APOL1 into clinical practice (16). Bugeja A, Knoll GA, Dipchand C, Monroy-Cuadros M, Lentine KL;
DONOR Network: Gestational hypertension and preeclampsia in living
Other Complications kidney donors. N Engl J Med 372: 124–133, 2015 PubMed
Lam et al. (18) reported an increased risk of gout 11. Kasiske BL, Anderson-Haag T, Israni AK, Kalil RS, Kimmel PL, Kraus
ES, Kumar R, Posselt AA, Pesavento TE, Rabb H, Steffes MW, Snyder
in donors compared with nondonors. They studied JJ, Weir MR: A prospective controlled study of living kidney donors:
1988 donors and 19,880 matched nondonors who were Three-year follow-up. Am J Kidney Dis 66: 114–124, 2015 PubMed
12. Grams ME, Sang Y, Levey AS, Matsushita K, Ballew S, Chang AR,
followed for a median of 8.4 years. Donors compared Chow EK, Kasiske BL, Kovesdy CP, Nadkarni GN, Shalev V, Segev
with nondonors were more likely to be given a diagnosis DL, Coresh J, Lentine KL, Garg AX; Chronic Kidney Disease
of gout (3.4% versus 2.0%; 3.5 versus 2.1 events per Prognosis Consortium: Kidney-failure risk projection for the living
kidney-donor candidate. N Engl J Med 374: 411–421, 2016 PubMed
1000 person-years, respectively; HR, 1.6; 95% CI, 1.2 to 13. Ross LF: Living kidney donors and ESRD. Am J Kidney Dis 66: 23–27,
2.1; P,0.001). Similarly, donors compared with non- 2015 PubMed
donors were more likely to receive a prescription for 14. Anjum S, Muzaale AD, Massie AB, Bae S, Luo X, Grams ME, Lentine
KL, Garg AX, Segev DL: Patterns of end-stage renal disease caused by
allopurinol or colchicine (3.8% versus 1.3%; OR, 3.2; diabetes, hypertension, and glomerulonephritis in live kidney donors.
95% CI, 1.5 to 6.7; P50.002). Results were consistent in Am J Transplant 16: 3540–3547, 2016 PubMed
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 293

15. Wainright JL, Kucheryavaya AY, Klassen DK, Stewart DE: The impact preprocurement hypotension persists (odds ratio [OR],
of the new kidney allocation system on prior living kidney donors’
access to deceased donor kidney transplants: An early look. Am J
1.42; 95% confidence interval [95% CI], 1.06 to 1.90).
Transplant 17: 1103–1111, 2017 PubMed Donor hypoxia during the first 10 minutes after ex-
16. Newell KA, Formica RN, Gill JS, Schold JD, Allan JS, Covington SH, tubation was also associated with graft failure (hazard
Wiseman AC, Chandraker A: Integrating APOL1 gene variants into
renal transplantation: Considerations arising from the American Society ratio [HR], 1.30; 95% CI, 1.03 to 1.64), with 5-year
of Transplantation Expert Conference. Am J Transplant 17: 901–911, graft survival of 70.0% (95% CI, 64.5% to 74.8%) for
2017 PubMed donors above the median versus 61.4% (95% CI,
17. Zwang NA, Shetty A, Sustento-Reodica N, Gordon EJ, Leventhal J,
Gallon L, Friedewald JJ: APOL1-associated end-stage renal disease in 55.5% to 66.7%) for those below the median. Patel
a living kidney transplant donor. Am J Transplant 16: 3568–3572, 2016 et al. (3) described the impact of premortem donor
PubMed hydroxyethyl starch (HES) use on allograft outcomes.
18. Lam NN, McArthur E, Kim SJ, Prasad GV, Lentine KL, Reese PP,
Kasiske BL, Lok CE, Feldman LS, Garg AX; Donor Nephrectomy Kidneys from donors who received HES had a higher
Outcomes Research (DONOR) Network; Donor Nephrectomy Out- crude rate of DGF (41% versus 31%; P,0.001). In this
comes Research DONOR Network: Gout after living kidney donation:
report, independent predictors of DGF included donor age
A matched cohort study. Am J Kidney Dis 65: 925–932, 2015 PubMed
19. Lin SY, Lin CL, Liu YL, Hsu WH, Lin CC, Wang IK, Jeng LB, Kao CH: (OR, 1.02; 95% CI, 1.01 to 1.04 per year), cold ischemia
Peptic ulcer disease in living liver donors: A longitudinal population- time (OR, 1.04; 95% CI, 1.02 to 1.06 per hour), creatinine
based study. Am J Transplant 16: 2925–2931, 2016 PubMed
elevation (OR, 1.5; 95% CI, 1.32 to 1.72 per mg/dl), and
HES use (OR, 1.41; 95% CI, 1.02 to 1.95). The broader
applicability of this remains unclear, because there are no
Delayed Graft Function
published data on the use of HES in kidney donors.
Delayed graft function (DGF) after kidney trans-
plantation is variably defined as the need for dialysis
during the first postoperative week, anuria, or failure of Donor, peritransplant, and recipient factors
prompt azotemia resolution. Importantly, DGF develops are implicated in the pathogenesis of delayed
in up to 25%–30% of kidney transplant recipients and is graft function after transplant, which in turn,
associated with increased cost, length of stay, and inferior impacts patient and graft survival.
long-term graft survival. There has been concern that the
Kidney Allocation System (KAS), a major change to
organ sharing implemented by the United Network for Premortem AKI
Organ Sharing in December of 2014, may be associated What about the use of kidneys from deceased
with increased DGF rates. Goals of the KAS included donors with premortem AKI? Heilman et al. (4) de-
directing the highest quality organs to younger/healthier scribed the outcome and phenotype of kidneys with and
recipients and increasing access to deceased donor kidney without donor AKI. DGF was more common in the
transplantation organs for highly sensitized patients and AKI group, but eGFR, graft survival at 1 year, and
racial/ethnic minorities at the expense of increased cold fibrosis scores at 12 months were similar for the two
ischemic time. Massie et al. (1), using national registry groups. At 1 month, there were 898 differentially
data, compared kidney distribution, transplant rates, and expressed genes in the AKI group (P,0.01), but at
recipient characteristics between January 1, 2013 and 4 months, there were no differences. The authors
December 3, 2014 (pre-KAS) with those between De- concluded that transplanting selected kidneys from
cember 4, 2014 and August 31, 2015 (post-KAS). deceased donors with AKI is safe and associates with
Regionally imported organs increased from 8.8% pre- excellent outcomes. Batra et al. (5) added to this lit-
KAS to 12.5% post-KAS; national imports increased erature by studying the clinical and histologic outcomes
from 12.7% pre-KAS to 19.1% post-KAS (P,0.001). In related to 61 transplanted kidneys from deceased
parallel, DGF increased from 24.8% pre-KAS to 29.9% donors with glomerular fibrin thrombi (GFT) conse-
post-KAS (P,0.001). The impact of this increase in DGF quent on disseminated intravascular coagulation before
rates nationwide on outcomes remains to be determined. organ procurement. DGF occurred in 49% of the GFT
Donor, peritransplant, and recipient factors have group and 39% of the control group (P50.14). Serum
been implicated in the pathogenesis of DGF. For ex- creatinine values at 1, 4, and 12 months and eGFR at 12
ample, Allen et al. (2) recently showed that donation months were similar in the two groups. Estimated 1-year
after cardiac death kidneys have inferior survival when graft survival was 93.2% in the GFT group and 95.1% in
294 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

the control group (P50.22 by log rank test). The authors donation service areas were randomly assigned to one of
concluded that GFT resolves rapidly after transplantation two targeted temperature ranges: 34C to 35C (hypother-
and that transplanting selected kidneys from deceased mia) or 36.5C to 37.5 C (normothermia). The study was
donors with GFT is a safe practice. Despite these terminated early on the recommendation of an independent
observations, deceased donor kidneys with AKI are often data safety monitoring board after the interim analysis
discarded due to fear of poor outcomes. Hall et al. (6) showed efficacy of hypothermia. DGF developed in 79
performed a multicenter study to determine associations recipients of kidneys from donors in the hypothermia
of AKI with kidney discard, DGF, and 6-month eGFRs. group (28%) and 112 recipients of kidneys from donors in
In 1632 donors, the kidney discard risk increased for AKI the normothermia group (39%; OR, 0.62; 95% CI, 0.43 to
stages 1, 2, and 3 compared with no AKI, with adjusted 0.92; P50.02). Feng (10), in an accompanying editorial,
relative risks (aRRs) of 1.28 (95% CI, 1.08 to 1.52), 1.82 describes some of the barriers that exist in conducting
(95% CI, 1.45 to 2.30), and 2.74 (95% CI, 2.0 to 3.75), research in deceased donors while issuing a cry to action:
respectively. The aRRs for DGF also increased by donor “The astonishing benefits of therapeutic hypothermia
AKI stage: aRR, 1.27; 95% CI, 1.09 to 1.49; aRR, 1.70; should inspire governmental regulatory and funding agen-
95% CI, 1.37 to 2.12; and aRR, 2.25; 95% CI, 1.74 to cies, basic and clinical scientists, and the entire donation
2.91, respectively. The 6-month eGFRs were similar and transplantation communities to vigorously advocate
across AKI categories but lower for recipients with for innovative research interventions in deceased donors to
DGF: 48 (interquartile range, 31–61) ml/min per improve the quality and increase the quantity of organs
1.73 m2 versus 58 (interquartile range, 45–75) ml/min available for transplantation” (10).
per 1.73 m2 for no DGF (P,0.001).
What about the use of postprocurement biopsy on Procedure Times
outcomes? Wang et al. (7) carried out a systematic re- Perhaps intuitively, prolonged ischemia during
view of the medical literature on the utility of procurement procurement, storage, and time to complete the vas-
and implantation biopsies for predicting post-transplant cular anastomoses has long been thought to impact
outcomes. Between 1994 and 2014, 47 studies were re- DGF risk. Osband et al. (11) studied the extraction time
viewed that examined the association between pretrans- of 576 kidneys beginning with aortic crossclamp and
plant donor biopsy findings and post-transplant graft perfusion/cooling of the kidneys and ending with re-
failure, DGF, or graft function. In general, study quality moval of the kidneys and placement on ice on the back
was poor. All of them were retrospective or did not table. This cohort was compared with Scientific Reg-
indicate if they were prospective. Results were heteroge- istry of Transplant Recipients (SRTR) data for out-
neous, with authors as often as not concluding that biopsy comes. Extraction time ranged from 14 to 123 minutes,
results did not predict post-transplant outcomes. The with a mean of 44.7 minutes. In SRTR-adjusted analy-
percentage glomerular sclerosis was the most often ex- ses, longer extraction times and DGF were statistically
amined parameter, and it failed to predict graft failure associated (OR, 1.19 per 5 minutes beyond 60 minutes;
in seven of 14 studies. Of 15 proposed semiquantitative 95% CI, 1.02 to 1.39; P50.03). Up to 60 minutes of
scoring systems, none consistently predicted post-transplant extraction time, DGF incidence was 27.8%; by 120 min-
outcomes across studies. In summary, it seems that the utes, it doubled to nearly 60%. Primary nonfunction rate
routine use of biopsies to help determine whether to trans- also rose dramatically to nearly 20% by 120 minutes of
plant a kidney is of little value in predicting outcomes. extraction time. Heylen et al. investigated the effect of
anastomosis time on allograft outcome in 669 first
Predonation Interventions single-kidney transplantations from brain-dead donors
Preprocurement interventions to mitigate deceased (12). Anastomosis time independently increased the
donor recipient DGF risk have yielded inconsistent risk of DGF (OR per minute, 1.05; 95% CI, 1.02 to
results to date. For example, preliminary reports that 1.07; P,0.001) and impaired allograft function after
donor remote ischemic preconditioning reduced DGF transplantation (P,0.01). In a subgroup of transplant
have not been reproduced (8). Provocatively, Niemann recipients, protocol-specified biopsies at 3 months,
et al. (9) recently targeted mild hypothermia in organ 1 year, and 2 years were blindly reviewed. Prolonged
donors before organ recovery. Organ donors (after anastomosis time independently increased the risk of
declaration of death per neurologic criteria) from two large post-transplant interstitial fibrosis and tubular atrophy
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 295

of these protocol-specified biopsies (P,0.001). Tennankore Delayed Graft Function and Rejection Risk
et al. (13) similarly found that ischemia time is associated DGF is commonly considered to be a risk factor
with adverse long-term patient and graft survival after for acute rejection, a biologically plausible association
kidney transplantation. given upregulation of chemokines, cytokines, and MHC
class II with AKI. However, this association has not
Recipient Factors been uniformly observed across all studies. In addition,
The prevalence of obesity is increasing globally and
the link between DGF and acute rejection may have
associates with CKD and premature mortality. However,
changed over time due to advances in immunosuppres-
the impact of recipient obesity on kidney transplant
sion and medical management. Wu et al. (17) conducted
outcomes remains unclear. Hill et al. (14) investigated the
a cohort study of 645 patients over 12 years to evaluate
association between recipient obesity (defined as body
the association of DGF and biopsy-proven acute re-
mass index .30 kg/m2) and mortality, death-censored
jection (BPAR) in a modern cohort of kidney transplant
graft loss, and DGF after kidney transplantation. A
recipients. The 1-, 3-, and 5-year cumulative probabil-
systematic review and meta-analysis, including 17 studies
ities of BPAR were 16.0%, 21.8%, and 22.6% in the
with a total of 138,081 patients, were analyzed. After
DGF group, significantly different from 10.1%, 12.4%,
adjustment, there was no significant difference in mortal-
and 15.7% in the non-DGF group. In a multivariable
ity risk in obese recipients (HR, 1.24; 95% CI, 0.90 to
Cox proportional hazards model, the adjusted relative
1.70; studies 55; n583,416). However, obesity was
HR for BPAR in DGF (versus no DGF) was 1.55 (95%
associated with an increased risk of death-censored
CI, 1.03 to 2.32). This association was generally robust
graft loss (HR, 1.06; 95% CI, 1.01 to 1.12; studies 55;
to different definitions of DGF. The relative HR was also
n583,416) and an increased likelihood of DGF (OR,
similarly elevated for T cell– or antibody-mediated
1.68; 95% CI, 1.39 to 2.03; studies 54; n528,847).
BPAR (HR, 1.52; 95% CI, 0.92 to 2.51 and HR, 1.54;
Peräsaari et al. (15) examined the association of
95% CI, 0.85 to 2.77, respectively). Finally, the associ-
donor-specific antibodies (DSAs) with DGF. Patients
ation was consistent across clinically relevant subgroups.
with DSAs had a higher incidence of DGF compared
Thus, DGF remains an important risk factor for BPAR
with nonsensitized patients (48% and 26%, respectively;
in a contemporary cohort of kidney transplant recipients.
P,0.001). Third-party antibodies (anti-HLA antibodies
Interventions to reduce the risk of DGF and/or its after
that are not donor specific) had no effect on DGF in-
effects remain of paramount importance to improve
cidence. The relative risk of DGF for patients with DSAs
kidney transplant outcomes.
in the multivariate analysis was 2.04 (95% CI, 1.25 to 3.34;
P,0.01). Analyses of the cumulative mean fluorescent
intensity (MFI) value of the DSAs revealed a rate of DGF Delayed graft function increases the risk
more than two times higher in patients with cumulative of allograft rejection by 50% compared
values of 3000–5000 MFI compared with cumulative with prompt graft function.
values of 1000–3000 MFI (65% versus 31%; P50.04).
DSAs against any locus were associated with an elevated
DGF incidence of 44%–69% compared with patients References
1. Massie AB, Luo X, Lonze BE, Desai NM, Bingaman AW, Cooper M,
without DSA (27%), independent of rejection. Importantly, Segev DL: Early changes in kidney distribution under the new
neither the use of antibody induction nor nonuse of such allocation system. J Am Soc Nephrol 27: 2495–2501, 2016 PubMed
agents impact risk of DGF in allosensitized patients. 2. Allen MB, Billig E, Reese PP, Shults J, Hasz R, West S, Abt PL: Donor
hemodynamics as a predictor of outcomes after kidney transplantation from
Schold et al. (16), in a cohort of approximately 14,000 donors after cardiac death. Am J Transplant 16: 181–193, 2016 PubMed
kidney retransplant recipients reported to the SRTR, 3. Patel MS, Niemann CU, Sally MB, De La Cruz S, Zatarain J, Ewing T,
disclosed DGF rates of approximately 22% for all groups Crutchfield M, Enestvedt CK, Malinoski DJ: The impact of hydroxy-
ethyl starch use in deceased organ donors on the development of
noted above, except alemtuzumab (27%; P50.04). delayed graft function in kidney transplant recipients: A propensity-
adjusted analysis. Am J Transplant 15: 2152–2158, 2015 PubMed
4. Heilman RL, Smith ML, Kurian SM, Huskey J, Batra RK, Chakkera
HA, Katariya NN, Khamash H, Moss A, Salomon DR, Reddy KS:
Patients with donor specific antibodies be- Transplanting kidneys from deceased donors with severe acute kidney
fore transplant have twice the incidence of injury. Am J Transplant 15: 2143–2151, 2015 PubMed
delayed graft function after transplantation. 5. Batra RK, Heilman RL, Smith ML, Thomas LF, Khamash HA, Katariya
NN, Hewitt WR, Singer AL, Mathur AK, Huskey J, Chakkera HA,
296 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

Moss A, Reddy KS: Rapid resolution of donor-derived glomerular fibrin kidney transplantation, with lymphocyte-depleting
thrombi after deceased donor kidney transplantation. Am J Transplant
16: 1015–1020, 2016 PubMed
induction reserved for high-risk cases. Despite this
6. Hall IE, Schröppel B, Doshi MD, Ficek J, Weng FL, Hasz RD, recommendation, most United States transplant recip-
Thiessen-Philbrook H, Reese PP, Parikh CR: Associations of deceased ients receive depleting induction therapy (Figure 8). It
donor kidney injury with kidney discard and function after trans-
plantation. Am J Transplant 15: 1623–1631, 2015 PubMed has been suggested that the older studies referenced by
7. Wang CJ, Wetmore JB, Crary GS, Kasiske BL: The donor kidney Kidney Disease Improving Global Outcomes mainly
biopsy and its implications in predicting graft outcomes: A systematic used outdated maintenance regimens, and specifically, no
review. Am J Transplant 15: 1903–1914, 2015 PubMed
8. Krogstrup NV, Oltean M, Nieuwenhuijs-Moeke GJ, Dor FJ, Møldrup large randomized trial has examined the effect of IL2RA
U, Krag SP, Bibby BM, Birn H, Jespersen B: Remote ischemic or rATG induction versus no induction in patients
conditioning on recipients of deceased renal transplants does not receiving tacrolimus (TAC), mycophenolic acid (MPA),
improve early graft function: A multicenter randomized, controlled
clinical trial. Am J Transplant 17: 1042–1049, 2017 PubMed and steroids (2). It has been suggested that, with this triple
9. Niemann CU, Feiner J, Swain S, Bunting S, Friedman M, Crutchfield maintenance therapy, the addition of basiliximab induc-
M, Broglio K, Hirose R, Roberts JP, Malinoski D: Therapeutic
tion may achieve an absolute risk reduction for acute
hypothermia in deceased organ donors and kidney-graft function. N
Engl J Med 373: 405–414, 2015 PubMed rejection of only 1%–4% in standard risk patients without
10. Feng S: Got it! Let’s cool it! But what’s next in organ donor research? improving graft or patient survival compared with no
Am J Transplant 16: 5–6, 2016 PubMed
11. Osband AJ, James NT, Segev DL: Extraction time of kidneys from
induction. In contrast, rATG induction lowers the relative
deceased donors and impact on outcomes. Am J Transplant 16: 700–703, risk (RR) of acute rejection by almost 50% versus IL2RA
2016 PubMed in patients at higher immunologic risk.
12. Heylen L, Naesens M, Jochmans I, Monbaliu D, Lerut E, Claes K, Heye S,
Verhamme P, Coosemans W, Bammens B, Evenepoel P, Meijers B,
Tanriover et al. (3) evaluated the United Network
Kuypers D, Sprangers B, Pirenne J: The effect of anastomosis time on for Organ Sharing data of transplant recipients maintained
outcome in recipients of kidneys donated after brain death: a cohort study. on TAC/mycophenolate (MPA) from 2000 to 2012 to
Am J Transplant 15: 2900–2907, 2015
13. Tennankore KK, Kim SJ, Alwayn IP, Kiberd BA: Prolonged warm
compare outcomes of IL2RA and other induction agents.
ischemia time is associated with graft failure and mortality after kidney Acute rejection within the first year and overall graft failure
transplantation. Kidney Int 89: 648–658, 2016 PubMed within 5 years of transplantation were more common with
14. Hill CJ, Courtney AE, Cardwell CR, Maxwell AP, Lucarelli G, Veroux M,
Furriel F, Cannon RM, Hoogeveen EK, Doshi M, McCaughan JA: no induction therapy (13.3%; P,0.001 and 28%; P50.01,
Recipient obesity and outcomes after kidney transplantation: A sys- respectively) with steroids and in the IL2RA category
tematic review and meta-analysis. Nephrol Dial Transplant 30: 1403– (11.1%; P50.16 and 27.4%; P,0.001, respectively)
1411, 2015 PubMed
15. Peräsaari JP, Kyllönen LE, Salmela KT, Merenmies JM: Pre-transplant without steroids (Figure 8). Compared with IL2RA,
donor-specific anti-human leukocyte antigen antibodies are associated analyses showed that outcomes in the steroid group were
with high risk of delayed graft function after renal transplantation. similar among induction categories, except that acute
Nephrol Dial Transplant 31: 672–678, 2016 PubMed
16. Schold J, Poggio E, Goldfarb D, Kayler L, Flechner S: Clinical outcomes rejection was significantly lower with rATG (odds ratio
associated with induction regimens among retransplant kidney recipients in [OR], 0.68; 95% confidence interval [95% CI], 0.62 to
the United States. Transplantation 99: 1165–1171, 2015 PubMed
0.74). In the no steroid group compared with IL2RA, the
17. Wu WK, Famure O, Li Y, Kim SJ: Delayed graft function and the risk
of acute rejection in the modern era of kidney transplantation. Kidney Int probabilities of acute rejection with rATG (OR, 0.80; 95%
88: 851–858, 2015 PubMed CI, 0.60 to 1.00) and AZ (OR, 0.68; 95% CI, 0.53 to 0.88)
were lower. rATG was associated with better graft survival
(hazard ratio [HR], 0.86; 95% CI, 0.75 to 0.99). The
Immunosuppression
authors concluded that, in deceased donor transplantation,
Induction Therapy compared with IL2RA induction, no induction was
Rabbit antithymocyte globulin (rATG)–induced associated with similar outcomes when TAC/MPA/
lymphocyte depletion and basiliximab-induced IL-2 steroids were used. However, rATG seems to offer better
receptor blockade (IL2RA) are the induction strategies graft survival over IL2RA in steroid avoidance proto-
in general use within the United States (1). Alemtuzumab cols. The same authors reached similar conclusions
(AZ) use is dwindling due to restricted access by when studying living donor transplant recipients (4).
the manufacturer. Prior randomized trials revealed Rituximab is a monoclonal antibody that targets
that rATG or IL2RA induction reduced early acute CD20 leading to B cell depletion. It is licensed for
rejection, prompting recommendations by the Kidney various non-transplant indications and has been stud-
Disease Improving Global Outcomes group that IL2RA ied off label both as an induction and rescue immu-
induction be used routinely as first-line therapy after notherapeutic. The efficacy and safety of rituximab as
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 297

Figure 8. Allograft rejection rates within the first post transplant year are lowest for those who receive lymphocyte depletion
(rabbit antithymocyte globulin [rATG] or alemtuzumab) compared with interleukin-2 receptor antibody [IL2r Ab] (basiliximab)
induction therapy versus no induction therapy. Observed frequencies of outcomes by induction type with or without steroids at
discharge. Reprinted with permission from Tanriover B, Jaikaransingh V, MacConmara MP, Parekh JR, Levea SL, Ariyamuthu
VK, Zhang S, Gao A, Ayvaci MU, Sandikci B, Rajora N, Ahmed V, Lu CY, Mohan S, Vazquez MA: Acute rejection rates and
graft outcomes according to induction regimen among recipients of kidneys from deceased donors treated with tacrolimus and
mycophenolate. Clin J Am Soc Nephrol 11: 1650–1661, 2016.

induction therapy in renal transplant patients were authors concluded that a single dose of rituximab was
evaluated by van den Hoogen et al. (5). In a double- safe as induction therapy but did not reduce the over-
blind, placebo-controlled study, 280 adult kidney trans- all incidence of biopsy-proven acute rejection, and it
plant patients were randomized to either a single dose of could benefit immunologically high-risk patients. Treat-
rituximab (375 mg/m2) or placebo. Maintenance immu- ment with rituximab was deemed safe.
nosuppression consisted of TAC, mycophenolate mofetil AZ is a humanized anti-CD52 mAb that is infre-
(MMF), and steroids. The incidence of rejection was quently used as induction therapy due to access restric-
comparable between rituximab- (23 of 138; 16.7%) and tions imposed by the manufacturer (Bayer, Wayne, NJ).
placebo-treated (30 of 142; 21.2%; P50.25) patients. Serrano et al. (6) reported a retrospective cohort study of
Immunologically high-risk patients (panel reactive anti- primary kidney transplant recipients receiving induction
body .6% or retransplant) not receiving rituximab had with AZ (n55521) or antithymocyte globulin (n58504)
a significantly higher incidence of rejection (13 of 34; and maintenance immunosuppression as TAC and MMF
38.2%) compared with other treatment groups (rituximab- with early steroid withdrawal. The transplant eras
treated immunologically high-risk patients and rituximab- were subcategorized. AZ was significantly associ-
or placebo-treated immunologically low-risk [panel ated with inferior death-censored graft survival in the
reactive antibody #6% or first transplant] patients: earliest 2003–2005 era (adjusted hazard ratio [aHR],
17.9%, 16.4%, and 15.7%, respectively; P50.004). 2.21; 95% CI, 1.72 to 2.84). However, these findings
Neutropenia (,1.53109/L) occurred more frequently were no longer significant more recently. Patient sur-
in rituximab-treated patients (24.3% versus 2.2%; vival and acute rejection with AZ were comparable
P,0.001). After 24 months, the cumulative incidence with antithymocyte globulin in the most recent era. The
of infections and malignancies was comparable. The effect of such findings is moot given the lack of general
298 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

availability of AZ. However, it may be distributed under substantially with time. Patient and donor character-
research protocols by the manufacturer with institutional istics explained only a limited amount of the observed
review board approval. variation in regimen use, whereas center choice explained
30%–46% of the use of nontriple-therapy immunosup-
Maintenance Therapy pression. The authors suggest that substantial variation in
Although maintenance immunosuppression man- center practice exists beyond that explained by differ-
agement in kidney transplantation has evolved to ences in patient and donor characteristics.
include a diverse repertoire of agents, TAC, mycophe- Regardless of regimen, medication nonadherence
nolate, and prednisone continue to represent the most increases the risk for kidney transplant loss after trans-
commonly used therapies (Figure 9) (1). plantation. Reese et al. (8) studied the utility of elec-
Despite this, the choice of immunosuppres- tronic reminders to enhance adherence in real time
sion regimen varies across transplant programs. Using using wireless-enabled pill bottles. Mean participant
a database integrating national transplant registry and age was 50 years old; 60% were men, and 40% were
pharmacy fill records, immunosuppression use after black. Mean adherence rates were 78%, 88%, and 55%
transplant was evaluated for 22,453 patients transplanted in groups with reminders, reminders plus notification,
in 249 United States programs between 2005 and 2010 or neither (control), respectively (P,0.001 for com-
by Axelrod et al. (7). Use of triple immunosuppression parison of each intervention with control). Interest-
composed of TAC, MPA, or azathioprine (AZA), and ingly, mean TAC levels were not significantly different
steroids varied widely (0%–100% of patients per pro- between groups.
gram), as did use of steroid-sparing regimens (0%–
77%), sirolimus-based regimens (0%–100%), and Tacrolimus
cyclosporine-based regimens (0%–78%) (Figure 10). Long-term transplant outcomes are clearly limited,
Use of triple therapy was more common in highly at least in part, by adverse effects of immunosuppressive
sensitized patients, women, and recipients with dialysis therapies, which have led to the search for minimization
duration .5 years. TAC has largely replaced cyclo- or withdrawal strategies. Many have been reported, and
sporine due to improved efficacy in preventing re- although some result in improved renal function, this is
jection, although registry data indicate that patient and often at increased risk of rejection (9). Dugast et al. (10)
graft survival rates are equivalent when a calcineurin conducted a prospective, randomized, multicenter, double-
inhibitor (CNI) is utilized. Sirolimus use has decreased blind, placebo-controlled clinical study to analyze the

Figure 9. Immunosuppression in adult kidney transplant recipients. First post transplant year immunosuppression trends in US
transplant recipients reveals in (A) most patients receive lymphocyte depleting induction therapy (rATG or alemtuzumab), (B)
Tacrolimus has almost completely replaced cyclosporine as the calcineurin inhibitor of choice, (C) Mycophenolate has almost
completely replaced Azathioprine as the antimetabolite of choice, (D) the use of mTOR inhibitors was never widespread and has
dwindled to almost zero, and in (E) most patients continue to receive corticosteroids. One-year post-transplant data are limited to
patients alive with graft function at 1 year post-transplant. Mycophenolate includes MMF and mycophenolate sodium. Reprinted
with permission from Hart A, Smith JM, Skeans MA, Gustafson SK, Stewart DE, Cherikh WS, Wainright JL, Boyle G, Snyder JJ,
Kasiske BL, Israni AK: Kidney. Am J Transplant 16[Suppl 2]: 11–46, 2016.
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 299

Figure 10. While most patients continue to receive the combination of tacrolimus (Tac)/mycophenolate mofetil (MMF)/
Prednisone, a degree of intercenter and regional variability exists regarding the specific alternate combinations that are in use.
The graphic represents the proportion of patients who receive one of six mutually exclusive immunosuppression (ISx) regimens
during months 6–12 after transplant. Each horizontal bar represents an individual center within United States regions ordered by
the proportion of patients who received triple ISx (TAC 1 mycophenolic acid (MPA)/azathioprine (AZA) 1 prednisone [Pred];
orange). Overall percentages of regimen use at patient level across centers: TAC 1 MPA/AZA 1 Pred, 33.8%; TAC 1 MPA/AZA
(no Pred), 25.8%; TAC without MPA/AZA, 11.3%; sirolimus (SRL) based, 9.9%; cyclosporine (CSA) based, 7.8%; and other
regimens (including CSA withdrawal or other trial medications), 11.6%. Proportion of patients receiving one of six mutually
300 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

impact of TAC withdrawal on patients .4 years post- steroid-free kidney transplant recipients receiving a moder-
transplantation with baseline normal histology, stable ate dose of MPA.
graft function, and no anti-HLA immunization. Only
ten of 52 eligible patients were randomized. Five patients
were assigned to the placebo arm, and five were assigned Tacrolimus remains the single most effec-
to the TAC maintenance arm. In the TAC maintenance tive immunosuppressive agent available to
arm, all patients maintained stable graft function, and prevent acute allograft rejection. Prospec-
no immunologic events occurred. By contrast, all five tive elimination studies in low risk kidney
patients in the placebo arm had to reintroduce TAC, allograft recipients have yielded unaccept-
because three presented with an acute rejection episode able results. Maintenance of therapeutic
(one humoral, one mixed, and one borderline) and two levels remains of critical importance for all
showed anti-HLA antibodies without histologic lesion transplant recipients.
(one with donor-specific antibodies [DSAs] and one
non-DSA). The authors concluded that TAC withdrawal
must be avoided long-term, even in highly selected stable Pharmacokinetics
kidney recipients. Another nail in the TAC elimination TAC blood-level variability associates with in-
coffin was provided by Hricik et al. (11), who reported ferior graft survival (13). TAC trough blood concen-
the results of the Clinical Trials in Organ Transplantation- trations for black kidney allograft recipients are lower
09 Trial. This was a randomized, prospective study of than those observed in white patients. This finding can
nonsensitized primary recipients of living donor kid- be associated with increased rejection and graft failure
ney transplants. Subjects received rATG, TAC, MMF, risks (14). Oetting et al. (14) identified two CYP3A5
and prednisone. Six months post-transplantation, subjects variants uniquely found in black recipients that in-
without de novo DSAs, acute rejection, or inflammation dependently associated with TAC troughs: CYP3A5*6
at protocol biopsy were randomized to wean off or remain (rs10264272) and CYP3A5*7 (rs41303343). These
on TAC. The study was terminated prematurely because variants and various clinical factors accounted for
of unacceptable rates of acute rejection (four of 14) and/or 53.9% of the observed variance in trough levels, with
de novo DSAs (five of 14) in the TAC withdrawal arm. 19.8% of the variance originating from demographic
Gatault et al. (12) took a different approach and and clinical factors (14). It has been suggested that
sought to determine the efficacy and safety of two dif- pretransplantation adaptation of the daily dose of TAC
ferent doses of extended release tacrolimus (LCPT) in to cytochrome genotype may be associated with im-
steroid-free kidney transplant recipients. More rejection proved achievement of target trough levels. CYP3A4*22
episodes occurred in the lower target TAC–level group is an allelic variant of the cytochrome P450 3A4 that
(50% dose reduction at 6 months with target level associates with decreased enzymatic degradatory activ-
.3 ng/ml; group A) than the higher target-level group ity, leading to reduced dosing. Pallet et al. (15) tested this
B (continued full-dose therapy with target level of 7– concept in a population of 186 kidney transplants, of
12 ng/ml; 11 versus three; P50.02): subclinical in- whom 9.3% (18 patients) were heterozygous for the
flammation (21.4% versus 8.8%; P50.05) and DSAs CYP3A4*22 genotype and none were homozygous (allele
(six versus zero patients; P,0.01). The authors concluded frequency, 4.8%). Ten days after transplantation, 11% of
that TAC levels should be maintained .7 mg/L during the the CYP3A4*22 carriers were within the target range,
first year after transplantation in low-immunologic risk, whereas among the CYP3A4*1/*1 carriers, 40% were

exclusive immunosuppression (ISx) regimens during months 6–12 after transplant. Each horizontal bar represents an individual
center within United States regions ordered by the proportion of patients who received triple ISx (TAC 1 MPA/AZA 1
prednisone [Pred]; orange). Overall percentages of regimen use at patient level across centers: TAC 1 MPA/AZA 1 Pred,
33.8%; TAC 1 MPA/AZA (no Pred), 25.8%; TAC without MPA/AZA, 11.3%; SRL based, 9.9%; cyclosporine (CSA) based,
7.8%; and other regimens (including CSA withdrawal or other trial medications), 11.6%. Reprinted with permission from
Axelrod DA, Naik AS, Schnitzler MA, Segev DL, Dharnidharka VR, Brennan DC, Bae S, Chen J, Massie A, Lentine KL:
National variation in use of immunosuppression for kidney transplantation: A call for evidence-based regimen selection. Am J
Transplant 16: 2453–2462, 2016.
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 301

within the target range (P50.02). The authors concluded Program (NephSAP) transplantation issue regarding
that the CYP3A4*22 allelic variant is associated with other LCPT preparations.
a significantly altered TAC metabolism, and carriers of
this polymorphism often reach supratherapeutic concen- Antimetabolite Therapy
trations. The same group evaluated the long-term clinical Mycophenolate (either MMF or enteric-coated
impact of the adaptation of initial TAC dosing according MPA/mycophenolate) has largely supplanted AZA as a
to the cytochrome genotype (16). The outcomes of 236 first-line agent in primary immunosuppression. Wagner
kidney transplant recipients included in the Tactique Study et al. (19) performed a Cochrane review of randomized,
were retrospectively investigated over a period of 5 years. controlled trials to evaluate the benefits and risks of
These patients were randomly assigned to receive TAC at MPA versus AZA as primary immunotherapy after
either a fixed dosage or one determined by genotype. The kidney transplantation. Data from 23 studies involving
incidence rates of biopsy-proven acute rejection and graft 3301 participants were included. MMF reduced the risk
survival were similar between the control and the adapted for graft loss, including death (RR, 0.82; 95% CI, 0.67
TAC dose groups. Patient deaths, cancer, cardiovascular to 1.0), and death-censored graft loss (RR, 0.78; 95%
events, and infections were also similar, and renal function CI, 0.62 to 0.99; P,0.05). No statistically significant
did not change. The authors concluded that optimization difference for MMF versus AZA treatment was found
of initial TAC dose using pharmacogenetic testing did not for all-cause mortality (16 studies, 2987 participants:
improve clinical outcomes. Shuker et al. (17) similarly RR, 0.95; 95% CI, 0.70 to 1.29). The risks for any acute
studied 244 living donor kidney transplant recipients. rejection (22 studies, 3301 participants: RR, 0.65; 95%
They concluded that pharmacogenetic adaptation of the CI, 0.57 to 0.73; P,0.01), biopsy-proven acute re-
TAC starting dose did not increase the number of patients jection (12 studies, 2696 participants: RR, 0.59; 95%
having therapeutic TAC exposure early after transplanta- CI, 0.52 to 0.68), and antibody-treated acute rejection
tion and did not lead to improved clinical outcome in (15 studies, 2914 participants: RR, 0.48; 95% CI, 0.36
a low-immunologic risk population. to 0.65; P,0.01) were reduced in MMF-treated patients.
Metaregression analysis suggested that the magnitude of
risk reduction of acute rejection may be dependent on
Pharmacogenetic testing is not associated the control rejection rate (relative risk reduction [RRR],
with improved outcomes after kidney 0.34; 95% CI, 0.10 to 1.09; P50.08), AZA dose (RRR,
transplantation. 1.01; 95% CI, 1.00 to 1.01; P50.10), and use of
cyclosporine A microemulsion (RRR, 1.27; 95% CI,
0.98 to 1.65; P50.07). Pooled analyses failed to show
a significant and meaningful difference between MMF
Extended Release Tacrolimus and AZA in kidney function measures. The risk for
The safety and efficacy of LCPT (Envarsus XR; cytomegalovirus (CMV) viremia in 13 studies with 2880
Veloxis Pharmaceuticals, Edison, NJ) compared with participants was not statistically significantly different
immediate release tacrolimus (IR-Tac) twice daily after between MMF- and AZA-treated patients (RR, 1.06;
kidney transplantation were studied by Rostaing et al. 95% CI, 0.85 to 1.32). The likelihood of tissue-invasive
(18). This group carried out a 2-year, double-blind, CMV disease was greater with MMF therapy in seven
multicenter, noninferiority design, phase 3 trial involving studies with 1510 participants (RR, 1.70; 95% CI, 1.10
543 de novo kidney recipients. Patients were randomly to 2.61). Adverse event profiles varied. Gastrointestinal
assigned to LCPT (n5268) or IR-Tac (n5275), and 507 symptoms were more likely in MMF-treated patients.
(93.4%) completed the 24-month study. LCPT tablets Thrombocytopenia and elevated liver enzymes were
were administered at 0.17 mg/kg daily, and IR-Tac was more common during AZA therapy.
administered twice daily at 0.1 mg/kg per day to maintain Adverse MMF-related effects often prompt dose
target trough ranges (first 30 days, 6–11 ng/ml; thereafter, reduction or discontinuation, which can lead to rejection
4–11 ng/ml). There were no differences in efficacy failure and possibly, graft loss. McAdams-DeMarco et al. (20)
or adverse effects between groups, including tremor and reported on risk factors for MMF dose reduction. Frailty
new-onset diabetes mellitus. Similar reports were pre- measures assessed included sarcopenia, weakness, ex-
viously reviewed in the Nephrology Self-Assessment haustion, low physical activity, and slowed walking
302 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

speed along with other patient and donor characteristics, onset diabetes after transplantation was significantly
including longitudinal MMF doses, and graft loss in 525 lower with belatacept plus TAC and belatacept alone
kidney transplantation recipients. Frail recipients were versus TAC alone (1.7%, 2.2%, and 3.8%, respectively;
1.29 times (95% CI, 1.01 to 1.66; P50.04) more likely P50.01). The authors suggested the addition of short-term
to experience MMF dose reduction, as were deceased TAC in the first year after transplant and lymphocyte-
donor recipients (aHR, 1.92; 95% CI, 1.44 to 2.54; depleting induction may be advisable when using
P,0.001) and older adults (age $65 versus ,65 years belatacept. Similarly, Cohen et al. (22) performed a co-
old: aHR, 1.47; 95% CI, 1.10 to 1.96; P50.01). Impor- hort study of adult kidney transplant recipients trans-
tantly, MMF dose reduction was independently associated planted between May 1, 2001 and December 31, 2015
with a substantially increased risk of death-censored graft using national transplant registry data to compare patient
loss (aHR, 5.24; 95% CI, 1.97 to 13.98; P50.001). and allograft survival in patients discharged on belatacept
versus TAC-based regimens. They found that belatacept
was not associated with a significant difference in risk of
MMF dose reduction after kidney trans- patient death (HR, 0.84; 95% CI, 0.61 to 1.15; P50.28)
plantation occurs more frequently in older or allograft loss (HR, 0.83; 95% CI, 0.62 to 1.11;
and frailer individuals and deceased donor P50.20), despite an increased risk of acute rejection
allograft recipients. MMF dose reduction in the first year post-transplant (OR, 3.12; 95% CI, 2.13
significantly associates with increased risk to 4.57; P,0.001). These findings were confirmed in
of graft loss. additional sensitivity analyses that accounted for use of
belatacept in combination with TAC, transplant center
effects, and differing approaches to matching.
Belatacept Because of the increased incidence of rejection
Belatacept is a cytotoxic T lymphocyte–associated seen with belatacept compared with CNI therapy, Xu
antigen-4 fusion protein that inhibits costimulation et al. (23) combined belatacept with AZ and rapamycin.
that was Food and Drug Administration–approved for Compared with conventional immunosuppression,
prevention of kidney allograft rejection in 2011 on the lymphocyte depletion evoked substantial homeostatic
basis of the Belatacept Evaluation of Nephroprotection lymphocyte activation balanced by regulatory T and B
and Efficacy as First-Line Immunosuppression Trial cell phenotypes. The reconstituted T cell repertoire was
and the Belatacept Evaluation of Nephroprotection and enriched for CD281 naı̈ve cells, notably diminished in
Efficacy as First-Line Immunosuppression Trial Ex- belatacept-resistant CD282 memory subsets, and depleted
tended Criteria Donors Trial previously discussed in prior of polyfunctional donor-specific T cells. Importantly,
NephSAP transplantation issues. The impact of belatacept in the latter cells could respond to third party and latent
a real clinical setting has been recently reported. Wen et al. herpes viruses. B cell responses were similarly favor-
(21) performed a retrospective cohort study using registry able without alloantibody development and a reduction
data comparing clinical outcomes between belatacept- in memory subsets, changes not seen in conventionally
and TAC-treated adult kidney transplant recipients from treated patients. The combined belatacept with AZ and
January 6, 2011 to January 12, 2014. Of 50,244 total rapamycin regimen uniquely altered the immune pro-
participants, 417 received belatacept plus TAC, 458 re- file, producing a repertoire enriched for CD281 T cells,
ceived belatacept alone, and 49,369 received TAC alone at hyporesponsive to donor alloantigen, and competent in
discharge. Belatacept alone was associated with a higher its protective immune capabilities. The resulting repertoire
risk of 1-year acute rejection, with the highest rates as- was permissive for control of rejection with belatacept
sociated with nonlymphocyte-depleting induction (aHR, monotherapy. The broader applicability of such an ap-
2.65; 95% CI, 1.90 to 3.70; P,0.001). There were no proach remains to be determined.
significant differences in rejection rates between belatacept
plus TAC and TAC alone. In expanded criteria donor Steroid Therapy
allograft recipients, 1-year kidney function was higher with Haller et al. (24) performed a Cochrane review to
belatacept plus TAC and belatacept alone versus TAC evaluate the benefits and harms of steroid withdrawal or
alone (mean eGFRs 565.6, 60.4, and 54.3 ml/min per avoidance for kidney transplant recipients. Forty-eight
1.73 m2, respectively; P,0.001). The incidence of new- studies (224 reports) that involved 7803 randomized
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 303

participants were included. These studies evaluated randomized, double-blind, placebo-controlled study pro-
three different comparisons: steroid avoidance or with- tocol that evaluated the impact of ramipril on urinary
drawal compared with steroid maintenance and steroid protein excretion in renal transplant patients converted to
avoidance compared with steroid withdrawal. For adult sirolimus from CNI treatment. Patients received ramipril
studies, there was no significant difference in patient or placebo for up to 6 weeks before conversion and 52
mortality in either studies comparing steroid withdrawal weeks afterward. Doses were increased if patients de-
with steroid maintenance (ten studies, 1913 participants, veloped proteinuria (urine protein-to-creatinine ratio
death at 1 year post-transplantation: RR, 0.68; 95% CI, $0.5); losartan was given as rescue therapy for persis-
0.36 to 1.30) or studies comparing steroid avoidance with tent proteinuria. The primary end point was time to
steroid maintenance (ten studies, 1462 participants, death losartan initiation. Of 295 patients randomized, 264 met
at 1 year post-transplantation: RR, 0.96; 95% CI, 0.52 to the criteria for sirolimus conversion (ramipril, n5138;
1.80). Similarly, no significant difference in graft loss was placebo, n5126). At 52 weeks, the cumulative rate of
found comparing steroid withdrawal with steroid main- losartan initiation was significantly lower with ramipril
tenance (eight studies, 1817 participants) or graft loss (6.2%) versus placebo (23.2%; P,0.001). No signifi-
(excluding death with functioning graft at 1 year after cant differences were observed between ramipril and
transplantation: RR, 1.17; 95% CI, 0.72 to 1.92) and placebo for change in GFR or rejection. Treatment-
comparing steroid avoidance with steroid maintenance emergent adverse events were consistent with the known
(seven studies, 1211 participants, graft loss excluding safety profile of sirolimus and were not potentiated by
death with functioning graft at 1 year after transplantation: ramipril coadministration. Ramipril was deemed effec-
RR, 1.09; 95% CI, 0.64 to 1.86). The risk of acute tive in reducing the incidence of proteinuria for up to 1
rejection significantly increased in patients treated with year after conversion to sirolimus in maintenance renal
steroids for ,14 days post-transplantation (seven studies, transplant patients.
835 participants: RR, 1.58; 95% CI, 1.08 to 2.30) and pa- Sirolimus has antineoplastic properties compared
tients withdrawn from steroids later post-transplantation with CNI therapy. Yanik et al. (26) investigated
(ten studies, 1913 participants; RR, 1.77; 95% CI, 1.20 to sirolimus effects on cancer incidence among kidney
2.61). There was no evidence to suggest a difference in recipients. Overall, the incidence was not significantly
harmful events, such as infection and malignancy, in adult lowered by sirolimus use (HR, 0.88; 95% CI, 0.70 to
kidney transplant recipients. The effect of steroid with- 1.11). However, the frequency of prostate cancer was
drawal in children is unclear. The authors concluded that greater during sirolimus use (HR, 1.86; 95% CI, 1.15 to
steroid avoidance and withdrawal after kidney transplan- 3.02). Incidence of other cancers was similar or lower
tation significantly increase the risk of acute rejection. with sirolimus use, with a 26% decrease overall (HR,
0.74; 95% CI, 0.57 to 0.96; excluding prostate cancer).
Results were similar after adjustment for demographic
Steroid avoidance and withdrawal after kid- and clinical characteristics. Despite such properties, mor-
ney transplantation significantly increase the tality rates for patients with cancer treated with mTORi are
risk of acute rejection. higher than for similar patients maintained on CNI therapy
as discussed in a prior transplantation NephSAP.
There are occasional reports of the utility of
everolimus in combination with cyclosporine. Oh et al.
Mammalian Target of Rapamycin Inhibitor (27) performed an open label study to compare the
Therapy efficacy and tolerability of everolimus and reduced
The use of mammalian target of rapamycin in- exposure to cyclosporine (investigational group) with
hibitor (mTORi) therapy, sirolimus and everolimus, enteric-coated mycophenolate sodium and standard ex-
continues to decline due to inferior outcomes compared posure cyclosporine (control group) in combination
with conventional immunotherapy. In parallel, new with basiliximab and steroids. Rejection rates were
reports of the use of such therapy are dwindling. One similar between groups (7.5% [investigational group]
of the potential complications of mTORi therapy is versus 11.1% [control group]; P50.57). The mean
proteinuria from nephrin inhibition that associates with eGFRs of the investigational group at 12 months after
FSGS. Mandelbrot et al. (25) reported on a prospective, transplantation were significantly higher (68.1616.8
304 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

ml/min per 1.73 m2) than those of the control group References
(60.6615.8 ml/min per 1.73 m2; P50.02). There were 1. Hart A, Smith JM, Skeans MA, Gustafson SK, Stewart DE, Cherikh
WS, Wainright JL, Boyle G, Snyder JJ, Kasiske BL, Israni AK: Kidney.
no significant differences (P.0.05) in the frequency of Am J Transplant 16[Suppl 2]: 11–46, 2016 PubMed
discontinuations or serious adverse events between 2. Hellemans R, Bosmans JL, Abramowicz D: Induction therapy for
groups. kidney transplant recipients: Do we still need Anti-IL2 receptor mono-
clonal antibodies? Am J Transplant 17: 22–27, 2017 PubMed
Tedesco-Silva et al. (28) studied the incidence of 3. Tanriover B, Jaikaransingh V, MacConmara MP, Parekh JR, Levea SL,
CMV infection/disease in de novo kidney transplant Ariyamuthu VK, Zhang S, Gao A, Ayvaci MU, Sandikci B, Rajora N,
recipients receiving everolimus or mycophenolate with Ahmed V, Lu CY, Mohan S, Vazquez MA: Acute rejection rates and
graft outcomes according to induction regimen among recipients of
no CMV pharmacologic prophylaxis. Patients were kidneys from deceased donors treated with tacrolimus and mycophe-
randomized to receive rATG/TAC/everolimus/prednisone nolate. Clin J Am Soc Nephrol 11: 1650–1661, 2016 PubMed
(group 1), basiliximab/TAC/everolimus/prednisone (group 4. Tanriover B, Zhang S, MacConmara M, Gao A, Sandikci B, Ayvaci
MU, Mete M, Tsapepas D, Rajora N, Mohan P, Lakhia R, Lu CY,
2), or basiliximab/TAC/mycophenolate/prednisone (group Vazquez M: Induction therapies in live donor kidney transplantation on
3). Patients in group 1 had a 90% proportional reduction in tacrolimus and mycophenolate with or without steroid maintenance.
the incidence of CMV infection (4.7% versus 37.6%; HR, Clin J Am Soc Nephrol 10: 1041–1049, 2015 PubMed
5. van den Hoogen MW, Kamburova EG, Baas MC, Steenbergen EJ,
0.10; 95% CI, 0.04 to 0.29; P,0.001). Group 2 had a 75% Florquin S, M Koenen HJ, Joosten I, Hilbrands LB: Rituximab as
proportional reduction in frequency of CMV infection/ induction therapy after renal transplantation: A randomized, double-
disease compared with the control group (10.8% versus blind, placebo-controlled study of efficacy and safety. Am J Transplant
15: 407–416, 2015 PubMed
37.6%; HR, 0.25; 95% CI, 0.13 to 0.48; P,0.001). No 6. Serrano OK, Friedmann P, Ahsanuddin S, Millan C, Ben-Yaacov A,
differences were observed in the frequencies of acute re- Kayler LK: Outcomes associated with steroid avoidance and alemtuzumab
jection, wound-healing complications, delayed graft func- among kidney transplant recipients. Clin J Am Soc Nephrol 10: 2030–
2038, 2015 PubMed
tion, or proteinuria. 7. Axelrod DA, Naik AS, Schnitzler MA, Segev DL, Dharnidharka VR,
The salutary effect of mTORi therapy on GFR com- Brennan DC, Bae S, Chen J, Massie A, Lentine KL: National variation in
pared with CNI therapy has been well described. Rostaing use of immunosuppression for kidney transplantation: A call for evidence-
based regimen selection. Am J Transplant 16: 2453–2462, 2016 PubMed
et al. (29) provide a cautionary note. They report the results 8. Reese PP, Bloom RD, Trofe-Clark J, Mussell A, Leidy D, Levsky S,
of a randomized, open label, 12-month trial, in which Zhu J, Yang L, Wang W, Troxel A, Feldman HI, Volpp K: Automated
de novo kidney transplant patients received cyclosporine, reminders and physician notification to promote immunosuppression
adherence among kidney transplant recipients: A randomized trial. Am J
enteric-coated mycophenolate sodium, and steroids. Pa- Kidney Dis 69: 400–409, 2017 PubMed
tients were stratified by an epithelial-mesenchymal transi- 9. Sawinski D, Trofe-Clark J, Leas B, Uhl S, Tuteja S, Kaczmarek JL, French
tion profile based on their month 3 biopsies and then B, Umscheid CA: Calcineurin inhibitor minimization, conversion, with-
drawal, and avoidance strategies in renal transplantation: A systematic
randomized to start everolimus with half-dose enteric- review and meta-analysis. Am J Transplant 16: 2117–2138, 2016 PubMed
coated mycophenolate sodium (EC1; 720 mg/d) and 10. Dugast E, Soulillou JP, Foucher Y, Papuchon E, Guerif P, Paul C,
cyclosporine withdrawal (CNI2) or continue cyclospor- Riochet D, Chesneau M, Cesbron A, Renaudin K, Dantal J, Giral M,
Brouard S: Failure of calcineurin inhibitor (tacrolimus) weaning
ine with standard enteric-coated mycophenolate sodium randomized trial in long-term stable kidney transplant recipients. Am
(CNI). The primary end point was progression of graft J Transplant 16: 3255–3261, 2016 PubMed
fibrosis that developed in 46.2% (12 of 26) of CNI2/EC1 11. Hricik DE, Formica RN, Nickerson P, Rush D, Fairchild RL, Poggio ED,
Gibson IW, Wiebe C, Tinckam K, Bunnapradist S, Samaniego-Picota M,
patients versus 51.6% (16 of 31) of CNI/EC1 patients Brennan DC, Schröppel B, Gaber O, Armstrong B, Ikle D, Diop H, Bridges
(P50.68). Biopsy-proven acute rejection and subclinical ND, Heeger PS; Clinical Trials in Organ Transplantation-09 Consortium:
events occurred in 25.0% and 5.1% of CNI2 and CNI Adverse outcomes of tacrolimus withdrawal in immune-quiescent kidney
transplant recipients. J Am Soc Nephrol 26: 3114–3122, 2015 PubMed
patients, respectively (P,0.001). The authors concluded 12. Gatault P, Kamar N, Büchler M, Colosio C, Bertrand D, Durrbach A,
that early CNI withdrawal with everolimus initiation does Albano L, Rivalan J, Le Meur Y, Essig M, Bouvier N, Legendre C,
not prevent interstitial fibrosis. Moulin B, Heng AE, Weestel PF, Sayegh J, Charpentier B, Rostaing L,
Thervet E, Lebranchu Y: Reduction of extended-release tacrolimus dose
in low-immunological-risk kidney transplant recipients increases risk of
rejection and appearance of donor-specific antibodies: A randomized
study. Am J Transplant 17: 1370–1379, 2017 PubMed
The use of the mTOR inhibitors, sirolimus 13. Rozen-Zvi B, Schneider S, Lichtenberg S, Green H, Cohen O, Gafter U,
and everolimus, continues to decline due Chagnac A, Mor E, Rahamimov R: Association of the combination of
time-weighted variability of tacrolimus blood level and exposure to low
to unfavorable efficacy, adverse event and drug levels with graft survival after kidney transplantation. Nephrol
outcome measures compared with calci- Dial Transplant 32: 393–399, 2017 PubMed
neurin inhibition. 14. Oetting WS, Schladt DP, Guan W, Miller MB, Remmel RP, Dorr C,
Sanghavi K, Mannon RB, Herrera B, Matas AJ, Salomon DR, Kwok
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 305

PY, Keating BJ, Israni AK, Jacobson PA; DeKAF Investigators: 29. Rostaing L, Hertig A, Albano L, Anglicheau D, Durrbach A, Vuiblet V,
Genomewide association study of tacrolimus concentrations in African Moulin B, Merville P, Hazzan M, Lang P, Touchard G, Hurault deLigny
American kidney transplant recipients identifies multiple CYP3A5 B, Quéré S, Di Giambattista F, Dubois YC, Rondeau E; CERTITEM
alleles. Am J Transplant 16: 574–582, 2016 PubMed Study Group: Fibrosis progression according to epithelial-mesenchymal
15. Pallet N, Jannot AS, El Bahri M, Etienne I, Buchler M, de Ligny BH, transition profile: A randomized trial of everolimus versus CsA. Am J
Choukroun G, Colosio C, Thierry A, Vigneau C, Moulin B, Le Meur Y, Transplant 15: 1303–1312, 2015 PubMed
Heng AE, Subra JF, Legendre C, Beaune P, Alberti C, Loriot MA, Thervet
E: Kidney transplant recipients carrying the CYP3A4*22 allelic variant
have reduced tacrolimus clearance and often reach supratherapeutic
tacrolimus concentrations. Am J Transplant 15: 800–805, 2015 PubMed
16. Pallet N, Etienne I, Buchler M, Bailly E, Hurault de Ligny B, Rejection
Choukroun G, Colosio C, Thierry A, Vigneau C, Moulin B, Le Meur
Y, Heng AE, Legendre C, Beaune P, Loriot MA, Thervet E: Long-term The incidence of acute rejection within the first
clinical impact of adaptation of initial tacrolimus dosing to CYP3A5
genotype. Am J Transplant 16: 2670–2675, 2016 PubMed year post-transplant decreased for living and deceased
17. Shuker N, Bouamar R, van Schaik RH, Clahsen-van Groningen MC, donor transplant recipients (Figure 11) from 10% during
Damman J, Baan CC, van de Wetering J, Rowshani AT, Weimar W, 2009 and 2010 to 7.9% during 2013 and 2014 (1).
van Gelder T, Hesselink DA: A randomized controlled trial comparing
the efficacy of Cyp3a5 genotype-based with body-weight-based tacrolimus The Banff consortium that classifies allograft
dosing after living donor kidney transplantation. Am J Transplant 16: rejection recently reported the outcomes of the 12th
2085–2096, 2016 PubMed
Banff Conference on Allograft Pathology (2). Some of
18. Rostaing L, Bunnapradist S, Grinyó JM, Ciechanowski K, Denny JE,
Silva HT Jr., Budde K; Envarsus Study Group: Novel once-daily the key issues discussed and reported at this meeting
extended-release tacrolimus versus twice-daily tacrolimus in de novo included C4d2 antibody-mediated rejection (AMR),
kidney transplant recipients: Two-year results of phase 3, double-blind,
randomized trial. Am J Kidney Dis 67: 648–659, 2016 PubMed
the relationships of donor-specific antibody (DSA) tests
19. Wagner M, Earley AK, Webster AC, Schmid CH, Balk EM, Uhlig K: with transplant histopathology, molecular transplant di-
Mycophenolic acid versus azathioprine as primary immunosuppression agnostics, and transcriptome gene sets to supplement the
for kidney transplant recipients. Cochrane Database Syst Rev 12:
CD007746, 2015 PubMed
diagnosis and classification of rejection. Newly intro-
20. McAdams-DeMarco MA, Law A, Tan J, Delp C, King EA, Orandi B, duced concepts include the inflammation within areas of
Salter M, Alachkar N, Desai N, Grams M, Walston J, Segev DL: Frailty, Interstitial Fibrosis and Tubular Atrophy (i-IFTA) score
mycophenolate reduction, and graft loss in kidney transplant recipients.
Transplantation 99: 805–810, 2015 PubMed
that describes inflammation within areas of interstitial
21. Wen X, Casey MJ, Santos AH, Hartzema A, Womer KL: Comparison fibrosis and tubular atrophy and acceptance of transplant
of utilization and clinical outcomes for belatacept- and tacrolimus-based arteriolopathy within the descriptions of chronic active
immunosuppression in renal transplant recipients. Am J Transplant 16:
3202–3211, 2016 PubMed
22. Cohen JB, Eddinger KC, Forde KA, Abt PL, Sawinski D: Belatacept
compared to tacrolimus for kidney transplantation: A propensity score
matched cohort study [published online ahead of print December 8,
2016]. Transplantation PubMed
23. Xu H, Samy KP, Guasch A, Mead SI, Ghali A, Mehta A, Stempora L,
Kirk AD: Postdepletion lymphocyte reconstitution during belatacept
and rapamycin treatment in kidney transplant recipients. Am J Trans-
plant 16: 550–564, 2016 PubMed
24. Haller MC, Royuela A, Nagler EV, Pascual J, Webster AC: Steroid
avoidance or withdrawal for kidney transplant recipients. Cochrane
Database Syst Rev 8: CD005632, 2016 PubMed Figure 11. Allograft rejection rates within the first post
25. Mandelbrot DA, Alberú J, Barama A, Marder BA, Silva HT Jr., transplant year are steadily declining for both deceased and
Flechner SM, Flynn A, Healy C, Li H, Tortorici MA, Schulman SL:
Effect of ramipril on urinary protein excretion in maintenance renal
living donor transplant recipients. Incidence of acute re-
transplant patients converted to sirolimus. Am J Transplant 15: 3174– jection by 1 year post-transplant among adult kidney trans-
3184, 2015 PubMed plant recipients by donor type. Acute rejection is defined as
26. Yanik EL, Gustafson SK, Kasiske BL, Israni AK, Snyder JJ, Hess GP, Engels a record of acute or hyperacute rejection as reported on the Organ
EA, Segev DL: Sirolimus use and cancer incidence among US kidney Procurement and Transplantation Network (OPTN) Transplant
transplant recipients. Am J Transplant 15: 129–136, 2015 PubMed
27. Oh CK, Huh KH, Ha J, Kim YH, Kim YL, Kim YS: Safety and efficacy Recipient Registration or Transplant Recipient Follow-Up Form.
of the early introduction of everolimus with reduced-exposure cyclo- Only the first rejection event is counted. Cumulative incidence is
sporine a in de novo kidney recipients. Transplantation 99: 180–186, estimated using the Kaplan–Meier competing risk method.
2015 PubMed SRTR, Scientific Registry of Transplant Recipients. Reprinted
28. Tedesco-Silva H, Felipe C, Ferreira A, Cristelli M, Oliveira N, Sandes- with permission from Hart A, Smith JM, Skeans MA, Gustafson
Freitas T, Aguiar W, Campos E, Gerbase-DeLima M, Franco M,
Medina-Pestana J: Reduced incidence of cytomegalovirus infection in
SK, Stewart DE, Cherikh WS, Wainright JL, Boyle G, Snyder
kidney transplant recipients receiving everolimus and reduced tacrolimus JJ, Kasiske BL, Israni AK: Kidney. Am J Transplant 16[Suppl
doses. Am J Transplant 15: 2655–2664, 2015 PubMed 2]: 11–46, 2016.
306 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

Table 1. Updated Banff classification categories


Updated Banff Classification Categories
Category 1: Normal biopsy or nonspecific changes
Category 2: Antibody-mediated changes
Acute/active ABMR: Three features required
Histologic evidence of acute tissue injury (inflammation, TMA, ATN)
Linear C4d staining
Serologic evidence of DSA
Chronic active ABMR: Three features required
Histologic evidence of chronic tissue injury (interstitial fibrosis and tubular atrophy/TG, arterial sclerosis)
Linear C4d staining
Serologic evidence of DSA
C4d staining without evidence of rejection
Category 3: Borderline changes
Category 4: TCMR
Acute TCMR: Grades
IA. Significant interstitial inflammation (.25% of nonsclerotic cortical parenchyma, i2 or i3) and foci of moderate
tubulitis (t2)
IB. Significant interstitial inflammation (.25% of nonsclerotic cortical parenchyma, i2 or i3) and foci of severe
tubulitis (t3)
IIA. Mild to moderate intimal arteritis
IIB. Severe intimal arteritis
III. Transmural arteritis
Chronic TCMR
Chronic allograft arteriopathy
Category 5: Interstitial fibrosis and tubular atrophy
Grades
I. Mild interstitial fibrosis and tubular atrophy (#25% of cortical area)
II. Moderate interstitial fibrosis and tubular atrophy (26%–50% of cortical area)
III. Severe interstitial fibrosis and tubular atrophy (.50% of cortical area)
Category 6: Other changes not considered to be rejection
BK virus nephropathy
Post-transplant lymphoproliferative disorders
Calcineurin inhibitor nephrotoxicity
Acute tubular injury
Recurrent disease
De novo glomerulopathy (other than transplant glomerulopathy)
Pyelonephritis
Drug-induced interstitial nephritis
ABMR, antibody mediated rejection; TMA, thrombotic microangiopathy; ATN, acute tubular necrosis; TG, transplant glomerulopathy; i2 and i3, scores that represent increasing
degrees of inflammatory cell infiltrate in the interstitium; t2 and t3, scores that represent increasing degrees of inflammatory cell infiltrate in the tubules. Modified from Loupy
A, Haas M, Solez K, Racusen L, Glotz D, Seron D, Nankivell BJ, Colvin RB, Afrouzian M, Akalin E, Alachkar N, Bagnasco S, Becker JU, Cornell L, Drachenberg C, Dragun D, de
Kort H, Gibson IW, Kraus ES, Lefaucheur C, Legendre C, Liapis H, Muthukumar T, Nickeleit V, Orandi B, Park W, Rabant M, Randhawa P, Reed EF, Roufosse C, Seshan SV, Sis B,
Singh HK, Schinstock C, Tambur A, Zeevi A, Mengel M: The Banff 2015 Kidney Meeting Report: Current challenges in rejection classification and prospects for adopting
molecular pathology. Am J Transplant 17: 28–41, 2017.

T cell–mediated rejection (TCMR) or chronic AMR. A conducted a worldwide survey among members of
mixed pattern of TCMR and AMR was increasingly the Renal Pathology Society. A web-based survey
recognized. The current iteration of the classification was sent out to all 503 current members with 153
system follows (Table 1). responses. Among the 139 nephropathologists using
Although the Banff consortium initially convened the borderline category, 67% use the Banff 1997
to standardize renal histology definitions for research, definition. Thirty-seven percent admitted to some-
the classification is now broadly clinically utilized. To times exaggerating Banff in the presence of tubulitis
evaluate adherence with the system, Becker et al. (3) to reach a diagnosis of borderline. Forty-eight percent
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 307

were dissatisfied with the definition of borderline. with adverse impact on both access to transplantation
Most of the influential manuscripts used the 1997 and outcomes; thus, the current practice is transfusion
definition, contrary to the current one. The authors avoidance for transplants candidates. However, little is
concluded that there is considerable dissatisfaction known about the impact of post-transplant blood trans-
with the borderline category and that practice is fusion on the sensitization of anti-HLA antibodies and
variable. It is unclear whether the more recent update the formation of DSAs. Ferrandiz et al. (6) determined
described above addresses these issues. Halloran et al. the 1-year incidence of DSAs and AMR in kidney
(4) used microarray-derived molecular AMR scores as transplant patients who had or had not received a blood
a histology-independent estimate of AMR in 703 transfusion during the first year post-transplantation.
biopsies to re-evaluate AMR criteria and determine There were 390 non–HLA-sensitized patients who had
the relative importance of various lesions. They con- received an ABO-compatible kidney transplant and had
firmed that the important features for AMR diagnosis not previously or simultaneously received a nonrenal
were peritubular capillaritis (PTC), glomerulitis, glo- organ transplant. A surprisingly high proportion of
merular double contours, DSA, and C4d staining. The patients, 64%, received a red blood cell transfusion
group questioned some features, including arterial fibro- within the first year after transplantation, most within
sis, vasculitis, and acute tubular injury. the first month. Importantly, the overall 1-year in-
cidence of DSAs was significantly higher in patients
who had undergone transfusion (7.2% versus 0.7% in
The incidence of acute renal allograft re- patients with no transfusion; P,0.001). More impor-
jection declined from 10% in 2009 and tantly, AMR also occurred more often in the transfusion
2010 to 7.9% in 2013 and 2014. group (n515; 6%) compared with the nontransfusion
group (n52; 1.4%; P50.04). Blood transfusion was an
independent predictive factor for de novo donor-specific
Temporal Dynamics antibody (dnDSA) formation but not for AMR. Patients
Halloran et al. (5) used microarrays and conven- who had a transfusion and developed DSAs were more
tional methods to study rejection in 703 unselected often treated with cyclosporine (n510; 55.5%) rather
biopsies taken between 3 days and 35 years post- than tacrolimus (n545; 19.4%; P,0.001). The authors
transplant. Rejection was conventionally diagnosed in concluded that post-transplant blood transfusion may
205 biopsy specimens (28%): 67 pure TCMR, 110 pure increase immunologic risk, especially in underimmuno-
AMR, and 28 mixed (89 designated borderline). Using suppressed patients. Although this finding is indeed
microarrays, rejection was diagnosed in 228 biopsy cautionary, the generalizability of the observation is less
specimens (32%): 76 pure TCMR, 124 pure AMR, and clear, because most transplant recipients are treated with
28 mixed. Molecular assessment confirmed most con- tacrolimus and are not transfused.
ventional diagnoses (agreement was 90% for TCMR
and 83% for AMR), but it revealed some variation,
particularly in mixed rejection, and improved predic- Transfusion post-transplantation may be
tion of graft failure. AMR was strongly associated with associated with new-onset donor specific
increased graft loss, but TCMR was not. AMR became antibodies and increased risk of antibody-
common in biopsy specimens obtained .1 year post- mediated rejection.
transplant and continued to appear in all subsequent
intervals. TCMR was common early but progressively
disappeared over time. Provocatively, TCMR defined
by molecular and conventional features was never Novel Rejection Mediators
observed after 10 years in this cohort. Allocation algorithms for deceased donor kidney
transplantation often incorporate HLA mismatches at
Transfusion the HLA A, B8, and DR loci but not HLA mismatches
The purported benefits of blood transfusion pre- at other loci, including HLA-DQ. Lim et al. (7) studied
transplantation (the so-called transfusion effect) have the impact of HLA-DQ mismatches on renal allograft
long been overshadowed by the risk of allosensitization, outcomes. Seven hundred eighty-eight recipients who
308 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

received zero, one, or two HLA-DQ mismatched Adherence


kidneys were followed for a median of 2.8 years. Nonadherence and HLA mismatch have been
Compared with no HLA-DQ mismatched allograft independently associated with poorer long-term outcomes
recipients, those who had received one or two HLA- post-transplantation. Wiebe et al. (11) prospectively cor-
DQ mismatched kidneys experienced more rejection related HLA mismatching with medication adherence
(50 of 321 [15.6%] versus 117 of 467 [25%]; P,0.01), measurement via electronic monitors in medication vial
late rejections (occurring .6 months post-transplant; caps. Recipients were grouped by medication adherence
eight of 321 [2.4%] versus 27 of 467 [5.8%]; P50.03), and high- or low-epitope mismatch load. They found that
and AMRs (12 of 321 [3.7%] versus 38 of 467 [8.1%]; the combination of higher epitope mismatch and poor
P50.01). Compared with recipients of zero HLA-DQ adherence acted synergistically to determine the risk of
mismatched kidneys, the adjusted hazard ratios [HRs] rejection or graft loss. For example, nonadherent recipients
for any and late rejections in recipients who had with an HLA-DR epitope mismatch experienced increased
received one or two HLA-DQ mismatched kidneys graft loss (35% versus 8%; P,0.01) compared with
were 1.54 (95% confidence interval [95% CI], 1.08 to adherent recipients with low epitope mismatch.
2.19) and 2.85 (95% CI, 1.05 to 7.75), respectively.
Bachelet et al. (8) similarly reported that preformed
anti–HLA-Cw and anti–HLA-DP DSAs are as delete- Nonadherent patients are at increased risk
rious as anti–HLA-A/B/DR/DQ DSAs. for rejection as HLA mismatches increase.
Jackson et al. (9) suggested a pathogenic role for
non-HLA antiendothelial cell antibodies (AECAs) in
allograft rejection. High-density protein arrays that
identified AECA target antigens were developed, and Gene Activation
four antigenic targets expressed on endothelial cells Acute kidney rejection is a major risk factor for
were identified: endoglin; fms-like tyrosine kinase-3 chronic allograft dysfunction and long-term graft failure.
ligand; EGF-like repeats, discoidin I-like domains 3; and Ghisdal et al. (12) performed a genome-wide study to
intercellular adhesion molecule 4. These AECAs were detect loci associated with acute, biopsy-proven TCMR
detected in 24% of pretransplant sera by ELISA, and they occurring within the first year after kidney transplanta-
were associated with post-transplant donor-specific HLA tion. In a discovery cohort of 4127 European renal
antibodies, AMR, and early transplant glomerulopathy. allograft recipients, Ghisdal et al. (12) utilized a DNA
The role of the lectin pathway of complement acti- pooling approach that reduced the cost of large-scale
vation and its recognition molecules in acute rejection association studies to compare an initial cohort of cases
and outcome after transplantation was studied by and controls before utilizing an independent replication
Golshayan et al. (10). Polymorphisms and serum levels cohort of 2765 allograft recipients. Two loci were
of lectin pathway components in 710 consecutive consistently and significantly associated with acute re-
kidney transplant recipients together with all biopsy- jection in univariate and multivariate analyses. One locus
proven rejection episodes and 1-year outcomes were encompassed protein tyrosine phosphatase, receptor type
studied. Low mannose binding lectin levels were O, which encodes a receptor-type tyrosine kinase essential
associated with a higher incidence of acute cellular for B cell receptor signaling. The other locus involved
rejection during the first year, especially in recipients a ciliary gene coiled-coil domain containing 67. The clinical
of deceased donor kidneys. This association remained implications of these findings remain to be determined
significant (HR, 1.75; 95% CI, 1.18 to 2.60) in a Cox
regression model after adjustment for relevant cova- Subclinical Rejection
riates. In contrast, there were no significant associa- The use of protocol biopsies is generally restricted
tions with rates of AMR, patient death, early graft to patients undergoing experimental immunosuppressive
dysfunction, or loss. Such studies underscore the protocols, because the incidence of subclinical rejection
multiplicity of molecules, pathways, and potential in patients treated with tacrolimus, mycophenolate mofetil,
biomarkers that continue to be elucidated. However, and steroids is extremely low as discussed in a previous
these scientific observations have not yet achieved Nephrology Self-Assessment Program transplantation
clinical applicability. issue. That said, the impact of subclinical rejection in
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 309

patients taking lower-intensity immunotherapy has re- post-dnDSA (23.63 versus 22.89 ml/min per 1.73 m2
mained unclear. Loupy et al. (13) studied whether early per year; P,0.001), suggesting that dnDSA is a marker
recognition of subclinical rejection impacted long-term and contributor to ongoing alloimmunity. Time to 50%
consequences for kidney allograft survival. Participants post-dnDSA graft loss was greater in recipients with
underwent prospective screening biopsies at 1 year post- subclinical versus clinical dnDSA phenotype (8.3
transplant, with concurrent evaluations of graft comple- versus 3.3 years; P,0.001). Analysis of 1091 allograft
ment deposition and circulating anti-HLA antibodies. biopsies found that dnDSA and time independently
Three distinct groups of patients were identified at the predicted chronic glomerulopathy but not interstitial
1-year screening: 727 (73%) patients without rejec- fibrosis and tubular atrophy.
tion, 132 (13%) patients with subclinical TCMR, and The diagnosis of AMR in the absence of peri-
142 (14%) patients with subclinical AMR. Patients tubular capillary C4d staining has recently been in-
with subclinical AMR had the poorest graft survival corporated into the Banff classification system. Ono et al.
8 years post-transplant (56%) compared with subclin- (16) quantified allograft loss risk in patients with C4d2
ical TCMR (88%) and nonrejection (90%) groups AMR compared with C4d1 AMR patients and AMR-
(P,0.001). In a multivariate Cox model, subclinical free controls. C4d2 AMR patients were not different
AMR at 1 year was independently associated with a 3.5- from C4d1 AMR patients regarding baseline charac-
fold increase in graft loss (95% CI, 2.1 to 5.7) along teristics, including immunologic risk factors (panel
with eGFR and proteinuria (P,0.001). Subclinical reactive antibody, prior transplant, HLA mismatch,
AMR was associated with more rapid progression to donor type, DSA class, and anti-HLA/ABO incompat-
transplant glomerulopathy. Of patients with subclinical ibility). C4d1 AMR patients were more likely to have a
TCMR at 1 year, only those who subsequently devel- clinical presentation (85.3% versus 54.9%; P,0.001),
oped dnDSAs and transplant glomerulopathy had a and those patients presented substantially earlier post-
greater risk of graft loss compared with patients without transplantation (median, 14 days; interquartile range, 8–32
rejection. days versus median, 46; interquartile range, 20–191 days;
P,0.001) and were three times more common (7.8%
Antibody-Mediated Rejection versus 2.5%); 1- and 2-year post–AMR-defining biopsy
Schinstock et al. (14) retrospectively studied adult graft survival rates in C4d2 AMR patients were 93.4%
conventional solitary kidney transplant recipients to and 90.2%, respectively, versus 86.8% and 82.6%, re-
define histologic features associated with new-onset spectively, in C4d1 AMR patients (P50.40). C4d2 AMR
dnDSA (defined by mean fluorescent intensity [MFI] was associated with a 2.56-fold (95% CI, 1.08 to 6.05;
.1000). The incidence of dnDSA was 7.0% (54 of P50.03) increased risk of graft loss compared with AMR-
771) over a mean follow-up of 4.261.9 years. Patients free matched controls. No clinical characteristics reliably
with dnDSA had reduced death-censored allograft distinguished C4d2 from C4d1 AMR. However, both
survival (87.0% versus 97.0% no dnDSA; P,0.01). phenotypes are associated with increased graft loss.
AMR was present in 25.0% and 52.9% of patients at The impact of subclinical AMR was recently
dnDSA detection and 1 year. Patients with both classes studied by Orandi et al. (17). This group compared 219
I and II dnDSAs had the highest rates of allograft loss. patients with AMR (77 subclinical; 142 clinical) with
Antibody levels correlated with the incidence of AMR. controls matched on HLA/ABO compatibility, donor
The authors concluded that patients with dnDSA type, prior transplant, panel reactive antibody, and age;
without AMR at time of detection may benefit from 1- and 5-year graft survival rates in subclinical AMR
a follow-up biopsy within 1 year, because AMR can be were 95.9% and 75.7%, respectively, compared with
missed initially. Similarly, Wiebe et al. (15) studied 96.8% and 88.4%, respectively, in matched controls
a consecutive cohort of 508 renal transplant recipients, (P50.01). Subclinical AMR was independently asso-
of whom 64 developed dnDSA. Recipients without ciated with a 2.15-fold increased risk of graft loss (95%
dnDSA or allograft dysfunction had an eGFR decline CI, 1.19 to 3.91; P50.01) compared with matched
of 20.65 ml/min per 1.73 m2 per year. In recipients controls but was not different from clinical AMR
with dnDSA, the rate of eGFR decline was significantly (defined as AKI with DSA demonstration and patho-
increased before dnDSA onset (22.89 versus 20.65 logic evidence of tissue injury and C4d deposition;
ml/min per 1.73 m2 per year; P,0.001) and accelerated P50.13). Most subclinical AMR patients were treated
310 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

with plasmapheresis within 3 days of their AMR- iDSA with C1q positivity. The distributions of iDSA
defining biopsy. The impact of AMR on graft loss was IgG1, -2, -3, and -4 subclasses among the population
heterogeneous when stratified by compatible deceased were 75.2%, 44.0%, 28.0%, and 26.4%, respectively.
donor (HR, 4.73; 95% CI, 1.57 to 14.26; P,0.01), AMR was mainly driven by IgG3 iDSA, whereas sub-
HLA-incompatible deceased donor (HR, 2.39; 95% CI, clinical AMR was driven by IgG4 iDSA. IgG3 iDSA was
1.10 to 5.19; P50.03), compatible live donor (no AMR associated with a shorter time to rejection (P,0.001),
patients experienced graft loss), ABO-incompatible live increased microcirculation injury (P50.002), and C4d
donor (HR, 6.13; 95% CI, 0.55 to 67.70; P50.14), and capillary deposition (P,0.001). IgG4 iDSA was associ-
HLA-incompatible live donor (HR, 6.29; 95% CI, 3.81 ated with later allograft injury, with increased allograft
to 10.39; P,0.001) transplant. The authors concluded glomerulopathy and interstitial fibrosis/tubular atrophy
that subclinical AMR substantially increases graft loss, lesions (P,0.001 for all comparisons). IgG3 iDSA and
and treatment seems warranted. Underscoring the re- C1q binding iDSA were strongly and independently
fractory nature of some AMR cases, the Johns Hopkins associated with allograft failure.
investigators reported the utility of splenic irradiation for The C1q complex is a complement protein in-
refractory AMR in two patients. The generalizability of volved in the innate immune system. C1q binds to IgM
this approach remains unknown (18). and IgG/antigen complexes, leading to activation of the
The impact of microvascular inflammation and classical complement pathway. Calp-Inal et al. (22)
extent of PTC has also been recently reported. Gupta reported that the incidence of acute AMR was higher
et al. (19) classified kidney allograft biopsies into in C1q1/DSA1 patients compared with C1q2/DSA1
groups based on microvascular inflammation scores and Cq12/DSA2 patients. In two different temporal
of zero, one, two, or more. Gene expression profiles were cohorts, the frequency of chronic AMR ranged from 36%
also assessed. The incidence of donor-specific anti-HLA to 51% in C1q1/DSA1 patients. In contrast, chronic
antibodies increased from 25% in group 1 to 36% in AMR ranged from 5% to 25% in C1q2/DSA1 patients
group 2 and 54% in group 3. Acute and chronic AMRs and from 2% to 6% of DSA2 patients (P,0.001). The
were significantly more frequent in group 3 (15% and authors concluded that the C1q2/DSA1 phenotype was
35%, respectively) compared with group 2 (3% and 15%, associated with acute and chronic AMR.
respectively) and group 1 (0% and 5%, respectively).
Gene expression profiling revealed more immune activa- References
tion in the third group as well. Kozakowski et al. (20) 1. Hart A, Smith JM, Skeans MA, Gustafson SK, Stewart DE, Cherikh
studied the degree of PTC with allograft loss rates. They WS, Wainright JL, Boyle G, Snyder JJ, Kasiske BL, Israni AK: Kidney.
Am J Transplant 16[Suppl 2]: 11–46, 2016 PubMed
found that a score of three (HR, 2.57; 95% CI, 1.25 to 2. Loupy A, Haas M, Solez K, Racusen L, Glotz D, Seron D, Nankivell
5.28) and diffuse PTC (HR, 1.67; 95% CI, 1.1 to 2.54) BJ, Colvin RB, Afrouzian M, Akalin E, Alachkar N, Bagnasco S,
were significant impartial risk factors for allograft loss. Becker JU, Cornell L, Drachenberg C, Dragun D, de Kort H, Gibson
IW, Kraus ES, Lefaucheur C, Legendre C, Liapis H, Muthukumar T,
The development of donor specific antibody and antibody Nickeleit V, Orandi B, Park W, Rabant M, Randhawa P, Reed EF,
mediated rejection, whether clinical or subclinical, im- Roufosse C, Seshan SV, Sis B, Singh HK, Schinstock C, Tambur A,
Zeevi A, Mengel M: The Banff 2015 Kidney Meeting Report: Current
pacts graft survival adversely.
challenges in rejection classification and prospects for adopting molec-
ular pathology. Am J Transplant 17: 28–41, 2017 PubMed
Immunoglobulin Subtypes 3. Becker JU, Chang A, Nickeleit V, Randhawa P, Roufosse C: Banff
borderline changes suspicious for acute T cell-mediated rejection:
Antibodies may have different pathogenicities Where do we stand? Am J Transplant 16: 2654–2660, 2016 PubMed
per IgG subclass. Lefaucheur et al. (21) investigated 4. Halloran PF, Famulski KS, Chang J: A probabilistic approach to
the association between anti-human HLA antibodies histologic diagnosis of antibody-mediated rejection in kidney transplant
biopsies. Am J Transplant 17: 129–139, 2017 PubMed
IgG subclasses and rejection. This group studied 125 of 5. Halloran PF, Chang J, Famulski K, Hidalgo LG, Salazar ID, Merino
635 patients with donor-specific anti-human HLA anti- Lopez M, Matas A, Picton M, de Freitas D, Bromberg J, Serón D,
bodies (DSAs) detected in the first year post-transplant; Sellarés J, Einecke G, Reeve J: Disappearance of T cell-mediated
rejection despite continued antibody-mediated rejection in late kidney
51 (40.8%) patients had acute AMR, 36 (28.8%) pa- transplant recipients. J Am Soc Nephrol 26: 1711–1720, 2015 PubMed
tients had subclinical AMR, and 38 (30.4%) patients 6. Ferrandiz I, Congy-Jolivet N, Del Bello A, Debiol B, Trébern-Launay K,
were AMR free. The MFI of the immunodominant Esposito L, Milongo D, Dörr G, Rostaing L, Kamar N: Impact of early
blood transfusion after kidney transplantation on the incidence of
donor-specific antibody (iDSA; the DSA with the highest donor-specific anti-HLA antibodies. Am J Transplant 16: 2661–2669,
MFI level) was 67246464, and 41.6% of patients had 2016 PubMed
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 311

7. Lim WH, Chapman JR, Coates PT, Lewis JR, Russ GR, Watson N, donor-specific anti-human HLA antibody subclasses and kidney allograft
Holdsworth R, Wong G: HLA-DQ mismatches and rejection in antibody-mediated injury. J Am Soc Nephrol 27: 293–304, 2016 PubMed
kidney transplant recipients. Clin J Am Soc Nephrol 11: 875–883, 22. Calp-Inal S, Ajaimy M, Melamed ML, Savchik C, Masiakos P, Colovai
2016 PubMed A, Akalin E: The prevalence and clinical significance of C1q-binding
8. Bachelet T, Martinez C, Del Bello A, Couzi L, Kejji S, Guidicelli G, donor-specific anti-HLA antibodies early and late after kidney trans-
Lepreux S, Visentin J, Congy-Jolivet N, Rostaing L, Taupin JL, Kamar plantation. Kidney Int 89: 209–216, 2016 PubMed
N, Merville P: Deleterious impact of donor-specific anti-HLA anti-
bodies toward HLA-Cw and HLA-DP in kidney transplantation.
Transplantation 100: 159–166, 2016 PubMed
9. Jackson AM, Sigdel TK, Delville M, Hsieh SC, Dai H, Bagnasco S,
Montgomery RA, Sarwal MM: Endothelial cell antibodies associated
with novel targets and increased rejection. J Am Soc Nephrol 26: 1161–
Histoincompatibility Strategies
1171, 2015 PubMed
10. Golshayan D, Wójtowicz A, Bibert S, Pyndiah N, Manuel O, Binet I, Buhler
Approximately 30% of medically suitable living
LH, Huynh-Do U, Mueller T, Steiger J, Pascual M, Meylan P, Bochud PY; donor candidates are incompatible with their intended
Swiss Transplant Cohort Study: Polymorphisms in the lectin pathway of recipients due to preformed ABO or HLA antibodies.
complement activation influence the incidence of acute rejection and graft
outcome after kidney transplantation. Kidney Int 89: 927–938, 2016 PubMed Historically, such candidate recipients had little option
11. Wiebe C, Nevins TE, Robiner WN, Thomas W, Matas AJ, Nickerson other than waiting for a suitable deceased donor offer.
PW: The synergistic effect of class II HLA epitope-mismatch and Consequentially, those with highest levels of HLA anti-
nonadherence on acute rejection and graft survival. Am J Transplant 15:
2197–2202, 2015 PubMed bodies often incurred long waiting times with reduced
12. Ghisdal L, Baron C, Lebranchu Y, Viklický O, Konarikova A, Naesens access to transplantation. To alleviate such problems, three
M, Kuypers D, Dinic M, Alamartine E, Touchard G, Antoine T, Essig strategies have emerged:
M, Rerolle JP, Merville P, Taupin JL, Le Meur Y, Grall-Jezequel A,
Glowacki F, Noël C, Legendre C, Anglicheau D, Broeders N, Coppieters
W, Docampo E, Georges M, Ajarchouh Z, Massart A, Racapé J,
• Change in kidney allocation to prioritize highly
Abramowicz D, Abramowicz M: Genome-wide association study of acute sensitized patients,
renal graft rejection. Am J Transplant 17: 201–209, 2017 PubMed • Kidney paired donation (KPD), and
13. Loupy A, Vernerey D, Tinel C, Aubert O, Duong van Huyen JP, Rabant
M, Verine J, Nochy D, Empana JP, Martinez F, Glotz D, Jouven X,
• Desensitization to either ABO or HLA antibodies.
Legendre C, Lefaucheur C: Subclinical rejection phenotypes at 1 year
post-transplant and outcome of kidney allografts. J Am Soc Nephrol 26:
1721–1731, 2015 PubMed These approaches are not necessarily mutually
14. Schinstock CA, Cosio F, Cheungpasitporn W, Dadhania DM, Everly MJ, exclusive. For candidates with an incompatible donor,
Samaniego-Picota MD, Cornell L, Stegall MD: The value of protocol KPD can improve the prospects of finding a compatible
biopsies to identify patients with de novo donor-specific antibody at high
risk for allograft loss. Am J Transplant 17: 1574–1584, 2017 PubMed living donor, but for many highly sensitized patients,
15. Wiebe C, Gibson IW, Blydt-Hansen TD, Pochinco D, Birk PE, Ho J, the probability of finding a match in the relatively
Karpinski M, Goldberg A, Storsley L, Rush DN, Nickerson PW: Rates small pools of donors in such programs remains
and determinants of progression to graft failure in kidney allograft
recipients with de novo donor-specific antibody. Am J Transplant 15: limited. Keith and Vranic (1) suggest that desensitiza-
2921–2930, 2015 PubMed tion of a living donor/recipient pair with low levels of
16. Ono E, Dos Santos AM, Viana PO, Dinelli MI, Sass N, De Oliveira L, incompatibility is another reasonable approach. Note
Goulart AL, de Moraes-Pinto MI: Immunophenotypic profile and increased
risk of hospital admission for infection in infants born to female kidney that desensitization has never been a large-volume
transplant recipients. Am J Transplant 15: 1654–1665, 2015 PubMed endeavor and may decline due to the prior changes in
17. Orandi BJ, Chow EH, Hsu A, Gupta N, Van Arendonk KJ, Garonzik-
Wang JM, Montgomery JR, Wickliffe C, Lonze BE, Bagnasco SM,
deceased donor allocation and KPD.
Alachkar N, Kraus ES, Jackson AM, Montgomery RA, Segev DL: An unexpectedly positive impact of the new
Quantifying renal allograft loss following early antibody-mediated Kidney Allocation System (KAS), implemented De-
rejection. Am J Transplant 15: 489–498, 2015 PubMed
18. Orandi BJ, Lonze BE, Jackson A, Terezakis S, Kraus ES, Alachkar N,
cember 4, 2014, is the degree of increase in allograft
Bagnasco SM, Segev DL, Orens JB, Montgomery RA: Splenic recipients with the highest calculated panel reactive
irradiation for the treatment of severe antibody-mediated rejection. antibody (cPRA) levels (2). Historically, such candi-
Am J Transplant 16: 3041–3045, 2016 PubMed
19. Gupta A, Broin PO, Bao Y, Pullman J, Kamal L, Ajaimy M, Lubetzky
dates have had extremely low transplant rates and are
M, Colovai A, Schwartz D, de Boccardo G, Golden A, Akalin E: consistently under-represented among transplant recip-
Clinical and molecular significance of microvascular inflammation in ients relative to their prevalence on the waiting list.
transplant kidney biopsies. Kidney Int 89: 217–225, 2016 PubMed
20. Kozakowski N, Herkner H, Böhmig GA, Regele H, Kornauth C, Bond Priority point allocations for cPRAs changed signifi-
G, Kikić Ž: The diffuse extent of peritubular capillaritis in renal allograft cantly from a flat four-point allocation for cPRAs#80%
rejection is an independent risk factor for graft loss. Kidney Int 88: 332– to a sliding scale that progressively increases priority
340, 2015 PubMed
21. Lefaucheur C, Viglietti D, Bentlejewski C, Duong van Huyen JP, Vernerey D, with cPRA elevations. Transplants among patients with
Aubert O, Verine J, Jouven X, Legendre C, Glotz D, Loupy A, Zeevi A: IgG cPRAs of 99%–100% increased from 2.2% pre-KAS,
312 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

peaked at 13.3% in the first quarter of 2015, and tapered those who received RTX or underwent splenectomy,
off to 10.1% by the end of 2015. In summary, these data the overall graft survival rates were 94.5% (95% CI,
represent a 366% increase, slightly above these candi- 91.6% to 96.5%) and 79.7% (95% CI, 72.9% to
dates’ waitlist prevalence of 8.1% (Figure 12). Trans- 85.1%), respectively. Data on other longer-term out-
plants among candidates with cPRAs of 98% also comes, including malignancy, were sparse. The authors
doubled from 1.2% pre-KAS to 2.4% post-KAS, with concluded that RTX and IA were potentially promising
no appreciable bolus effect. The greater than anticipated preconditioning strategies before ABOi kidney trans-
effect of the KAS on high-cPRA transplants may be plantation. However, the overall quality of evidence
explained by an underestimate of the acceptance of high- and the confidence in the observed treatment effects are
quality regional and national offers for these highly low. Note that IA is not readily available in the United
sensitized candidates. In some regions, the proportion States, although it has been used successfully in Europe
of such patients awaiting transplantation has declined (see below).
by .50%. Okumi et al. (4) studied changes in outcomes of
ABOi transplants during the last 25 years. Patients were
ABO-Incompatible Transplantation divided into four groups by transplantation era and
ABO-incompatible (ABOi) living kidney trans- ABO compatibility. In the past decade, ABOi live
plantation is a low-volume endeavor internationally donor transplant and ABO-compatible (ABOc) recipi-
that includes a preconditioning regimen that variously ents yielded almost equivalent outcomes with respect to
involves combinations of plasmapheresis, immunoad- 9-year graft survival rates: 86.9% and 92.0%, respec-
sorption (IA), intravenous Ig (IVIg), rituximab (RTX), tively (hazard ratio [HR], 1.38; 95% CI, 0.59 to 3.22;
conventional immunosuppression, and less commonly, P50.46). The graft survival rate for ABO-incompatible
splenectomy. Lo et al. (3) performed a meta-analysis of living kidney transplants (ABO-ILKTs) conducted
83 published reports involving 4810 ABOi transplant between 2005 and 2013 was superior compared with
recipients. During a mean follow-up time of 28 months, that of ABO-ILKTs conducted between 1998 and 2004
the overall graft survival rates for recipients who (HR, 0.30; 95% CI, 0.13 to 0.72; P,0.01). ABO-ILKT
received either IA or apheresis were 94.1% (95% recipients showed substantial improvements in graft
confidence interval [95% CI], 88.2% to 97.1%) and survival rate over time. Graft survival was almost
88.0% (95% CI, 82.6% to 91.8%), respectively. For identical over the past decade, regardless of ABO

Figure 12. There has been a marked increase in the number of deceased donor transplant recipients in the most highly
allosensitized recipients (defined as cPRA.98%) since implementation of KAS 12/4/14. There was an initial bolus effect that is
now declining. This has been at the expense of a small reduction in proportion of kidneys allocated to non sensitized individuals.
Distribution of kidney transplants by cPRA. Kidney-alone transplant recipients from January 1, 2013 to December 31, 2015.
KAS was implemented December 4, 2014. OPTN, Organ Procurement and Transplantation Network; SRTR, Scientific Registry
of Transplant Recipients. Reprinted with permission from Hart A, Gustafson SK, Skeans MA, Stock P, Stewart D, Kasiske BL,
Israni AK: OPTN/SRTR 2015 Annual Data Report: Early effects of the new kidney allocation system. Am J Transplant 17
[Suppl 1]: 543–564, 2017.
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 313

compatibility. Similarly, Zschiedrich et al. (5) reported The rates of overall infectious complications among the
excellent graft survival rates equivalent to ABOc three groups were comparable. However, serious infections
kidney transplantation outcome in Germany. developed in 13 of the ABOc kidney transplants
(11.0%), 20 from the standard RTX group (26.3%),
and two from the reduced RTX group (10.5%;
Patient and graft survival rates for ABO P50.02). Standard dose RTX was found to be an
compatible and preconditioned ABO in- independent risk factor for serious infections (HR,
compatible transplantation are generally 2.59; 95% CI, 1.33 to 5.07; P,0.01). There were no
equivalent. significant differences in rejection, renal function, graft
survival, and patient survival between standard and
reduced RTX groups.
It has been suggested that the infrequent use of IA was previously reviewed in prior editions of
ABOi kidney transplantation in the United States may the Nephrology Self-Assessment Program on trans-
reflect concern about the costs of necessary precondi- plantation and continues to be reported outside of the
tioning and post-transplant care (6). Axelrod et al. (6) United States. Becker et al. (8) compared the clinical
studied Medicare data for 26,500 live donor kidney outcomes of 34 ABOi living donor kidney recipients
transplant recipients (2000 to March of 2011), in- who were transplanted using this protocol with those of
cluding 271 ABOi and 62 A2-incompatible (A2i) re- 68 matched ABOc patients. Before desensitization, the
cipients, and these data were analyzed to assess the median titer of 34 ABOi patients was 1:64. Patients
impact of pretransplant, peritransplant, and 3-year post- received a median of seven preoperative IA treatments.
transplant costs. Compared with ABOc transplantation, Twenty-four patients had a median of two additional
3-year rates of patient (93.2% versus 88.15%; P,0.001) plasmapheresis treatments to reach the preoperative
and death-censored graft survival (85.4% versus 76.1%; target isoagglutinin titer of 1:8 or less. After a median
P,0.05) were lower after ABOi transplant. Excluding postoperative follow-up of 22 months, overall graft
the cost of organ acquisition, the average overall cost of survival in the ABOi group was not significantly different
the transplant episode was significantly higher for ABOi from that in ABOc patients (log rank P50.20), whereas
($65,080) compared with blood type A2i ($36,752) and patient survival tended to be lower (log rank P50.05).
ABOc ($32,039) transplantation (P,0.001); ABOi trans- The incidence of rejection episodes was 15% in both
plant was associated with high adjusted post-transplant groups. The ABOi recipients had a higher incidence of
spending (marginal costs compared with ABOc: year 1, BK virus replication (P50.04) and nephropathy (P50.01)
$25,044; year 2, $10,496; year 3, $7307; P,0.01). The and showed colonization with multidrug-resistant bacte-
authors concluded that ABOi transplantation provides ria (P50.02) more frequently. In comparison with blood
a clinically effective method to expand access to trans- group antigen–specific IA, nonantigen-specific IA showed
plantation, albeit with the risks of greater cost and inferior, equal efficacy but was associated with cost reduction.
if acceptable, outcomes. The histologic findings and incidence of antibody-
RTX, an mAb that targets B cell CD20, induces mediated rejection (AMR) after ABOi preconditioning
lymphodepletion and is frequently a part of the ABOi remain unclear. Masutani et al. (9) reviewed the histo-
conditioning regimen. Lee et al. (7) provide a note of logic findings at 3 and 12 months after ABOc and ABOi
caution regarding the safety of this approach after transplantation. Subclinical acute rejection occurred in
analyzing 213 recipients (n5118 ABOc kidney trans- 6.9% and 9.9% of patients in the ABOc and ABOi
plants and n595 ABOi kidney transplants) who un- groups, respectively, at 3 months (P50.40) and 12.4%
derwent living donor transplantation. The effect of and 10.1%, respectively, at 12 months (P50.50). The
RTX dose on infectious complications in ABOi kidney cumulative incidence rates of acute rejection were
transplantation was studied; ABOi kidney transplant 20.5% and 19.6%, respectively (P50.80). The degrees
patients were categorized by RTX dose: standard dose of microvascular inflammation (MVI) and interstitial
(n576; 375 mg/m2) versus reduced RTX dose (n519; fibrosis/tubular atrophy were comparable. Polyoma-
200 mg). All participants received basiliximab induc- virus BK nephropathy was found in 2.7% and 3.0%
tion for nondepleting IL-2 receptor blockade and were of patients in the ABOc and ABOi groups, respectively
maintained on conventional triple immunosuppression. (P.0.99). The incidence of other infections and the
314 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

graft/patient survival rates were not different. It has AMR episodes were treated with IVIg/RTX
been reported that C4d deposition is common after (n525) or in more severe AMR, the addition of plasma
ABOi transplantation, although the prognostic signifi- exchange (n520). Graft survival for patients treated
cance remains unclear. Ishihara et al. (10) investigated with IVIg plus RTX was superior (P50.03). Increased
MVI scores greater than or equal to two within 1 year mortality was seen in AMR patients experiencing graft loss
after ABOi transplantation as a predictor of graft out- after AMR treatment (P50.004). Plasmapheresis and
come. They determined that 5-year graft survival was eculizumab improved graft survival for TMA patients
significantly lower (P50.01) in MVI patients (89.8%) (P50.04). The authors concluded that patients desensitized
than in non-MVI patients. Graft function, characterized with IVIg plus RTX who remained AMR free experienced
by serum eGFR, was also significantly worse for patients long-term graft and patient survival. However, AMR
with MVI than it was for patients without MVI from 3 patients had significantly reduced graft survival and GFRs
months to 10 years post-transplantation (P50.05). at 5 years, especially after development of TMA.
Multivariate analysis indicated that HLA class II mis- Jackson et al. (13) studied 256 post-transplant
match (P,0.01) was an independent marker of MVI. HLA antibody levels in 25 recipients desensitized with
The authors concluded that MVI score greater than or and 25 without RTX induction to determine the impact
equal to two is significantly associated with poor graft of B cell depletion. They found significantly less HLA
outcome after ABOi kidney transplantation. antibody rebound in the RTX-treated patients (7% of
DSAs and 33% of non-DSAs) compared with in
HLA Antibody Desensitization a control cohort desensitized and transplanted without
Preconditioning regimens to permit HLA-incompatible RTX (32% DSAs and 55% non-DSAs). Importantly,
transplantation generally follow protocols similar in prin- although RTX induction in HLA-incompatible recipi-
ciple to those for ABOi regimens. However, these pro- ents reduced the incidence and magnitude of HLA
tocols are associated with inferior outcomes; even intense antibody rebound, it did not affect DSA elimination,
immunomodulation may not completely abrogate HLA AMR, or 5-year allograft survival compared with
AMR. Schwaiger et al. (11) evaluated transplant outcomes recipients desensitized and transplanted without RTX.
in 101 donor-specific antibody–positive (DSA1) deceased
donor kidney transplant recipients with a median follow-
up of 24 months who were subjected to IA-based de- Desensitization to permit HLA-incompatible
sensitization. Treatment included a single pretransplant IA transplantation continues to be associated
session followed by antilymphocyte antibody and serial with antibody-mediated rejection and infe-
post-transplant IA. Three-year death-censored graft sur- rior outcomes compared with either HLA-
compatible or ABO incompatible transplants.
vival in DSA1 patients was significantly worse than that
in 513 DSA2 recipients transplanted during the same
period (79% versus 88%; P,0.01). Thirty-three DSA1
Novel Desensitization Strategies
recipients (33%) had AMR.
The limitations of existing desensitization strate-
Vo et al. (12) described their updated results of
gies continue to drive the search for alternate options.
IVIg and RTX in 226 highly sensitized patients who
Bortezomib (BTZ), a proteasome inhibitor that causes
received transplants after desensitization. Most received
plasma cell depletion, has been studied in this regard.
alemtuzumab induction and standard immunosuppression.
Moreno Gonzalez et al. (14) studied the safety and
Significant risks for AMR included previous transplants efficacy of 32 doses of BTZ (1.3 mg/m body surface
and pregnancies as sensitizing events, DSA relative area) in ten highly sensitized kidney transplant candi-
intensity scores, presence of both classes I and II DSAs dates with alloantibodies against their intended living
at transplant, and time on the waitlist. Banff pathology donor. Dose reduction was needed in two patients, and
characteristics for AMR patients with or without graft two others completely discontinued therapy for adverse
loss did not differ. C4d1 versus C4d2 AMR did not events. Anti-HLA antibody mean fluorescence inten-
predict graft loss (P50.09); however, thrombotic micro- sity (MFI) values were stable before BTZ administra-
angiopathy (TMA) significantly predicted graft failure tion (P50.96) but decreased after therapy (mean
(P50.05). decrease, 1916; SEM5425 MFI; P,0.01). No patients
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 315

developed a negative crossmatch against their original Evolution of outcomes and immunosuppressive management. Am J
Transplant 16: 886–896, 2016 PubMed
intended donor, and the cPRAs based on MFIs of 2000, 5. Zschiedrich S, Kramer-Zucker A, Jänigen B, Seidl M, Emmerich F,
4000, and 8000 were unchanged in all patients. The Pisarski P, Huber TB: An update on ABO-incompatible kidney trans-
authors concluded that 32 doses of BTZ monotherapy plantation. Transpl Int 28: 387–397, 2015 PubMed
6. Axelrod D, Segev DL, Xiao H, Schnitzler MA, Brennan DC, Dharnid-
were not well tolerated and resulted in only a modest harka VR, Orandi BJ, Naik AS, Randall H, Tuttle-Newhall JE, Lentine
reduction in anti-HLA antibodies. By way of contrast, KL: Economic impacts of ABO-incompatible live donor kidney trans-
Woodle et al. (15) performed a prospective iterative plantation: A national study of medicare-insured recipients. Am J
Transplant 16: 1465–1473, 2016 PubMed
trial of proteasome inhibitor–based therapy for reducing 7. Lee J, Lee JG, Kim S, Song SH, Kim BS, Kim HO, Kim MS, Kim SI,
HLA antibody. Nineteen of 44 patients (43.2%) were Kim YS, Huh KH: The effect of rituximab dose on infectious
transplanted with low acute rejection rates (18.8%) and complications in ABO-incompatible kidney transplantation. Nephrol
Dial Transplant 31: 1013–1021, 2016 PubMed
de novo DSA formation (12.5%). Although the authors 8. Becker LE, Siebert D, Süsal C, Opelz G, Leo A, Waldherr R, Macher-
concluded that proteasome inhibition–based desensiti- Goeppinger S, Schemmer P, Schaefer SM, Klein K, Beimler J, Zeier M,
Schwenger V, Morath C: Outcomes following ABO-incompatible kidney
zation consistently and durably reduced HLA antibody
transplantation performed after desensitization by nonantigen-specific
levels and provided an alternative to IVIg-based de- immunoadsorption. Transplantation 99: 2364–2371, 2015 PubMed
sensitization, most patients were still not transplanted. 9. Masutani K, Tsuchimoto A, Kurihara K, Okabe Y, Kitada H, Okumi M,
Tanabe K, Nakamura M, Kitazono T, Tsuruya K; Japan Academic
IL-6 may also be an attractive target, because it Consortium of Kidney Transplantation (JACK) investigators: Histolog-
promotes B cell differentiation to plasma cells, is im- ical analysis in ABO-compatible and ABO-incompatible kidney trans-
portant for Ig production, and induces Th17 cells, a plantation by performance of 3- and 12-month protocol biopsies.
Transplantation 101: 1416–1422, 2017 PubMed
subset of proinflammatory T helper cells defined by 10. Ishihara H, Ishida H, Unagami K, Hirai T, Okumi M, Omoto K,
their production of IL-17 that is related to T regulatory Shimizu T, Tanabe K: Evaluation of microvascular inflammation in
cells. Tocilizumab (TCZ) is an anti–IL-6 receptor ABO-incompatible kidney transplantation. Transplantation 101: 1423–
1432, 2017 PubMed
antibody undergoing investigation as a potential treat- 11. Schwaiger E, Eskandary F, Kozakowski N, Bond G, Kikić Ž, Yoo D,
ment for AMR. Vo et al. (16) performed a phase I/II Rasoul-Rockenschaub S, Oberbauer R, Böhmig GA: Deceased donor
pilot study using TCZ with IVIg to assess safety and kidney transplantation across donor-specific antibody barriers: Predic-
tors of antibody-mediated rejection. Nephrol Dial Transplant 31: 1342–
limited efficacy for clinical AMR. Ten patients, un- 1351, 2016 PubMed
responsive to desensitization with IVIg 1 RTX, were 12. Vo AA, Sinha A, Haas M, Choi J, Mirocha J, Kahwaji J, Peng A,
treated with IVIg 1 TCZ. No differences in baseline Villicana R, Jordan SC: Factors predicting risk for antibody-mediated
rejection and graft loss in highly human leukocyte antigen sensitized
characteristics were seen in patients not transplanted patients transplanted after desensitization. Transplantation 99: 1423–
versus transplanted. Two patients in each group de- 1430, 2015 PubMed
veloped serious adverse events: not transplanted, pul- 13. Jackson AM, Kraus ES, Orandi BJ, Segev DL, Montgomery RA,
Zachary AA: A closer look at rituximab induction on HLA antibody
monary congestion with status epilepticus; transplanted, rebound following HLA-incompatible kidney transplantation. Kidney
infectious colitis with colonic perforation and Bell palsy. Int 87: 409–416, 2015 PubMed
14. Moreno Gonzales MA, Gandhi MJ, Schinstock CA, Moore NA, Smith
Five of ten patients were transplanted. Six-month pro-
BH, Braaten NY, Stegall MD: 32 Doses of bortezomib for desensitization
tocol biopsies showed no AMR. DSA strength and is not well tolerated and is associated with only modest reductions in anti-
number were reduced by TCZ treatment. Renal function HLA antibody. Transplantation 101: 1222–1227, 2017 PubMed
15. Woodle ES, Shields AR, Ejaz NS, Sadaka B, Girnita A, Walsh RC,
at 12 months was 60625 ml/min. The authors con- Alloway RR, Brailey P, Cardi MA, Abu Jawdeh BG, Roy-Chaudhury P,
cluded that TCZ and IVIg seem safe; however, larger Govil A, Mogilishetty G: Prospective iterative trial of proteasome inhibitor-
controlled studies are required. based desensitization. Am J Transplant 15: 101–118, 2015 PubMed
16. Vo AA, Choi J, Kim I, Louie S, Cisneros K, Kahwaji J, Toyoda M, Ge
S, Haas M, Puliyanda D, Reinsmoen N, Peng A, Villicana R, Jordan SC:
References A phase I/II trial of the interleukin-6 receptor-specific humanized
1. Keith DS, Vranic GM: Approach to the highly sensitized kidney trans- monoclonal (tocilizumab) 1 intravenous immunoglobulin in difficult
plant candidate. Clin J Am Soc Nephrol 11: 684–693, 2016 PubMed to desensitize patients. Transplantation 99: 2356–2363, 2015 PubMed
2. Hart A, Gustafson SK, Skeans MA, Stock P, Stewart D, Kasiske BL,
Israni AK: OPTN/SRTR 2015 Annual Data Report: Early effects of the
new kidney allocation system. Am J Transplant 17[Suppl 1]: 543–564,
2017 PubMed
3. Lo P, Sharma A, Craig JC, Wyburn K, Lim W, Chapman JR, Palmer SC, Clinical Tolerance
Strippoli GF, Wong G: Preconditioning therapy in ABO-incompatible
living kidney transplantation: A systematic review and meta-analysis. The overwhelming majority of conventionally
Transplantation 100: 933–942, 2016 PubMed
4. Okumi M, Toki D, Nozaki T, Shimizu T, Shirakawa H, Omoto K, Inui immunosuppressed allograft recipients who stop therapy
M, Ishida H, Tanabe K: ABO-incompatible living kidney transplants: develop symptomatic rejection and premature allograft
316 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

failure. However, the occasional fortunate patient who in tolerant patients. Taken together, the data show that
conducts this experiment fails to undergo rejection, operationally tolerant patients mobilize an array of
having developed operational tolerance. Furthermore, it potentially suppressive cells, including regulatory B
is possible that there are more such patients who may cells and Tregs.
forego immunosuppression (IS) who never take the risk
with ongoing implications for cost and adverse effects. Clinical Tolerance Induction
The reader is reminded that the clinical trials of calcineurin It has been lightheartedly suggested that clinical
inhibitor withdrawal in low-immunologic risk allograft tolerance is the “holy grail” of transplantation and
recipients discussed above revealed significant rejection always will be. This viewpoint has been challenged by
risk, underscoring the importance of ongoing adherence reports from three centers in the United States with
for most patients. For all of these reasons, the cellular and proof of concept data that are described in greater detail
immunologic events that underlie tolerance events are of below. Note that tolerance induction is deemed exper-
interest, because prospective identification could poten- imental and has met with varying degrees of success.
tially identify that subset of individuals who may undergo Rejections and graft losses, although infrequent, have
immunotherapy reduction or withdrawal. occurred using regimens that generally rely on in-
Baron et al. (1) collected samples and clinical duction of either transient or persistent donor chime-
data from 96 tolerant patients from five different centers rism. This comment is made to refocus the readers’
with specific reference to measurable transcription attention on the experimental nature of these regimens
patterns. Common gene signatures and discriminating in evolution, not to undermine the importance of such
biomarkers for tolerance after kidney transplantation work. The three protocols below share common themes
were identified. A robust gene signature involving that include the use of lymphocyte depletion, cellular
proliferation of B and CD41 T cells with inhibition therapy with either donor bone marrow or CD34-
of CD14 monocyte functions was confirmed and cross- positive stem cells, and various combinations of con-
validated with almost 92% accuracy. The data also ventional immunotherapy that are subsequently tapered
indicated participation of other cell subsets in tolerance. off with time. A more recent clinical trial planned to
The authors concluded that the use of the top 20 evaluate administration of Tregs to harness the immune
biomarkers, mostly centered on B cells, may provide response has recently been initiated in eight European
a standardized tool toward personalized medicine for and American centers. (4).
monitoring of tolerant or low-risk patients among Leventhal and coworkers (5–7) described results
kidney allotransplant recipients. Chesneau et al. (2) of a nonchimeric, operational tolerance protocol in
analyzed the role of B cells from 12 operationally HLA-identical living donor renal transplants. Recipi-
tolerant patients compared with healthy volunteers and ents given alemtuzumab and tacrolimus/mycopenolic
stable kidney transplant recipients on IS. Proliferation, acid with early sirolimus conversion were repeatedly
apoptosis, and type I proinflammatory cytokine produc- infused with donor hematopoietic CD341 stem cells.
tion by effector CD41CD25– T cells were measured. IS was withdrawn by 24 months. Twelve months later,
They reported that B cells inhibit the CD41CD25– operational tolerance was confirmed by rejection-free
effector T cell response in a dose-dependent manner. transplant biopsies. Five of the first eight enrollees were
This effect required B cells to interact with T cell targets initially tolerant 1 year off IS. Biopsies of three others
and was achieved through a granzyme B–dependent after total withdrawal showed Banff 1A acute cellular
pathway. Tolerant recipients harbored a higher number rejection without allograft dysfunction. At 5 years post-
of B cells expressing granzyme B and displayed a partic- transplantation, four of five tolerant recipients remained
ular plasma cell phenotype. The investigators concluded rejection free, whereas one developed Banff 1A re-
that these data provide insights into the characterization of jection without allograft dysfunction. Significant time-
B cell–mediated immunoregulation in clinical tolerance. dependent increases of circulating Tregs were seen in
The role of Foxp31 regulatory T cells (Tregs) in a more recent cohort of tolerant versus nontolerant
operational tolerance was studied by Braza et al. (3). subjects (P,0.001). Gene expression signatures, de-
They observed a higher proportion of CD41 T cells veloped using global RNA expression, were highly
with demethylated Foxp3 and a specific expansion of associated with operational tolerance as early as 1 year
CD41CD45RA– Foxp3(hi) memory Tregs exclusively post-transplant.
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 317

Kawai et al. (8) previously reported the long-term 12 months by increasing the dose of T cells and CD341
results of HLA-mismatched kidney transplantation cells infused compared with matched recipients in a dose
without maintenance IS in ten subjects after combined escalation study. Success of drug withdrawal in chimeric-
kidney and bone marrow transplantation. All subjects mismatched patients remains to be determined. None
were treated with nonmyeloablative conditioning and of the 38 patients had kidney graft loss or graft versus
an 8- to 14-month course of calcineurin inhibitor host disease in up to 14 years of observation. The au-
with or without rituximab. All ten subjects developed thors concluded that complete immunosuppressive
transient chimerism, and in seven of these, IS was drug withdrawal could be achieved with the tolerance
successfully discontinued for 4 or more years. At the induction regimen in HLA-matched patients with uni-
time of publication, four subjects remained IS free for form, long-term graft survival in all patients. In a sub-
periods of 4.5–11.4 years, whereas three required sequent publication, Scandling and coworkers (11)
reinstitution of IS after 5–8 years due to recurrence of contended that, although tolerance conditioning is ex-
original disease or chronic antibody-mediated rejection pensive, the negation of the IS requirement may be less
(AMR). Of the ten renal allografts, three failed due to costly long term. This group estimated costs for a hypo-
thrombotic microangiopathy or rejection. In collabora- thetical 40-year-old two-haplotype match living donor
tion with the Massachusetts General Hospital group, kidney transplant recipient. The standard IS therapy
Newell et al. (9) theorized that biomarkers of transplant cost is $179,000 over a patient’s lifetime, whereas the
tolerance could enhance the safety and feasibility of tolerance induction cost is $88,000, yielding an expected
clinical tolerance trials and potentially facilitate man- lifetime savings of $92,000. If all 5170 two-haplotype
agement of patients receiving IS. To this end, blood match living donor kidney transplant recipients in the
from spontaneously tolerant renal transplant recipients United States from 2003 to 2012 had received tolerance
and patients enrolled in two interventional tolerance induction, this would have potentially yielded approxi-
trials using flow cytometry and gene expression pro- mately $350 million United States dollars in aggregate
filing was examined. They found that tolerance was lifetime savings.
associated with increased expression of B cell–associated Regarding tolerance induction, kidney allografts
genes relative to immunosuppressed patients. Moreover, behave differently from heart allografts, which behave
serial measurements of gene expression showed that this differently from lung allografts for reasons that remain
pattern persisted over several years, although patients unclear. Madariaga et al. (12) remind us that inducing
receiving IS also displayed an increase in the two most tolerance in experimental recipients of heart and lung
dominant tolerance-related B cell genes, IGKV1D-13 allografts has proven more challenging than for kidney
and IGLL-1, over time. Importantly, patients rendered allografts.
tolerant via induction of transient mixed chimerism and
those weaned to minimal IS showed similar increases in
IGKV1D-13 as did spontaneously tolerant individuals. Experimental tolerance induction utilizing
Collectively, these findings support the notion that transient or persistent donor chimerism
alterations in B cells may be a common theme for has led to durable, IS-free engraftment
tolerant kidney transplant recipients. for a limited number of human transplant
Scandling et al. (10) have provided iterative recipients.
reports, the most recent of which describes 38 HLA-
matched and -mismatched living kidney donor recipients
enrolled in tolerance protocols using post-transplant Biomarkers
conditioning with total lymphoid irradiation, enriched Despite almost 20 years of research, the use of
CD341 hematopoietic cell, and antithymocyte globulin. individual biomarkers (e.g., granzyme B, perforin, kidney
Persistent chimerism for at least 6 months was associated injury molecule-1, and neutrophil gelatinase–associated
with successful complete withdrawal of immunosup- lipocalin) to noninvasively ascertain allograft stability or
pressive drugs in 16 of 22 matched patients without differentiate among different causes of kidney injury has
rejection episodes or kidney disease recurrence with up failed to transition from the laboratory to the clinic due to
to 5 years follow-up. Persistent mixed chimerism was cost, availability, reproducibility, and performance
achieved in some haplotype-matched patients for at least characteristics. The clinical nephrologist still utilizes
318 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

more conventional allograft markers, including serum injury repair in the microcirculation induced by cognate
creatinine, urine protein-to-creatinine ratio, BK virus recognition between antibody and CD16a that triggers
nucleic acid testing, donor-specific antibody testing for IFN-g release and antibody-dependent natural killer
surveillance, and the allograft biopsy for definitive cell–mediated cytotoxicity.
diagnostics, with all of their associated limitations. O’Connell et al. (16) studied gene sets to predict
Overall, the increasing availability of sophisticated renal allografts at risk of progressive injury due to fibrosis.
and complex genomic and proteomic testing provides They prospectively collected biopsies from renal allograft
novel avenues of investigation. Sigdel et al. (13) aimed to recipients (n5204) with stable renal function 3 months
map changes in the urine proteome after kidney trans- after transplantation. The group identified a set of 13
plantation by comparing urine and biopsy changes in 396 genes independently predictive for the development of
transplant recipients. Blinded histologic data were used to fibrosis at 1 year. The gene set had high predictive
classify urine samples into diagnostic categories of acute capacity (AUC, 0.967), which was superior to that of
rejection (AR), chronic allograft nephropathy, BK virus baseline clinical variables (AUC, 0.706) and clinical and
nephritis, and stable graft function. Ultimately, a minimal pathologic variables (AUC, 0.806). Furthermore, routine
set of 35 proteins was identified for its ability to segregate pathologic variables were unable to identify which histo-
the three major transplant injury clinical groups, making logically normal allografts would progress to fibrosis
up the final panel of 11 urinary peptides for AR (93% area (AUC, 0.754), whereas the predictive gene set accurately
under the curve [AUC]), 12 urinary peptides for chronic discriminated between transplants at high and low risk of
allograft nephropathy (99% AUC), and 12 urinary pep- progression (AUC, 0.916). The 13 genes also accurately
tides for BK virus nephritis (83% AUC). Thus, urinary predicted early allograft loss (AUC, 0.842 at 2 years and
proteome discovery and targeted validation may identify AUC, 0.844 at 3 years). The predictive value of this gene
peptidomes that provide rapid, noninvasive differentiation set was established in an independent cohort. The authors
of causes of kidney transplant injury without an invasive concluded that this set of 13 genes could be used
kidney biopsy. Modena et al. (14) examined gene to identify kidney transplant recipients at risk of allograft
expression profiles of 234 graft biopsy samples matched loss before the onset of irreversible damage, potentially
with their clinical and outcome data to study interstitial allowing therapy to be modified to prevent progression to
fibrosis and tubular atrophy (IFTA). The biopsies were fibrosis.
divided into subphenotypes by degree of histologic in- Oetting et al. (17) provide a cautionary note about
flammation: IFTA with AR, IFTA with inflammation, and using such technology for prognostic purposes, warning
IFTA without inflammation. A novel analysis using gene about the requirement for validation beyond the initial
coexpression networks revealed that all IFTA phenotypes dataset. One goal of a genome-wide association study
were strongly enriched for dysregulated gene pathways was to associate the allele(s) of a single-nucleotide poly-
that were shared with the biopsy profiles of AR, including morphism with phenotype. Although there have been
IFTA samples without histologic evidence of inflamma- many successful associations documented, there have
tion. Thus, by molecular profiling, most IFTA samples been many more genome-wide association study–based
have ongoing immune-mediated injury or chronic rejection results that were later deemed false positives. As an
more sensitively detected by gene expression profiling. example, Pihlstrøm et al. (18) were unable to reproduce
Venner et al. (15) used microarray analysis of prior study data suggesting that two single-nucleotide
kidney transplant biopsies to identify changes of pure polymorphisms (rs3811321 and rs6565887) associated
AMR. In a set of 703 biopsies, 2603 transcripts were with serum creatinine and clinical outcome.
significantly changed (P,0.05) in AMR versus all
other biopsies. In cultured cells, the transcripts strongly
associated with AMR were expressed in endothelial
Allograft biomarkers have not impacted
cells. Pathway analysis of AMR transcripts identified
clinical practice to date. Novel genomic
angiogenesis, with roles for angiopoietin and vascular
and proteomic analyses are elucidating
endothelial growth factors, leukocyte-endothelial inter-
molecular events underlying allograft injury
actions, and natural killer signaling, including evidence
and tolerance. However, broader clinical
for CD16a Fc receptor–signaling elements shared with
applicability remains to be determined.
T cells. These data support a model of AMR involving
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 319

References Evidence for NK involvement through CD16a Fc receptors. Am J


1. Baron D, Ramstein G, Chesneau M, Echasseriau Y, Pallier A, Paul C, Transplant 15: 1336–1348, 2015 PubMed
Degauque N, Hernandez-Fuentes MP, Sanchez-Fueyo A, Newell KA, 16. O’Connell PJ, Zhang W, Menon MC, Yi Z, Schröppel B, Gallon L,
Giral M, Soulillou JP, Houlgatte R, Brouard S: A common gene signature Luan Y, Rosales IA, Ge Y, Losic B, Xi C, Woytovich C, Keung KL,
across multiple studies relate biomarkers and functional regulation in Wei C, Greene I, Overbey J, Bagiella E, Najafian N, Samaniego M,
tolerance to renal allograft. Kidney Int 87: 984–995, 2015 PubMed Djamali A, Alexander SI, Nankivell BJ, Chapman JR, Smith RN,
2. Chesneau M, Michel L, Dugast E, Chenouard A, Baron D, Pallier A, Colvin R, Murphy B: Biopsy transcriptome expression profiling to
Durand J, Braza F, Guerif P, Laplaud DA, Soulillou JP, Giral M, identify kidney transplants at risk of chronic injury: A multicentre,
Degauque N, Chiffoleau E, Brouard S: Tolerant kidney transplant prospective study. Lancet 388: 983–993, 2016 PubMed
patients produce B cells with regulatory properties. J Am Soc Nephrol 17. Oetting WS, Jacobson PA, Israni AK: Validation is critical for genome-
26: 2588–2598, 2015 PubMed wide association study-based associations. Am J Transplant 17: 318–
3. Braza F, Dugast E, Panov I, Paul C, Vogt K, Pallier A, Chesneau M, 319, 2017 PubMed
Baron D, Guerif P, Lei H, Laplaud DA, Volk HD, Degauque N, Giral 18. Pihlstrøm HK, Mjøen G, Mucha S, Haraldsen G, Franke A, Jardine A,
M, Soulillou JP, Sawitzki B, Brouard S: Central role of CD45RA- Fellström B, Holdaas H, Melum E: Single nucleotide polymorphisms
Foxp3hi memory regulatory T cells in clinical kidney transplantation and long-term clinical outcome in renal transplant patients: A validation
tolerance. J Am Soc Nephrol 26: 1795–1805, 2015 PubMed study. Am J Transplant 17: 528–533, 2017 PubMed
4. Elias N, Cosimi AB, Kawai T: Clinical trials for induction of renal allograft
tolerance. Curr Opin Organ Transplant 20: 406–411, 2015 PubMed
5. Leventhal JR, Mathew JM, Salomon DR, Kurian SM, Friedewald JJ, Glomerular Injury Post-Transplant
Gallon L, Konieczna I, Tambur AR, Charette J, Levitsky J, Jie C,
Kanwar YS, Abecassis MM, Miller J: Nonchimeric HLA-identical renal Proteinuria in the Post-Transplant Setting
transplant tolerance: Regulatory immunophenotypic/genomic bio-
markers. Am J Transplant 16: 221–234, 2016 PubMed Measures of proteinuria post-transplant may signal
6. Yolcu ES, Leventhal JR, Ildstad ST: Facilitating cells in tolerance induction for de novo or recurrent glomerular disease or a manifesta-
kidney transplantation. Curr Opin Organ Transplant 20: 57–63, 2015 PubMed tion of alloantibody-mediated injury, also referred to as
7. Leventhal JR, Elliott MJ, Yolcu ES, Bozulic LD, Tollerud DJ, Mathew
JM, Konieczna I, Ison MG, Galvin J, Mehta J, Badder MD, Abecassis transplant glomerulopathy (see above). A recent study
MM, Miller J, Gallon L, Ildstad ST: Immune reconstitution/immuno- showed the predictive value of proteinuria (measured
competence in recipients of kidney plus hematopoietic stem/facilitating
3 months and 1, 2, 5, and 10 years post-transplantation
cell transplants. Transplantation 99: 288–298, 2015 PubMed
8. Kawai T, Sachs DH, Sprangers B, Spitzer TR, Saidman SL, Zorn E, and at the time of renal allograft biopsy) for graft
Tolkoff-Rubin N, Preffer F, Crisalli K, Gao B, Wong W, Morris H, failure in 1518 patients followed for over 7 years post-
LoCascio SA, Sayre P, Shonts B, Williams WW Jr., Smith RN, Colvin RB,
Sykes M, Cosimi AB: Long-term results in recipients of combined HLA-
transplant. Protocol biopsies correlated proteinuria with
mismatched kidney and bone marrow transplantation without maintenance graft histology (1). Although proteinuria of 300–1000 mg/d
immunosuppression. Am J Transplant 14: 1599–1611, 2014 PubMed was associated with a twofold increased risk of graft
9. Newell KA, Asare A, Sanz I, Wei C, Rosenberg A, Gao Z, Kanaparthi
S, Asare S, Lim N, Stahly M, Howell M, Knechtle S, Kirk A, Marks
loss, proteinuria of 1–3 g/d was associated with a three-
WH, Kawai T, Spitzer T, Tolkoff-Rubin N, Sykes M, Sachs DH, Cosimi fold increased risk of graft loss, independent of GFR
AB, Burlingham WJ, Phippard D, Turka LA: Longitudinal studies of and histology. The predictive value, in terms of sensi-
a B cell-derived signature of tolerance in renal transplant recipients. Am
J Transplant 15: 2908–2920, 2015 PubMed
tivity, specificity, and negative and positive predictive
10. Scandling JD, Busque S, Shizuru JA, Lowsky R, Hoppe R, Dejbakhsh- values of proteinuria, was relatively low for abnormal
Jones S, Jensen K, Shori A, Strober JA, Lavori P, Turnbull BB, glomerular histology or graft loss. The best predictor for
Engleman EG, Strober S: Chimerism, graft survival, and withdrawal of
immunosuppressive drugs in HLA matched and mismatched patients biopsy findings of transplant glomerulopathy, glomerular
after living donor kidney and hematopoietic cell transplantation. Am J disease, and/or microcirculatory inflammation was pro-
Transplant 15: 695–704, 2015 PubMed teinuria .1 g/d measured 3 months post-transplant. The
11. Erickson KF, Winkelmayer WC, Busque S, Lowsky R, Scandling JD,
Strober S: A cost analysis of tolerance induction for two-haplotype positive predictive value of this parameter was 58.2%,
match kidney transplant recipients. Am J Transplant 16: 371–373, whereas the negative predictive value was 78.7%. The
2016 PubMed
12. Madariaga ML, Kreisel D, Madsen JC: Organ-specific differences in achiev-
research group suggested that kidney biopsy was indi-
ing tolerance. Curr Opin Organ Transplant 20: 392–399, 2015 PubMed cated when a proteinuria threshold of $1 g/d was
13. Sigdel TK, Gao Y, He J, Wang A, Nicora CD, Fillmore TL, Shi T, achieved to better characterize the nature of glomerular
Webb-Robertson BJ, Smith RD, Qian WJ, Salvatierra O, Camp DG
2nd, Sarwal MM: Mining the human urine proteome for monitoring
injury. Beyond these modestly predictive findings, a sep-
renal transplant injury. Kidney Int 89: 1244–1252, 2016 PubMed arate retrospective study of 805 kidney transplant recipi-
14. Modena BD, Kurian SM, Gaber LW, Waalen J, Su AI, Gelbart T, ents concluded that the resolution of low-grade proteinuria
Mondala TS, Head SR, Papp S, Heilman R, Friedewald JJ, Flechner SM,
Marsh CL, Sung RS, Shidban H, Chan L, Abecassis MM, Salomon DR:
(from 0.15–1000 mg/d to ,0.15 g/d from years 1 and 2 to
Gene expression in biopsies of acute rejection and interstitial fibrosis/tubular years 2 and 3 post-transplant) was associated with a higher
atrophy reveals highly shared mechanisms that correlate with worse long- graft survival rate compared with those with persistent
term outcomes. Am J Transplant 16: 1982–1998, 2016 PubMed
15. Venner JM, Hidalgo LG, Famulski KS, Chang J, Halloran PF: The proteinuria, independent of BP (2). The combination of
molecular landscape of antibody-mediated kidney transplant rejection: persistent low-grade proteinuria and systolic BP .140
320 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

mmHg during the first 3 years was associated with inferior to 1.60) and type II (HR, 0.90; 95% CI, 0.32 to 2.52).
patient survival and death-censored graft survival (60.2% The authors concluded that living donation should be not
and 87.6%, respectively) compared with in normotensive be discouraged across all GN subtypes, with informed
patients without proteinuria (87.9% and 97.0% with decision making in cases involving MPGN.
P,0.001 and P,0.01, respectively). One potential contrib-
utor to the development of proteinuria is mammalian target Focal and Segmental Glomerulosclerosis
of rapamycin inhibitor (mTORi). Prevention of proteinuria The risk for recurrent of FSGS post-transplant is
related to mTORi may be reasonably achieved using renin- substantial and frequently leads to early graft loss. In
angiotensin system blockade. A prospective, randomized, an Australian and New Zealand Dialysis and Trans-
double-blind, placebo-controlled study of 264 renal trans- plant Registry analysis of 666 adults and 70 pediatric
plant patients converted from a calcineurin inhibitor to kidney transplant recipients with ESRD eventuating
sirolimus compared ramipril with placebo in primary preven- from biopsy-proven FSGS, the disorder recurred in
tion of proteinuria (defined as urinary protein-to-creatinine 10.3% of cases, with a 52% 5-year graft survival versus
ratio $0.5) (3). Losartan was administered if proteinuria 83% in those without recurrent disease (P,0.001) (7).
was not controlled, despite uptitration of ramipril. At 52 Risk factors for recurrence included younger age, non-
weeks, the cumulative rate of losartan initiation was signif- white ethnicity, and living donor versus deceased donor
icantly lower in the ramipril arm (6.2%) than placebo transplant. Overall, median graft survival was still superior
(23.2%; P,0.001), with no difference between study arms with living donor transplant than deceased donor transplant
with respect to changes in eGFR over the study period. A (14.8 versus 12.1 years; P,0.01), again suggesting that
recent review provides the reader with greater depth of the living donor transplant should not be avoided in this
topic of proteinuria in the post-transplant setting (4). cohort. Recurrence rates strongly depend on etiology of
FSGS. In pediatric patients with known or suspected
Glomerular Disease: Post-Transplant Outcomes genetic causes of FSGS, the post-transplant recurrence
Patients with GN tend to have excellent outcomes rate is low. For example, only one of 25 patients in a
after kidney transplantation when assessed as a grouped National Israeli Kidney Registry Study developed recur-
cohort. However, there is variation within individual rent FSGS versus 17 of 22 patients (64%), in whom FSGS
subtypes. A United States registry analysis of outcomes was considered “primary/nongenetic” (8).
in 32,131 patients with one of six GN subtypes showed Treatment for recurrent post-transplant FSGS
highest graft survival in IgA nephropathy and lowest graft remains a challenge. A review and meta-analysis of the
survival in vasculitis, the latter driven entirely due to death role of plasma exchange (PLEX) as a primary intervention
as an end point rather than graft loss (5). Regarding the pooled 34 case reports and 43 case series totaling 423
hazard for death-censored graft loss, the IgA nephropathy patients (9). Defining remission as ,3.5 g/d (partial) or
and vasculitis subgroups were similar (hazard ratio [HR], ,0.5 g/d (complete), the overall remission rate was 71%.
0.94; P value not significant) and superior to FSGS (HR, Neither age nor donor type were associated with remission.
1.20; 95% CI, 1.12 to 1.28), membranous nephropathy Proteinuria .7 g/d was inversely associated with remission,
(MN; HR, 1.27; 95% CI, 1.14 to 1.41), membranoproli- whereas men and early treatment within 2 weeks of
ferative GN (MPGN; HR, 1.50; 95% CI, 1.36 to 1.66), recurrence were associated with higher rates of remission.
and lupus nephritis (HR, 1.11; 95% CI, 1.02 to 1.20), Rituximab (RTX) administration to individuals who failed
with all HRs statistically significant, compared with IgA or could not be weaned from PLEX and high-dose
as reference. A European registry analysis of 41,383 calcineurin inhibitors was reported in a series of 19 patients.
patients with GN showed similar findings of highest Ten patients underwent RTX readministration after failure
death-censored graft survival in the IgA subtype and to respond to PLEX (10): five achieved partial (n54) or
lowest survival in the MPGN subtypes (6). Addressing complete (n51) remission, and eight developed serious
the potential concern of using living donors for individ- infections within 1 year, suggesting that adding RTX to
uals with primary GNs, all subtypes had superior graft achieve control of FSGS must be done with close monitor-
survival with living versus deceased donor kidney ing and expectation for infectious complications. Finally,
transplant (e.g., IgA: HR, 0.74; 95% confidence interval initial optimism regarding targeted B7-1 (putative role in
[95% CI], 0.59 to 0.92; FSGS: HR, 0.69; 95% CI, 0.45 aberrant podocyte integrity) treatment with the costimulation
to 1.06), except MPGN type I (HR, 1.08; 95% CI, 0.73 blockers, abatacept or belatacept, has been tempered by
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 321

a recent report of nine patients with recurrent, post-transplant donors with two APOL1 risk variants may confer risk of
FSGS. All failed to respond to this therapy, despite decreased allograft survival in the recipient. In a two-
combination with other therapies that included PLEX and/ center report of 675 kidneys transplanted, the presence
or immunoadsorption in all patients and RTX in most (11). of two donor risk variants was associated with graft loss
(HR, 2.26; P50.001) (17). In an extension of this study
Membranous Nephropathy including an additional 478 deceased donor kidney
Membranous Nephropathy (MN) often recurs post- transplants, in whom APOL1 genotyping was per-
transplant but typically, with a far less aggressive course formed, donor APOL1 high-risk genotype was associ-
than FSGS. In the largest and most comprehensive case ated with inferior allograft survival (Figure 13) (HR,
series to date, 63 recipients with biopsy-proven primary 2.05; 95% 95% CI 1.39 to 3.02; P,0.001) (18). Despite
MN pretransplant were followed for a median of 77 this risk, most recipients maintained excellent function
months post-transplant. Recurrent disease developed in (72.6% of grafts functioning for .5 years and 56.9% of
30 of 63 (48%) cases, and 16 were diagnosed by protocol those functioning .10 years). It is possible that the risk
biopsy. In 17 of the 30 cases, there was progressive conferred by APOL1 risk variant donors is only loosely
proteinuria (12). Progressors were treated with RTX, approximated by the inclusion of black race in the
with complete or partial remission in 14 (82%) cases. calculation of the deceased donor risk index that
Importantly, graft and patient survival rates were not determines the donor profile index for deceased donor
different among those who developed recurrence versus kidney allocation. A recent assessment of the impact of
those who did not, similar to a concurrently transplanted substituting APOL1 status for race described the poten-
autosomal dominant polycystic kidney disease cohort. tial for refinement of risk for graft loss and may more
The degree of pretransplant proteinuria was associated precisely classify the impact of black race as a donor risk
with recurrence (approximately 50% recurrence rate for factor (19). Current understanding of the role and value
proteinuria of 5–10 g/d; approximately 80% for .10 of testing for APOL1 risk variants is rapidly evolving
g/d). The presence of pretransplant antiphopholipase A2 and includes consideration for not only recipients but
receptor antibodies (anti-PLA2R Abs) was also associ- also, potential living donors. These concepts were
ated with recurrence (HR, 3.76; 95% CI, 1.64 to 8.65; recently summarized in proceedings from an American
P50.002). In a separate study, the positive predictive Society of Transplantation Expert Conference (20).
value of pretransplant anti-PLA2R for recurrence was
83%, and negative predictive value was 42%, suggesting
a role in screening for anti-PLA2R Ab to risk stratify The presence of two APOL1 renal risk
recurrence. Namely, a positive value may indicate a need variants in deceased donors is associated
for heightened surveillance, whereas a negative value with a 1.5- to twofold increased risk of graft
does not necessarily rule out the possibility of recurrent loss. APOL1 renal risk variants may explain
disease (13). This finding was corroborated by another the disparity in allograft outcomes from
single-center analysis of 16 patients (14). A more detailed black kidney donors.
review of the value of anti-PLA2R Ab screening and
monitoring in pre- and post-transplant settings has been
recently published (15). Complement-Mediated Glomerular Disease
Complement-mediated diseases, including atypi-
APOL1 and Kidney Transplantation cal hemolytic uremic syndrome (aHUS) and C3
APOL1 gene polymorphisms (risk variants) are glomerulopathies, have high recurrence rates in renal
enriched in blacks due to the trypanolytic effects allografts and substantially impact graft survival. Despite
conferred by these ApoL1 proteins. Individuals with this, a recent registry analysis reported the persistence of
two APOL1 risk variants (10%–15% of the black survival benefit of transplant in adult patients with
population) are more likely to develop glomerular a diagnosis of hemolytic uremic syndrome (not explicitly
diseases, including FSGS, HIV-associated nephropa- defined as aHUS) compared with those remaining on the
thy, lupus nephritis, and hypertensive nephrosclerosis, waiting list (5-year mortality risk with living donor: HR,
in the nontransplant setting (16). In the transplant setting, 0.27; 95% CI, 0.11 to 0.65; with deceased donor: HR,
recent clinical observations suggest that kidneys from 0.52; 95% CI, 0.29 to 0.94) (21). In general, data are
322 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

Figure 13. Adjusted Kaplan–Meier survival plots (full model) in 1153 deceased donor kidney transplantations from black
donors based on donor APOL1 genotypes. Plots compare survival of kidneys from donors with two renal risk variants with
that for kidneys from donors with fewer than two renal risk variants. The numbers within the parentheses below the curves
reflect the following: (number of functioning allografts at the start of each year, number of allograft failures within
that year). Reprinted with permission from Freedman BI, Pastan SO, Israni AK, Schladt D, Julian BA, Gautreaux
MD, Hauptfeld V, Bray RA, Gebel HM, Kirk AD, Gaston RS, Rogers J, Farney AC, Orlando G, Stratta RJ, Mohan S,
Ma L, Langefeld CD, Bowden DW, Hicks PJ, Palmer ND, Palanisamy A, Reeves-Daniel AM, Brown WM, Divers J:
APOL1 genotype and kidney transplantation outcomes from deceased African American donors. Transplantation 100:
194–202, 2016.

sparse for the management of patient with complement- rence. These recommendations may help in prediction,
mediated injury, but examples can be gleaned from the but in practice, they may unfortunately not change
nontransplant population. In a recent open label trial of management. For example, when a high-risk comple-
41 patients treated with eculizumab for aHUS, 73% had ment mutation is identified, therapy may be considered to
a complete thrombotic microangiopathic response, with prevent or treat recurrence; however, expense may limit
15 of 19 patients dialysis dependent at initiation of access to therapies, such as eculizumab, except in cases
therapy discontinuing dialysis (22). Of the overall cohort, where clinical diagnosis of aHUS has been established.
nine patients had a history of prior kidney post-transplant
(three on dialysis); data were not reported for this small References
cohort. A recent review describes current considerations 1. Naesens M, Lerut E, Emonds MP, Herelixka A, Evenepoel P, Claes K,
Bammens B, Sprangers B, Meijers B, Jochmans I, Monbaliu D, Pirenne
for diagnosis and management of these rare diseases in J, Kuypers DR: Proteinuria as a noninvasive marker for renal allograft
the transplant setting (23). Pertinent recommendations histology and failure: An observational cohort study. J Am Soc Nephrol
include careful consideration of C3 glomerulopathy in 27: 281–292, 2016 PubMed
2. Cherukuri A, Tattersall JE, Lewington AJ, Newstead CG, Baker RJ:
any renal transplant candidate whose original disease Resolution of low-grade proteinuria is associated with improved
was reported as MPGN, mesangial proliferative GN, or outcomes after renal transplantation-a retrospective longitudinal study.
endocapillary proliferative GN. If C3 glomerulopathy Am J Transplant 15: 741–753, 2015 PubMed
3. Mandelbrot DA, Alberú J, Barama A, Marder BA, Silva HT Jr.,
can be verified, detailed complement testing can be Flechner SM, Flynn A, Healy C, Li H, Tortorici MA, Schulman SL:
performed to identify and predict the risk of recur- Effect of ramipril on urinary protein excretion in maintenance renal
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 323

transplant patients converted to sirolimus. Am J Transplant 15: 3174– 18. Freedman BI, Pastan SO, Israni AK, Schladt D, Julian BA, Gautreaux
3184, 2015 PubMed MD, Hauptfeld V, Bray RA, Gebel HM, Kirk AD, Gaston RS, Rogers J,
4. Tsampalieros A, Knoll GA, Molnar AO, Fergusson N, Fergusson DA: Farney AC, Orlando G, Stratta RJ, Mohan S, Ma L, Langefeld CD,
Corticosteroid use and growth after pediatric solid organ transplanta- Bowden DW, Hicks PJ, Palmer ND, Palanisamy A, Reeves-Daniel AM,
tion: A systematic review and meta-analysis. Transplantation 101: 694– Brown WM, Divers J: APOL1 genotype and kidney transplantation
703, 2017 PubMed outcomes from deceased African American donors. Transplantation
5. O’Shaughnessy MM, Liu S, Montez-Rath ME, Lenihan CR, Lafayette 100: 194–202, 2016 PubMed
RA, Winkelmayer WC: Kidney transplantation outcomes across 19. Julian BA, Gaston RS, Brown WM, Reeves-Daniel AM, Israni AK,
GN subtypes in the United States. J Am Soc Nephrol 28: 632–644, Schladt DP, Pastan SO, Mohan S, Freedman BI, Divers J: Effect of
2017 PubMed replacing race with apolipoprotein L1 genotype in calculation of kidney
6. Pippias M, Stel VS, Aresté-Fosalba N, Couchoud C, Fernandez- donor risk index. Am J Transplant 17: 1540–1548, 2017 PubMed
Fresnedo G, Finne P, Heaf JG, Hoitsma A, De Meester J, Pálsson R, 20. Newell KA, Formica RN, Gill JS, Schold JD, Allan JS, Covington SH,
Ravani P, Segelmark M, Traynor JP, Reis×ter AV, Caskey FJ, Jager KJ: Wiseman AC, Chandraker A: Integrating APOL1 gene variants into
Long-term kidney transplant outcomes in primary glomerulonephritis: renal transplantation: Considerations arising from the American Society
Analysis from the ERA-EDTA Registry. Transplantation 100: 1955– of Transplantation Expert Conference. Am J Transplant 17: 901–911,
1962, 2016 PubMed 2017 PubMed
7. Francis A, Trnka P, McTaggart SJ: Long-term outcome of kidney 21. Santos AH Jr., Casey MJ, Wen X, Zendejas I, Rehman S, Womer KL,
transplantation in recipients with focal segmental glomerulosclerosis. Andreoni KA: Survival with dialysis versus kidney transplantation in
Clin J Am Soc Nephrol 11: 2041–2046, 2016 PubMed adult hemolytic uremic syndrome patients: A fifteen-year study of the
8. Cleper R, Krause I, Bar Nathan N, Mor M, Dagan A, Weissman I, waiting list. Transplantation 99: 2608–2616, 2015 PubMed
Frishberg Y, Rachamimov R, Mor E, Davidovits M: Focal segmental 22. Fakhouri F, Hourmant M, Campistol JM, Cataland SR, Espinosa M,
glomerulosclerosis in pediatric kidney transplantation: 30 Years’ Gaber AO, Menne J, Minetti EE, Provôt F, Rondeau E, Ruggenenti P,
experience. Clin Transplant 30: 1324–1331, 2016 PubMed Weekers LE, Ogawa M, Bedrosian CL, Legendre CM: Terminal
9. Kashgary A, Sontrop JM, Li L, Al-Jaishi AA, Habibullah ZN, complement inhibitor eculizumab in adult patients with atypical
Alsolaimani R, Clark WF: The role of plasma exchange in treating hemolytic uremic syndrome: A single-arm, open-label trial. Am J
post-transplant focal segmental glomerulosclerosis: A systematic Kidney Dis 68: 84–93, 2016 PubMed
review and meta-analysis of 77 case-reports and case-series. BMC 23. Barbour S, Gill JS: Advances in the understanding of complement
Nephrol 17: 104, 2016 PubMed mediated glomerular disease: Implications for recurrence in the trans-
10. Garrouste C, Canaud G, Büchler M, Rivalan J, Colosio C, Martinez F, plant setting. Am J Transplant 15: 312–319, 2015 PubMed
Aniort J, Dudreuilh C, Pereira B, Caillard S, Philipponnet C, Anglicheau
D, Heng AE: Rituximab for recurrence of primary focal segmental
glomerulosclerosis after kidney transplantation: Clinical outcomes.
Transplantation 101: 649–656, 2017 PubMed
11. Delville M, Baye E, Durrbach A, Audard V, Kofman T, Braun L, Infection after Transplant
Olagne J, Nguyen C, Deschênes G, Moulin B, Delahousse M, Kesler-
Roussey G, Beaudreuil S, Martinez F, Rabant M, Grimbert P, Gallazzini Infectious complications after transplant are a ma-
M, Terzi F, Legendre C, Canaud G: B7-1 blockade does not improve
post-transplant nephrotic syndrome caused by recurrent FSGS. J Am jor cause of morbidity and mortality and may interpose
Soc Nephrol 27: 2520–2527, 2016 PubMed allograft injury and graft loss (1). Newer infections,
12. Grupper A, Cornell LD, Fervenza FC, Beck LH Jr., Lorenz E, Cosio such as Zika virus, are incompletely understood in the
FG: Recurrent membranous nephropathy after kidney transplantation:
Treatment and long-term implications [published online ahead of print transplant population (2). Only a small case series from
December 30, 2015]. Transplantation PubMed Brazil described the courses of two kidney and two
13. Kattah A, Ayalon R, Beck LH Jr., Sethi S, Sandor DG, Cosio FG, liver transplant recipients. All cases were identified
Gandhi MJ, Lorenz EC, Salant DJ, Fervenza FC: Anti-phospholipase
A2 receptor antibodies in recurrent membranous nephropathy. Am J concurrent with confounding bacterial infections that
Transplant 15: 1349–1359, 2015 PubMed do not permit greater understanding of the impact of
14. Gupta G, Fattah H, Ayalon R, Kidd J, Gehr T, Quintana LF, Kimball P,
Zika infection per se (3). Preventing common infec-
Sadruddin S, Massey HD, Kumar D, King AL, Beck LH Jr.: Pre-
transplant phospholipase A2 receptor autoantibody concentration is tions, such as influenza and common urinary tract
associated with clinically significant recurrence of membranous ne- infections (UTIs), are worthy goals but have limita-
phropathy post-kidney transplantation. Clin Transplant 30: 461–469,
2016 PubMed
tions. After an outbreak of influenza A in one inpatient
15. De Vriese AS, Glassock RJ, Nath KA, Sethi S, Fervenza FC: A proposal transplant unit, 23 kidney transplant patients were
for a serology-based approach to membranous nephropathy. J Am Soc identified as at risk: 17 had adequate seasonal influenza
Nephrol 28: 421–430, 2017 PubMed
16. Riella LV, Sheridan AM: Testing for high-risk APOL1 alleles
vaccination, of whom two tested positive without
in potential living kidney donors. Am J Kidney Dis 66: 396–401, clinical consequences. Of six unvaccinated recipients,
2015 PubMed five developed influenza, and three died from severe
17. Freedman BI, Julian BA, Pastan SO, Israni AK, Schladt D, Gautreaux
MD, Hauptfeld V, Bray RA, Gebel HM, Kirk AD, Gaston RS, Rogers J, respiratory failure (4). In a study of 60 patients after
Farney AC, Orlando G, Stratta RJ, Mohan S, Ma L, Langefeld CD, influenza vaccination, adequate seroconversion to at
Hicks PJ, Palmer ND, Adams PL, Palanisamy A, Reeves-Daniel AM, least one of the three influenza antigens occurred in
Divers J: Apolipoprotein L1 gene variants in deceased organ donors are
associated with renal allograft failure. Am J Transplant 15: 1615–1622, 63% without an increase in HLA alloantibody titers (5).
2015 PubMed In conclusion, vaccination is less efficacious in ESRD
324 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

and kidney transplant populations than the general Forty-eight high-risk (CMV IgG1 donor/CMV IgG–
population but is still considered valuable. recipient) recipients who received thymoglobulin induc-
The same utility cannot be ascribed to screening tion and 6 months of valganciclovir underwent CMV
for and treating asymptomatic bacteriuria to prevent PCR monitoring during prophylaxis and for 3 additional
UTI. In 112 kidney transplant recipients with at least months of surveillance (10). Nineteen (40%) developed
one episode of asymptomatic bacteriuria after the second CMV infection: nine occurred during prophylaxis (five
month, post-transplant routine screening was conducted, were associated with suboptimal valganciclovir dosing);
with subsequent randomization in 1:1 fashion to antimi- six occurred during the surveillance period post-transplant,
crobial therapy or no therapy. No difference in the acute and four occurred during the 3-month postsurveillance
pyelonephritis rate at 24 months was found (7.5% versus period. The authors suggest that identification of viremia
8.4%). There was no difference in lower UTI frequency, can ideally permit immunosuppression adjustment and
allograft function, or mortality (6). reduce morbidity and hospitalizations.
Beyond CMV PCR monitoring, a number of other
Cytomegalovirus markers that may identify CMV-specific immunity are
To prevent cytomegalovirus (CMV) infection/ under study, including T cell subsets (11), chemokine
disease, most transplant centers in the United States (C-C motif) ligand 8 (CCL8) (12), and single-nucleotide
use a CMV prophylaxis strategy using valganciclovir polymorphisms in genes involved in CMV immune
for 3–6 months post-transplant. High-dose valacyclovir response (13). These references are provided for the
has also been recommended. A randomized trial of 119 interested reader (11–13).
patients receiving either valganciclovir 900 mg daily
or valacyclovir 2 g four times daily showed similar efficacy BK Virus
in prevention of CMV viremia at 12 months (31% versus Control of BK virus (BKV) reactivation from the
43%; P50.36). Notably, the valganciclovir group had transplanted kidney remains an important challenge in
a lower incidence of acute rejection (17% versus 31%; optimizing graft survival. Recent advances in the
P50.03) and higher BK virus (BKV) viremia rates (36% understanding of the mechanisms that contribute to
versus 18%; P50.04), supporting an additive immuno- viral reactivation include measures of BKV-specific
suppressive effect of valganciclovir (7). Conversely, im- immunity, viral relationship to specific immunosup-
munosuppressive medications (specifically, sirolimus and pressive agents, and contribution of donor versus
everolimus) may have additive antiviral effects. In the recipient pretransplant BKV exposure. In a study com-
absence of CMV prophylaxis, 288 kidney transplant paring BKV-specific and alloreactive T cells from 24
recipients were randomized to receive tacrolimus (TAC), kidney transplant patients who developed BKV vire-
prednisone, and everolimus (n5102 with antithymocyte mia with those from 127 patients without BKV
globulin induction, n585 with basiliximab induction) or (sampled pretransplant and 1, 2, and 3 months post-
mycophenolate (n5101 with basiliximab induction) (8). transplant), changes in BKV-specific T cells from the
Rates of CMV infection/disease at 12 months were 4.7%, pre- to post-transplant period were associated with risk
10.8%, and 37.6%, respectively (P,0.001 for each ever- for subsequent BKV replication (14). Another study of
olimus versus mycophenolate group). This finding was cytokine profiles from T cell subsets from individuals
corroborated in a multicenter registry analysis of 311 who cleared BKV versus those who did not revealed
pediatric kidney transplant recipients that revealed an an increase in a specific polyfunctional T cell subset that
83% lower risk of CMV replication overall and a lower secreted multiple cytokines only in the former group
CMV disease rate in high-risk (CMV IgG1 donor/CMV (15). Thus, measures of T cell functionality may be
IgG2 recipient) recipients when prescribed immuno- predictive of risk for BKV infection, and the likelihood
suppression consisting of cyclosporine (CsA)/everolimus of clearance after replication has been established.
versus CsA or TAC in combination with mycophenolate Mammalian target of rapamycin inhibitors (mTORis)
(9). These findings support experimental data that docu- may not only preserve BKV-specific T cell immunity but
ment inhibition of CMV replication by mammalian target may also inhibit BKV replication in renal tubular epithelial
of rapamycin inhibition. cells more than calcineurin inhibitors. The former hy-
Monitoring for CMV infection may be warranted pothesis was supported by a study showing reductions in
even after prophylaxis in certain high-risk circumstances. cytotoxic T cell function using CsA but not mTORi (16),
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 325

whereas the latter hypothesis was shown in an in vitro antibody-depleting induction therapy. Factors associated
study, in which sirolimus exerted inhibitory effects on with reduced risk included recipient diabetes and sirolimus
BKV replication in renal tubular epithelial cells (17). TAC use. The rate of BKV infection in paired transplanted
increased BKV replication and counteracted sirolimus kidneys was threefold higher than expected, implying that
inhibition when the two medications were combined. This donor factors, such as serostatus, are of greater impor-
suggests a possible advantage of mTORi compared with tance in BKV risk.
TAC for prevention of BKV infection. Allograft biopsy findings and other recipient
The measurement of pretransplant anti-BKV sero- antibodies during BKV infection and BKV may elicit
status from the donor and recipient has been proposed enhanced understanding of the subsequent clinical
as a measure to determine risk of post-transplant BKV course. Associations of de novo donor-specific anti-
infection, analogous to measuring CMV IgG status in bodies (DSAs) with persistent BKV viremia (22) and
donors and recipients pretransplant to define CMV risk. increased risk of rejection with antibodies to self-
Recent studies increasingly support this concept as a antigens fibronectin and collagen type IV with BKV
potential risk stratification tool. In 407 living kidney viremia (23) have recently been reported. Regarding
donor-recipient pairs, donor but not recipient BKV the risk (or association) of BKV viremia and de novo
seroreactivity was strongly associated with BKV vire- DSA formation, in a large, single-center study, 785
mia within 1 year post-transplant. Coupling donors with patients were screened for BKV viremia, of which 132
high seroreactivity with recipients of low seroreactivity (17%) developed BKV viremia. BKV viremia that
was associated with a tenfold increased risk of BKV persisted .140 days despite immunosuppression dose
viremia (18). Another study of 116 donor/recipient pairs reduction was not associated with inferior graft or patient
reported that kidneys from donors with significant survival. However, de novo HLA DSAs were identified
serum neutralizing activity (donor positive) had a five- in 30% with persistent viremia compared with 23% in
fold elevated risk for BKV viremia, regardless of recip- those who rapidly cleared the virus and 15% in those
ient serostatus (donor positive versus donor negative: without BKV viremia (on multivariate analysis: hazard
odds ratio, 5.0; 95% confidence interval [95% CI], 1.9 ratio [HR], 2.53; 95% CI, 1.40 to 4.59; P,0.01 for the
to 12.7; P,0.001), and those donor-recipient pairs with development of de novo DSA with persistent BKV
donor-positive/recipient-negative neutralizing serostatus viremia). Development of de novo DSAs generally
had the greatest risk for BKV viremia (odds ratio, 4.9; occurred after BKV viremia had been identified (me-
95% CI, 1.7 to 14.6; P50.004) (19). These data together dian 344 days post-transplant versus 137 days post-
with another study that described a correlation of spe- transplant, respectively). Beyond serologic assessment,
cific donor-derived BKV subtypes with post-transplant other studies have examined biopsies of BKV nephrop-
BKV infection (20) implicate donor source and immu- athy and found a higher frequency of allospecific T cell
nity as important risk factors for clinically significant BKV clones rather than BKV-specific clones (24). These
infection post-transplant. Transplant recipients who are findings highlight the interplay between BKV-related
seronegative for BK virus who receive seropositive grafts inflammation/viral immunity and rejection/alloimmunity
(BK D1/R-) are at increased risk of BK viremia. and make the clinical strategy immunosuppression dose
Other possible risk factors for BKV susceptibility reduction for BKV viremia a more complex decision
were clarified in a registry analysis using a paired than previously understood.
kidney analysis that identified recipient pairs receiving
kidneys from the same deceased donor. By stratifying HIV
those cases where one of two recipients were treated Although kidney transplantation for HIV1 candi-
for BKV (discordant treatment) from those cases where dates is associated with generally good outcomes, there is
both recipients were treated for BKV (concordant treat- an increased risk of acute rejection along with other donor
ment), associations could be made to determine risk and recipient risk factors that may contribute to inferior
invoked by the donor (concordant treatment) versus the graft and patient survival compared with in the HIV–
recipient (discordant treatment) (21). Recipient factors recipient. In a recent analysis of 499 HIV1 recipients,
that were associated with increased risk for treated improvements in graft survival and mortality were
BKV infection included age ,18 or $60 years old, noted over time (2008–2011 versus 2004–2007). En-
men, four or more HLA mismatches, acute rejection, and couragingly, center experience (performance of less than
326 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

ients. In a national cohort study in the United Kingdom


composed of 78 patients (31 treated with CsA and 47
treated with TAC), 1-year acute rejection rates were
58% versus 21%, respectively (HR for acute rejection
with TAC versus CsA, 0.25; 95% CI, 0.11 to 0.57;
P50.001) (29). Beyond immunosuppression, a major
factor that strongly contributes to inferior outcomes in
HIV1 recipients is coinfection with HCV. Two recent
studies highlighted that monoinfected HIV1 transplant
recipients have similar graft and patient survival
compared with HIV– reference groups (30, 31). How-
ever, coinfection with hepatitis C virus (HCV; HIV1
/HCV1) is associated with marked reductions in graft
Figure 14. Three-year allograft survival was higher in the and patient survival (Figure 14). An updated analysis
HIV monoinfected (81%) and uninfected groups (86%) than not only examined the relationship of HCV coinfection
in the HCV monoinfected (78%) and HIV/HCV coinfected on outcomes but also, focused on other factors that may
patients (60%; P,0.001). Reprinted with permission from contribute to inferior graft loss compared with HIV–
Sawinski D, Forde KA, Eddinger K, Troxel AB, Blumberg
recipients (32). In total, 526 HIV1 kidney transplant
E, Tebas P, Abt PL, Bloom RD: Superior outcomes in HIV-
positive kidney transplant patients compared with HCV-
recipients were compared with 82,236 HIV– recipients
infected or HIV/HCV-coinfected recipients. Kidney Int 88: in a registry analysis of the era from 2001 to 2013. Two
341–349, 2015. primary risk factors, HCV coinfection and more than
three HLA mismatches, were identified as risk amplifiers
or more than six HIV1 transplants) was not associated when examining risk factors for graft loss (adjusted HR,
with outcomes, a point that should encourage further 3.86; 95% CI, 2.37 to 6.30; P,0.001).
expansion to a greater number of centers (25). In other
registry studies, factors that were predictive of acute
rejection included short duration of viral suppression Transplant recipients co-infected with HIV
pretransplant (greater than twofold risk in those with HIV and HCV have markedly inferior graft sur-
RNA suppression for 6 months; 2 years versus .2 years) vival rates compared with either mono-
(26) and nonuse of antithymocyte globulin (27). In the infected or non-infected individuals.
latter study, antithymocyte globulin use was associated
with a 31% reduction in risk for rejection (95% CI, 0.35
to 0.99) (27). Antithymocyte globulin induction was also With the passage of the HIV Organ Policy Equity
associated with a reduction in delayed graft function and Act, HIV1 recipients may now consent in clinical trials
graft loss and a trend to lower mortality without an to receive organs from deceased donors who were
increase in risk of infection. Such data support the relative HIV1. The precedent for this originated from experi-
safety and efficacy of the counterintuitive strategy of T ences in South Africa, where an updated report of 27
cell–depleting therapy in the HIV1 recipient. However, HIV1 recipients of kidneys from HIV1 donors was
a word of caution is offered by a single-center series that recently published (33). Death-censored graft survival
reported that, in 38 HIV1 recipients treated with antithy- rates were 93%, 84%, and 84% at 1, 3, and 5 years
mocyte globulin induction, pretransplant CD4 counts of (median follow-up, 2.4 years). Updated reviews on
200–349 versus $350 cells per 1 mm3 were associated medication interactions and considerations for drug
with severe lymphopenia (,200 cells per mm3) at 4 resistance, opportunistic infection, and HIV-associated
weeks post-transplant (75% versus 30%; relative risk, 2.6; kidney dysfunction are available (34, 35).
95% CI, 1.3 to 5.1). The lymphopenia was associated Although any increase in organ donation is of
with a more than twofold greater rate of severe infection benefit, an estimate of the impact on the organ shortage
during the first 6 months (28). in the greater Philadelphia area by examination of the
TAC versus CsA seems to be the preferred medical records of HIV-infected patients dying in
calcineurin inhibitor in HIV1 kidney transplant recip- care at six HIV clinics implies a small impact locally
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 327

Table 2. Main characteristics of the approved direct-acting antivirals that are currently used for the treatment of HCV
Direct-Acting Antiviral Dose Adjustment
(Commercial Name), in Renal CNIs
Dose Category Impairment Antiviral Activity Coadministration
Sofosbuvir (Sovaldi), Nucleotide analogue Contraindicated in Genotypes 1-6 No change
tablet 400 mg, once NS5B polymerase patients with High genetic barrier
daily inhibitor GFR,30 ml/min
Simeprevir (Olysio), tablet NS3/4A protease inhibitor No change in renal Genotypes 1 and 4 Contraindicated
150 mg, once daily with impairment Low genetic barrier with CsA
food
Daclatasvir (Daklinza), NS5A inhibitor No change in renal Genotypes 1–4 No change
tablet 60 mg, once daily impairment Low genetic barrier
Ledipasvir/sofosbuvir NS5A inhibitor 1 Contraindicated in Genotypes 1 and 4–6 No change
(Harvoni), tablet nucleotide analogue patients with High genetic barrier
90/400 mg, once daily NS5B polymerase GFR,30 ml/min
inhibitor
Ombitasvir/paritaprevir/ NS5A inhibitor 1 NS3/4A No change in renal Genotypes 1 and 4 CsA: 20% of
ritonavir (Viekirax), protease inhibitor dysfunction Genetic barrier pretreatment total
tablet 12.5/75/50 mg, boosted by ritonavir depending on daily dose; TAC:
two once daily with food boosted HCV genotype 0.2 mg/72 h or
Dasabuvir (Exviera), tablet Non-nucleos(t)ide No change in renal Genotype 1 0.5 mg once weekly
250 mg, every 12 h analogue NS5B dysfunction Low genetic barrier
polymerase inhibitor
Elbasvir/grazoprevir NS5A inhibitor 1 NS3/4A No change in renal Genotypes 1 and 4 Coadministration
(Zepatier), tablet inhibitor dysfunction increases TAC
100/50 mg, once daily concentrations
Velpatasvir/sofosbuvir/ NS5A inhibitor 1 Contraindicated in Genotypes 1–6 No change
(Epclusa), tablet 100/ nucleotide analogue patients with
400 mg, once daily NS5B polymerase GFR,30 ml/min
inhibitor
CNI, calcineurin inhibitor. Modified from Cholongitas E, Pipili C, Papatheodoridis GV: Interferon-free regimens in patients with hepatitis C infection and renal dysfunction or
kidney transplantation. World J Hepatol 9: 180–190, 2017.

(four to five organ donors annually). When extrapolated transplant, a number of case series are now available
nationally, the authors estimated an additional 192 that show efficacy similar to that in the general
kidneys transplanted annually (36). population with a number of different DAA regimens
that include all six HCV genotypes (38, 39). One small
Hepatitis C Virus case series describes successful clearance of HCV in
As described above, HCV1 recipients have in- six HIV/HCV-coinfected recipients, a population that
ferior kidney graft and patient survival compared with had been excluded from study in the C-SURFER Trial
non–HCV-infected recipients. With the new generation (40). The American Association for the Study of Liver
direct-acting antiviral (DAA) agents, cure for HCV is Diseases continues to update its recommendations for
achievable. However, until recently, there were no data use of DAA in HCV (41). A recent review of literature
supporting their use in advanced CKD or after kidney for DAA in CKD and after kidney transplant forms
transplantation. In CKD, a large, multicenter, double- a good reference (Table 2) (42). In patients with CKD,
blind, randomized trial (the C-SURFER Trial) using drug-drug interactions between calcineurin inhibitors
a 12-week course of elbasvir-grazoprevir in 235 pa- and DAAs should always be considered and reviewed
tients with creatinine clearances ,30 ml/min (n5179, (43). These data have encouraged the expansion of use
dialysis dependent) achieved viral eradication in 94% of HCV1 organ donors for not only HCV1 recipients,
of patients with an excellent safety profile. There were thereby shortening the waiting time to transplant, but
no discontinuations due to adverse events (37). After also, the possibility of using HCV1 donors for HCV–
328 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

recipients who have no living donors and are expected 3. Nogueira ML, Estofolete CF, Terzian AC, Mascarin do Vale EP, da
Silva RC, da Silva RF, Ramalho HJ, Fernandes Charpiot IM, Vasilakis
to fare poorly on dialysis and/or have an excessively N, Abbud-Filho M: Zika virus infection and solid organ transplantation:
long expected waiting time for transplant when con- A new challenge. Am J Transplant 17: 791–795, 2017 PubMed
sidering the candidate’s comorbidities (44). Pilot stud- 4. Helanterä I, Anttila VJ, Lappalainen M, Lempinen M, Isoniemi H:
Outbreak of influenza A(H1N1) in a kidney transplant unit-protective
ies in this regard are underway. effect of vaccination. Am J Transplant 15: 2470–2474, 2015 PubMed
5. Kumar D, Campbell P, Hoschler K, Hidalgo L, Al-Dabbagh M, Wilson
L, Humar A: Randomized controlled trial of adjuvanted versus non-
adjuvanted influenza vaccine in kidney transplant recipients. Trans-
Direct-acting antiviral agents can effectively plantation 100: 662–669, 2016 PubMed
clear HCV pre- or post-transplantation. Be- 6. Origüen J, López-Medrano F, Fernández-Ruiz M, Polanco N, Gutiérrez
fore treating a potential transplant candidate, E, González E, Mérida E, Ruiz-Merlo T, Morales-Cartagena A, Pérez-
Jacoiste Asín MA, García-Reyne A, San Juan R, Orellana MA, Andrés
carefully consider the patient’s expected A, Aguado JM: Should asymptomatic bacteriuria be systematically
waiting time and willingness to accept treated in kidney transplant recipients? Results from a randomized
a kidney from an HCV1 deceased donor. controlled trial. Am J Transplant 16: 2943–2953, 2016 PubMed
7. Reischig T, Kacer M, Jindra P, Hes O, Lysak D, Bouda M: Randomized
trial of valganciclovir versus valacyclovir prophylaxis for prevention of
cytomegalovirus in renal transplantation. Clin J Am Soc Nephrol 10:
294–304, 2015 PubMed
Hepatitis B Virus 8. Tedesco-Silva H, Felipe C, Ferreira A, Cristelli M, Oliveira N, Sandes-
Freitas T, Aguiar W, Campos E, Gerbase-DeLima M, Franco M,
Similar to efforts to utilize HIV1 and HCV1 Medina-Pestana J: Reduced incidence of cytomegalovirus infection in
deceased donors in a thoughtful manner and expand kidney transplant recipients receiving everolimus and reduced tacroli-
organ utilization, recent guidelines have been published mus doses. Am J Transplant 15: 2655–2664, 2015 PubMed
9. Höcker B, Zencke S, Pape L, Krupka K, Köster L, Fichtner A, Dello
that clarify the risk of transmission from organ donors Strologo L, Guzzo I, Topaloglu R, Kranz B, König J, Bald M, Webb NJ,
with serologic evidence of hepatitis B virus infection Noyan A, Dursun H, Marks S, Ozcakar ZB, Thiel F, Billing H, Pohl M,
(45). This guideline specifically reviews considerations Fehrenbach H, Schnitzler P, Bruckner T, Ahlenstiel-Grunow T,
Tönshoff B: Impact of everolimus and low-dose cyclosporin on
when using kidneys from deceased donors who are cytomegalovirus replication and disease in pediatric renal transplanta-
hepatitis B surface antigen positive or hepatitis B core tion. Am J Transplant 16: 921–929, 2016 PubMed
antibody positive. In general, kidneys from an hepatitis 10. Puttarajappa C, Bhattarai M, Mour G, Shen C, Sood P, Mehta R, Shah
N, Tevar AD, Humar A, Wu C, Hariharan S: Cytomegalovirus infection
B surface antigen positive donor can be considered in in high-risk kidney transplant recipients receiving thymoglobulin
hepatitis B virus–vaccinated recipients with high hep- induction-a single-center experience. Clin Transplant 30: 1159–1164,
atitis B surface antibody (HbSAb) titers, provided that 2016 PubMed
11. Kaminski H, Garrigue I, Couzi L, Taton B, Bachelet T, Moreau JF,
lifelong chronic antiviral therapy is administered. Déchanet-Merville J, Thiébaut R, Merville P: Surveillance of gd T cells
Additionally, hepatitis B surface antigen positive donor predicts cytomegalovirus infection resolution in kidney transplants. J
kidneys can be considered for recipients with low/ Am Soc Nephrol 27: 637–645, 2016 PubMed
12. Lisboa LF, Egli A, Fairbanks J, O’Shea D, Manuel O, Husain S, Kumar
absent HbSAb titer with a combination of hepatitis B D, Humar A: CCL8 and the immune control of cytomegalovirus in organ
immune globulin administration and continuation of transplant recipients. Am J Transplant 15: 1882–1892, 2015 PubMed
13. Fernández-Ruiz M, Corrales I, Arias M, Campistol JM, Giménez E,
chronic antiviral therapy. Conversely, kidneys from an
Crespo J, López-Oliva MO, Beneyto I, Martín-Moreno PL, Llamas-
hepatitis B core antibody positive donor pose negligible Fuente F, Gutiérrez A, García-Álvarez T, Guerra-Rodríguez R, Calvo
risks and should be considered for all recipients, with N, Fernández-Rodríguez A, Tabernero-Romo JM, Navarro MD,
Ramos-Verde A, Aguado JM, Navarro D; OPERA Study Group:
recommendations for a period of antiviral prophylaxis Association between individual and combined SNPs in genes related to
for up to 1 year only for the recipient with low HbSAb innate immunity and incidence of CMV infection in seropositive kidney
titer. transplant recipients. Am J Transplant 15: 1323–1335, 2015 PubMed
14. Schachtner T, Stein M, Babel N, Reinke P: The loss of BKV-specific
immunity from pretransplantation to posttransplantation identifies
Epstein Barr Virus kidney transplant recipients at increased risk of bkv replication. Am J
Issues related to Epstein Barr virus infection are Transplant 15: 2159–2169, 2015 PubMed
discussed below. 15. Schaenman JM, Korin Y, Sidwell T, Kandarian F, Harre N, Gjertson D,
Lum EL, Reddy U, Huang E, Pham PT, Bunnapradist S, Danovitch GM,
Veale J, Gritsch HA, Reed EF: Increased frequency of BK virus-specific
References polyfunctional CD81 T cells predict successful control of BK viremia after
1. Martin-Gandul C, Mueller NJ, Pascual M, Manuel O: The impact of kidney transplantation. Transplantation 101: 1479–1487, 2017 PubMed
infection on chronic allograft dysfunction and allograft survival after solid 16. Weist BJ, Wehler P, El Ahmad L, Schmueck-Henneresse M, Millward
organ transplantation. Am J Transplant 15: 3024–3040, 2015 PubMed JM, Nienen M, Neumann AU, Reinke P, Babel N: A revised strategy for
2. Blumberg EA, Fishman JA: Zika virus in transplantation: Emerging monitoring BKV-specific cellular immunity in kidney transplant pa-
infection and opportunities. Am J Transplant 17: 599–600, 2017 PubMed tients. Kidney Int 88: 1293–1303, 2015 PubMed
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 329

17. Hirsch HH, Yakhontova K, Lu M, Manzetti J: BK polyomavirus 32. Locke JE, Shelton BA, Reed RD, MacLennan PA, Mehta S, Sawinski
replication in renal tubular epithelial cells is inhibited by sirolimus, D, Segev DL: Identification of optimal donor-recipient combinations
but activated by tacrolimus through a pathway involving FKBP-12. Am among human immunodeficiency virus (HIV)-positive kidney trans-
J Transplant 16: 821–832, 2016 PubMed plant recipients. Am J Transplant 16: 2377–2383, 2016 PubMed
18. Wunderink HF, van der Meijden E, van der Blij-de Brouwer CS, Mallat 33. Muller E, Barday Z, Mendelson M, Kahn D: HIV-positive-to-HIV-
MJ, Haasnoot GW, van Zwet EW, Claas EC, de Fijter JW, Kroes AC, positive kidney transplantation–results at 3 to 5 years. N Engl J Med
Arnold F, Touzé A, Claas FH, Rotmans JI, Feltkamp MC: Pretrans- 372: 613–620, 2015 PubMed
plantation donor-recipient pair seroreactivity against BK polyomavirus 34. Patel SJ, Kuten SA, Musick WL, Gaber AO, Monsour HP, Knight RJ:
predicts viremia and nephropathy after kidney transplantation. Am J Combination drug products for HIV-a word of caution for the transplant
Transplant 17: 161–172, 2017 PubMed clinician. Am J Transplant 16: 2479–2482, 2016 PubMed
19. Abend JR, Changala M, Sathe A, Casey F, Kistler A, Chandran S, 35. Boyarsky BJ, Durand CM, Palella FJ Jr., Segev DL: Challenges and
Howard A, Wojciechowski D: Correlation of BK virus neutralizing clinical decision-making in HIV-to-HIV transplantation: Insights from
serostatus with the incidence of BK viremia in kidney transplant the HIV literature. Am J Transplant 15: 2023–2030, 2015 PubMed
recipients. Transplantation 101: 1495–1505, 2017 PubMed 36. Richterman A, Sawinski D, Reese PP, Lee DH, Clauss H, Hasz RD,
20. Schwarz A, Linnenweber-Held S, Heim A, Framke T, Haller H, Schmitt Thomasson A, Goldberg DS, Abt PL, Forde KA, Bloom RD, Doll SL,
C: Viral origin, clinical course, and renal outcomes in patients with BK Brady KA, Blumberg EA: An assessment of HIV-infected patients
virus infection after living-donor renal transplantation. Transplantation dying in care for deceased organ donation in a United States urban
100: 844–853, 2016 PubMed center. Am J Transplant 15: 2105–2116, 2015 PubMed
21. Thangaraju S, Gill J, Wright A, Dong J, Rose C, Gill J: Risk factors for 37. Roth D, Nelson DR, Bruchfeld A, Liapakis A, Silva M, Monsour H Jr.,
BK polyoma virus treatment and association of treatment with kidney Martin P, Pol S, Londoño MC, Hassanein T, Zamor PJ, Zuckerman E,
transplant failure: Insights from a paired kidney analysis. Transplanta- Wan S, Jackson B, Nguyen BY, Robertson M, Barr E, Wahl J, Greaves W:
tion 100: 854–861, 2016 PubMed Grazoprevir plus elbasvir in treatment-naive and treatment-experienced
22. Sawinski D, Forde KA, Trofe-Clark J, Patel P, Olivera B, Goral S, patients with hepatitis C virus genotype 1 infection and stage 4-5 chronic
Bloom RD: Persistent BK viremia does not increase intermediate-term kidney disease (the C-SURFER study): A combination phase 3 study.
graft loss but is associated with de novo donor-specific antibodies. J Am Lancet 386: 1537–1545, 2015 PubMed
Soc Nephrol 26: 966–975, 2015 PubMed 38. Kamar N, Marion O, Rostaing L, Cointault O, Ribes D, Lavayssière L,
23. Seifert ME, Gunasekaran M, Horwedel TA, Daloul R, Storch GA, Esposito L, Del Bello A, Métivier S, Barange K, Izopet J, Alric L:
Mohanakumar T, Brennan DC: Polyomavirus reactivation and immune Efficacy and safety of sofosbuvir-based antiviral therapy to treat
responses to kidney-specific self-antigens in transplantation. J Am Soc hepatitis c virus infection after kidney transplantation. Am J Transplant
Nephrol 28: 1314–1325, 2017 PubMed 16: 1474–1479, 2016 PubMed
24. Zeng G, Huang Y, Huang Y, Lyu Z, Lesniak D, Randhawa P: Antigen- 39. Sawinski D, Kaur N, Ajeti A, Trofe-Clark J, Lim M, Bleicher M, Goral
specificity of T cell infiltrates in biopsies with t cell-mediated rejection S, Forde KA, Bloom RD: Successful treatment of hepatitis C in renal
and BK polyomavirus viremia: Analysis by next generation sequencing. transplant recipients with direct-acting antiviral agents. Am J Transplant
Am J Transplant 16: 3131–3138, 2016 PubMed 16: 1588–1595, 2016 PubMed
25. Locke JE, Reed RD, Mehta SG, Durand C, Mannon RB, MacLennan P, 40. Sawinski D, Lee DH, Doyle AM, Blumberg EA: Successful posttrans-
Shelton B, Martin MY, Qu H, Shewchuk R, Segev DL: Center-level plant treatment of hepatitis C with ledipasvir-sofosbuvir in HIV1 kidney
experience and kidney transplant outcomes in HIV-infected recipients. transplant recipients. Transplantation 101: 974–979, 2017 PubMed
Am J Transplant 15: 2096–2104, 2015 PubMed 41. Panel AIHG; AASLD/IDSA HCV Guidance Panel: Hepatitis C guidance:
26. Husson J, Stafford K, Bromberg J, Haririan A, Sparkes T, Davis C, Redfield AASLD-IDSA recommendations for testing, managing, and treating adults
R, Amoroso A: Association between duration of human immunodeficiency infected with hepatitis C virus. Hepatology 62: 932–954, 2015 PubMed
virus (HIV)-1 viral suppression prior to renal transplantation and acute 42. Cholongitas E, Pipili C, Papatheodoridis GV: Interferon-free regimens
cellular rejection. Am J Transplant 17: 551–556, 2017 PubMed in patients with hepatitis C infection and renal dysfunction or kidney
27. Kucirka LM, Durand CM, Bae S, Avery RK, Locke JE, Orandi BJ, transplantation. World J Hepatol 9: 180–190, 2017 PubMed
McAdams-DeMarco M, Grams ME, Segev DL: Induction immunosup- 43. Badri P, Dutta S, Coakley E, Cohen D, Ding B, Podsadecki T, Bernstein B,
pression and clinical outcomes in kidney transplant recipients infected Awni W, Menon R: Pharmacokinetics and dose recommendations for
with human immunodeficiency virus. Am J Transplant 16: 2368–2376, cyclosporine and tacrolimus when coadministered with ABT-450, ombitasvir,
2016 PubMed and dasabuvir. Am J Transplant 15: 1313–1322, 2015 PubMed
28. Suarez JF, Rosa R, Lorio MA, Morris MI, Abbo LM, Simkins J, Guerra 44. Reese PP, Abt PL, Blumberg EA, Goldberg DS: Transplanting hepatitis
G, Roth D, Kupin WL, Mattiazzi A, Ciancio G, Chen LJ, Burke GW, C-positive kidneys. N Engl J Med 373: 303–305, 2015 PubMed
Goldstein MJ, Ruiz P, Camargo JF: Pretransplant CD4 count influences 45. Huprikar S, Danziger-Isakov L, Ahn J, Naugler S, Blumberg E, Avery
immune reconstitution and risk of infectious complications in human RK, Koval C, Lease ED, Pillai A, Doucette KE, Levitsky J, Morris MI,
immunodeficiency virus-infected kidney allograft recipients. Am J Lu K, McDermott JK, Mone T, Orlowski JP, Dadhania DM, Abbott K,
Transplant 16: 2463–2472, 2016 PubMed Horslen S, Laskin BL, Mougdil A, Venkat VL, Korenblat K, Kumar V,
29. Gathogo E, Harber M, Bhagani S, Levy J, Jones R, Hilton R, Davies G, Grossi P, Bloom RD, Brown K, Kotton CN, Kumar D: Solid organ
Post FA; UK HIV Kidney Transplantation Study Group: Impact of transplantation from hepatitis B virus-positive donors: Consensus
tacrolimus compared with cyclosporin on the incidence of acute guidelines for recipient management. Am J Transplant 15: 1162–
allograft rejection in human immunodeficiency virus-positive kidney 1172, 2015 PubMed
transplant recipients. Transplantation 100: 871–878, 2016 PubMed
30. Sawinski D, Forde KA, Eddinger K, Troxel AB, Blumberg E, Tebas P,
Abt PL, Bloom RD: Superior outcomes in HIV-positive kidney trans-
plant patients compared with HCV-infected or HIV/HCV-coinfected
recipients. Kidney Int 88: 341–349, 2015 PubMed
31. Locke JE, Mehta S, Reed RD, MacLennan P, Massie A, Nellore A, Malignancy and Kidney Transplantation
Durand C, Segev DL: A national study of outcomes among HIV-
infected kidney transplant recipients. J Am Soc Nephrol 26: 2222–2229, There is an increased risk of certain malignancies
2015 PubMed after kidney transplant compared with in the general
330 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

2000 to 2011, the incidence of nonmelanoma skin cancer


was diagnosed in 5.7%, viral-linked cancer was diagnosed
in 1.9%, and other cancers were diagnosed in 6.3%. The
mortality risk was more than three times higher in those
with viral-linked malignancy compared with individuals
who were malignancy free by 3 years. Overall, malig-
nancy accounted for 3%–5.5% of total inpatient Medicare
expenditures and 1.5%–3.3% of outpatient expenditures in
the first 3 years after transplant (4). A separate analysis
described a twofold increased risk of death in transplant
recipients after they were diagnosed with de novo cancer
compared with in the general population, despite the
competing risk of death that transplant recipients face
otherwise (5). These data help define the magnitude of the
Figure 15. Observed cumulative incidence of lung cancer ongoing problem of cancer in the transplant population and
during the first 10 years after renal transplantation. All support ongoing screening of CKD patients as they pass
between-group comparisons were significant by pairwise
from phases of ESRD to transplant and back to ESRD.
log rank test (P,0.001). Reprinted with permission from
Opelz G, Döhler B: Influence of current and previous
smoking on cancer and mortality after kidney transplanta- Risk Factors for Malignancy in the Transplant
tion. Transplantation 100: 227–232, 2016. Recipient
To define appropriate surveillance strategies, one
population. Quantifying the excess risk of malignancy must first be able to identify potential risk factors in
specifically due to kidney transplantation and attendant addition to common malignancy patterns. Perhaps the
immunosuppression has been challenging. One issue is most compelling recent analysis regarding risk factor
the difficulty of accurate reporting. When comparing identification and management comes from a large
malignancy rates of the Scientific Registry of Transplant registry study addressing the risk of smoking and the
Recipients (SRTR) with those of 15 linked cancer benefits of smoking cessation. In 45,548 kidney trans-
registries, only 36.8% of cancers were identified in plant recipients from 1995 to 2012, smoking cessation
both sources, with 47.5% additional cases documented before transplant led to a significant reduction in fre-
only in cancer registries and 15.7% only identified in quency of many malignancies, most prominently lung
the SRTR (1). Thus, malignancy data solely utilizing cancer (Figure 15 and Table 3) (6). Persistent smoking
the SRTR are likely an underestimate of the magnitude was associated with 50% increase in all-cause graft
of the difference. When using linked data, certain loss, with similar findings for death and death-censored
malignancies have a clear relationship with transplanta- graft loss (P,0.001), whereas those who stopped
tion: Kaposi sarcoma, non-Hodgkin lymphoma (NHL), smoking before transplant had only a modest 10%
lip cancer, and nonepithelial skin cancer. Others are increase in all-cause graft loss compared with non-
linked more strongly with renal failure itself (kidney smokers. Thus, smoking cessation should be strongly
cancer and thyroid cancer) (2). The risk of cancer in recommended at the time of transplant evaluation and
retransplanted recipients compared with primary kid- while on the waiting list.
ney transplant recipients, again using linked databases,
is similar, except for renal cell carcinoma (RCC) of the
native kidneys (twofold increased risk; adjusted in- Smoking cessation before transplant is as-
cidence risk ratio [aIRR], 2.03; 95% confidence interval sociated with a significant reduction in risk
[95% CI], 1.45 to 2.77), perhaps due to increased kidney for lung cancer and other virus-associated
disease exposure or dialysis time (3). The incidence, malignancies after transplant.
clinical consequences, and economic costs of malignancy
after transplantation are significant. When segregating
malignancies into three types, nonmelanoma skin cancer, A history of malignancy or specific types of cancer
viral linked, and other cancers, in 67,157 recipients from before transplant may compel heightened screening after
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 331

Table 3. The standardized incidence ratio (SIR; the ratio of the incidence of malignancy in the kidney transplant
population to the incidence in the general population) of various malignancy subtypes during the first 10 years after
renal transplantation by smoking status

ICD10, International statistical classification of diseases and related health problems, 10th revision; CNS, central nervous system. Reprinted with permission from Opelz G,
Döhler B: Influence of current and previous smoking on cancer and mortality after kidney transplantation. Transplantation 100: 227–232, 2016.

transplant. A recent United States registry analysis of kidney transplant recipients from 1987 to 2010 (10).
showed that pretransplant skin cancer was associated There was a more than fivefold increased risk (standard-
with post-transplant malignancy (PTM) for not only skin ized incidence ratio, 5.68; 95% CI, 5.27 to 6.13) of RCC,
cancer (subhazard ratio [SHR], 2.92; 95% CI, 2.52 to particularly of the papillary subtype, compared with in
3.39) but also, solid tumors (SHR, 1.44; 95% CI, 1.04 to the general population. Additive risk was noted with
1.99) and post-transplant lymphoproliferative disease blacks, men, increasing dialysis time, and the use of
(PTLD; SHR, 1.93; 95% CI, 1.01 to 3.66) (7). The 5- induction therapy but was not noted with specific
year cumulative incidence of any malignancy in those maintenance immunosuppression regimen. Interest-
with pretransplant skin cancer was 31.6% versus 7.4% ingly, for colon cancer, an increased risk was present
(P,0.001) in those with no such history. Conversely, an in kidney transplant patients treated with cyclosporine
Australian and New Zealand Dialysis and Transplant and azathioprine but not in those treated with tacrolimus
Registry analysis of prior malignancy history (excluding and mycophenolate, suggesting differential effects of
melanocytic skin cancers) failed to find an association immunosuppression on different types of solid tumors
between pretransplant malignancy history and post- (11).
transplant recurrence or a second primary malignancy The question of which immunosuppression reg-
(8). A systematic meta-analysis of solid organ transplant imen confers the greatest risk (or the least risk) for
recipients reported an association of pretransplant malig- development of PTM is an ongoing subject of study.
nancy (of any type) with all-cause mortality (ten studies, Two recent studies implicate azathioprine as a primary
1.5-fold increased risk), cancer-specific mortality (three offender in the development of skin cancer, but one
studies, threefold increased risk), and post-transplant de provocatively acquits mycophenolate use as a risk
novo malignancy (seven studies, twofold increased risk) factor. In a meta-analysis of 27 studies that evaluated
(9). Perhaps the Australian/New Zealand data are dis- the risk of skin cancer in relation to azathioprine
tinctly different from the other analyses due to differ- treatment, there was a significant risk of squamous cell
ences in transplant eligibility criteria or differences in carcinoma (SCC; hazard ratio [HR], 1.56; 95% CI,
point prevalence of nonmelanoma skin cancer. Regard- 1.11 to 2.18) but not of other types of skin cancers
ing risk factors for specific types of PTM, the risk of (12). In a second report of solid organ transplant
RCC was specifically studied in a linked registry analysis recipients (not limited to kidney transplant), a matched
332 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

Table 4. Association between induction therapy and incident virus–related cancers


Malignancy Type Cancers, N Incidencea aIRR (95% CI) P Value
NHL
No induction 377 142.1 Reference
Polyclonal 125 131.6 0.96 (0.77 to 1.20) 0.70
Muromonab-CD3 80 210.9 1.37 (1.06 to 1.76) 0.02
Alemtuzumab 15 216.2 1.79 (1.02 to 3.14) 0.04
Anti-IL2R 96 114.9 0.82 (0.65 to 1.05) 0.10
Non-NHL VRCs
No induction 164 61.8 Reference
Polyclonal 56 60.0 1.11 (0.82 to 1.53) 0.50
Muromonab-CD3 25 65.9 1.02 (0.65 to 1.58) 0.90
Alemtuzumab 4 57.6 2.05 (0.66 to 6.33) 0.20
Anti-IL2R 53 63.5 1.09 (0.78 to 1.51) 0.60
All VRCs
No induction 541 203.9 Reference
Polyclonal 181 190.6 1.01 (0.84 to 1.21) 0.90
Muromonab-CD3 104 276.8 1.26 (1.01 to 1.57) 0.04
Alemtuzumab 19 273.8 1.84 (1.11 to 3.03) 0.02
Anti-IL2R 149 178.4 0.90 (0.74 to 1.10) 0.30
aIRRs for VRCs after kidney transplantation. The number of cancers diagnosed in each stratum is shown (N). VRCs included NHL, Hodgkin lymphoma, human papillomavirus-
related cancers (cancers of the cervix, vagina, vulva, anus, penis, oropharynx, and tonsil), Kaposi sarcoma, and liver cancer. All models were adjusted for recipient age, sex, race,
retransplantation, zero HLA mismatch status, receipt of living donor kidney, and year of transplant. Only transplants completed in years with at least 20 recipients receiving the
induction therapy of interest were included: polyclonal: 1987–2009; muromonab-CD3: 1987–2003; alemtuzumab: 2001–2009; and anti-IL2R: 1995–2009. Anti-IL2R, anti–IL-2
receptor (daclizumab and basiliximab); VRC, virus-related cancer. Modified from Hall EC, Engels EA, Pfeiffer RM, Segev DL: Association of antibody induction
immunosuppression with cancer after kidney transplantation. Transplantation 99: 1051–1057, 2015.
a
Per 100,000 person-years.

control study of 170 affected patients and 324 controls, (aIRR, 1.50; 95% CI, 1.06 to 2.14) (Tables 4 and 5).
azathioprine use again was associated with increased Both thyroid and colorectal cancer, incidence risk ratios
risk of SCC (odds ratio, 2.67; 95% CI; 1.23 to 5.76), were also higher in the cohort treated with alemtuzumab.
whereas an inverse relationship was noted with mycophe- In general, these data point to the safety of IL-2 receptor
nolate use with or without concomitant use of tacrolimus antagonists, the lack of increased risk of NHL with
(approximately 50% reduction in odds ratio) (13). The polyclonal depleting therapy, and the need for tailored
authors concluded that the excess risk of SCC his- monitoring for other malignancies based on the specific
torically noted with azathioprine is substantially reduced induction agent.
using modern immunosuppressive regimens. Beyond Finally, the question of the effects on PTM of
azathioprine use, the association of induction therapy mammalian target of rapamycin inhibitors (mTORis)
and malignancy risk has remained a question. Using has been further clarified. A large linked registry
linked registry analyses, aIRRs of NHL, melanoma, analysis of cancer registries and national pharmacy
and the solid tumors lung, colorectal, kidney, and thyroid claims examined cancer incidence in 32,604 kidney
cancers were calculated based on use of different types of transplant recipients. Non-melanoma skin cancers were
induction therapy (14). Nondepleting induction agents excluded from this analysis (15). Overall, cancer
(IL-2 receptor antagonists) were not associated with a incidence was modestly and nonsignificantly lower
higher aIRR of any malignancy compared with kidney with sirolimus use (HR, 0.88; 95% CI, 0.70 to 1.11). A
transplant recipients who did not receive induction. second analysis that included skin cancer data from
Conversely, muromonab-CD3 (OKT3) and alemtuzumab a single-registry analysis from the Collaborative Trans-
were associated with an increased risk of NHL (aIRR, plant Study of 78,146 primary deceased donor kidney
1.37; 95% CI, 1.06 to 1.76 and aIRR, 1.79; 95% CI, transplant recipients from 1999 to 2013 showed only
1.02 to 3.14, respectively). Polyclonal induction thera- a reduction in the incidence of basal cell carcinoma
pies, such as antithymocyte globulin preparations, with mTORi (HR, 0.56; P50.004), with no reduction
were associated with an increased risk of melanoma in incidence or risk of any other type of tumor, including
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 333

Table 5. Associations of induction agents with nonvirus-related cancers


Malignancy Type Cancers, N Incidencea aIRR (95% CI) P Value
Lung cancer
No induction 297 111.9 Reference
Polyclonal 127 133.7 1.14 (0.91 to 1.43) 0.20
Muromonab-CD3 46 121.3 1.03 (0.75 to 1.41) 0.80
Alemtuzumab 7 100.9 0.93 (0.42 to 2.07) 0.90
Anti-IL2R 77 92.2 0.73 (0.56 to 0.95) 0.02
Kidney cancer
No induction 261 98.4 Reference
Polyclonal 121 127.4 1.08 (0.85 to 1.38) 0.50
Muromonab-CD3 34 89.6 0.93 (0.65 to 1.34) 0.70
Alemtuzumab 11 158.5 1.20 (0.63 to 2.30) 0.60
Anti-IL2R 102 122.1 1.03 (0.81 to 1.32) 0.80
Colorectal cancer
No induction 164 61.8 Reference
Polyclonal 60 63.2 1.14 (0.82 to 1.58) 0.40
Muromonab-CD3 20 52.7 0.77 (0.48 to 1.24) 0.30
Alemtuzumab 7 100.9 2.46 (1.03 to 5.91) 0.04
Anti-IL2R 47 56.3 0.88 (0.62 to 1.25) 0.50
Thyroid cancer
No induction 103 38.8 Reference
Polyclonal 26 27.4 0.65 (0.40 to 1.04) 0.07
Muromonab-CD3 5 131.8 0.32 (0.12 to 0.80) 0.01
Alemtuzumab 10 144.1 3.37 (1.55 to 7.33) 0.002
Anti-IL2R 27 32.3 0.81 (0.51 to 1.29) 0.40
Melanomab
No induction 107 68.7 Reference
Polyclonal 56 103.4 1.50 (1.06 to 2.14) 0.02
Muromonab-CD3 11 51.1 0.77 (0.41 to 1.45) 0.40
Alemtuzumab 3 81.1 1.14 (0.33 to 3.95) 0.80
Anti-IL2R 38 88.1 1.15 (0.77 to 1.72) 0.50
aIRRs for nonvirus-related cancers after kidney transplantation. The number of cancers diagnosed in each stratum is shown (N). All models were adjusted for recipient age, sex,
race, retransplantation, zero HLA mismatch status, receipt of living donor kidney, and year of transplant. Only transplants completed in years with at least 20 recipients receiving
the induction therapy of interest were included: polyclonal: 1987–2009; muromonab-CD3: 1987–2003; alemtuzumab: 2001–2009; and anti-IL2R: 1995–2009. Anti-IL2R, anti–IL-
2 receptor (daclizumab and basiliximab). Modified from Hall EC, Engels EA, Pfeiffer RM, Segev DL: Association of antibody induction immunosuppression with cancer after
kidney transplantation. Transplantation 99: 1051–1057, 2015.
a
Per 100,000 person-years.
b
Among non-Hispanic white kidney recipients.

SCC (16). A final large, longitudinal analysis from tially different from those used in the general population.
Australia and New Zealand, where utilization of mTORi For prostate cancer screening, prostate-specific antigen
has been higher due to its hypothesized potential (PSA)–based screening per American Urological As-
impact on skin cancer, showed a higher risk of all-cause sociation screening guidelines of 3782 men undergoing
mortality with mTORi use during a median follow-up of 7 transplant evaluation did not lead to improved patient
years (17). Taken collectively, the application of mTORi, survival after transplant but did increase the time to
specifically for its antineoplastic properties in the pre- listing for those with a positive test result (18).
vention of PTM, is not supported by existing data. Furthermore, those screened had a reduced likelihood
of receiving a transplant, regardless of PSA value.
Transplant Candidate Cancer Screening Thus, routine PSA screening should be discouraged,
In general, there have been few studies specific to because it may create an unnecessary barrier to trans-
the transplant candidate population to guide whether plantation. For colon cancer, the findings of 70 patients
screening practices for malignancies should be substan- with ESRD undergoing colonoscopy as a function of
334 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

kidney transplant evaluation were compared with those Recipient Screening for Malignancy
from 70 non-ESRD controls matched for age, sex, and As in the setting of transplant candidate malig-
endoscopist (19). Adenomatous polyps were present in nancy screening, perhaps even less has been published
54.3% of ESRD subjects versus 32.9% of controls specific to screening practices after transplant. A sys-
(P,0.01), supporting the practice of colonoscopy tematic review of the literature identified 13 relevant
screening in the candidate transplant evaluation clinical practice guidelines for all organ transplants, of
similar to the general population. Regarding the use which eight were specific to kidney transplant recip-
of mammography for breast cancer screening, a study ients (24). The reviewers noted specifically that “there
of 541 women $40 years old undergoing kidney is a lack of direct evidence to support cancer screen-
transplant evaluation from 2006 to 2012 showed that ing recommendations” (24). Recommendations were
patients ages 40–49 and $50 years old had similar based primarily on the authors’ interpretations of risk
biopsy rates and breast malignancy (20). Although in the context of general population risk rather than
not a cost-effectiveness analysis, these data generally evidence in a transplant population, without input
support the American Cancer Society recommenda- from oncology specialists, primary care providers,
tions to start annual mammogram at age 40 years old and public health experts. Despite these limitations, the
in transplant candidates rather than the US Preven- relationship of immunosuppression and virus-mediated
tative Services Task Force recommendation of bi- cancers is strong. The appropriate screening for Epstein
ennial mammogram beginning at age 50 years old. In Barr virus (EBV)–related cancer remains an area of active
cases where breast cancer is identified, genomic research.
profiling may be increasingly possible to identify Routine screening for EBV by PCR has been
low-risk cancers and shorten the “disease-free in- generally ineffective in the prediction of risk for PTLD
active status” that is often 2–5 years long without this in the adult kidney transplant population: a positive
information (21). EBV PCR is insensitive and nonspecific, hindered by
Regarding dermatology clearance in individuals the lack of standardized assay or source (i.e., whole
who have had a prior skin cancer, there has been a blood or plasma) (reviewed in ref. 25). An additional
consensus report from the International Transplant Skin confounder is the increase in the number of cases of
Cancer Collaborative that describes a waiting time con- EBV-related PTLD over time (from 10% during the
struct for each skin cancer subtype based on its staging period 1990–1995 to 48% of cases during the period
(22). Using a 60% 5-year post-transplant survival as a 2008–2013 in one large series) (26). Parenthetically,
threshold for acceptable expected survival post-transplant, remission rates and survival were found to be similar
the International Transplant Skin Cancer Collaborative whether PTLD was EBV1 or EBV–. Although the data
advocates for a 2-year waiting time for SCC that is high supporting EBV PCR screening for PTLD risk have
risk (.2 cm, high-risk location [e.g., ear, lip, scalp, or been inconsistent, a positive test may generally predict
temple], recurrent lesion, or poorly differentiated), and an overimmunosuppressed state that is associated with
waiting times of 2 years for melanoma stage Ia, 5 years for a higher malignancy risk in general. In 669 patients
stage Ib, and 5–10 years for stages IIa, IIb, and IIIa; screened annually with EBV PCR from whole blood (if
patients with stages IIc, IIIb, IIIc, and IV would not be positive, repeated on each routine visit), 147 patients
candidates for transplant. Finally, the question of screen- had a persistently high EBV viral load (27). Persistent
ing for RCC via ultrasound remains controversial. There is EBV replication was associated with a 70% increase in
clearly an increased risk of RCC in patients with ESRD as the risk of post-transplant cancer (of all types) during
described previously. In a surveillance study of 2642 a mean follow-up of 94 months, with an incidence of
transplant candidates who underwent renal ultrasound at cancer of 26.7% in 45 recipients with a persistent PCR
the time of evaluation followed by annual screening (23) of .10,000 copies per 1 ml. These findings are pro-
71 RCCs were identified (52 on initial screening), vocative, and a standard of care regarding EBV PCR
with independent associations with men and black screening remains undefined. In the absence of screen-
race. Whether this 2%–3% detection rate warrants ing, prophylactic antiviral treatment of those at high risk
such a monitoring strategy remains unanswered, and (EBV-seronegative recipients of organs from seroposi-
no firm consensus has been adopted across the tive donors) has been considered a strategy in minimiz-
transplant community. ing risk for PTLD. Unfortunately, a meta-analysis of 31
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 335

studies found no effect of this strategy on the prediction pembrolizumab, with no allograft rejection with ipilimumab
or the incidence of PTLD (28). (but potentially lesser antitumor effect [33] than the PD-1
mAbs [34]).
Cancer Management in the Transplant Recipient
A common question in the transplant patient with
malignancy is whether altering immunosuppression is
safe and/or effective. Regarding the safety of immuno- Checkpoint inhibitors targeting PD-1
suppression withdrawal during chemotherapy for PTLD, (nivolumab and pembrolizumab) and CTLA-
a study of 24 patients investigated immunosuppression 4 (ipilimumab) are used to treat metastatic
cessation during chemotherapy and resumption at lower malignancies. These agents restore T cell
dose (calcineurin inhibitor at 50%, prednisolone #10 function; PD-1 inhibition can place the trans-
mg daily, no third agent) approximately 6 weeks after plant at very high risk for rejection.
chemotherapy (29). Compared with 83 matched controls,
no difference was noted in time to $25% increase in
serum creatinine. Only one patient with PTLD experi- References
1. Yanik EL, Nogueira LM, Koch L, Copeland G, Lynch CF, Pawlish KS,
enced a $25% increase in serum creatinine within 6
Finch JL, Kahn AR, Hernandez BY, Segev DL, Pfeiffer RM, Snyder JJ,
months of withdrawal, which was associated with pro- Kasiske BL, Engels EA: Comparison of cancer diagnoses between the
gressive PTLD and death. This provides some comfort in US solid organ transplant registry and linked central cancer registries.
Am J Transplant 16: 2986–2993, 2016 PubMed
eliminating immunosuppression for short intervals during 2. Yanik EL, Clarke CA, Snyder JJ, Pfeiffer RM, Engels EA: Variation in
periods of aggressive chemotherapy. There are no recent cancer incidence among patients with ESRD during kidney function
publications addressing immunosuppression reduction/ and nonfunction intervals. J Am Soc Nephrol 27: 1495–1504, 2016 PubMed
3. Kalil RS, Lynch CF, Engels EA: Risk of cancer in retransplants
withdrawal/transition during treatment for solid tumors. compared to primary kidney transplants in the United States. Clin
Novel therapeutic targets for control of a variety Transplant 29: 944–950, 2015 PubMed
of malignancies include the programmed cell death 4. Dharnidharka VR, Naik AS, Axelrod D, Schnitzler MA, Xiao H,
Brennan DC, Segev DL, Randall H, Chen J, Kasiske B, Lentine KL:
pathway (programmed cell death protein 1 [PD-1] Clinical and economic consequences of early cancer after kidney
ligand and its receptor PD-1) and the cytotoxic T transplantation in contemporary practice. Transplantation 101: 858–
lymphocyte–associated antigen-4 (CTLA-4) pathway 866, 2017 PubMed
5. Acuna SA, Fernandes KA, Daly C, Hicks LK, Sutradhar R, Kim SJ,
(CTLA-4 binding to its ligand B7-1/2) (30). PD-1 is Baxter NN: Cancer mortality among recipients of solid-organ trans-
expressed on T cells (as well as B cells, and natural plantation in Ontario, Canada. JAMA Oncol 2: 463–469, 2016 PubMed
killer cells) and binds PD-1 ligand expressed by tumor 6. Opelz G, Döhler B: Influence of current and previous smoking on
cancer and mortality after kidney transplantation. Transplantation 100:
cells. This ligand binding leads to T cell exhaustion, 227–232, 2016 PubMed
creating a state of immunosuppression. Like the PD-1 7. Kang W, Sampaio MS, Huang E, Bunnapradist S: Association of
pathway, CTLA-4 is expressed on T cells, and ligand pretransplant skin cancer with posttransplant malignancy, graft failure
and death in kidney transplant recipients. Transplantation 101: 1303–
binding leads to T cell exhaustion. Two mAbs to PD-1, 1309, 2017 PubMed
nivolumab and pembrolizumab, inhibit T cell exhaus- 8. Viecelli AK, Lim WH, Macaskill P, Chapman JR, Craig JC, Clayton P,
Cohney S, Carroll R, Wong G: Cancer-specific and all-cause mortality
tion and increase tumor surveillance. These agents have
in kidney transplant recipients with and without previous cancer.
shown efficacy in advanced melanoma and nonsmall cell Transplantation 99: 2586–2592, 2015 PubMed
lung cancer, and they are under investigation in multiple 9. Acuna SA, Huang JW, Daly C, Shah PS, Kim SJ, Baxter NN: Outcomes
of solid organ transplant recipients with preexisting malignancies in
other tumor types. In transplant recipients (kidney and remission: A systematic review and meta-analysis. Transplantation
other organs), these agents have been associated with 101: 471–481, 2017 PubMed
a near-universal incidence of rejection, often leading to 10. Karami S, Yanik EL, Moore LE, Pfeiffer RM, Copeland G, Gonsalves
L, Hernandez BY, Lynch CF, Pawlish K, Engels EA: Risk of renal cell
dialysis dependence in the kidney transplant setting (31, carcinoma among kidney transplant recipients in the United States. Am J
32). The mAb ipilimumab binds and inhibits this CTLA- Transplant 16: 3479–3489, 2016 PubMed
4–induced T cell suppression. Although effective in 11. Safaeian M, Robbins HA, Berndt SI, Lynch CF, Fraumeni JF Jr., Engels
EA: Risk of colorectal cancer after solid organ transplantation in the
preventing tumor progression in metastatic disease, re- United States. Am J Transplant 16: 960–967, 2016 PubMed
jection has not been reported in solid organ transplant 12. Jiyad Z, Olsen CM, Burke MT, Isbel NM, Green AC: Azathioprine and
recipients treated with ipilimumab. A recent review of risk of skin cancer in organ transplant recipients: Systematic review and
meta-analysis. Am J Transplant 16: 3490–3503, 2016 PubMed
case reports was published describing 15 cases (in kidney 13. Coghill AE, Johnson LG, Berg D, Resler AJ, Leca N, Madeleine MM:
and liver transplants) using ipilimumab, nivolumab, and Immunosuppressive medications and squamous cell skin carcinoma:
336 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

Nested case-control study within the Skin Cancer after Organ Trans- organ transplant recipients: A systematic review. Am J Transplant 17:
plant (SCOT) Cohort. Am J Transplant 16: 565–573, 2016 PubMed 770–781, 2017 PubMed
14. Hall EC, Engels EA, Pfeiffer RM, Segev DL: Association of antibody 29. Taylor E, Jones M, Hourigan MJ, Johnson DW, Gill DS, Isbel N, Hawley
induction immunosuppression with cancer after kidney transplantation. CM, Marlton P, Gandhi MK, Campbell SB, Mollee P: Cessation of
Transplantation 99: 1051–1057, 2015 PubMed immunosuppression during chemotherapy for post-transplant lymphopro-
15. Yanik EL, Gustafson SK, Kasiske BL, Israni AK, Snyder JJ, Hess GP, liferative disorders in renal transplant patients. Nephrol Dial Transplant
Engels EA, Segev DL: Sirolimus use and cancer incidence among 30: 1774–1779, 2015 PubMed
US kidney transplant recipients. Am J Transplant 15: 129–136, 30. Postow MA, Callahan MK, Wolchok JD: Immune checkpoint blockade
2015 PubMed in cancer therapy. J Clin Oncol 33: 1974–1982, 2015 PubMed
16. Opelz G, Unterrainer C, Süsal C, Döhler B: Immunosuppression with 31. Lipson EJ, Bagnasco SM, Moore J Jr., Jang S, Patel MJ, Zachary AA,
mammalian target of rapamycin inhibitor and incidence of post-transplant Pardoll DM, Taube JM, Drake CG: Tumor regression and allograft
cancer in kidney transplant recipients. Nephrol Dial Transplant 31: 1360– rejection after administration of anti-PD-1. N Engl J Med 374: 896–898,
1367, 2016 PubMed 2016 PubMed
17. Badve SV, Pascoe EM, Burke M, Clayton PA, Campbell SB, Hawley 32. Boils CL, Aljadir DN, Cantafio AW: Use of the PD-1 pathway inhibitor
CM, Lim WH, McDonald SP, Wong G, Johnson DW: Mammalian nivolumab in a renal transplant patient with malignancy. Am J Trans-
target of rapamycin inhibitors and clinical outcomes in adult plant 16: 2496–2497, 2016 PubMed
kidney transplant recipients. Clin J Am Soc Nephrol 11: 1845– 33. Robert C, Schachter J, Long GV, Arance A, Grob JJ, Mortier L, Daud
1855, 2016 PubMed A, Carlino MS, McNeil C, Lotem M, Larkin J, Lorigan P, Neyns B,
18. Vitiello GA, Sayed BA, Wardenburg M, Perez SD, Keith CG, Canter Blank CU, Hamid O, Mateus C, Shapira-Frommer R, Kosh M, Zhou H,
DJ, Ogan K, Pearson TC, Turgeon N: Utility of prostate cancer Ibrahim N, Ebbinghaus S, Ribas A; KEYNOTE-006 investigators:
screening in kidney transplant candidates. J Am Soc Nephrol 27: Pembrolizumab versus ipilimumab in advanced melanoma. N Engl J
2157–2163, 2016 PubMed Med 372: 2521–2532, 2015 PubMed
19. Saumoy M, Jesudian AB, Aden B, Serur D, Sundararajan S, Sivananthan 34. Maggiore U, Pascual J: The bad and the good news on cancer
G, Gambarin-Gelwan M: High prevalence of colon adenomas in end- immunotherapy: Implications for organ transplant recipients. Adv
stage kidney disease patients on hemodialysis undergoing renal transplant Chronic Kidney Dis 23: 312–316, 2016 PubMed
evaluation. Clin Transplant 30: 256–262, 2016 PubMed
20. Stoecker JB, Cote DR, Augustine JJ, Sarabu N, Schulak JA, Sanchez
EQ, Humphreville VR, Ammori JB, Woodside KJ: Utility of mam- Cardiovascular Disease
mography for chronic kidney disease patients undergoing kidney
transplant evaluation. Clin Transplant 30: 445–451, 2016 PubMed Pretransplant Screening
21. Mukhtar RA, Piper ML, Freise C, Van’t Veer LJ, Baehner FL, There is ongoing debate regarding the utility of
Esserman LJ: The novel application of genomic profiling assays to
shorten inactive status for potential kidney transplant recipients with testing for ischemic heart disease for risk stratification
breast cancer. Am J Transplant 17: 292–295, 2017 PubMed purposes in potential kidney transplant candidates. An
22. Zwald F, Leitenberger J, Zeitouni N, Soon S, Brewer J, Arron S, updated meta-analysis of a total of 52 cohort and
Bordeaux J, Chung C, Abdelmalek M, Billingsley E, Vidimos A, Stasko
T: Recommendations for solid organ transplantation for transplant randomized, controlled studies that included 7401
candidates with a pretransplant diagnosis of cutaneous squamous cell participants evaluated major adverse cardiac events
carcinoma, merkel cell carcinoma and melanoma: A consensus opinion and cardiovascular and all-cause mortality after trans-
from the International Transplant Skin Cancer Collaborative (ITSCC).
Am J Transplant 16: 407–413, 2016 PubMed plant after undergoing a preoperative cardiac screening
23. Klein JA, Gonzalez SA, Fischbach BV, Yango AF, Rajagopal A, Rice test (1). Cardiac screening tests were either noninva-
KM, Saim M, Barri YM, Melton LB, Klintmalm GB, Chandrakantan A:
sive (segregated by dobutamine stress echocardiogra-
Routine ultrasonography surveillance of native kidneys for renal cell
carcinoma in kidney transplant candidates. Clin Transplant 30: 946– phy or myocardial perfusion scintigraphy) or invasive
953, 2016 PubMed (angiography). The authors determined that the prog-
24. Acuna SA, Huang JW, Scott AL, Micic S, Daly C, Brezden-Masley C,
Kim SJ, Baxter NN: Cancer screening recommendations for solid organ
nostic value of an abnormal noninvasive test was
transplant recipients: A systematic review of clinical practice guidelines. similar to that of abnormal coronary angiography for
Am J Transplant 17: 103–114, 2017 PubMed predicting cardiovascular mortality and major adverse
25. Dharnidharka VR: Peripheral blood epstein-barr viral nucleic acid
surveillance as a marker for posttransplant cancer risk. Am J Transplant
cardiac events, with a trend suggesting lower all-cause
17: 611–616, 2017 PubMed mortality in the cohort screened by angiography. Based
26. Luskin MR, Heil DS, Tan KS, Choi S, Stadtmauer EA, Schuster SJ, on these findings, the authors suggested that initial
Porter DL, Vonderheide RH, Bagg A, Heitjan DF, Tsai DE, Reshef R:
The impact of EBV status on characteristics and outcomes of post-
investigation using coronary angiography is not indicated
transplantation lymphoproliferative disorder. Am J Transplant 15: in the absence of a conventional indication, and they did
2665–2673, 2015 PubMed not find superiority of one noninvasive test over another.
27. Bamoulid J, Courivaud C, Coaquette A, Crépin T, Carron C, Gaiffe E,
Roubiou C, Rebibou JM, Ducloux D: Late persistent positive EBV viral Disappointingly, even in large meta-analyses, such as
load and risk of solid cancer in kidney transplant patients. Trans- this one, large numbers of patients with negative test
plantation 101: 1473–1478, 2017 PubMed results still have adverse cardiac outcomes, and large
28. AlDabbagh MA, Gitman MR, Kumar D, Humar A, Rotstein C, Husain
S: The role of antiviral prophylaxis for the prevention of epstein-barr numbers of patients with abnormal test results do not
virus-associated posttransplant lymphoproliferative disease in solid have adverse cardiac outcomes; thus, the answer to the
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 337

question of whether any testing should be routinely 1.51), transplant failure (RR, 0.76; 95% CI, 0.49 to
performed remains unanswered. Regarding this latter 1.18), or creatinine level doubling (RR, 0.84; 95% CI,
point, the reader is referred to a review of the evidence 0.51 to 1.39) compared with in the control group, with
for screening for cardiovascular disease in kidney trans- a more than twofold greater risk of hyperkalemia using
plant candidates (2). Despite this ongoing debate, it is RAAS agents (6).
heartening to note that, even with increasing rates of The small n in the case of the randomized trial, the
transplant of higher-risk candidates (older with more uncontrolled nature of the single-center trial, and the
cardiovascular events before transplant), one regional meta-analysis inclusion of studies with generally short
Canadian study described that the risk of death or major follow-up summarize the difficulties in proving (or
cardiovascular event post-transplant has remained stable disproving) benefit of RAAS agents in kidney trans-
over time when comparing 3-year outcomes in a recent era plant recipients. It has been estimated that it would
(2006–2009) with those from a prior era (1994–1997) (3). require a study of over 10,000 subjects to determine if
graft failure was impacted by RAAS inhibition, a very
Hypertension unlikely circumstance. A thoughtful editorial of the
Although renin-angiotensin-aldosterone system Lancet Trial concludes that, although RAAS blockade
(RAAS) blockade is a cornerstone of antihypertensive is unlikely to show a benefit in hard end points, there
management in the CKD population, the nephropro- are no data to suggest avoiding RAAS inhibitors (7). A
tective benefits have not been shown in the kidney recently published expert review also concludes that no
transplant population. Recently, a large, single-center specific antihypertensive medication has been proven
series; a multicenter, randomized clinical trial; and to be more efficacious than others in reducing mortal-
a meta-analysis have been published that frame the ity or preventing graft loss (8). As an aside, sodium
issues surrounding RAAS use after kidney transplanta- restriction remains important in the efficacy of RAAS
tion. In the single-center series, 2684 primary kidney blockade in transplant recipients as shown in a series of
transplant recipients from 1994 to 2010 with a function- 22 subjects treated with RAAS blockers whose systolic
ing graft at 6 months were retrospectively evaluated for BP declined from a mean of 140614 to 129612 mmHg
RAAS agent use or nonuse in a time-dependent analysis, and diastolic BP declined from 8668 to 7968 mmHg
and the outcomes of mortality and graft loss were as-
sessed (4). A total of 638 graft failures occurred at a
mean of 5.4 years post-transplant. RAAS blockade was
associated with a statistically significant 37% reduction
in hazard for overall graft loss (adjusted hazard ratio
[aHR], 0.63; 95% CI, 0.53 to 0.75), a 31% reduction in
hazard for death (aHR, 0.69; 95% CI, 0.55 to 0.86),
and a 38% reduction in death-censored graft failure
(aHR, 0.62; 95% CI, 0.49 to 0.78). Conversely, a
double-blind, placebo-controlled, randomized trial (the
first in kidney transplant recipients to investigate
RAAS blockade) was conducted at 14 centers and en-
rolled 213 patients .3 months post-transplant who had
eGFR$20 ml/min per 1.73 m2 and proteinuria $200 Figure 16. Ramipril versus placebo after kidney transplan-
mg/24 h in a 1:1 ratio to receive placebo versus ramipril tation. Time to the primary outcome of doubling serum
5 mg twice daily (5). Ramipril did not reduce the creatinine, ESRD, or death during the extension phase of the
primary end point, a composite of doubling serum study. Reprinted with permission from Knoll GA, Fergusson
creatinine, ESRD, or death. Despite .4 years of D, Chassé M, Hebert P, Wells G, Tibbles LA, Treleaven D,
Holland D, White C, Muirhead N, Cantarovich M, Paquet M,
follow-up, the absolute difference between the groups
Kiberd B, Gourishankar S, Shapiro J, Prasad R, Cole E,
was only 0.5% (Figure 16). Finally, a meta-analysis of Pilmore H, Cronin V, Hogan D, Ramsay T, Gill J: Ramipril
eight randomized, controlled trials (1502 participants) versus placebo in kidney transplant patients with proteinuria:
with .1 year of follow-up failed to show reductions in A multicentre, double-blind, randomised controlled trial.
risk of death (risk ratio [RR], 0.96; 95% CI, 0.62 to Lancet Diabetes Endocrinol 4: 318–326, 2016.
338 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

after initiation of a low-sodium diet (total daily sodium a time-varying model (HR, 1.78; 95% CI, 1.29 to 2.45;
intake change from164650 mmol/24 h during the regular P,0.001) and a rolling average model (HR, 1.83; 95%
sodium diet to 87655 mmol/24 h) (9). CI, 1.30 to 2.57; P50.001). The authors conjecture
that the risk of PTDM imparted by hypomagnese-
Post-Transplant Diabetes Mellitus mia is due to alterations in cellular transport of glucose
Post-transplant diabetes mellitus (PTDM) remains and/or reduced b cell secretion of insulin. Interventional
an often-encountered clinical syndrome with strong trials are needed to determine if this association can
associations for patient and graft survival. Multiple be remedied with magnesium supplementation. Finally,
definitions and multiple risk factors clearly play a role an ongoing concern is the role of chronic corticosteroids
in the frequency of diagnosis of PTDM. Regarding the in the development of PTDM. In a recent report
incidence of PTDM, perhaps the best point of reference stemming from a 5-year double-blind study compar-
is a randomized, controlled trial of 1083 patients using ing early CSWD with corticosteroid maintenance
the most objective definitions of diabetes (the Amer- tapered to 5 mg/d (CCS) from 6 months onward, no
ican Diabetes Association definitions of hemoglobin difference in PTDM rates was noted: 36.3% of CCS
A1c $6.5%, fasting blood glucose $7.0 mmol/L patients versus 35.9% of CSWD patients were di-
($126 mg/dl), 2-hour plasma glucose $11.1 mmol/L agnosed with PTDM by 5 years. Notably, insulin
($200 mg/dl) during an oral glucose tolerance test, or therapy was more prevalent in the CCS cohort versus
symptomatic hyperglycemia and a random plasma CSWD (11.6% versus 3.7%; P50.05) (15). Thus,
glucose $11.1 mmol/L) and examining a low-risk traditional risk factors (obesity) and nontraditional risk
population using tacrolimus-containing regimens factors (hypomagnesemia) may help predict PTDM,
(rapid corticosteroid withdrawal [CSWD], predomi- whereas immunosuppression-related risk factors (low-
nantly white, and mean body mass index [BMI] of 25 dose corticosteroids and tacrolimus dose/trough con-
kg/m2) that reported an incidence of PTDM of approx- centrations) may be less valuable.
imately 17% 24 weeks post-transplant (10). Identifica-
tion of risk factors for PTDM may facilitate mitigation Hyperlipidemia
of its development and/or severity. In a single-center As reviewed in the last Nephrology Self-Assessment
study of 672 patients without preexisting diabetes, 32% Program, the Kidney Disease Improving Global Out-
developed PTDM (11). A bimodal incidence (#3 months comes Clinical Practice Guideline for Lipid Manage-
and .3 months post-transplant) was identified, in which ment in CKD states that “in adult kidney transplant
31% of early PTDM reverted, and a normal oral glucose recipients, we suggest treatment with a statin,” implying
tolerance test and normal hemoglobin A1c at 3 months that all kidney transplant patients may derive benefit
predicted a low risk (4% incidence) of later-onset from therapy, regardless of lipid profile (grade 2B
PTDM. Obesity was strongly associated with PTDM, evidence) (16, 17). A recent provocative study evaluated
but tacrolimus levels were not. In a separate series of HDL cholesterol (HDL-C) function and its association
150 patients, visceral adipose tissue content measured with cardiovascular and graft outcomes. HDL-C is
via dual energy X-ray absorptiometry scans was the thought to protect from the development of atheroscle-
strongest predictor of PTDM 1 year after transplant rosis in part due to its role in removal of cholesterol from
(12). In another series in which 102 of 481 (22.5%) macrophage foam cells (cholesterol efflux) (18). In
patients developed PTDM, with mean follow-up of a prospective study, 495 stable kidney transplant recip-
57 months post-transplant, obesity (BMI$25 kg/m2) ients underwent assessment of macrophage cholesterol
was the strongest predictor of PTDM followed by efflux capacity and were followed for a mean of 7.0
planned maintenance with sirolimus, nonwhite recipi- years. When examined by tertile, cholesterol efflux
ent status, and older recipient age (13). A novel risk capacity was independently associated with increased
factor for PTDM, hypomagnesemia, was reported in a long-term kidney allograft survival. Interestingly, it was
retrospective series of 948 recipients (14). A serum not associated with cardiovascular outcomes or mortal-
magnesium of ,0.74 mmol/L at 1 month post-transplant ity. The association with graft survival was independent
(1.8 mg/dl) was significantly associated with increased of serum HDL-C levels, suggesting a novel potential
risk of PTDM (hazard ratio [HR], 1.58; 95% CI, 1.07 to therapeutic target for the prevention of graft failure may
2.34; P50.02) and also associated with PTDM using be to increase HDL-C function rather than quantity.
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 339

Obesity to minimize the substantial risk of wound complications


Although obesity is not a traditional cardiovas- known to occur in this patient segment (23). There were
cular risk factor, its association with PTDM, and its no graft losses due to graft thrombosis or infection. The
frequent consideration for transplant candidacy justifies authors compared their outcomes with registry data
discussion as a separate cardiovascular risk factor in (a total of 612 living donor transplants in recipients
kidney transplantation. Recently published trials con- with BMI$40 kg/m2 were performed during the period
tinue to show the survival advantage of transplantation 2009–2014) and found similar rates of DGF and equiv-
versus remaining on dialysis in obese candidates but alent graft function and patient survival. However, 2%
with greater clarity regarding risks. Using the United of morbidly obese recipients who underwent the open
Kingdom renal and transplant registry data from 2004 technique had graft loss due to infection or graft thrombo-
to 2010, 1- and 5-year survival rates after transplant sis. Perhaps expansion of these surgical approaches will
were superior to those of waitlisted candidates in all lead to a greater comfort in evaluating and transplanting the
BMI subgroups, including BMI535–40 and .40 kg/m2 obese transplant candidate.
(19). Further comparisons of obese with nonobese trans-
plant recipients (BMI518.5–25 kg/m2) showed no Nontraditional and Novel Cardiovascular Risk
differences in mortality with increasing BMI. However, Factors under Investigation
conclusions regarding the latter subgroups are difficult Structural Parameters. Left ventricular (LV) hy-
to extrapolate, because only 540 of the 13,526 total pertrophy has been described as a risk factor for poor
patients available for analysis had BMI.35 kg/m2. cardiovascular and renal outcomes in patients with
Using statistical methodology to control for the com- stages 2–5 native CKD (24). After transplant, LV mass
peting risk of death, a United States registry analysis of index and left atrial volume index often improve
108,654 primary kidney transplant recipients from coincident with improvement in LV ejection fraction
2001–2009 showed worse graft survival with increas- (25). In a long-term follow-up study of 100 patients
ing BMI (20). With BMI518.5–25 kg/m2 as the from two randomized, controlled trials (mean of 8.4
reference, the sub-HRs were 1.15 for 30–35 kg/m2 years of follow-up), LV hypertrophy regression was
(P,0.001), 1.21 for 35–40 kg/m2 (P,0.001), and 1.13 significantly associated with reduction in risk for
for .40 kg/m2 (P50.002). A meta-analysis that did not cardiovascular events and mortality, and it was associ-
account for the competing risk of death similarly found ated with higher GFR (26). Beyond echocardiographic
an increased risk of death-censored graft loss (HR, parameters, aortic stiffness has also been proposed as
1.06; 95% CI, 1.01 to 1.12), an increased likelihood of a cardiovascular risk predictor. A recent study of 1497
delayed graft function (DGF; odds ratio, 1.68; 95% CI, kidney transplant recipients evaluated aortic stiff-
1.39 to 2.03), and no significant difference in mortality ness measured by carotid-femoral pulse wave velocity
risk in obese recipients (defined as BMI$30 kg/m2; (PWV) 8 weeks after transplant, with median follow-up
HR, 1.24; 95% CI, 0.90 to 1.70) (21). Taken together, of 4.2 years (27). For every 1-m/s increase in PWV up
kidney transplant can be considered effective therapy to 12 m/s, an increase in mortality risk was identified
from the obese patient’s perspective compared with (HR, 1.36; 95% CI, 1.14 to 1.62; P50.001). Each
dialysis, but there are higher risks of morbidity (DGF interquartile range increase was similar in magnitude
and graft loss) than in nonobese transplant recipients for risk of death as an age increase of 20 years. Thus,
that transplant programs must reconcile. structural assessments of LV hypertrophy and PWV
This additive risk noted in obese patients has may provide greater opportunity for post-transplant risk
prompted transplant centers to explore surgical options stratification, and intervention strategies with LV hy-
to optimize outcomes. In one series, laparoscopic sleeve pertrophy and PWV regression as surrogate or primary
gastrectomy was performed in 52 renal transplant candi- end points should be considered.
dates with a mean BMI of 43.0 kg/m2 (range, 35.8– Biochemical Markers. Substudies of the Folic Acid
67.7 kg/m2), with 29 achieving goal BMI of ,35 kg/m2 for Vascular Outcome Reduction in Transplantation
at a mean of 92 days (range, 13–420 days) and six (FAVORIT) Trial have examined serum and urinary
undergoing successful transplant (22). A single-center markers as potential indicators of cardiovascular out-
series used minimally invasive robotic surgery in 67 living comes. In a population of 1131 participants enriched
donor kidney transplants for patients with BMI$40 kg/m2 for adverse events from the overall cohort of 4110
340 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

participants, combined elevations of B-type natriuretic CI, 1.18 to 3.39), and the death aHR was 3.49 (95% CI,
peptide and cardiac troponin I exceeding clinical cutoffs 1.95 to 6.24). These differential rates of mortality and graft
were associated with future cardiovascular adverse loss persisted at 1 year and are clinically meaningful given
events (HR, 6.3; 95% CI, 2.7 to 15.0). For kidney graft modern transplant success rates (Figure 17). Conversely,
loss, the HR was 8.8 (95% CI, 3.4 to 23.1) (28). A severe hypertension before transplant may not associate
separate post hoc study evaluated urine neutrophil with poor transplant outcomes. In a single-center study
gelatinase–associated lipocalin, kidney injury molecule- comparing recipients with severe preoperative hyper-
1, IL-18, and liver-type fatty acid binding protein as tension, defined as systolic BP .180 mmHg or di-
potential biomarkers for adverse events (29). The urine astolic BP .110 mmHg (n5111), with those without
neutrophil gelatinase–associated lipocalin-to-creatinine severe hypertension (n598), there were no differences
ratio was strongly and independently associated with all in cardiac events, patient or graft survival, or postoperative
three end points of mortality, graft loss, and cardiovascular complications, such as DGF, length of stay, or nadir of
events. Urine kidney injury molecule-1–to-creatinine and serum creatinine (33).
IL-18–to-creatinine ratios were only independently asso-
ciated with greater mortality risk.
Beyond these exploratory biomarkers, another Midodrine use before transplant was re-
novel serum assay predictive of cardiovascular end points cently identified as a risk factor for delayed
was recently reported, a measure of vascular calcification graft function, early graft loss, and death
propensity referred to as T50. T50 represents the time point after transplant.
of the half-maximal generation of secondary calciprotein
particles from primary calciprotein particles, and is a re-
flection of the capacity of serum to inhibit the amorphous-
to-crystalline transformation of calcium phosphate. In 1455 Interventions and Outcomes
kidney transplant recipients, T50 measured at 10 weeks Applying cardioprotective principles garnered
post-transplant was associated with all-cause mortality, from the general and high-risk populations to the trans-
with low (faster) versus high (slower) T50 quartile (HR, plant population has been a challenge. For example,
1.60; 95% CI, 1.00 to 2.57) and cardiac mortality (HR, aspirin use as primary cardiovascular prevention in
3.60; 95% CI, 1.10 to 11.83) (30). These findings were kidney transplant recipients with hyperhomocysteine-
replicated in a longitudinal cohort of 699 patients with mia was assessed in a post hoc analysis of 981 aspirin
a median follow-up of 3.1 years, in which lower T50 users versus a matched cohort of 981 nonusers in the
values were independently associated with increased FAVORIT Trial (34). After 4 years of follow-up, no
all-cause mortality and increased cardiovascular mor- differences in risk for cardiovascular disease events,
tality, independent of age, sex, eGFR, or other Fra- all-cause mortality, kidney failure, composite of kid-
mingham risk factors (31). Standardization of test ney failure or mortality, or composite of primary
characteristics and replication of these findings are cardiovascular disease events or mortality were noted
warranted in the future. with aspirin use. Thus, there is no evidence that
Clinical Features. Clinical assessments of BP at the supports its use in primary prevention. Conversely,
extremes in the transplant candidate may predict kidney the survival advantage of drug-eluting stents (DESs)
allograft outcomes. Midodrine, often used for symp- versus bare metal stents (BMS) for percutaneous
tomatic hypotension in the chronic dialysis setting, was coronary intervention in the high-risk general popula-
evaluated as a risk factor for post-transplant outcomes tion was mirrored in a retrospective study of kidney
by linking prescription use 1 year before transplant to transplant recipients (35). Using propensity score–
3-month and 1-year outcomes post-transplant (32). matched registry data, 3245 kidney transplant recipi-
DGF was much more common in the midodrine cohort ents underwent percutaneous coronary intervention, in
(32% versus 19%; P,0.05). More remarkably, at 3 months, whom a 20% lower risk of death and 14% lower risk of
graft failure (5% versus 2%), death (4% versus 1%), and death, myocardial infarction, or repeat revasculariza-
acute myocardial infarction and cardiac arrest (all P,0.05) tion at 3 years of follow-up were noted with DES
were more common in the midodrine cohort. In the latter, versus BMS. Similarly, patency after endovascular
the 3-month death-censored graft failure aHR was 2.0 (95% intervention of transplant renal artery stenosis was
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 341

management of hypertension in transplant patients. J Am Soc Nephrol


26: 1248–1260, 2015 PubMed
9. de Vries LV, Dobrowolski LC, van den Bosch JJ, Riphagen IJ, Krediet
CT, Bemelman FJ, Bakker SJ, Navis G: Effects of dietary sodium
restriction in kidney transplant recipients treated with renin-angiotensin-
aldosterone system blockade: A randomized clinical trial. Am J Kidney
Dis 67: 936–944, 2016 PubMed
10. Mourad G, Glyda M, Albano L, Viklický O, Merville P, Tydén G,
Mourad M, Lõhmus A, Witzke O, Christiaans MHL, Brown MW,
Undre N, Kazeem G, Kuypers DRJ; Advagraf-based immunosuppres-
sion regimen examining new onset diabetes mellitus in kidney
transplant recipients (ADVANCE) study investigators: Investigators,
Advagraf-based immunosuppression regimen examining new onset
diabetes mellitus in kidney transplant recipients study: Incidence of
posttransplantation diabetes mellitus in de novo kidney transplant
recipients receiving prolonged-release tacrolimus-based immunosup-
Figure 17. Incidence of complications at 3 and 12 months post- pression with 2 different corticosteroid minimization strategies:
transplant according to pretransplant midodrine use. AMI, ADVANCE, a randomized controlled trial. Transplantation 101:
acute myocardial infarction. Reprinted with permission from 1924–1934, 2017 PubMed
Alhamad T, Brennan DC, Brifkani Z, Xiao H, Schnitzler 11. Porrini EL, Díaz JM, Moreso F, Delgado Mallén PI, Silva Torres I,
MA, Dharnidharka VR, Axelrod D, Segev DL, Lentine KL: Ibernon M, Bayés-Genís B, Benitez-Ruiz R, Lampreabe I, Lauzurrica
R, Osorio JM, Osuna A, Domínguez-Rollán R, Ruiz JC, Jiménez-Sosa
Pretransplant midodrine use: A newly identified risk marker for A, González-Rinne A, Marrero-Miranda D, Macía M, García J, Torres
complications after kidney transplantation. Transplantation 100: A: Clinical evolution of post-transplant diabetes mellitus. Nephrol Dial
1086–1093, 2016. Transplant 31: 495–505, 2016 PubMed
12. von Düring ME, Jenssen T, Bollerslev J, Åsberg A, Godang K,
Hartmann A: Visceral fat is strongly associated with post-transplant
diabetes mellitus and glucose metabolism 1 year after kidney trans-
shown to be superior with DES versus BMS, and both plantation. Clin Transplant 31: e12869, 2017 PubMed
13. Gaynor JJ, Ciancio G, Guerra G, Sageshima J, Hanson L, Roth D,
were superior to percutaneous transluminal angioplasty Goldstein MJ, Chen L, Kupin W, Mattiazzi A, Tueros L, Flores S,
alone in a trial of 45 patients who underwent a total of Barba LJ, Lopez A, Rivas J, Ruiz P, Vianna R, Burke GW 3rd:
50 primary endovascular interventions (DES, 18; BMS, Multivariable risk of developing new onset diabetes after transplant-
results from a single-center study of 481 adult, primary kidney trans-
26; percutaneous transluminal angioplasty, six) (36). plant recipients. Clin Transplant 29: 301–310, 2015 PubMed
14. Huang JW, Famure O, Li Y, Kim SJ: Hypomagnesemia and the risk of
References new-onset diabetes mellitus after kidney transplantation. J Am Soc
1. Wang LW, Masson P, Turner RM, Lord SW, Baines LA, Craig JC, Nephrol 27: 1793–1800, 2016 PubMed
Webster AC: Prognostic value of cardiac tests in potential kidney 15. Pirsch JD, Henning AK, First MR, Fitzsimmons W, Gaber AO,
transplant recipients: A systematic review. Transplantation 99: 731– Reisfield R, Shihab F, Woodle ES: New-onset diabetes after trans-
745, 2015 PubMed plantation: Results from a double-blind early corticosteroid withdrawal
2. Hart A, Weir MR, Kasiske BL: Cardiovascular risk assessment in trial. Am J Transplant 15: 1982–1990, 2015 PubMed
kidney transplantation. Kidney Int 87: 527–534, 2015 PubMed 16. Kidney Disease: Improving Global Outcomes Transplant Work Group:
3. Lam NN, Kim SJ, Knoll GA, McArthur E, Lentine KL, Naylor KL, Li AH, KDIGO clinical practice guideline for the care of kidney transplant
Shariff SZ, Ribic CM, Garg AX: The risk of cardiovascular disease is not recipients. Am J Transplant 9[Suppl 3]: S1–S155, 2009 PubMed
increasing over time despite aging and higher comorbidity burden of kidney 17. Wanner C, Tonelli M; Kidney Disease: Improving Global Outcomes
transplant recipients. Transplantation 101: 588–596, 2017 PubMed Lipid Guideline Development Work Group Members: KDIGO Clinical
4. Shin JI, Palta M, Djamali A, Kaufman DB, Astor BC: The association Practice Guideline for Lipid Management in CKD: Summary of
between renin-angiotensin system blockade and long-term outcomes in recommendation statements and clinical approach to the patient. Kidney
renal transplant recipients: The Wisconsin Allograft Recipient Database Int 85: 1303–1309, 2014 PubMed
(WisARD). Transplantation 100: 1541–1549, 2016 PubMed 18. Annema W, Dikkers A, de Boer JF, Dullaart RP, Sanders JS, Bakker SJ,
5. Knoll GA, Fergusson D, Chassé M, Hebert P, Wells G, Tibbles LA, Tietge UJ: HDL cholesterol efflux predicts graft failure in renal
Treleaven D, Holland D, White C, Muirhead N, Cantarovich M, Paquet M, transplant recipients. J Am Soc Nephrol 27: 595–603, 2016 PubMed
Kiberd B, Gourishankar S, Shapiro J, Prasad R, Cole E, Pilmore H, Cronin 19. Krishnan N, Higgins R, Short A, Zehnder D, Pitcher D, Hudson A,
V, Hogan D, Ramsay T, Gill J: Ramipril versus placebo in kidney transplant Raymond NT: Kidney transplantation significantly improves patient
patients with proteinuria: A multicentre, double-blind, randomised con- and graft survival irrespective of BMI: A cohort study. Am J Transplant
trolled trial. Lancet Diabetes Endocrinol 4: 318–326, 2016 PubMed 15: 2378–2386, 2015 PubMed
6. Hiremath S, Fergusson DA, Fergusson N, Bennett A, Knoll GA: Renin- 20. Naik AS, Sakhuja A, Cibrik DM, Ojo AO, Samaniego-Picota MD,
angiotensin system blockade and long-term clinical outcomes in kidney Lentine KL: The impact of obesity on allograft failure after kidney
transplant recipients: A meta-analysis of randomized controlled trials. transplantation: A competing risks analysis. Transplantation 100:
Am J Kidney Dis 69: 78–86, 2017 PubMed 1963–1969, 2016 PubMed
7. Cross NB, Webster AC: Angiotensin-converting enzyme inhibitors- 21. Hill CJ, Courtney AE, Cardwell CR, Maxwell AP, Lucarelli G, Veroux
beneficial effects seen in many patient groups may not extend to kidney M, Furriel F, Cannon RM, Hoogeveen EK, Doshi M, McCaughan JA:
transplant recipients. Transplantation 100: 472–473, 2016 PubMed Recipient obesity and outcomes after kidney transplantation: A sys-
8. Weir MR, Burgess ED, Cooper JE, Fenves AZ, Goldsmith D, McKay tematic review and meta-analysis. Nephrol Dial Transplant 30: 1403–
D, Mehrotra A, Mitsnefes MM, Sica DA, Taler SJ: Assessment and 1411, 2015 PubMed
342 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

22. Freeman CM, Woodle ES, Shi J, Alexander JW, Leggett PL, Shah SA, 36. Biederman DM, Fischman AM, Titano JJ, Kim E, Patel RS, Nowakowski
Paterno F, Cuffy MC, Govil A, Mogilishetty G, Alloway RR, Hanseman FS, Florman S, Lookstein RA: Tailoring the endovascular management
D, Cardi M, Diwan TS: Addressing morbid obesity as a barrier to renal of transplant renal artery stenosis. Am J Transplant 15: 1039–1049,
transplantation with laparoscopic sleeve gastrectomy. Am J Transplant 2015 PubMed
15: 1360–1368, 2015 PubMed
23. Garcia-Roca R, Garcia-Aroz S, Tzvetanov I, Jeon H, Oberholzer J,
Benedetti E: Single center experience with robotic kidney transplanta-
tion for recipients with BMI of 40 kg/m2 or greater: A comparison with
the UNOS registry. Transplantation 101: 191–196, 2017 PubMed Bone and Mineral Disorders
24. Paoletti E, De Nicola L, Gabbai FB, Chiodini P, Ravera M, Pieracci L,
Marre S, Cassottana P, Lucà S, Vettoretti S, Borrelli S, Conte G, Hyperparathyroidism
Minutolo R: Associations of left ventricular hypertrophy and geometry Hyperparathyroidism is commonly recognized
with adverse outcomes in patients with CKD and hypertension. Clin J after transplant, but the natural history of disease,
Am Soc Nephrol 11: 271–279, 2016 PubMed
25. Kensinger C, Hernandez A, Bian A, Fairchild M, Chen G, Lipworth L, response to therapy, and long-term ramifications are
Ikizler TA, Birdwell KA: Longitudinal assessment of cardiac morphol- less well understood. To shed light on the severity and
ogy and function following kidney transplantation [published online
natural history of disordered mineral metabolism after
ahead of print January, 2017]. Clin Transplant doi:10.1111/ctr.
12864PubMed transplant, a multicenter, prospective study cohort of
26. Paoletti E, Bellino D, Signori A, Pieracci L, Marsano L, Russo R, 246 recipients with parathyroid hormone (PTH) levels
Massarino F, Ravera M, Fontana I, Carta A, Cassottana P, Garibotto G:
Regression of asymptomatic cardiomyopathy and clinical outcome of
.65 pg/ml was intensively monitored in the first year
renal transplant recipients: A long-term prospective cohort study. post-transplant (1). In those untreated, 86.2% had
Nephrol Dial Transplant 31: 1168–1174, 2016 PubMed persistent hyperparathyroidism .65 pg/ml, and 40%
27. Dahle DO, Eide IA, Åsberg A, Leivestad T, Holdaas H, Jenssen TG,
Fagerland MW, Pihlstrøm H, Mjøen G, Hartmann A: Aortic stiffness in
had PTH.130 pg/ml at 1 year. In those with persistent
a mortality risk calculator for kidney transplant recipients. Trans- hyperparathyroidism, hypercalcemia rates (albumin-
plantation 99: 1730–1737, 2015 PubMed corrected serum calcium .10.2 mg/dl) peaked at week
28. Jarolim P, Claggett BL, Conrad MJ, Carpenter MA, Ivanova A, Bostom
AG, Kusek JW, Hunsicker LG, Jacques PF, Gravens-Mueller L, Finn P,
8 (48%), decreasing to 25% at 1 year. Similarly,
Solomon SD, Weiner DE, Levey AS, Pfeffer MA: B-type natriuretic hypophosphatemia (,2.5 mg/dl) rates peaked at week
peptide and cardiac troponin I are associated with adverse outcomes in 2 post-transplant (54%) and improved progressively
stable kidney transplant recipients. Transplantation 101: 182–190,
2017 PubMed over 1 year. Fibroblast growth factor 23 reached a nadir by
29. Bansal N, Carpenter MA, Weiner DE, Levey AS, Pfeffer M, Kusek JW, the third month post-transplant. This study together with
Cai J, Hunsicker LG, Park M, Bennett M, Liu KD, Hsu CY: Urine another longitudinal study of 1237 children, in whom
injury biomarkers and risk of adverse outcomes in recipients of
prevalent kidney transplants: The Folic Acid for Vascular Outcome hyperparathyroidism was observed in 41% of recipients
Reduction in Transplantation Trial. J Am Soc Nephrol 27: 2109–2121, (2), provide perspectives on the need for and timing of
2016 PubMed potential interventions in the post-transplant setting.
30. Dahle DO, Åsberg A, Hartmann A, Holdaas H, Bachtler M, Jenssen TG,
Dionisi M, Pasch A: Serum calcification propensity is a strong and Potential interventions include medical (cinacalcet
independent determinant of cardiac and all-cause mortality in kidney and paricalcitol) and surgical (subtotal parathyroidec-
transplant recipients. Am J Transplant 16: 204–212, 2016 PubMed
tomy) options. A recent randomized, controlled trial
31. Keyzer CA, de Borst MH, van den Berg E, Jahnen-Dechent W,
Arampatzis S, Farese S, Bergmann IP, Floege J, Navis G, Bakker SJ, compared cinacalcet with parathyroidectomy for 30
van Goor H, Eisenberger U, Pasch A: Calcification propensity and patients with hypercalcemia and hyperparathyroidism 6
survival among renal transplant recipients. J Am Soc Nephrol 27: 239–
248, 2016 PubMed
months or more from time of transplantation (mean, 3.7
32. Alhamad T, Brennan DC, Brifkani Z, Xiao H, Schnitzler MA, years) (3). After 1 year, ten of 15 patients treated with
Dharnidharka VR, Axelrod D, Segev DL, Lentine KL: Pretransplant cinacalcet had normal serum calcium levels versus 15 of
midodrine use: A newly identified risk marker for complications after
kidney transplantation. Transplantation 100: 1086–1093, 2016 PubMed
15 in the parathyroidectomy arm. The latter group had
33. Ajaimy M, Lubetzky M, Kamal L, Gupta A, Dunn C, de Boccardo G, lower PTH levels and significant increases in femoral
Akalin E, Kayler L: Kidney transplantation in patients with severe neck bone mineral density (BMD). In a separate single-
preoperative hypertension. Clin Transplant 29: 781–785, 2015 PubMed
34. Dad T, Tighiouart H, Joseph A, Bostom A, Carpenter M, Hunsicker L,
center, crossover trial of 43 patients that examined the
Kusek JW, Pfeffer M, Levey AS, Weiner DE: Aspirin use and incident effect of a 6-month treatment with paricalcitol (1 mg/d
cardiovascular disease, kidney failure, and death in stable kidney for 3 months and uptitrated to 2 mg/d if tolerated), PTH
transplant recipients: A post hoc analysis of the Folic Acid for Vascular
Outcome Reduction in Transplantation (FAVORIT) Trial. Am J Kidney declined during paricalcitol treatment from a mean of
Dis 68: 277–286, 2016 PubMed 115.6 to 63.3 pg/ml, with vertebral mineral bone density
35. Lenihan CR, Montez-Rath ME, Winkelmayer WC, Chang TI: Drug- increasing and bone turnover markers decreasing with
eluting stents versus bare metal stents for percutaneous coronary
intervention in kidney transplant recipients. Transplantation 101: therapy for 6 months (4). Both a reduction in proteinuria
851–857, 2017 PubMed and an increase of serum creatinine were detected during
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 343

therapy. The latter observation was hypothesized to findings included a 3-year cumulative incidence of
result from decreased tubular secretion or increased nonvertebral fracture of 1.6% (95% confidence interval
generation of creatinine rather than a true change in [95% CI], 1.3% to 2.0%), a rate higher than that in the
GFR. In a separate 3-month study, paricalcitol increased general population of 0.5% (95% CI, 0.4% to 0.6%;
serum fibroblast growth factor 23 levels and slightly P,0.001), which is only slightly higher than that in the
increased serum KLOTHO levels, raising the possibility nondialysis CKD population (1.1%; 95% CI, 0.9% to
of longer-term benefits from this active vitamin D 1.2%; P50.03). This may be due to changes over time in
receptor analogue. This latter point is highlighted by patient management as suggested by a single-center study
a follow-up study of 1840 transplant recipients recruited from France comparing patients transplanted from 2004
in the Lescol in Renal Transplantation Trial with 6–7 to 2006 with a group transplanted from 2009 to 2011 (9).
years of follow-up (5). Persistent elevation in PTH.65 Cinacalcet, vitamin D, and bisphosphonates were used
pg/ml was associated with a 46% (P,0.01) higher risk more frequently. Median PTH levels were lower, and
of death and an 85% higher risk of graft loss (P,0.001). vitamin D deficiency declined from 64% to 20%, whereas
Stepwise PTH elevations were associated with all-cause fracture incidence decreased from 9.1% to 3.1% (P50.05).
mortality and graft loss in multivariate analysis. Taken
together, these findings support more aggressive monitor-
ing and management of hyperparathyroidism and its effect The incidence of bone fracture post trans-
on mineral metabolism. plantation has declined over time concur-
rent with increasing use of prophylaxis.
Osteoporosis
The measure of bone integrity in CKD and after
transplant often involves noninvasive markers as surrogates Prevention of osteoporosis post-transplant remains
for hard end points, such as fracture or bone biopsy data. an active area of study. A recent meta-analysis of 12
Although dual energy X-ray absorptiometry is commonly studies evaluating the efficacy of bisphosphonates in-
used to determine BMD, other novel methods include the cluded 621 patients and described improvements in
trabecular bone score and the in vivo microindentation BMD of 6.0% at the femoral neck and 7.4% at the
measurement that evaluates trabecular microarchitecture lumbar spine in those taking bisphosphonates com-
and the mechanical properties of bone to predict fracture pared with those not taking bisphosphonate. However,
risk. In 40 kidney transplant recipients, all three modal- this was not associated with a difference in fracture
ities were compared with those in 94 healthy controls incidence between groups (10). In the general population,
(6). BMD by dual energy X-ray absorptiometry was treatment of low bone mineral density becomes cost
lower compared with in controls. Trabecular bone score effective when the 10-year probability of hip fracture
and bone strength did not differ from controls, despite reaches approximately 3% as assessed by the Fracture
longstanding kidney transplant status of .10 years. This Risk Assessment Tool (11). In the post-transplant setting,
observation suggests that declines in bone health may it is reasonable to consider treatment with bisphospho-
not necessarily be a forgone conclusion after transplant. nates in individuals with suspected osteoporosis when
That bone disease is less severe in the current the 10-year probability of hip fracture reaches this
transplantation era compared with prior eras is further threshold in the absence of other contributing factors in
supported by three recent studies from different geo- the first 12 months after transplant when the eGFR is
graphic regions. A database analysis examining 21,769 .30 ml/min per 1.73 m2 (12). Another potential treat-
patients transplanted between 2001 and 2013 disclosed ment for the prevention of osteoporosis and fractures
a 3.8% incidence of hip fracture: a rate of 1.54/1000 post-transplant is the receptor activator of NF-kB ligand
patient-years for hip fracture and a rate of 9.99/1000 inhibitor denosumab. An open label, prospective, ran-
patient-years for any fracture. On balance, these con- domized trial of 90 patients compared two denosumab
temporary rates are approximately one half of prior re- doses of 60 mg at 2 weeks and 6 months after transplant
ports from earlier eras (7). A Canadian analysis of with no treatment (13). At 12 months, lumbar spine BMD
fracture rates from 4821 kidney transplant recipients increased by 4.6% in the denosumab group and decreased
from 1994 to 2008 showed a similar 10-year cumula- by 20.5% in the control group (between-group differ-
tive incidence of hip fracture of 1.7% (8). Additional ence, 5.1%; 95% CI, 3.1% to 7.0%; P,0.001), and total
344 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

Table 6. Mean lifetimes up to 10 years for SPK and KA and their differences based on graft loss and death for
recipients with ages at transplants from 18 to 55, from 18 to 39, and from 40 to 55 years old
SPK KA Difference between SPK and KA P Value
Graft loss
Age group, 18–55 y 8.00 (0.05) 7.82 (0.09) 0.18 (0.09) 0.05
Age group, 18–39 y 7.99 (0.07) 7.68 (0.11) 0.30 (0.13) 0.02
Age group, 40–55 y 8.00 (0.08) 7.93 (0.10) 0.08 (0.12) 0.52
Death
Age group, 18–55 y 8.68 (0.04) 8.51 (0.07) 0.17 (0.08) 0.03
Age group, 18–39 y 8.77 (0.06) 8.59 (0.10) 0.18 (0.11) 0.10
Age group, 40–55 y 8.59 (0.06) 8.40 (0.09) 0.19 (0.11) 0.08
Numbers in parentheses are SEMs. Modified from Sung RS, Zhang M, Schaubel DE, Shu X, Magee JC: A Reassessment of the Survival Advantage of Simultaneous Kidney-
Pancreas Versus Kidney-Alone Transplantation. Transplantation 99: 1900–1906, 2015.

hip BMD increased by 1.9% (95% CI, 0.1% to 3.7%; References


P50.04) compared with in the control group. Other 1. Wolf M, Weir MR, Kopyt N, Mannon RB, Von Visger J, Deng H, Yue
S, Vincenti F: A prospective cohort study of mineral metabolism after
biochemical measures of bone turnover all significantly kidney transplantation. Transplantation 100: 184–193, 2016 PubMed
decreased with denosumab. The safety of denosumab 2. Bonthuis M, Busutti M, van Stralen KJ, Jager KJ, Baiko S, Bakkaloğlu S,
Battelino N, Gaydarova M, Gianoglio B, Parvex P, Gomes C, Heaf JG,
within this trial was reported separately, with more Podracka L, Kuzmanovska D, Molchanova MS, Pankratenko TE,
infections in the denosumab group (n5146) compared Papachristou F, Reusz G, Sanahuja MJ, Shroff R, Groothoff JW, Schaefer
with in the control group (n599), primarily driven by F, Verrina E: Mineral metabolism in European children living with a renal
transplant: A European Society for Paediatric Nephrology/European
a higher incidence of urinary tract infection/cystitis (51 Renal Association-European Dialysis and Transplant Association Regis-
versus 25 episodes in 24 denosumab-treated versus 11 try Study. Clin J Am Soc Nephrol 10: 767–775, 2015 PubMed
control patients; P,0.01). The frequencies of opportu- 3. Cruzado JM, Moreno P, Torregrosa JV, Taco O, Mast R, Gómez-Vaquero
C, Polo C, Revuelta I, Francos J, Torras J, García-Barrasa A, Bestard O,
nistic viral infections, pyelonephritis, and urosepsis were Grinyó JM: A randomized study comparing parathyroidectomy with
not more common (14). cinacalcet for treating hypercalcemia in kidney allograft recipients with
hyperparathyroidism. J Am Soc Nephrol 27: 2487–2494, 2016 PubMed
4. Trillini M, Cortinovis M, Ruggenenti P, Reyes Loaeza J, Courville K,
Ferrer-Siles C, Prandini S, Gaspari F, Cannata A, Villa A, Perna A,
The use of bisphosphonates after trans- Gotti E, Caruso MR, Martinetti D, Remuzzi G, Perico N: Paricalcitol for
secondary hyperparathyroidism in renal transplantation. J Am Soc
plant is associated with improvements in Nephrol 26: 1205–1214, 2015 PubMed
BMD but has not been associated with 5. Pihlstrøm H, Dahle DO, Mjøen G, Pilz S, März W, Abedini S, Holme I,
reduction in risk for fracture. Fellström B, Jardine AG, Holdaas H: Increased risk of all-cause
mortality and renal graft loss in stable renal transplant recipients with
hyperparathyroidism. Transplantation 99: 351–359, 2015 PubMed
6. Pérez-Sáez MJ, Herrera S, Prieto-Alhambra D, Nogués X, Vera M,
Redondo-Pachón D, Mir M, Güerri R, Crespo M, Díez-Pérez A, Pascual
J: Bone density, microarchitecture and tissue quality long-term after
Finally, another potential contributor to the risk for
kidney transplant. Transplantation 101: 1290–1294, 2017 PubMed
hip fracture after transplant is the frequent use of proton 7. Ferro CJ, Arnold J, Bagnall D, Ray D, Sharif A: Fracture risk and mortality
pump inhibitors (PPIs). This specific risk has been de- post-kidney transplantation. Clin Transplant 29: 1004–1012, 2015 PubMed
8. Naylor KL, Jamal SA, Zou G, McArthur E, Lam NN, Leslie WD,
scribed in the general population but until recently, was Hodsman AB, Kim SJ, Knoll GA, Fraser LA, Adachi JD, Garg AX:
not reported in the transplant population. A case-control fracture incidence in adult kidney transplant recipients. Transplantation
study compared 231 kidney transplant patients with a first 100: 167–175, 2016 PubMed
9. Perrin P, Kiener C, Javier RM, Braun L, Cognard N, Gautier-Vargas G, Heibel
hip fracture with 15,575 kidney transplant controls using F, Muller C, Olagne J, Moulin B, Ohlmann S: Recent changes in chronic
registry data matched for age, sex, race, and transplant kidney disease-mineral and bone disorders and associated fractures after
year (15). PPI use and persistence of use in the year kidney transplantation. Transplantation 101: 1897–1905, 2017 PubMed
10. Toth-Manikowski SM, Francis JM, Gautam A, Gordon CE: Outcomes
before fracture (the index date) were greater in the hip of bisphosphonate therapy in kidney transplant recipients: A systematic
fracture cohort. After multivariable adjustment, the odds review and meta-analysis. Clin Transplant 30: 1090–1096, 2016 PubMed
ratio of hip fracture with PPI use was 1.39 (95% CI, 1.04 11. Tosteson AN, Melton LJ 3rd, Dawson-Hughes B, Baim S, Favus MJ,
Khosla S, Lindsay RL: National Osteoporosis Foundation Guide
to 1.84), consistent with but not higher than the risk of Committee: Cost-effective osteoporosis treatment thresholds: the
PPI and fracture noted in the general population (16). United States perspective. Osteoporos Int 19: 437–447, 2008
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 345

12. Kidney Disease: Improving Global Outcomes (KDIGO) CKD-MBD measures of coronary calcification at the time of evaluation
Update Work Group: KDIGO 2017 Clinical Practice Guideline Update
for the Diagnosis, Evaluation, Prevention, and Treatment of Chronic
and again .7 years (median, 10.1 years) after transplant
Kidney Disease–Mineral and Bone Disorder (CKD-MBD). Kidney Int (4). Coronary artery disease progression, calcification, and
Suppl 7: 39–40, 2017 PubMed systolic function were no different between groups, sug-
13. Bonani M, Frey D, Brockmann J, Fehr T, Mueller TF, Saleh L, von
Eckardstein A, Graf N, Wüthrich RP: Effect of twice-yearly denosumab on gesting that the addition of the pancreas transplant did not
prevention of bone mineral density loss in de novo kidney transplant recipients: impact these parameters, despite differences in hemoglobin
A randomized controlled trial. Am J Transplant 16: 1882–1891, 2016 PubMed A1c levels (mean, 5.5% versus 8.3%) between groups. In
14. Bonani M, Frey D, de Rougemont O, Mueller NJ, Mueller TF, Graf N,
Wüthrich RP: Infections in de novo kidney transplant recipients treated another single-center study, health-related quality of life
with the RANKL inhibitor denosumab [published online ahead of print was assessed in 126 SPK recipients pre- and post-transplant
October 28, 2016]. TransplantationPubMed (mean, 5 years) (5). On multiple measures, including gastro-
15. Lenihan CR, Sukumaran Nair S, Vangala C, Ramanathan V, Montez-Rath
ME, Winkelmayer WC: Proton pump inhibitor use and risk of hip fracture in intestinal quality of life, psychological status, social func-
kidney transplant recipients. Am J Kidney Dis 69: 595–601, 2017 PubMed tion, and burden of medical treatment, these measures
16. Kwok CS, Yeong JK, Loke YK: Meta-analysis: Risk of fractures with
improved with transplant but did not improve in those with
acid-suppressing medication. Bone 48: 768–776, 2011 PubMed
only one organ functioning (15.9% of patients).
In 2015, 9.7% of SPK transplants were for in-
dications of type 2 diabetes mellitus (1). This reflects
Multiorgan Transplant
recent clarifications in eligibility criteria for SPK: (1)
Simultaneous Pancreas-Kidney Transplantation renal insufficiency with an eGFR,20 ml/min, (2)
Simultaneous pancreas-kidney transplantation (SPK) requirement for insulin for treatment of diabetes, and
rates have remained stable over the past 2 years after (3) C peptide ,2 ng/ml or C peptide .2 ng/ml with
falling significantly from 2004 to 2013. The number of body mass index cutoff #30 kg/m2. Patient survival in
active new patients on the waiting list fell over 50% SPK recipients with type 1 diabetes mellitus was higher
from 2004 to 2013 as transplant rates fell 25% during than in patients with type 2 diabetes mellitus in the most
the same period. This decline stabilized with 1160 patients recent registry analysis (90% versus 86.4% at 5 years)
added to the waiting list and 719 SPK transplants (1), likely due to older age and comorbidity in the
performed in 2015 (1). The waiting time to transplant type 2 diabetes mellitus cohort. Using these eligibility
remains significantly shorter for SPK transplants than criteria, with the additional criterion of pretransplant
for deceased donor transplants, with an average esti- insulin use ,1 U/kg per day, a comparison of glucose
mated waiting time of slightly over 18 months versus homeostasis in seven type 2 diabetes mellitus SPK
.7 years for deceased donor transplant (2). Five-year recipients with that in nine type 1 diabetes mellitus SPK
survival was 89.6% for SPK recipients transplanted in recipients within 1 year of transplant determined no dif-
2010, a steady improvement over time. An ongoing ferences in measures of insulin resistance and secretion (6).
question is the relative advantage of SPK over kidney Recurrent autoimmunity was recently shown to
transplant alone (KA). A recent analysis of 7308 SPK contribute to pancreas allograft failure. In a cohort of
and 4653 KA adult patients with type 1 diabetes trans- 223 SPK recipients with a history of type 1 diabetes
planted in 1998–2009 simulated a randomized, con- mellitus, antibodies to the autoantigens GAD65, IA-2,
trolled trial by using propensity score matching and and ZnT8 were measured yearly (7). Biopsy-confirmed
transplant center volume to compare survival rates of SPK type 1 diabetes mellitus recurrence (insulitis and ab-
versus KA (3). There was a marginal advantage in patient sence of insulin staining) was found in 5.8% over a
survival (0.17 years) and graft survival (0.18 years) with mean follow-up of 6.264.1 years, and it strongly as-
SPK, findings contrary to previous comparisons (Table 6). sociated with autoantibody conversion but not with
The KA cohort was not stratified by donor type (deceased autoantibody persistence. Recurrent autoimmunity was
versus living donor), suggesting that the marginal benefit not correlated with the type or degree of immunosup-
of SPK may disappear if limited to comparison with living pression. Finally, for patients who experience pancreas
donor KA. Benefits of pancreas transplant beyond survival graft failure, outcomes of a repeat pancreas transplant
rates have also been explored in smaller exploratory are encouraging. A recent comparison of 18 pancreas
studies. Patients with type 1 diabetes who received SPK retransplant recipients with 78 primary pancreas after
(n525) or living donor KA (n517) underwent angiogra- kidney transplant recipients found comparable 3-year
phy, echocardiography, and computed tomography for pancreas graft survival rates: 85.1% in both groups (8).
346 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

Simultaneous Liver-Kidney Transplantation and included matching for dialysis needs and serum
In 2016, eligibility criteria for simultaneous liver- creatinine $2.0 mg/dl pretransplant (12). A small sur-
kidney transplants (SLKs) were developed and ap- vival benefit of 3.7 months at 5 years was noted in
proved by the Organ Procurement and Transplantation patients with pretransplant CKD (no dialysis) who had
Network, and they are in the process of implementa- an SLK versus an LTA. In those requiring dialysis
tion (Table 7). This policy was in response to the (acute and chronic; unspecified duration), there was no
increase in SLK transplants performed over time (626 benefit to SLK over LTA at 5 years post-transplant.
in 2015 compared with 362 in 2009). A summary of the Given the difficulties in predicting who truly re-
rationale has been recently published (9). Some of quires a kidney transplant at the time of liver trans-
the more controversial aspects include the definition plant, a recent report described the use of native renal
of CKD (,30 ml/min on one occasion with a history biopsy before liver transplant to determine if a concur-
of GFR,60 for 90 days) and the duration of AKI rent kidney transplant should be performed (13). In 59
(.6 weeks, GFR,25 ml/min, or dialysis dependence) liver transplant candidates with renal impairment of
that permit eligibility. In the meantime, many single- unclear etiology, SLK transplant was recommended for
center reports and registry analyses have described the patients with .40% global glomerular sclerosis, with
difficulties in defining the best utility of kidneys trans- interstitial fibrosis .30%, or requiring hemodialysis for
planted to either the kidney waiting list or the more .2 months. Based on these criteria, 70% of patients
morbid cohort of liver transplant candidates with renal listed for transplant did not undergo SLK: 23 ultimately
disease (10, 11). Perhaps the analysis most reflective of underwent LTA and ten underwent SLK. Renal function
clinical practice is one that compared 1884 SLK and survival were comparable at 1 year post-transplant.
recipients with 31,882 liver transplant alone (LTA) Importantly, biopsy-related complications were uncom-
recipients transplanted from 2002 to 2009. Patients mon (n52). Thus, biopsy may provide useful informa-
were compared based on propensity score matching tion to minimize unnecessary use of kidneys for SLK.

Table 7. Medical eligibility criteria for adult simultaneous liver-kidney transplantation


Then the Transplant Program Must Report
If the Candidate’s Transplant in the UNOS Computer System and
Nephrologist Confirms a Diagnosis of Document in the Candidate’s Medical Record
CKD with a measured or calculated At least one of the following
GFR#60 ml/min for .90 consecutive d That the candidate has begun regularly administered dialysis as an ESRD
patient in a hospital-based, independent nonhospital-based, or home setting
At the time of registration on the kidney waiting list, that the candidate’s
most recent measured or calculated CrCl or GFR is #30 ml/min
On a date after registration on the kidney waiting list, that the candidate’s
measured or calculated CrCl or GFR is #30 ml/min
Sustained AKI At least one of the following or a combination of both of the following
for the last 6 wk
That the candidate has been on dialysis at least once every 7 d
That the candidate has a measured or calculated CrCl or GFR that is
#25 ml/min at least once every 7 d
If the candidate’s eligibility is not confirmed at least once every 7 d for
the last 6 wk, the candidate is not eligible to receive a liver and a kidney
from the same donor
Metabolic disease A diagnosis of at least one of the following
Hyperoxaluria
Atypical HUS from mutations in factor H or factor I
Familial non-neuropathic systemic amyloidosis
Methylmalonic aciduria
For the adult SLK candidate to be prioritized ahead of all KA candidates at the time of their liver offer, the candidate must meet one of the criteria related to kidney function. UNOS,
United Network for Organ Sharing; CrCl, creatinine clearance; HUS, hemolytic uremic syndrome. Modified from https://optn.transplant.hrsa.gov/media/1240/05_slk_allocation.pdf.
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 347

Parenthetically, the most common renal diagnoses 3. Sung RS, Zhang M, Schaubel DE, Shu X, Magee JC: A reassessment of
the survival advantage of simultaneous kidney-pancreas versus kidney-
from these biopsies were membranoproliferative GN alone transplantation. Transplantation 99: 1900–1906, 2015 PubMed
(23%), IgA nephropathy (19%), and acute tubular 4. Lindahl JP, Massey RJ, Hartmann A, Aakhus S, Endresen K, Günther
necrosis (19%). A, Midtvedt K, Holdaas H, Leivestad T, Horneland R, Øyen O, Jenssen
T: Cardiac assessment of patients with type 1 diabetes median 10 years
Finally, the question of whether the liver is after successful simultaneous pancreas and kidney transplantation
immunoprotective for the kidney in the SLK trans- compared with living donor kidney transplantation. Transplantation
plant setting has been asked. In a relatively large, 101: 1261–1267, 2017 PubMed
5. Martins LS, Outerelo C, Malheiro J, Fonseca IM, Henriques AC, Dias
single-center cohort of 68 SLK recipients, 14 with LS, Rodrigues AS, Cabrita AM, Noronha IL: Health-related quality of
pretransplant donor-specific antibodies (DSAs), pro- life may improve after transplantation in pancreas-kidney recipients.
tocol post-transplant renal biopsies were performed Clin Transplant 29: 242–251, 2015 PubMed
6. Chakkera HA, Kudva YC, Chang YH, Heilman RL, Singer AL, Mathur
(14). Compared with a control group of patients who AK, Hewitt WR, Khamash HA, Huskey JL, Katariya NN, Moss AA,
had undergone KA (n5136), 28 KA individuals with Behmen S, Reddy KS: Glucose homeostasis after simultaneous pan-
creas and kidney transplantation: A comparison of subjects with C-
pretransplant DSAs experienced higher incidence rates
peptide-positive non-type 1 diabetes mellitus and type 1 diabetes
of antibody-mediated rejection (46.4% versus 7.1%) mellitus. Clin Transplant 30: 52–59, 2016 PubMed
and chronic transplant glomerulopathy (53.6% versus 7. Vendrame F, Hopfner YY, Diamantopoulos S, Virdi SK, Allende G,
Snowhite IV, Reijonen HK, Chen L, Ruiz P, Ciancio G, Hutton JC,
0%). KA patients without DSAs had higher frequencies Messinger S, Burke GW 3rd, Pugliese A: Risk factors for type 1 diabetes
of T cell–mediated rejection (30.6% versus 7.4%) and recurrence in immunosuppressed recipients of simultaneous pancreas-
decline in renal function, whereas the SLK cohort had kidney transplants. Am J Transplant 16: 235–245, 2016 PubMed
8. Seal J, Selzner M, Laurence J, Marquez M, Bazerbachi F, McGilvray I, Schiff
stable GFRs. A concurrent liver transplant was strongly J, Norgate A, Cattral MS: Outcomes of pancreas retransplantation after
associated with lower cellular- and antibody-mediated simultaneous kidney-pancreas transplantation are comparable to pancreas after
rejection and stable GFR, suggesting that the liver may kidney transplantation alone. Transplantation 99: 623–628, 2015 PubMed
9. Formica RN, Aeder M, Boyle G, Kucheryavaya A, Stewart D, Hirose R,
indeed protect against chronic immunologic injury. Mulligan D: Simultaneous liver-kidney allocation policy: A proposal to
optimize appropriate utilization of scarce resources. Am J Transplant
Multiorgan Transplantation: Utility 16: 758–766, 2016 PubMed
Considerations 10. Brennan TV, Lunsford KE, Vagefi PA, Bostrom A, Ma M, Feng S:
Although kidney transplant coupled with a second Renal outcomes of simultaneous liver-kidney transplantation compared
to liver transplant alone for candidates with renal dysfunction. Clin
organ may lead to improved outcomes for other organ Transplant 29: 34–43, 2015 PubMed
transplant cohorts, this diminishes use of kidneys for the 11. Hmoud B, Kuo YF, Wiesner RH, Singal AK: Outcomes of liver
kidney transplant candidate. A recent paired kidney anal- transplantation alone after listing for simultaneous kidney: Comparison
to simultaneous liver kidney transplantation. Transplantation 99: 823–
ysis evaluated kidney transplant outcomes where a kidney 828, 2015 PubMed
was transplanted to one recipient (KA) and the contralat- 12. Sharma P, Shu X, Schaubel DE, Sung RS, Magee JC: Propensity score-
based survival benefit of simultaneous liver-kidney transplant over liver
eral kidney was transplanted to an SLK (n51998) or
transplant alone for recipients with pretransplant renal dysfunction.
simultaneous heart-kidney (SHK; n5276) recipient (15). Liver Transpl 22: 71–79, 2016 PubMed
The 5-year survival was higher in the KA cohort (81% 13. Pichler RH, Huskey J, Kowalewska J, Moiz A, Perkins J, Davis CL,
versus 66% in SLK; P,0.001; 84% versus 75% in SHK; Leca N: Kidney biopsies may help predict renal function after liver
transplantation. Transplantation 100: 2122–2128, 2016 PubMed
P50.02). Graft survival was significantly higher in KA 14. Taner T, Heimbach JK, Rosen CB, Nyberg SL, Park WD, Stegall MD:
versus SLK but not SHK. The authors suggest that these Decreased chronic cellular and antibody-mediated injury in the kidney
following simultaneous liver-kidney transplantation. Kidney Int 89:
outcomes are contrary to allocation policies that are meant 909–917, 2016 PubMed
to optimize survival overall but instead, favor multiorgan 15. Choudhury RA, Reese PP, Goldberg DS, Bloom RD, Sawinski DL, Abt
transplantation. For greater details regarding allocation PL: A paired kidney analysis of multiorgan transplantation: Implications
for allograft survival. Transplantation 101: 368–376, 2017 PubMed
issues and outcomes in multiorgan transplant, the reader 16. Stites E, Wiseman AC: Multiorgan transplantation. Transplant Rev
is referred to a recent review (16). (Orlando) 30: 253–260, 2016 PubMed

References
1. Kandaswamy R, Stock PG, Gustafson SK, Skeans MA, Curry MA,
Pregnancy
Prentice MA, Israni AK, Snyder JJ, Kasiske BL: OPTN/SRTR 2015
Annual Data Report: Pancreas. Am J Transplant 17[Suppl 1]: 117–173,
A greater understanding of the risks of pregnancy
2017 PubMed for the kidney transplant recipient and fetus has been
2. Hart A, Smith JM, Skeans MA, Gustafson SK, Stewart DE, Cherikh advanced by recent large, single-center experiences,
WS, Wainright JL, Kucheryavaya A, Woodbury M, Snyder JJ, Kasiske
BL, Israni AK: OPTN/SRTR 2015 Annual Data Report: Kidney. Am J voluntary registry analyses, and patient surveys. In a
Transplant 17[Suppl 1]: 21–116, 2017 PubMed single-center experience of 138 pregnancies in 89 renal
348 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

transplant patients from 1973 to 2013, 74% resulted in the first trimester (6). Rather than interpret this as safe to
live births (1). Of these, 61% births were premature, continue MPA during the first trimester, a more appropri-
52% had low birth weight, and 14% were associated ate interpretation is that MPA should be discontinued as
with preeclampsia. Long-term graft function (eGFR at soon as awareness of pregnancy (or planning) is present.
1, 5, and 10 years) and 10-year patient survival were A general guideline is that mycophenolate should be
no different compared with those in a paired, nonpregnant transitioned to azathioprine .6 weeks prior to conception.
control group. Thus, although obstetric complications An editorial on this manuscript, which includes reference
were common, most pregnancies were successful, with to a separate National Transplantation Pregnancy Registry
no overt risk to transplant function. A similar registry analysis, describes pregnancy outcomes in 302 recipients
report from Italy in 1978–2013 compared 222 successful who discontinued MPA before conception and 142 who
pregnancies with 1418 low-risk controls, with general took MPA in the first trimester. Higher incidences of
findings of higher rates of preterm delivery (57.3% versus miscarriages (20% versus 48%; P,0.001) and birth defects
6.3%; P,0.001) and babies small for gestational age (5.7% versus 11.6%; P50.10) were noted in the latter
(14.0% versus 4.5%; P,0.001) (2). In a subset of this group (7). With respect to other immunosuppressive agents,
cohort, 121 successful pregnancies from 2000 to 2014 a recent study of changes in calcineurin inhibitor dose and
were compared with the same control group (n51418) blood trough concentrations highlighted changes that
and an additional control group of 610 births in mothers occur in pregnancy (8). In 75 women (88 deliveries),
with CKD, risk for preterm delivery was similar in the cyclosporine (n560) and tacrolimus (n528) trough
CKD and transplant cohorts, suggesting that degree of concentrations fell by the second trimester compared
CKD rather than transplant status was the primary with 12 months before delivery. These declines necessi-
determinant for maternal-fetal outcomes (3). tated an increase in calcineurin inhibitor dose of 20%–25%,
Greater insight into the appropriate timing post- a reasonable rule of thumb to expect during pregnancy.
transplant to pursue pregnancy for the health of the When considering risk factors for untoward out-
transplant was provided by a registry analysis of 21,814 comes during pregnancy, one concern is that highly
women ages 15–45 years old transplanted from 1990 to sensitized patients may be more susceptible to rejection,
2010 with 729 pregnancies occurring within 3 years of posing a greater risk to graft outcomes. In a retrospective,
transplant (4). When examining risk of death-censored single-center study of 11 pregnant patients, eight had
graft loss after pregnancy in years 1–3 post-transplant, an detectable panel reactive antibody levels (.0%) (9). Two
increased risk was noted in those with pregnancies during of the eight had second trimester miscarriages, one had
the first post-transplant year (hazard ratio, 1.25; 95% a stillbirth, and after a median follow-up of 2.3 years after
confidence interval [95% CI], 1.04 to 1.50) and year 2 delivery, three of the eight developed antibody-mediated
(hazard ratio, 1.26; 95% CI, 1.06 to 1.50) but not in year rejection leading to graft loss. Although impossible to
3. Thus, current recommendations to wait at least 1 year control for the many risk factors for rejection, this small
post-transplant before pursuing pregnancy (5) perhaps report raises concern for fetal and allograft outcomes in
should be tempered with the understanding that increased sensitized patients.
risk persists into the second year post-transplant. Lastly, whether organ transplant immunosuppres-
Beyond timing of pregnancy, the type and degree of sion in the father impacts pregnancy outcomes was
immunosuppression are also an important consideration recently explored. A Norwegian registry analysis iden-
for individuals who plan to conceive or become pregnant. tified 2463 men who fathered babies of 4614 deliveries
Mycophenolate (MPA) is teratogenic and carries a US before and 474 deliveries after solid organ transplanta-
Food and Drug Administration advisory to discontinue tion (396 after kidney transplant) from 1967 to 2009 and
this agent when planning for pregnancy. A recent analysis compared these cohorts with the general population of
of 382 pregnancies in patients with a history of MPA use 2,511,506 births (10). They found no increased risk of
from the National Transplantation Pregnancy Registry congenital malformations or other fetal or maternal
reported that birth defects and miscarriages were similar outcomes compared with in the general population.
among kidney transplant recipients who discontinued Higher rates of preeclampsia (adjusted odd ratio, 7.4;
MPA.6 versus ,6 weeks before pregnancy and during 95% CI, 1.1 to 51.4) occurred in pregnancies post-
the first trimester, whereas the risks of were significantly transplant compared with pretransplant. Given the wide
increased when MPA was continued or discontinued after 95% CI, this finding should be interpreted with caution.
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 349

References months post-transplant was associated with increased


1. Stoumpos S, McNeill SH, Gorrie M, Mark PB, Brennand JE, Geddes risk of death-censored graft failure (hazard ratio [HR],
CC, Deighan CJ: Obstetric and long-term kidney outcomes in renal
transplant recipients: A 40-yr single-center study. Clin Transplant 30: 1.66; 95% confidence interval [95% CI], 1.14 to 2.42),
673–681, 2016 PubMed and time-varying concentration of total CO2 ,22
2. Piccoli GB, Cabiddu G, Attini R, Gerbino M, Todeschini P, Perrino mmol/L was also associated with death-censored
ML, Manzione AM, Piredda GB, Gnappi E, Caputo F, Montagnino G,
Bellizzi V, Di Loreto P, Martino F, Montanaro D, Rossini M, Castellino graft loss (HR, 3.17; 95% CI, 2.12 to 4.73) and death
S, Biolcati M, Fassio F, Loi V, Parisi S, Versino E, Pani A, Todros T; (HR, 3.16; 95% CI, 1.77 to 5.62), even after control-
Italian Study Group on Kidney and Pregnancy of the Italian Society of
ling for eGFR. Whether intervention reduces this risk
Nephrology: Pregnancy outcomes after kidney graft in Italy: Are the
changes over time the result of different therapies or of different is yet unproven. Further research in this area is
policies? A nationwide survey (1978-2013). Nephrol Dial Transplant required (2).
31: 1957–1965, 2016 PubMed
3. Piccoli GB, Cabiddu G, Attini R, Gerbino M, Todeschini P, Perrino
Normal saline is commonly and liberally admin-
ML, Manzione AM, Piredda GB, Gnappi E, Caputo F, Montagnino G, istered in the perioperative period after kidney trans-
Bellizzi V, Di Loreto P, Martino F, Montanaro D, Rossini M, Castellino plantation. A theoretical concern is that this practice
S, Biolcati M, Fassio F, Loi V, Parisi S, Versino E, Pani A, Todros T;
Italian Study group on Kidney and Pregnancy of the Italian Society of
may provoke hyperchloremic metabolic acidosis and
Nephrology: Outcomes of pregnancies after kidney transplantation: the risk of hyperkalemia with need for dialysis. A
lessons learned from CKD. A comparison of transplanted, nontrans- meta-analysis of lower-chloride solutions versus nor-
planted chronic kidney disease low-risk pregnancies: A multicenter
nationwide analysis [published online ahead of print January 21, 2017].
mal saline identified six randomized trials with 477
Transplantation doi:10.1097/TP.0000000000001645 patients (3). No difference in risk of hyperkalemia or
4. Rose C, Gill J, Zalunardo N, Johnston O, Mehrotra A, Gill JS: Timing delayed graft function was noted, and there were no
of pregnancy after kidney transplantation and risk of allograft failure.
Am J Transplant 16: 2360–2367, 2016 PubMed significant differences in rejection or graft loss. The
5. Kidney Disease: Improving Global Outcomes Transplant Work Group: use of balanced electrolyte solutions was associated
KDIGO clinical practice guideline for the care of kidney transplant with higher pH, higher bicarbonate, and lower serum
recipients. Am J Transplant 9[Suppl 3]: S1–S155, 2009 PubMed
6. King RW, Baca MJ, Armenti VT, Kaplan B: Pregnancy outcomes chloride concentrations. A strong case cannot be made
related to mycophenolate exposure in female kidney transplant recipi- for the use of one solution versus the other.
ents. Am J Transplant 17: 151–160, 2017 PubMed The ongoing question of whether elevated uric
7. Moritz MJ, Constantinescu S, Coscia LA, Armenti D: Mycophenolate
and pregnancy: Teratology principles and national transplanta- acid is a marker for or contributor to progressive CKD
tion pregnancy registry experience. Am J Transplant 17: 581–582, and graft loss was addressed in a novel retrospective
2017 PubMed
8. Kim H, Jeong JC, Yang J, Yang WS, Ahn C, Han DJ, Park JS, Park SK:
analysis of 1170 transplants from 2000 to 2010 (4).
The optimal therapy of calcineurin inhibitors for pregnancy in kidney Using time-dependent marginal structural Cox propor-
transplantation. Clin Transplant 29: 142–148, 2015 PubMed tional hazards models that controlled for kidney func-
9. Ajaimy M, Lubetzky M, Jones T, Kamal L, Colovai A, de Boccardo G,
Akalin E: Pregnancy in sensitized kidney transplant recipients: A
tion, with eGFR as a time-varying confounder, uric acid
single-center experience. Clin Transplant 30: 791–795, 2016 PubMed was not associated with graft failure (HR, 0.90; 95% CI,
10. Morken NH, Diaz-Garcia C, Reisaeter AV, Foss A, Leivestad T, Geiran 0.85 to 0.94 for every 10-mmol/L increase in uric acid),
O, Hervás D, Brännström M: Obstetric and neonatal outcome of
pregnancies fathered by males on immunosuppression after solid organ
death-censored graft loss, or death. This supports the
transplantation. Am J Transplant 15: 1666–1673, 2015 PubMed concept that hyperuricemia is not an independent risk
factor for progressive renal dysfunction post-transplant
but rather, is a marker for lower eGFR.
Acid-Base and Electrolyte Disorders after Transplant
A few recent studies were published on acid-base References
and electrolyte metabolism in transplantation during 2015 1. Park S, Kang E, Park S, Kim YC, Han SS, Ha J, Kim DK, Kim S, Park
and 2016. A summary of key studies includes (1) me- SK, Han DJ, Lim CS, Kim YS, Lee JP, Kim YH: Metabolic acidosis and
long-term clinical outcomes in kidney transplant recipients. J Am Soc
tabolic acidosis and transplant outcomes, (2) fluid man- Nephrol 28: 1886–1897, 2017 PubMed
agement post-kidney transplant, and (3) a novel analysis of 2. Messa PG, Alfieri C, Vettoretti S: Metabolic acidosis in renal trans-
the interplay of uric acid and allograft outcomes. plantation: Neglected but of potential clinical relevance. Nephrol Dial
Transplant 31: 730–736, 2016 PubMed
In a multicenter, retrospective cohort of 2318 3. Wan S, Roberts MA, Mount P: Normal saline versus lower-chloride
kidney transplant recipients from 1997 to 2015, the solutions for kidney transplantation. Cochrane Database Syst Rev 8:
relationship of serum bicarbonate (total CO2 con- CD010741, 2016 PubMed
4. Kim ED, Famure O, Li Y, Kim SJ: Uric acid and the risk of graft failure
centration) and outcomes was described (1) After in kidney transplant recipients: A re-assessment. Am J Transplant 15:
multivariable adjustment, total CO2 ,22 mmol/L at 3 482–488, 2015 PubMed
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

Transplantation
Claiming Credits and Evaluation Process
Accreditation Statement
The American Society of Nephrology (ASN) is accredited by the Accreditation Council for Continuing Medical Education to
provide continuing medical education for physicians.
AMA Credit Designation Statement
The ASN designates this enduring material for a maximum of 10 AMA PRA Category 1 Credits™. Physicians should claim
only the credit commensurate with the extent of their participation in the activity.
Original Release Date: November 2017
CME Credit Termination Date: October 31, 2019
Examination Available Online: On or before Wednesday, November 15, 2017
Estimated Time for Completion: 10 hours
Answers with Explanations
• Provided with a passing score after the first and/or after the second attempt
• November 2019: posted on the ASN website when the issue is archived.
Method of Participation
• Read the syllabus that is supplemented by original articles in the reference lists.
• Complete the online self-assessment examination.
• Each participant is allowed two attempts to pass the examination (.75% correct) for CME credit.
• Upon completion, review your score and incorrect answers and print your certificate.
• Answers and explanations are provided with a passing score or after the second attempt.
Activity Evaluation and CME Credit Instructions
• Go to www.asn-online.org/cme, and enter your ASN login on the right.
• Click the ASN CME Center.
• Locate the activity name and click the corresponding ENTER ACTIVITY button.
• Read all front matter information.
• On the left-hand side, click and complete the Demographics & General Evaluations.
• Complete and pass the examination for CME credit.
• Upon completion, click Claim Your CME Credits, check the Attestation Statement box, and enter the number of
CME credits commensurate with the extent of your participation in the activity.
• If you need a certificate, Print Your Certificate on the left.
For your complete ASN transcript, click the ASN CME Center banner, and click View/Print Transcript on the left.

Instructions to obtain American Board of Internal Medicine (ABIM) Maintenance of Certification (MOC) Points
Each issue of NephSAP provides 10 MOC points. Respondents must meet the following criteria:
• Be certified by ABIM in internal medicine and/or nephrology and enrolled in the ABIM–MOC program
• Enroll for MOC via the ABIM website (www.abim.org).
• Enter your (ABIM) Candidate Number and Date of Birth prior to completing the examination.
• Take the self-assessment examination within the timeframe specified in this issue of NephSAP.
• Upon completion, click Claim Your MOC points, the MOC points submitted will match your CME credits claimed,
check the Attestation Statement box and submit.
• ABIM will notify you when MOC points have been added to your record.

Maintenance of Certification Statement


Successful completion of this CME activity, which includes participation in the evaluation component, enables the participant to
earn up to 10 MOC points in the American Board of Internal Medicine’s (ABIM) Maintenance of Certification (MOC) program.
Participants will earn MOC points equivalent to the amount of CME credits claimed for the activity. It is the CME activity
provider’s responsibility to submit participant completion information to ACCME for the purpose of granting ABIM MOC credit.
MOC points will be applied to only those ABIM candidates who have enrolled in the MOC program. It is your responsibility to
complete the ABIM MOC enrollment process.

350
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

NephSAP, Volume 16, Number 4, November 2017—Transplantation

1. A 68-year-old black man with ESRD is waitlisted transplant. You begin a discussion about the bene-
for deceased donor kidney transplantation. His blood fits and risks of a Public Health Service Increased
type is A, and he has a calculated panel reactive Risk Donor (PHS-IRD) kidney. He is ambivalent
antibody (cPRA) level of 80%. about receiving a kidney from an intravenous drug
Since implementation of the new Kidney Allo- user and concerned about the risk of acquiring an
cation System (KAS) by the United Network for infection from the procedure.
Organ Sharing on December 4, 2014, which Which ONE of the following should you tell him
ONE of the following factors now associates about PHS-IRD kidneys?
with him having an INCREASED likelihood of A. Transplantation of PHS-IRD kidneys is as-
receiving a deceased donor kidney compared sociated with an increased waiting time
with the prior allocation system? B. Kidneys from individuals with increased
A. His race risk behaviors currently make up ,10% of
B. His cPRA all deceased donors
C. His age C. Kidneys from intravenous drug users are
D. His blood type never used for transplantation
2. A 66-year-old man is referred for kidney trans- D. His risk of contracting HIV, hepatitis C virus
plantation. He has stage 5 CKD due to membranous (HCV), or hepatitis B virus from a PHS-IRD
nephropathy (MN), and he has been maintained on donor is ,1%
hemodialysis for 3 months. He uses a walker and 4. A 59-year-old woman with stage 5 CKD is referred
needs help to get dressed each day. His examination for transplantation. She donated a kidney to her sister
shows a slow walking speed and difficulty standing 18 years ago and now has advanced CKD compli-
from a sitting position. He asks about his prognosis cating diabetes mellitus and hypertension. She has
regarding kidney transplantation. had multiple pregnancies and received several blood
Which ONE of the following should you tell him transfusions after a recent gastrointestinal bleed.
about the effect of his physical function on his Her blood type is A, and her cPRA is 9%. She is
transplantation outcomes? concerned about the effect of the new KAS on her
A. It will not affect his graft survival after chances of receiving a kidney. A discussion ensues
kidney transplantation about kidney transplantation options.
B. It is associated with increased waitlist Which ONE of the following should you tell her
mortality regarding her options for kidney
C. It is associated with a decreased risk for transplantation?
readmissions after transplantation A. The quality of decreased donor kidneys has
D. Physical therapy can mitigate the effects of deteriorated since implementation of Kidney
frailty on his transplant outcomes Allocation System (KAS), with a median
3. A 56-year-old man with ESRD due to IgA nephrop- kidney donor risk index of .80%
athy returns for his routine yearly re-evaluation visit B. Prior living donors have improved access to
with your local transplant program. He has been transplantation since implementation of KAS
listed for deceased donor transplantation for 4 C. Her projected waiting time for a deceased
years. His blood type is B. No living donors have donor kidney is likely to be ,4 months
been identified. He previously was listed for D. She can be listed for transplantation only
an expanded criteria donor kidney, and he asks after initiation of dialysis
whether there are any other options to increase 5. A 63-year-old man with ESRD and HCV infection
his chances of receiving a deceased donor kidney (genotype 1) as a consequence of a remote history
351
352 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

of intravenous drug use is listed for kidney trans- a low-level donor-specific antibody targeting HLA-
plantation. He is HCV treatment naı̈ve. He asks A2 (mean fluorescence intensity of 1582). The flow
whether it would be better to opt for an HCV-positive cytometry crossmatch is negative. She receives the
kidney compared with a standard criteria HCV- deceased donor kidney after 28 hours of cold
negative kidney. His HCV viral load is persistently ischemia time. The preimplantation biopsy shows
high at .450,000 IU/ml. no fibrin thrombi. Postoperatively, her urine output
Which ONE of the following statements should is 3 L over the initial 24 hours. However, the serum
you tell him regarding the expected outcomes of creatinine level rises from 5 mg/dl on postoperative
receiving an HCV-positive kidney and manage- day 1 to a peak of 6.2 mg/dl on postoperative day 7.
ment of HCV infection? She does not require dialysis, and the serum creatinine
A. Consent to receive an HCV-positive kidney subsequently begins to decline to 1.4 mg/dl over the
may dramatically reduce his waiting time next 2 weeks.
B. Receipt of an HCV-positive donor kidney is Which ONE of the following statements about
associated with worse allograft survival com- this patient’s delayed graft function (DGF) is
pared with waiting for an HCV-negative donor CORRECT?
kidney A. The presence of the donor-specific antibody
C. Consent to receive an HCV-positive donor increases her risk of DGF
kidney is not recommended in order to avoid B. Her BMI does not influence her risk of DGF
infection with a virus of different genotype C. Her DGF only affects allograft prognosis
D. HCV infection should be treated before when fibrin thrombi are seen on preimplan-
transplantation because direct-acting antivi- tation biopsy
ral agents are contraindicated in transplant D. Her DGF only affects long-term allograft
recipients receiving calcineurin inhibitors function if she develops rejection
6. Your transplant surgeon informs you that a de- 8. A 29-year-old woman is seen in consultation to
ceased donor kidney with premortem AKI has assess her candidacy to donate a kidney to her
been offered to a dialysis patient from an affiliated father. She is planning to have a family within the
practice. The donor was 28 years old. The terminal next 3 years. All of her donor testing is normal. She
serum creatinine level is 2.5 mg/dl. The preim- is ABO compatible and has a negative crossmatch
plantation allograft biopsy shows fibrin thrombi. with her father. They are a one-haplotype match.
Glomerulosclerosis is not present. Your nephrol- Which ONE of the following should you tell her
ogy colleague asks you about the expected post- about her pregnancy risk after live donor
transplant prognosis. nephrectomy?
Which ONE of the following should you tell A. Pregnancy after kidney donation is more likely
your colleague about the expected prognosis of to be complicated by low-birth weight infants
transplanting this kidney? B. Pregnancy after kidney donation is more
A. The risk of delayed graft function (DGF) is likely to be complicated by preterm births
equivalent to receiving a deceased donor C. Pregnancy after kidney donation is associated
kidney without AKI with an increased risk of gestational hyperten-
B. This donor’s preimplantation kidney biopsy sion and preeclampsia
predicts that DGF is certain D. The risk of neonatal mortality is higher among
C. The finding of fibrin thrombi on biopsy is kidney donors compared with nondonors
associated with .10% lower allograft sur- 9. A 23-year-old white man has been accepted as
vival at 1 year an altruistic living donor. He is completely healthy
D. AKI in the donor kidney does not affect and does not smoke. On physical examination, his
allograft function at 1 year BP is 110/80 mmHg. His creatinine clearance is 106
7. A 59-year-old woman with ESRD due to polycystic ml/min. He asks about his future risk of developing
kidney disease is admitted for transplantation. She kidney failure.
is asymptomatic. Her physical examination shows Which ONE of the following should you tell him
a body mass index (BMI) of 37 kg/m2. She has about his risk of ESRD after kidney donation?
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 353

A. It is higher compared with black donors B. The risk of cytomegalovirus (CMV) viremia
B. It is no higher than that of nondonors in the is greater with azathioprine compared with
general population mycophenolate
C. It is three to five times higher than non- C. Mycophenolate is associated with a reduced
donors in the general population risk of mortality compared with azathioprine
D. It may be lower than that of older donors D. Physical frailty does not affect the frequency
10. A 32-year-old woman with ESRD due to reflux of mycophenolate adverse events
nephropathy is seen in follow-up 6 weeks after E. Dose-limiting adverse effects are more com-
receiving a 6/6-antigen mismatch kidney transplant monly seen in living compared with de-
from a living unrelated donor. She received in- ceased donor allograft recipients
duction therapy with antithymocyte globulin and is 13. A 38-year-old kidney transplant recipient is admit-
now maintained on tacrolimus, mycophenolic acid ted for management of an upper gastrointestinal
(MPA), and prednisone 15 mg daily, with a plan bleed. She is found to have a bleeding gastric ulcer
to taper prednisone to 5 mg daily by week 10. that is cauterized endoscopically. She is fatigued
Her serum creatinine concentration is 1.2 mg/dl. but otherwise asymptomatic. Her allograft function
She wishes to taper entirely off prednisone because has been excellent since transplantation 4 months
of weight gain and onset of glucose intolerance. ago, and her serum creatinine has been stable at 1.3
Which ONE of the following should you advise mg/dl. Her medications include cyclosporine, my-
her about the risks of steroid elimination after cophenolate mofetil, prednisone, and valganciclo-
transplantation? vir. Her hemoglobin stabilizes at 6.7–6.9 g/dl. Her
A. It is associated with an increased risk of internist recommends transfusion with 1 U packed
mortality red blood cells. You discuss the benefits and risks
B. It will lower her risk of infection of transfusion as well as transfusion methods with
C. It will lower her risk of developing diabetes her internist.
mellitus Which ONE of the following statements should
D. It will lower her risk of malignancy you advise about the risks and methods of
E. It is associated with an increased risk of transfusion in this patient?
acute rejection A. There is no risk of allosensitization, because
11. Which ONE of the following statements about she is immunosuppressed
antibody induction therapy for kidney trans- B. Blood transfusion may induce donor-specific
plantation is CORRECT? antibody and an increased risk of rejection
A. Basiliximab therapy associates with marked C. Leukocyte filtration will eliminate the risk of
reduction in rejection risk in tacrolimus/ allosensitization
mycophenolate/steroid-treated nonsensi- D. If she receives blood, it should be CMV
tized patients negative
B. Rituximab induction reduces rejection risk E. If she receives blood, it should be irradiated
in nonsensitized allograft recipients 14. A 48-year-old woman with ESRD due to lupus
C. An alemtuzumab/prednisone-free regimen nephritis is evaluated during her annual visit while
is becoming the most widely used strategy waitlisted for transplantation. Her cPRA is 99%
in the United States as a result of prior pregnancies and remote trans-
D. Induction therapy associates with approxi- fusions. She has been on the waitlist for 3 years.
mately 50% rejection risk reduction com- Her sister is her only potential living donor.
pared with no induction therapy However, their complement-dependent cytotoxic-
12. Which ONE of the following statements about ity crossmatch is positive. Her nephrologist con-
use of mycophenolate after kidney transplanta- sults you about her candidacy for desensitization.
tion is CORRECT? She has already been listed for kidney paired
A. Dose reductions stemming from adverse donation nationally.
effects are associated with an increased risk Which ONE of the following should you advise
of rejection and graft failure about HLA antibody desensitization?
354 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

A. HLA antibody desensitization using plasma- Which ONE of the following is the MOST ap-
pheresis and intravenous Ig is associated with propriate management strategy for this woman?
the same excellent outcomes as desensitization A. No further treatment
to permit ABO-incompatible transplantation B. Denosumab
B. Treatment of subclinical rejection detected C. A bisphosphonate
by protocol biopsy abrogates the risk of allo- D. Teriparatide
graft failure after HLA antibody desensitization 17. A 56-year-old man has hypercalcemia and hyper-
C. The addition of rituximab to her regimen parathyroidism 18 months after a second deceased
will increase the risk of infection and ma- donor kidney transplant. Laboratory studies show
lignancies at 2 years a serum creatinine of 1.0 mg/dl, a PTH of 1208 pg/ml
D. HLA antibody desensitization is associated (reference range 512–88 pg/ml), a serum calcium of
with inferior patient and graft survival com- 10.6–11.2 mg/dl, a serum albumin of 4.2 g/dl, a serum
pared with HLA-compatible transplantation phosphorus of 1.4 mg/dl, and a serum 25-hydroxy-
E. Thrombotic microangiopathy after desensi- vitamin D level of 35 ng/ml. A dual energy x-ray
tization is mitigated by plasmapheresis and absorptiometry scan shows a Z score of 22.1 and a T
does not affect allograft outcome score of 22.7, which was a significant decrease of 7.1%
15. A 68-year-old woman with polycystic kidney disease compared with a previous measurement 1 year ago.
receives a kidney donor profile index .86% deceased Parathyroid scintigraphy shows four enlarged glands.
donor kidney. Her achieved serum creatinine level is Which ONE of the following is the MOST ap-
1.8 mg/dl. She is tolerating maintenance immuno- propriate next step in treatment?
suppression with tacrolimus and mycophenolate well. A. Subtotal parathyroidectomy
In addition to serum creatinine levels, which B. Cinacalcet 30 mg daily and titrate until the
ONE of the following biomarkers should be serum PTH level is normalized
used to best monitor her allograft stability? C. Paricalcitol 1 mg/d titrated to 2 mg/d as tolerated
A. Kidney injury marker-1 D. Denosumab 60 mg administered every 6
B. Neutrophil gelatinase–associated lipocalin months
C. Perforin 18. A 31-year-old woman is seen for preconception
D. Granzyme B counseling 5 years after successful kidney trans-
E. Urine protein-to-creatinine ratio and/or urine plantation. She is maintained on stable doses of
albumin-to-creatinine ratio tacrolimus, mycophenolate mofetil (MMF), and
16. A 54-year-old woman with a history of reflux prednisone. She had one episode of mild acute
nephropathy and recurrent urinary tract infection cellular rejection 3 months after transplantation that
is seen six months after living donor kidney trans- responded to pulse methylprednisolone. Her serum
plantation. She had prompt function of the allograft, creatinine level has been stable in the 1.4- to 1.6-
and her serum creatinine level has been stable at mg/dl range over the past year, and the urine protein-
1.2 mg/dl for the last six weeks. She is concerned to-creatinine ratio has averaged 400 mg/d. A biopsy
about her bone health because her 10-year proba- 2 years ago showed mild to moderate transplant
bility of a hip fracture is estimated to be 3.2% based glomerulopathy with no active inflammation. At that
on the Fracture Risk Assessment Tool (FRAXÒ). time, a donor-specific antibody to HLA-DQ7 was
Her immunosuppression includes tacrolimus, detected with a mean fluorescence intensity of 7000.
mycophenolate mofetil, and prednisone. The serum Which ONE of the following is the MOST ap-
creatinine level is 1.2 mg/dl (eGFR .60 ml/min per propriate management of her immunosuppres-
1.73 m2), the parathyroid hormone (PTH) level is sion in preparation for and during pregnancy?
116 pg/ml (reference range, 12 - 88 pg/ml), the serum A. Transition MMF to azathioprine .6 weeks prior
calcium is 9.1 mg/dl, the phosphorus is 2.6 mg/dl, the to attempts to conceive and plan to increase
serum alkaline phosphatase is normal, the 25-hydroxy- tacrolimus about 20%–25% during the sec-
vitamin D level is 30 ng/ml (reference range 5 30– ond trimester to maintain therapeutic levels
50 ng/ml), and the 1,25-dihydroxy-vitamin D level B. Wait until the first trimester to discontinue
is 32 pg/ml (reference range 5 25–65 pg/ml). mycophenolate mofetil
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 355

C. Transition mycophenolate mofetil to siroli- dialysis to maintain systolic BP .90 mmHg. Her
mus now systolic BP on nondialysis days is 90–100 mmHg.
D. Discontinue tacrolimus after pregnancy is Her medications include omeprazole, aspirin, seve-
confirmed lamer carbonate, magnesium oxide, and methoxy
19. A 54-year-old man with type 2 diabetes mellitus is polyethylene glycol epoetin-b. She had a normal
evaluated for persistent hypertension 2 years after exercise stress test within the past year. An echo-
a living related kidney transplant. He has no prior cardiogram shows severe left ventricular hypertro-
cardiovascular history. His medications are tacroli- phy. Her laboratory studies are significant for a low
mus, sirolimus (implemented because of gastroin- serum magnesium of 1.3 mg/dl.
testinal intolerance to mycophenolate), prednisone, In considering her clinical risk factors, which
amlodipine 10 mg daily, and hydrochlorothiazide 25 ONE of the following clinical features is MOST
mg daily. The systolic BP has consistently been in the highly associated with graft loss after
150- to 159-mmHg range using home BP monitoring. transplant?
On physical examination, the BP is 158/90 mmHg. A. Midodrine use
There is no lower extremity edema. Laboratory B. Hypomagnesemia
studies show a serum creatinine level of 0.9 mg/dl, C. Proton pump inhibitor use
potassium of 4.3 mEq/L, and a urine protein-to- D. Aspirin use
creatinine ratio of 200 mg/g. 22. A 44-year-old man with HIV-associated nephrop-
Which ONE of the following is the MOST ap- athy has been on dialysis for 4 years. His HIV viral
propriate agent to add for BP control? load is undetectable on highly active antiretroviral
A. Furosemide therapy, with a recent CD4 count of 480/ml and no
B. Hydralazine opportunistic infections. He wishes to pursue
C. Labetalol kidney transplantation.
D. Lisinopril When considering his transplant referral and
20. A 62-year-old man with ESRD due to type 2 management, which ONE of the following is the
diabetes mellitus currently maintained on hemodi- MOST appropriate management?
alysis for 1 year is seen in consultation to evaluate A. Use of cyclosporine rather than tacrolimus
his candidacy for kidney transplantation. His past for maintenance immunosuppression
medical history is significant for stable coronary B. Use of thymoglobulin induction therapy
artery disease that required placement of a drug- C. Referral to a transplant center with extensive
eluting stent in the left circumflex artery 3 years prior experience in HIV-positive transplantation
ago. His BMI is 36.5 kg/m2. D. Avoidance of HIV-positive deceased donors
Which ONE of the following is the MOST ap- 23. A 62-year-old man was found to have BK virus
propriate statement regarding his expected (BKV) viremia (16,240 copies/ml) 6 months after
transplant outcome? deceased donor transplant while on tacrolimus,
A. His risk of graft loss (death censored) is similar MPA, and prednisone. A biopsy of the allograft 6
to that of nonobese transplant recipients months after transplantation showed no evidence
B. His mortality risk is higher than nonobese of rejection or BKV nephropathy. His tacrolimus
transplant recipients and MPA doses were initially reduced by 25%.
C. His risk of DGF is higher than nonobese MPA was further reduced by an additional 50%
transplant recipients over the subsequent 3 months due to persistent
D. His mortality risk with transplant is higher viremia. Over the next 3 months, BKV viremia
than his nonobese waitlisted counterparts remained present at very low levels (1000–2000
21. A 64-year-old woman with ESRD due to chronic copies/ml). Renal allograft function remained sta-
GN and a remote history of a transient ischemia ble throughout the 6-month period, with serum
attack and peptic ulcer disease has been on hemo- creatinine of 1.2 mg/dl.
dialysis for 6 years. She is active on the kidney Compared with individuals who never de-
waitlist. She suffers from hypotension during di- veloped BKV viremia, which ONE of the fol-
alysis and requires regular use of midodrine during lowing is the MOST likely clinical outcome?
356 Nephrology Self-Assessment Program - Vol 16, No 4, November 2017

A. Development of de novo donor-specific D. Recurrent MN with progressive proteinuria


antibodies typically responds to rituximab
B. Graft loss due to BKV nephropathy 26. A 49-year-old woman with type 2 diabetes mellitus
C. Development of a second viral infection and stage 5 CKD undergoes an evaluation for kidney
D. Increased mortality transplantation. She has no identified living donors.
24. A 55-year-old man with ESRD secondary to type She has had insulin-requiring diabetes mellitus for 18
2 diabetes mellitus is evaluated for future kidney years. Her BMI is 28 kg/m2. Laboratory studies show
transplantation. He has been on dialysis for 6 months, an eGFR of 14 ml/min per 1.73 m2, a C-peptide level
and he has a potential living donor, his wife, who of 3.2 ng/ml, and a hemoglobin A1c of 6.8%. She
is otherwise healthy. He is a nonsmoker. He has wishes to discuss pancreas transplantation.
a history of squamous cell skin cancer on his scalp Which ONE of the following is the MOST ac-
that was diagnosed and successfully resected with curate when discussing simultaneous pancreas-
clear margins 1 year ago. The lesion was 2.1 cm in kidney (SPK) transplantation with her?
size. He had a 1-cm tubular adenomatous polyp A. She is an eligible candidate for SPK
resected 2 years ago during screening colonoscopy. transplantation
He has no symptoms of urinary frequency, urgency, B. SPK transplant in type 2 diabetes mellitus
or hesitancy. He does not have hematuria and has not recipients is associated with a higher risk of
worked in aluminum smelting, rubber manufacture, rejection compared with in type 1 diabetes
or leather industries. There is no family history of mellitus recipients
colon cancer or prostate cancer. C. SPK transplant in type 2 diabetes mellitus
Regarding his risk for malignancy after trans- recipients is associated with an increased
plant, which ONE of the following is the MOST risk of primary nonfunction compared with
appropriate management recommendation? in type 1 diabetes mellitus recipients
A. An additional 1-year waiting time (total D. SPK transplant in type 2 diabetes mellitus
waiting time 52 years) on the basis of prior recipients is associated with worse pancreas
skin cancer history graft survival compared with in type 1 diabetes
B. Repeat colonoscopy now before proceeding mellitus recipients
with a living donor transplant 27. A 55-year-old woman with type 1 diabetes and ESRD
C. Check a screening serum prostate-specific for 6 months is evaluated for kidney transplantation.
antigen level She has consistently achieved a hemoglobin A1c level
D. Refer him for screening cystoscopy of about 7.2 %–7.8% using an insulin pump. Her
25. A 55-year-old man with ESRD secondary to membra- BMI is 31 kg/m2. She has not had emergency care or
nous nephropathy (MN) is seen to assess his candi- third party assistance for hypoglycemia. She does not
dacy for kidney transplantation. He still produces urine have symptomatic cardiovascular disease. She has no
and has persistent proteinuria (6 g/d). An antiphospho- potential living kidney donors and wishes to consider
lipase A2 receptor autoantibody level drawn at the time pancreas transplantation. Her C peptide is ,2 ng/ml.
of evaluation is negative. He is scheduled to receive Which ONE of the following should you tell her
a living unrelated kidney transplant from a friend. about SPK transplantation?
Which ONE of the following is the MOST ac- A. She is not a candidate for SPK transplantation
curate regarding his risk for recurrent MN af- B. SPK transplant is associated with substantial
ter transplant? improvements in health-related quality of
A. He has a high (.80%) likelihood of de- life, even in the setting of early pancreas
veloping recurrent MN transplant failure
B. The absence of anti-phospholipase A2 re- C. SPK transplant is associated with substantial
ceptor (anti-PLA2R) antibody indicates improvements in long-term survival com-
a low risk (,20%) of recurrent MN pared with kidney transplant alone
C. Graft survival is significantly lower in those D. SPK transplant waiting time is substantially
who develop recurrent MN than the general shorter than for deceased donor kidney trans-
transplant population plant alone
Nephrology Self-Assessment Program - Vol 16, No 4, November 2017 357

28. In 2016, eligibility criteria for simultaneous liver- tacrolimus combined with the mammalian target of
kidney transplant were developed and approved by rapamycin inhibitor everolimus and prednisone as
the Organ Procurement and Transplantation Network. initial maintenance immunosuppression.
Which ONE of the following liver transplant Compared with tacrolimus/mycophenolate-
candidates is NOT eligible for simultaneous based immunosuppression, which ONE of the
liver-kidney transplantation? following is the MOST likely clinical outcome of
A. A 64-year-old man with stage 3 CKD using tacrolimus/everolimus/prednisone in the
(eGFR545–50 ml/min per 1.73 m2) docu- absence of CMV prophylaxis?
mented for 4 months and a recent fall in A. Lower incidence of CMV viremia
eGFR during a period of hypovolemia to 25 B. Lower incidence of acute rejection
ml/min per 1.73 m2 C. Lower incidence of mortality
B. A 19-year-old woman with primary hyper- D. Lower incidence of nonskin malignancy
oxaluria, nephrocalcinosis, and stage 5 CKD 30. A 62-year-old man with chronic GN is to undergo
C. A 54-year-old man with HCV infection, type living unrelated kidney transplant. His calculated
2 diabetes mellitus, an eGFR of 35–40 ml/min panel reactive antibody score is 0%, and the HLA
per 1.73 m2, and proteinuria of 1 g/d mismatch with the donor is four out of six. In addition
D. A patient with type 2 hepatorenal syndrome to tacrolimus/mycophenolate-based immunosup-
with an eGFR of 15–25 ml/min per 1.73 m2 pression with prednisone, the transplant center
over the past 6 weeks is planning use of rabbit antithymocyte globulin
29. A 62-year-old man with chronic GN is to undergo for induction therapy.
living unrelated kidney transplant. He has a history Compared with induction therapy with an IL-2
of squamous cell carcinoma successfully treated 5 receptor antagonist, which ONE of the following is
years ago. He is CMV naı̈ve (serum IgG for CMV the MOST likely clinical outcome of using rabbit
is negative), whereas his prospective donor is CMV antithymocyte globulin for induction therapy?
IgG positive. Other past medical history includes A. A lower incidence of acute rejection
chronic idiopathic diarrhea two to three times daily. B. A lower incidence of graft loss
An extensive evaluation of the diarrhea was unre- C. A lower incidence of mortality
vealing. The transplant center is considering use of D. A lower incidence of malignancy

Вам также может понравиться