242
Evaluation of transplacental pharmacology and toxicology from bench to bedside by Janine Rose Hutson A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Institute of Medical Science University of Toronto © Copyright by Janine Rose Hutson 2014

Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

Evaluation of transplacental pharmacology and toxicology from bench to bedside

by

Janine Rose Hutson

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy

Institute of Medical Science University of Toronto

© Copyright by Janine Rose Hutson 2014

Page 2: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

ii

Evaluation of transplacental pharmacology and toxicology from bench to bedside

Janine Rose Hutson

Doctor of Philosophy

Institute of Medical Science University of Toronto

2014

Abstract

Many women require pharmacologic treatment during pregnancy. Clinical studies of drug

safety in human pregnancy are often limited because of ethical considerations and thus a

theoretical framework to evaluate drug safety in pregnancy is needed. Dual perfusion of a single

placental lobule is the only experimental model to study human placental transfer of substances

in organized placental tissue. The objectives of this thesis were to systematically evaluate the

perfusion model in predicting placental drug transfer and to develop a model to account for non-

placental pharmacokinetic parameters in the perfusion results. In general, the fetal-to-maternal

drug concentration ratios matched well between placental perfusion experiments and in vivo

samples taken at the time of delivery. After modeling for differences in maternal and

fetal/neonatal protein binding and blood pH, the perfusion results were able to accurately predict

in vivo transfer at steady state (R2 = 0.85, P < 0.0001). We then utilized the perfusion model to

evaluate the placental transfer of 6-mercaptopurine (6-MP), a drug commonly used in pregnancy

for the treatment of inflammatory bowel disease as well as the toxic effects of alcohol and formic

acid on the placenta. Placental transfer of 6-MP is limited and binding to placental tissue and

maternal pharmacokinetic parameters are the main factors that restrict placental transfer.

Evaluation of the placental perfusion results together with our meta-analysis of clinical studies

supported other evidence that the benefit of 6-MP outweighs any fetal risk when indicated. The

placental transfer of folic acid was decreased in pregnancies with heavy alcohol exposure and

Page 3: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

iii

folic acid supplementation may decrease the toxic effects of formic acid on the placenta.

Furthermore, our validation of the ex vivo perfusion model showed that it is effective in

predicting fetal exposure to drugs and should have a place in clinical and regulatory

pharmacology and toxicology.

Page 4: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

iv

Acknowledgments

I would like to thank the following individuals for their support and contributions during

the course of my training.

Dr. Gideon Koren for his exceptional mentorship, the many opportunities he has provided

me, and the encouragement and support to follow my ambitions.

Dr. Bhushan Kapur for his mentorship and support during my training and for the

opportunity to work on a completely new area of research! Drs Suzanne Schuh and Micheline

Piquette-Miller, my advisors, for their time and invaluable guidance that has facilitated my

degree completion. Drs. Yaron Finkelstein and Cindy Woodland for also being great mentors and

for giving me amazing opportunities.

I would also thank to express my gratitude to those who have collaborated with me on my

studies. Angelika Lubetsky for her expertise in the placental perfusion experiment and for never

saying no when I wanted to try numerous experiments in a day! I thank you for sharing your

expertise with me. To Myla Moretti, Jeremy Matlow, Dr. Facundo Bournissen, Amy Davis, Jeff

Eichhorst, Dr. Kathy Aleksa, Dr. Asnat Walfisch, Dr. Rinat Hackmon, Dr. Brenda Stade, Dr.

Denis Lehotay, and Dr. Christine Collier for your time and contributions to these studies. I could

not have completed this work without you. To Sarah MacDougall and Irene Zelner for their

friendship and support. The other members of the Koren Laboratory and the Motherisk Program

for their contribution and companionship through the course of this work. To the lab ladies who

lunch for the good times and great memories!

The generous support from the Canadian Institutes of Health Research Vanier

Scholarship Program, University of Toronto Open Fellowships, Canadian Foundation on Fetal

Alcohol Research, and the McLaughlin Centre for Molecular Medicine.

Finally, my parents, family and close friends for their encouragement and understanding.

You have instilled in me the courage to succeed. Lastly, to all of my friends from both the MD

and PhD programs for your support throughout the ride.

Page 5: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

v

Contributions I am grateful to the following for the contributions they have made to the individual

studies as detailed below as well as their contributions to the preparation of the respective

manuscripts.

Amy Davis: A summer student who assisted me with the systematic searches for the in

vivo umbilical cord blood and maternal blood drug concentration values that are included in

Chapter 2.

Dr. Facundo Garcia-Bournissen who contributed to the study design and data

interpretation of Chapters 2 and 3.

Angelika Lubetsky for her technical assistance with the placental perfusion experiments

included in Chapters 3 and 6. Dr. Asnat Walfisch, Dr. Rinat Hackmon for recruitment of eligible

patients in Chapters 3 and 6, respectively. Dr. Brenda Stade and Dr. Christine Collier for

recruitment of patients and data collection in Chapter 5.

Jeremy Matlow who was the second reviewer to evaluate appropriateness of studies to

include for the meta-analysis in Chapter 4 and for data extraction from these studies. Myla

Moretti for data analysis in Chapter 4.

Jeff Eichhorst and Dr. Denis Lehotay for measurement of formic acid and folate in

Chapters 5 and 6.

Dr. Gideon Koren, my supervisor, for overseeing study design, data analysis,

interpretation of data, manuscript preparation, and obtaining funding. Dr. Bhushan Kapur for

obtaining funding for Chapters 5 & 6 and for overseeing these two studies as the principle

investigator.

Page 6: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

vi

Table of Contents

Acknowledgments .......................................................................................................................... iv!

Contributions ................................................................................................................................... v

Table of Contents ........................................................................................................................... vi!

List of Tables ................................................................................................................................. ix!

List of Figures ................................................................................................................................. x!

List of Publications ....................................................................................................................... xii!

List of Abbreviations ................................................................................................................... xiii!

List of Appendices ....................................................................................................................... xvi!

Chapter 1 General Introduction ...................................................................................................... 1!

! Drug Exposure In Pregnancy ..................................................................................................... 1!1

1.1! Factors Affecting Maternal-Fetal Pharmacokinetics and Transplacental Drug Transfer ... 1!

1.1.1! Maternal pharmacokinetic considerations .............................................................. 2!

1.1.2! Fetal Pharmacokinetic Considerations .................................................................... 5!

1.1.3! Transplacental Drug Transfer ................................................................................. 6!

1.1.4! Methods for Studying Transplacental Drug Transfer ........................................... 12!

1.2! Therapeutic Drug Use during Pregnancy: Inflammatory Bowel Disease and Treatment with Thiopurines ............................................................................................................... 19!

1.3! Recreational Drug Use During Pregnancy: Alcohol and Fetal Alcohol Spectrum Disorder ............................................................................................................................. 23!

1.4! Overall Rational and Primary Objectives ......................................................................... 29!

Chapter 2 ....................................................................................................................................... 32!

! The human placental perfusion model: a systematic review and development of a model to 2predict in vivo transfer of therapeutic drugs ............................................................................ 33!

2.1! Abstract ............................................................................................................................. 33!

2.2! Introduction ....................................................................................................................... 34!

2.3! Methods ............................................................................................................................. 36!

Page 7: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

vii

2.3.1! Literature Search ................................................................................................... 36!

2.3.2! Extraction of Data and Qualitative Comparisons ................................................. 36!

2.3.3! Quantitative Calculations ...................................................................................... 37!

2.3.4! Model Development .............................................................................................. 37!

2.4! Results ............................................................................................................................... 39!

2.4.1! Search Results ....................................................................................................... 39!

2.4.2! Qualitative Comparison of Perfusion Results to In Vivo Data ............................. 39!

2.4.3! Quantitative Comparison of Perfusion Results to In Vivo Data ........................... 42!

2.5! Discussion ......................................................................................................................... 45!

Chapter 3 ....................................................................................................................................... 52!

! The transfer of 6-mercaptopurine in the dually perfused human placenta ............................... 53!3

3.1! Abstract ............................................................................................................................. 53!

3.2! Introduction ....................................................................................................................... 54!

3.3! Methods ............................................................................................................................. 55!

3.4! Results ............................................................................................................................... 58!

3.5! Discussion ......................................................................................................................... 63!

3.6! Conclusions ....................................................................................................................... 65!

Chapter 4 ....................................................................................................................................... 67!

! The fetal safety of thiopurines for the treatment of inflammatory bowel disease in 4pregnancy: a meta-analysis ...................................................................................................... 68!

4.1! Abstract ............................................................................................................................. 68!

4.2! Introduction ....................................................................................................................... 69!

4.3! Methods ............................................................................................................................. 69!

4.4! Results ............................................................................................................................... 71!

4.5! Discussion ......................................................................................................................... 82!

Chapter 5 ....................................................................................................................................... 85!

Page 8: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

viii

! Folic acid transport to the human fetus is decreased in pregnancies with chronic alcohol 5exposure ................................................................................................................................... 86!

5.1! Abstract ............................................................................................................................. 86!

5.2! Introduction ....................................................................................................................... 87!

5.3! Methods ............................................................................................................................. 89!

5.4! Results and Discussion ..................................................................................................... 90!

Chapter 6 ....................................................................................................................................... 98!

! Adverse placental effect of formic acid on hCG secretion is mitigated by folic acid ............. 99!6

6.1! Abstract ............................................................................................................................. 99!

6.2! Introduction ..................................................................................................................... 100!

6.3! Methods ........................................................................................................................... 101!

6.3.1! Materials ............................................................................................................. 101!

6.3.2! Placental perfusion .............................................................................................. 101!

6.3.3! Sample analysis ................................................................................................... 102!

6.4! Results ............................................................................................................................. 103!

6.5! Discussion ....................................................................................................................... 108!

6.6! Conclusions ..................................................................................................................... 110!

Chapter 7 ..................................................................................................................................... 111!

! General Discussion ................................................................................................................. 111!7

7.1! Summary of Research Findings and Future Directions .................................................. 112!

7.2! Summary ......................................................................................................................... 127!

! References .............................................................................................................................. 128!8

Appendices .................................................................................................................................. 161!

Page 9: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

ix

List of Tables

Table 1-1. Advantages and disadvantages of different methods used to determine placental drug

transfer. ......................................................................................................................................... 15!

Table 1-2 The 1996 Institute of Medicine’s Diagnostic Criteria for FASD ................................. 24!

Table 1-3 Secondary Disabilities in Adolescents and Adults with FAS or FAE .......................... 25!

Table 2-1. Drugs that had the maternal-to-fetal transfer evaluated using the placental perfusion

model and where the fetal to maternal drug concentration ratios (F:M) from the perfusion agreed

with in vivo data from mother-infant pairs. .................................................................................. 41!

Table 2-2. Drugs that had the maternal-to-fetal transfer evaluated using the placental perfusion

model and where the fetal to maternal drug concentration ratios (F:M) from the perfusion

disagreed with in vivo data from mother-infant pairs. .................................................................. 41!

Table 2-3. The observed and calculated F:M ratios for data obtained from placental perfusion

experiments ................................................................................................................................... 43!

Table 3-1. Placental viability parameters and metabolic capacity throughout the perfusions ...... 59!

Table 4-1 Characteristics of studies included in the meta-analysis .............................................. 72!

Table 4-2 Pregnancy outcomes in thiopurine-exposed pregnancies versus controls .................... 74!

Table 4-3 Specific malformations and other drug use reported in the included studies ............... 75!

Table 5-1. Demographic information on alcohol-using women included in this study and fetal

parameters. .................................................................................................................................... 91!

Table 6-1 Placental viability parameters and metabolic capacity throughout the perfusion ...... 104!

Page 10: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

x

List of Figures

Figure 1-1. ABC Transport proteins present on the human placental barrier. ................................ 9!

Figure 1-2. The transfer of the basic drug, bupivacaine, across the placental membrane at

equilibrium is determined by the acid-base equilibrium effect and protein binding.. .................. 13!

Figure 1-3. Schematic presentation of the placental perfusion experimental set-up. ................... 16!

Figure 1-4 Metabolic pathway for 6-mercaptopurine and its prodrug azathioprine. .................... 22!

Figure 2-1. Flow diagram of the search strategy and articles retrieved and included in the review.

....................................................................................................................................................... 40!

Figure 2-2. Agreement between perfusion fetal to maternal (F:M) concentration ratios and the in

vivo umbilical cord to maternal blood (C:M) drug concentration ratios ...................................... 44!

Figure 2-3. Comparison of F:M ratios for morphine obtained from placental perfusion

experiments versus maternal blood and umbilical cord blood ...................................................... 46!

Figure 3-1. Maternal-to-fetal transfer of 6-mercaptopurine after dual perfusion of a single

placental lobule ............................................................................................................................. 61!

Figure 3-2 Concentrations of 6-MP in the fetal and maternal circulations after perfusion of a

single placental lobule under equilibrative concentrations ........................................................... 62!

Figure 4-1 Overall effect of the incidence of congenital abnormalities after in utero exposure to

thiopurines………………………………………………………………………………………..76

Figure 4-2 Overal effect of the incidence of prematurity (<37 weeks gestation) after in utero

exposure to thiopurines…………………………………………………………………………..78

Figure 4-3 Overall effect of the incidence of low birth weight (< 2500 g) after in utero exposure

to thiopurines. ............................................................................................................................... 80!

Figure 4-4. Overall effect of in utero exposure to thiopurines on mean birth weight .................. 81!

Page 11: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

xi

Figure 5-1. The transport proteins present on the syncytiotrophoblast that are involved in the

transfer of folate to the fetal circulation. ....................................................................................... 88!

Figure 5-2. Scatter-plot of the fetal to maternal (F:M) folate ratios as measured in cord blood and

maternal blood, respectively, at the time of delivery. ................................................................... 92!

Figure 5-3. Corresponding maternal and fetal folate concentrations at the time of delivery in

pregnancies with A) heavy alcohol exposure and B) in controls. ................................................. 93!

Figure 5-4. Mean (± SEM) cord and maternal plasma folate concentrations at the time of delivery

in alcohol-abusing women and controls ....................................................................................... 94!

Figure 6-1 Maternal-to-fetal transfer of formic acid after dual perfusion of a single placental

lobule in the presence or absence of folate ................................................................................. 105!

Figure 6-2 The rate of hCG secretion into the maternal circulation during the placental perfusion

was calculated in the pre-experimental control period (before addition of formic acid) and during

the experimental period in the presence and absenceof folate. ................................................... 106!

Figure 6-3 Tissue concentration of hCG in the perfused lobule at the end of the 180 minute

experimental period with formic acid expressed as a percentage of the initial hCG tissue

concentration in the same placenta in the presence and absence of folate. ................................ 107!

Figure 7-1 The interaction between alcohol use in pregnancy and the proposed mechanisms of

toxicity by reduced placental folate transfer to the fetus and by formic acid. ............................ 126!

Page 12: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

xii

List of Publications

Etwel F, Hutson JR, Madadi P, Gareri J, Koren G. Fetal and Perinatal Exposure to Drugs and Chemicals; Novel Biomarkers of Risk Assessment. Ann Rev Pharmacol Tox 2014;54:295-315.

Hutson JR, Lubetsky A, Eichhorst J, Hackmon R, Koren G, Kapur BM. Adverse placental effect of formic acid on hCG secretion is mitigated by folic acid. Alcohol Alcohol 2013;48(3):283-7.

Hutson JR, Matlow JN, Moretti M, Koren G. The fetal safety of thiopurines for the treatment of inflammatory bowel disease in pregnancy: a meta-analysis. J Obstet Gynecol 2013;33(1):1-8.

Hutson JR, Stade B, Lehotay DC, Collier CP, Kapur B. Folic acid transport to the human fetus is decreased in pregnancies with chronic alcohol exposure. PLoS ONE 2012:7(5);e38057.

Hutson JR, Garcia-Bournissen F, Davis A, Koren G. The human placental perfusion model: a

systematic review and development of a model to predict in vivo transfer of therapeutic drugs. Clin Pharmacol Ther 2011;90(1):67-76.

Hutson JR, Lubetsky A, Walfisch A, Ballios BG, Garcia-Bournissen F, Koren G. The transfer

of 6-mercaptopurine in the dually perfused human placenta. Reprod Toxicol 2011;32:349–353.

Hutson JR. Prediction of Placental Drug Transfer Using the Human Placental Perfusion Model.

J Popul Ther Clin Pharmacol 2011;18(3):e533-e543.

Hutson JR, Koren G, Matthews SG. Placental P-glycoprotein and breast cancer resistance protein: influence of polymorphisms on fetal drug exposure and physiology. Placenta 2010;31(5):351-357.

Page 13: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

xiii

List of Abbreviations

5-ASA 5-aminosalicylic acid

6-MMP 6-methylmercaptopurine

6-MP 6-mercaptopurine

6-TG 6-thioguanine

AAG α1-acid glycoprotein

ABC ATP-binding cassette

ADH alcohol dehydrogenase

AUC area under the curve

AZA azathioprine

B/F bound to free drug concentration ratio

BCRP breast cancer-resistance protein

C:M cord to maternal concentration ratio

CI confidence interval

CL clearance

Cmax peak serum concentration

CNT concentrative nucleoside transporter

Css steady state concentration

CYP 450 cytochrome P450

D dose

Page 14: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

xiv

ENT equilibrative nucleoside transporters

F:M fetal to maternal concentration ratio

FAS fetal alcohol syndrome

FASD fetal alcohol spectrum disorder

FR-α folate receptor-α

GFR glomerular filtration rate

GMPS guanine monophosphate synthetase

hCG human chorionic gonadotropin

HPRT hypoxanthine–guanine phosphoribosyltransferase

IBD inflammatory bowel disease

IMPDH inosine monophosphate dehydrogenase

ip intraperitoneal

LBW low birth weight

MCT monocarboxylate transporters

MDMA 3,4-methylenedioxy-N-methylamphetamine

meTGMP methylated TGMP

meTIMP methylated TIMP

MRP multidrug resistance-associated proteins

NOAEL no observable adverse effect level

NSAID nonsteroidal anti-inflammatory drug

Page 15: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

xv

OR odds ratio

PCFT proton-coupled folate transporter

PFR placental folate receptors

Pgp P-glycoprotein

RFC reduced folate carrier

RR relative risk

SA spontaneous abortion

SD standard deviation

SEM standard error of the mean

TGMP 6’thioguanosine 5’monophosphate

TGN thioguanine nucleotides

THC tetrahydrocannabinol

TIMP 6-thioinosine 5 ́ monophosphate

TPMT thiopurine S-methyltransferase

TXMP 6-thioxanthosine 5’monophosphate

UGT uridine diphosphate glucuronosyltransferase

Vd volume of distribution

WMD weighted mean difference

XO xanthine oxidase

Page 16: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

xvi

List of Appendices

Appendix A: Placental Perfusion System Materials and Set-up……………………………162

Materials…………………………………………………………………………….162

Fetal Circuit Hardware Set-up………………………………………………………163

Maternal Circuit Hardware Setup…………………………………………………...165

Perfusion Chamber….………………………………………………………………167

Composition of Medium 199 Culture Medium in Fetal and Maternal Perfusate….. 168

Appendix B: Systematic Review of Perfusion Data - Supplementary Table (Chapter 2)…..169

Appendix C: Placental P-glycoprotein and breast cancer resistance protein: influence of

polymorphisms on fetal drug exposure and physiology…………………………………….202

Page 17: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

1

Chapter 1 General Introduction

Drug Exposure In Pregnancy 1Historically, the placenta was believed to restrict transfer of xenobiotics from the

maternal to the fetal circulation, thereby protecting the fetus from any adverse effects of drugs

(Koren et al., 1998). The thalidomide disaster in the 1950s and early 1960s was the first large-

scale evidence that the fetus is not exempt from adverse drug effects as exposure during critical

periods during development resulted in severe limb defects and other organ dysgenesis (e.g.,

kidney and heart defects) (McBride, 1978). The thalidomide disaster prompted the belief that

every drug has a potential to be harmful to the fetus (Koren et al., 1998).

We now know that most small molecules cross the placenta in a measurable way,

sometimes achieving concentrations in fetal plasma as high as those in maternal plasma. Despite

most drugs reaching the fetal circulation, only a limited number of drugs have been proven to be

teratogenic (Koren et al., 1998). This is reassuring, as many women require pharmacologic

treatment during pregnancy for chronic medical conditions or illnesses that arise during

pregnancy. The benefit of drug treatment must be weighed against any potential risk to the

unborn child as well as risk of the untreated maternal disease. This potential maternal-fetal

conflict brings to light the need to identify drugs that can treat the mother without harming the

fetus. As the placenta is the interface between the maternal and fetal circulations, a thorough

understanding of the role of the placenta in maternal-fetal pharmacokinetics is vital to

understanding drug safety in pregnancy. Factors influencing placental drug transfer as well as

methods to study transfer will be reviewed. In addition, examples of therapeutic agents and

toxins that can be studied with respect to placental transfer will be reviewed.

1.1 Factors Affecting Maternal-Fetal Pharmacokinetics and Transplacental Drug Transfer

Dynamic changes in maternal, placental, and fetal pharmacokinetics occur throughout

pregnancy as a result of altered physiology during pregnancy and also by the growth and

Page 18: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

2

development of the placenta and fetus. These changes can influence the pharmacokinetic

processes of drug absorption, distribution, metabolism and elimination. Three factors influence

maternal and fetal drug exposure and thus also influence drug response in pregnancy: 1)

maternal pharmacokinetics influenced by pregnancy-induced physiological changes 2)

distribution, metabolism, and elimination by the fetus and 3) the amount of drug that can cross

the placenta (Hutson et al, 2012).

1.1.1 Maternal pharmacokinetic considerations

During pregnancy, a variety of anatomical, physiological, hormonal and biochemical

changes occur, which affect virtually every aspect of the pharmacokinetics of drugs, including

absorption, distribution, metabolism, and elimination (Loebstein et al., 1997).

Absorption - Pregnancy can theoretically affect oral absorption of drugs, mainly in late

pregnancy when gastrointestinal transit time is increased by 30 to 50% (Chiloiro et al., 2001;

Loebstein et al., 1997). Opioids administered during labor can also result in delayed gastric

emptying, which can slow absorption of the drug and also the therapeutic effect (Porter et al.,

1997). During pregnancy, gastric acid secretions are reduced and there is an increase in mucous

secretions, resulting in an increase in gastric pH (Loebstein et al., 1997). A change in gastric pH

affects absorption of weak acids and bases by changing the proportion of ionized to unionized

drug: only the unionized form can be absorbed across biological membranes. For most drugs, the

change in total bioavailability during pregnancy is relatively small, however, the rate of

absorption may be altered and be clinically relevant when a rapid effect is desired. Nausea and

vomiting in pregnancy during the first trimester may also influence drug absorption. Absorption

of drugs across the lungs can be influenced by the increased cardiac output and tidal volumes

that occur in pregnancy (Hutson et al., 2012). Furthermore, increased pulmonary blood flow and

alveolar ventilation occur. At term, alveolar ventilation is 70% above normal in the supine

position. These physiological changes may favour alveolar uptake when administering drug

aerosols.

Distribution – Both total body water and fat content increase during pregnancy and

consequently often affect the volume of distribution (Vd) of drugs (Loebstein et al., 1997). There

Page 19: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

3

is a mean increase of 8L in total body water, most (60%) of which is distributed to the fetus,

placenta, and amniotic fluid (Loebstein et al., 1997). Plasma volume during pregnancy expands

by 50% (Pirani et al., 1973) and can lead to a decrease in the peak serum concentration (Cmax)

of many drugs. For some drugs, a larger Vd could necessitate a higher initial dose to obtain

therapeutic plasma concentration. Adipose tissue content also increases by a mean of 19kg at the

end of pregnancy and can increase the Vd for lipophilic substances. For lipophilic drugs, such as

anesthetics, this can lead to persistence of a high drug concentration as the drug distributes back

out of the adipose tissue.

Albumin and α1-acid glycoprotein (AAG) are the two most important drug binding

plasma proteins. Albumin binds mainly weak acids and lipophilic drugs, whereas AAG binds

mainly basic drugs. Pregnant women may exhibit decreased protein binding because of

decreased levels of plasma albumin (Krauer et al., 1984). Throughout pregnancy, albumin

concentrations decrease, reaching approximately 70-80% of normal values at the time of

delivery. This decrease is a result of the plasma volume expanding at a greater rate than the

increase in albumin production. Furthermore, a 3-fold higher concentration of free fatty acids

towards term in the maternal circulation can displace drugs from binding to albumin (Nau et al.,

1984; Ridd et al., 1983). In contrast to albumin, plasma concentrations of AAG remain similar

throughout pregnancy. Since AAG is an acute phase protein that increases during trauma or

inflammation, it is expected to increase during labor; however, there was no increase in plasma

AAG concentrations during delivery when effective analgesia was provided using epidural

ropivacaine (Porter et al., 2001). Increases in placental and steroid hormones throughout

pregnancy can occupy and displace drugs from the protein binding sites. The clinical relevance

of this phenomenon is not clear. For example, protein binding of bupivacaine decreased during

maximum estrogen concentrations in patients undergoing in vitro fertilization (Tsen et al., 1997).

However, this decrease in protein binding is mainly attributed to estrogen decreasing the

concentration of plasma binding proteins rather than displacement of the drug from the binding

site. For highly protein-bound drugs, decreased protein binding has a variable effect on kinetics:

more drug is free and available for effect, yet more free drug is also available for metabolism

and/or elimination. Changes in protein binding during pregnancy can influence the transplacental

transfer of many drugs and will be further discussed in section 1.1.3.4.

Page 20: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

4

Metabolism and Elimination - Although maternal physiologic changes begin early in

pregnancy, they are most pronounced in the third trimester. A rise in cardiac output begins in the

first trimester of pregnancy and can be as high as 50% above normal non-pregnant values. An

increase in hepatic clearance may result from the increase in hepatic blood flow as a result of

increased cardiac output. This may be especially critical for drugs with high extraction ratios,

such as lidocaine and morphine, and less significant for low hepatic extraction drugs such as

acetaminophen. Changes in hepatic enzyme activity can have an important impact on drug

clearance, mainly in the phase I cytochrome P450 (CYP450) pathway. As steady state

concentrations (Css) correlate with clearance rate (CL) (according to the equation: Css = D/CL),

a dose of drug (D) would have to be increased proportionally to increase clearance rate to

maintain a similar pharmacologic effect. Specifically, pregnancy increases the hepatic activity of

CYP3A4, CYP2D6, CYP2C9 and CYP2A6. Drugs predominately metabolized by these

enzymes, such as nonsteroidal anti-inflammatory drugs (NSAIDs), may require an increased

dose to avoid loss of efficacy (Anderson, 2005). In contrast, CYP1A2 and CYP2C19 activity is

decreased during pregnancy, suggesting that dosage reductions may be needed to minimize

potential toxicity of their substrates. Phase II enzymes can also be increased (ie. uridine

diphosphate glucuronosyltransferase (UGT) isoenzymes) or decreased (ie. N-acetyltransferase 2)

during pregnancy.

Blood flow to the kidneys increases 60-80% during pregnancy. As a result, glomerular

filtration rate (GFR) increases 50% by the first trimester and continues to increase throughout

pregnancy compared with postpartum values (Davison & Hytten, 1974). Only during the last

three weeks of pregnancy does GFR begin to decrease (Anderson, 2005). The increase in renal

clearance during pregnancy is likely to have notable effects on drugs that are eliminated

predominately unchanged by the kidneys (e.g. β-lactam antibiotics, lithium, digoxin) and their

dosage may need to be increased by 20-65% in order to maintain pre-pregnancy concentrations

(Anderson, 2005). The influence of pregnancy on drug transporters involved in tubular secretion

and reabsorption has not been extensively studied.

Page 21: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

5

1.1.2 Fetal Pharmacokinetic Considerations

Distribution - Drugs that cross the placenta are carried to the fetus by the umbilical vein

and enter the unique fetal circulation. Drugs that readily transfer across the placenta often are

rapidly distributed to highly perfused fetal organs, such as anesthetics (Finster et al., 1972a).

Before entering the systemic circulation, the majority of blood from the umbilical vein flows

through the fetal liver. The fetal liver therefore has the potential to reduce the amount of drug

distributed to sensitive fetal organs. Lidocaine and thiopental were observed to bind and

accumulate in fetal liver (Finster et al., 1972a; Finster et al., 1972b). Drug distribution in the

fetus will also depend on plasma protein binding. Plasma concentrations of both albumin and

AAG in the fetus progressively increase throughout gestation. Between the 12th week of

gestation and term, albumin increases from 10.9 g/L to 34.2 g/L and AAG increases from 0.05

g/L to 0.21 g/L (Krauer et al., 1984). This dynamic process contributes to the change in the fetal

to maternal albumin/AAG concentration ratio at different gestational ages. The influence of this

change on drug transfer across the placenta is discussed in section 1.1.3.4. In fetuses with severe

cardiac insufficiency, elevated umbilical venous pressure may reduce placental drug transfer

(Schmolling et al., 2000).

Metabolism and Elimination - The fetal liver has generally less enzymatic activity for

drug metabolism compared to the adult. Throughout gestation, the amount and intra-lobular

distribution of drug metabolizing enzymes in the fetal liver undergo changes. Several drug-

metabolizing enzymes have been detected as early as 8 to 10 weeks of gestation (Hines, 2007).

However, CYP450 and some conjugating enzymes are located to the smooth endoplasmic

reticulum in the hepatocyte and a significant amount of smooth endoplasmic reticulum is not

present until mid-gestation (Myllynen et al., 2009). Activity of CYP1A1 has been detected in

first trimester fetal liver, but not in the second and third trimesters (Myllynen et al., 2009). No

significant activity of CYP1A2 has been detected (Sonnier & Cresteil, 1998). Activity of

CYP2C9 increases with gestational age, but CYP2C19 activity remains constant (Koukouritaki

et al., 2004). The expression of CYP2D6 is variable and activity has ranged from undetectable

to approximately 1% of adult activity (Myllynen et al., 2009). The major CYP3A isoform in the

fetal liver is CYP3A7 and it accounts for 50% of the total fetal hepatic CYP content (Ring et al.,

1999). After birth, CYP3A7 activity decreases and CYP3A4 becomes the predominant hepatic

CYP3A isoform. Hepatic activity of UGTs starts during the latter half of the second trimester

Page 22: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

6

(Myllynen et al., 2009). Drug-metabolizing enzymes present in the fetal liver may contribute to

a first-pass effect. A study in the baboon demonstrated that the fetal liver had a high intrinsic

clearance toward morphine (a UGT substrate) and that fetal hepatic clearance of morphine was

flow limited (Garland et al., 2006). This study demonstrates that for specific drugs, there may be

significant placenta-fetal hepatic first-pass metabolism when drugs are administered to the

mother. Drugs entering the fetal circulation may also be excreted into the amniotic fluid through

the fetal kidneys. However, most drug elimination from the fetus is generally considered to be a

result of transfer back to the maternal circulation.

1.1.3 Transplacental Drug Transfer

Drug transfer across the placenta depends on the physicochemical properties of the drug

as well as the mechanism of transfer across the placental barrier, blood flow to the placenta,

metabolism and binding by the placenta, protein binding in both the mother and fetus, and the pH

of the maternal and fetal circulations. Furthermore, drug transfer will relate to the developmental

stage of the placenta as the thickness of the interface between the maternal and fetal circulations

changes as the placenta develops (Enders, 1981). The expression and function of drug transport

proteins and metabolic enzymes in the placenta also can change throughout pregnancy (Hutson

et al., 2010).

1.1.3.1 Placental Structure

The functional vascular unit of the placenta is the cotyledon. Each cotyledon contains

highly branched villi suspended in the intervillous space (Enders, 1981). This space is filled

with maternal blood supplied by spiral arteries and carried away by uterine veins. Each fetal

villous tree consists of capillary endothelium, villous stroma and a trophoblast layer. Throughout

gestation, the syncytium of cells separating the maternal and fetal circulations thins and the

surface area for transfer increases (Enders, 1981). The rate-limiting barrier for transfer across the

mature placenta is the single layer of polarized multinucleated cells called syncytiotrophoblasts

(Audus, 1999; Ceckova-Novotna et al., 2006). Consequently, drug transfer at term may represent

Page 23: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

7

the highest fetal drug exposure compared to earlier gestational ages (Vahakangas & Myllynen,

2006).

1.1.3.2 Transfer Mechanisms

As in other lipid barriers, molecules that are relatively lipophilic, with molecular weights

< 600 Da, can readily cross the syncytiotrophoblast by diffusion in the unbound non-ionized

state. As a result, most drugs, being small molecules, cross the placenta by passive diffusion

(Ala-Kokko et al., 2000; Audus, 1999). Accordingly, the concentration gradient across the

placental barrier drives transfer for most drugs. A well-mixed double pool flow model best

describes transplacental kinetics where venous equilibrium exists between the maternal and fetal

compartments (Ala-Kokko et al., 2000).

Transfer of hydrophilic drugs is permeability-limited and transfer of lipophilic drugs is

usually considered flow-limited (Giroux et al., 1997; Syme et al., 2004). Blood flow to the

placenta increases during gestation from about 50 mL/min at 10 weeks to 600 mL/min at term

and factors that alter utero-placental blood flow may alter placental drug transfer (Syme et al.,

2004). These factors include anesthesia, nicotine, maternal position, and degenerative changes

within the placenta that are observed with conditions such as hypertension, diabetes, or renal

disease. During labour, uterine contractions can lead to intermittent decreases in uterine arterial

flow. These contractions can delay the transfer of drugs to the fetus and also delay the clearance

of drug from the fetal to the maternal circulation. For example, concentrations of diazepam were

higher in newborns born to mothers administered the drug at the onset of uterine contraction,

compared to those administered the drug during uterine diastole (Haram et al., 1978). The

steady-state fetal to maternal drug concentration ratio will not be affected by changes in blood

flow, however, it may accelerate or delay the rate of transfer.

A wide variety of drug transporters have been localized to the placenta. Although most

are found to have mainly physiological substrates, they can also transport several xenobiotics.

Active transport in the placenta is mediated by transport proteins expressed on both the

microvillous brush-border and basal membranes (maternal blood-facing and fetal endothelial

cell-facing, respectively) of the syncytiotrophoblast cells. Several members of the ATP-binding

Page 24: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

8

cassette (ABC) drug efflux proteins have been localized to the placenta, including P-glycoprotein

(Pgp), breast cancer-resistance protein (BCRP), and multidrug resistance-associated proteins

(MRPs) (Figure 1-1) (Atkinson et al., 2003; Evseenko et al., 2007; MacFarland et al., 1994;

Maliepaard et al., 2001; Meyer Zu Schwabedissen et al., 2005; Meyer zu Schwabedissen et al.,

2006; Nagashige et al., 2003; St-Pierre et al., 2000; Sugawara et al., 1997; Sun et al., 2006;

Yeboah et al., 2006). Depending on the location of the protein, it can facilitate drug-uptake or

drug-efflux from the fetal or maternal circulations. The placement of Pgp and BCRP on the

large surface area of the syncytiotrophoblast brush-border is optimal to limit transfer across the

placental barrier to the fetal side. Pharmacologic blockade of efflux transporters can lead to

disruption of the placental barrier and increase the transfer of respective substrates to the fetal

side by several-fold, which may increase the fetal drug response or also be a mechanism for

teratogenicity. Conversely, transport proteins may also facilitate drug transfer across the

placenta. The monocarboxylate transporters (MCTs) are proton-dependent transporters, and

have been suggested to mediate the placental transfer of diclofenac and salicylic acid (Shintaku

et al., 2007; Shintaku et al., 2009). The influence of polymorphisms in placental drug transport

proteins with respect to fetal drugs exposure and physiology is described in Appendix C.

Page 25: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

9

Figure 1-1. ABC Transport proteins present on the human placental barrier.

1.1.3.3 Metabolism and binding by the placenta

Both phase I and II drug metabolizing enzymes have been detected and characterized in

the placenta. For phase I enzymes, CYP1A1 is the major CYP isoform present in the placenta

and is induced by lifestyle factors such as smoking and medications such as glucocorticoids

(Myllynen et al., 2009). CYP1A1 is expressed throughout gestation, whereas CYP1A2 mRNA

has been detected in first trimester placenta but not in term placenta (Syme et al., 2004).

Functional activity of CYP2C, 2D6, and 3A has not been reported, whereas variable activity has

been reported for CYP2E1 (Myllynen et al., 2009; Syme et al., 2004). For phase II enzymes,

UGTs are suggested to have a significant role in placental metabolic activity and are present

throughout gestation (Syme et al., 2004). For example, morphine was observed to have a

placental clearance by UGTs slightly lower than that for total fetal clearance (Garland et al.,

Page 26: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

10

2006). However, for most drugs, placental metabolism is a relatively minor and likely non-

significant factor in limiting placental drug transfer (Syme et al., 2004).

Drug transfer across the placenta can also be influenced by drug binding to placental

tissue. For example, bupivacaine disappeared from the maternal circulation more rapidly than

less lipophilic lidocaine during dual perfusion of a placental lobule ex vivo; however, because

bupivacaine was more extensively bound to the placental tissue, less drug appeared in the fetal

compartment compared to lidocaine (Ala-Kokko et al., 1995). Because of the substantial

placental tissue binding, the maternal clearance of bupivacaine exceeds the transplacental

clearance into the fetal circulation (Johnson et al., 1995). High binding to placental tissue can

lead to the placenta acting as a drug depot. In this case, after discontinuation of drug

administration to the maternal circulation, the drug can continue to wash out into the fetal

circulation and potentially prolong drug exposure. Less binding to the placenta can lead to a

more rapid drug equilibration with the fetus and thus a more rapid fetal drug response.

1.1.3.4 Maternal and fetal protein binding

The two most important drug binding plasma proteins are albumin and AAG. These two

proteins differ in concentration in the maternal and fetal circulations. Maternal plasma albumin

gradually decreases towards term while fetal albumin progressively increases. Compared to the

maternal circulation, the fetal circulation has an appreciably higher albumin concentration at

term, with the fetal-to-maternal (F:M) albumin concentration ratio increasing from 0.28 in the

first trimester to 1.20 at term (Krauer et al., 1984). Furthermore, a 3-fold higher concentration of

free fatty acids in the maternal circulation towards term can displace drugs from binding to

albumin in the maternal circulation (Nau et al., 1984; Ridd et al., 1983). Maternal protein binding

can also be reduced by therapeutic maneuvers such as aggressive fluid hydration prior to regional

anesthesia because it leads to a physiological protein dilution (Herman et al., 2000). Taken

together, there is increased binding to albumin in the fetal circulation compared to the maternal

circulation at term; however, the difference in binding between the two circulations is dynamic

throughout gestation. Because it is the free concentration of the drug that will equilibrate across

the placenta, increased fetal protein binding can lead to increased placental drug transfer. Alpha-

fetoprotein is considered to be the fetal analogue of albumin and differs slightly in structure.

Page 27: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

11

However, the differences between maternal and fetal plasma protein concentrations influences

placental transfer to a greater degree than varying protein structures or competing endogenous

ligands (Hill & Abramson, 1988)

The other major plasma binding protein, AAG, stays fairly constant throughout gestation

in the maternal circulation while the fetal concentration gradually increases (Krauer et al., 1984).

The F:M AAG concentration ratio increases from 0.09 in the first trimester to 0.37 at term

(Krauer et al., 1984). Compared to maternal, there is lower binding to AAG in the fetal

circulation throughout pregnancy. Because of this, drugs bound primarily to AAG such as

bupivacaine and ropivacaine have a lower total concentration in the fetal circulation compared to

the maternal (Ala-Kokko et al., 1997; Datta et al., 1995). The F:M AAG concentration gradient

was found to be strongly associated with the distribution of ropivacaine and bupivacaine across

the placenta (Ala-Kokko et al., 1997) and is considered to be the main determinant in the steady-

state distribution across the maternal-placental-fetal unit for these agents (Figure 1-2).

1.1.3.5 Acid-Base Equilibrium Effect

Unionized lipid-soluble molecules are able to penetrate biological membranes more

quickly than ones that are ionized and less lipid-soluble. Maternal and fetal plasma pH are

important determinants in placental transfer for weak acid or basic drugs whose pKa is close to

physiologic pH. The fetal circulation is slightly more acidic than the maternal (7.35 vs 7.4

respectively) (Reynolds & Knott, 1989). Because of this difference, weak bases will become

more ionized in the fetal circulation and result in ion trapping (Figure 1-2). For the amide-type

anesthetics bupivacaine (pKa=8.1), lidocaine (pKa=7.9), ropivacaine (pKa=8.1), 2-

chloroprocaine (pKa=8.9), and mepivacaine (pKa=7.6), placental transfer to the fetal circulation

has been observed to increase with decreasing fetal pH (Johnson et al., 1996; Ueki et al., 2009).

For these agents, it is the basic uncharged drug concentration that determines placental transfer

(Ueki et al., 2009).

The difference between maternal and fetal pH is minimal (<0.1 pH units) under normal

circumstances. However, fetal pH may fall considerably in cases of fetal compromise.

Compared with general or epidural anesthesia, there is also a greater degree of fetal acidosis

Page 28: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

12

observed during spinal anesthesia for Caesarean section (Reynolds & Seed, 2005). As fetal pH

decreases, the effect of ion trapping on basic drugs in the fetal circulation is accentuated and less

drug is able to cross back from the fetal to the maternal circulation. For amide-type anesthetics,

this leads to increased fetal concentrations and is expected to lead to increased fetal effect (Biehl

et al., 1978; Pickering et al., 1981). However, this phenomenon may be less pronounced for

agents that are rapidly metabolized and do not have the chance to accumulate. For example, at

the time of delivery, there was no association found between umbilical cord concentrations of 2-

chloroprocaine and a fetal pH ≤7.25 (Philipson et al., 1985).

In addition to influencing the level of drug ionization, plasma pH can also influence other

pharmacokinetic factors, such as protein binding. With decreasing pH, protein binding for

lidocaine was shown to decrease (Burney et al., 1978). Thus, the effect of decreased fetal pH on

placental drug transfer will be the net result of changes in both drug ionization and protein

binding. During fetal distress and fetal acidosis, there is increased perfusion to the heart and

brain, which may also alter fetal drug response. This may lead to increased drug delivery to

these important organs.

1.1.4 Methods for Studying Transplacental Drug Transfer

There is often limited information available on drug safety for new pharmacologics in

pregnancy and specifically on maternal-to-fetal pharmacokinetics. The full time-course of

pharmacokinetic changes during pregnancy is rarely studied. Obtaining data on placental and

fetal pharmacokinetics is also difficult since the feto-placental unit in situ is not easily accessible

until delivery. In vivo human studies evaluating the transplacental kinetics of drugs have

numerous ethical concerns regarding fetal and maternal safety. Because the placenta is the most

species-specific mammalian organ, animal studies cannot always be extrapolated to humans

(Ala-Kokko et al., 2000). Limited pharmacokinetic data can be obtained via cord blood

sampling in neonates whose mothers receive the agent to be studied. By measuring the difference

in drug concentrations between the umbilical vein and artery at term, fetal uptake and/or

metabolism of a drug can be estimated. For lidocaine, concentrations in the fetal umbilical vein

were 1.5 times higher than the umbilical artery, suggesting fetal uptake or metabolism of the

Page 29: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

13

drug (De Barros Duarte et al., 2010). From these studies, it is difficult to characterize how the

drug distributes in the fetus and the amount/type of drug metabolism.

Figure 1-2. The transfer of the basic drug, bupivacaine, across the placental membrane at

equilibrium is determined by the acid-base equilibrium effect and protein binding. A)

Unbound bupivacaine (pKa=8.1) is more ionized in the more acidic fetal circulation as a result of

ion trapping. A fetal to maternal ratio of 1.08 for unbound bupivacaine results from ion trapping.

B) Bupivacaine is bound to α1-acid glycoprotein (AAG) and higher binding in the maternal

circulation results from the higher AAG concentration in plasma. As a result, the fetal to

maternal concentration ratio for total bupivacaine is 0.32.

Page 30: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

14

Dual perfusion of a single placental lobule ex vivo is the only experimental model to

study human placental transfer of substances in organized placental tissue and theoretically may

be able to better predict fetal exposure compared to other experimental methods (Hutson et al.,

2011). In vitro models utilizing cell culture and membrane vesicles are mostly limited to the

investigation of specific mechanisms of transfer, such as active transport or passive diffusion, but

they lack anatomic integrity and blood flow. In vitro cellular models have focused on transport in

trophoblast cells and have largely neglected the potential role of the fetal endothelial cells to

influence transfer (Elad et al., 2014). The advantages and disadvantages of different methods

used to study placental drug transfer are provided in Table1-1.

1.1.4.1 Description of the Dual Perfusion of a Single Placental Lobule ex vivo

The first perfusion of an isolated human placental lobule was described by Panigel et al.

in 1967 (Panigel et al., 1967) and was subsequently modified by Schneider and Miller (Miller et

al., 1985; Schneider et al., 1972). Use of a placenta after delivery obviates ethical problems

because tissue collection is non-invasive and harmless to both the mother and newborn. Figure 1-

3 provides a schematic description of the placental perfusion system. Between laboratories,

experimental systems can vary with respect to cannulation method, isolation procedure,

perfusion fluid, oxygenation, perfusion flow, and control measurements (Mathiesen et al., 2010).

The perfusion model described here is routinely used in our laboratory in Toronto.

Page 31: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

15

Table 1-1. Advantages and disadvantages of different methods used to determine placental

drug transfer.

Method Advantages Disadvantages

Placental Perfusion • Most closely resembles the

in vivo situation as structural integrity and cell-cell organization are maintained

• Can measure transfer over time

• Sampling is available from all compartments including placental tissue

• Often performed in term tissue and cannot be used to determine transfer in the first trimester

• No standardized criteria between laboratories

Measuring Umbilical

Cord Blood and

Maternal Blood

Drug Concentrations

at Delivery

• True in vivo measure in humans

• Sampling at delivery avoids ethical issues

• Timing of sampling usually restricted to delivery

• Represents only one point in time and large inter-individual variation

• New drugs or toxic substances can not be studied ethically

Animal Models • Transfer and toxicity at

different gestational ages can be studied

• Drug accumulation in fetal tissues can be studied

• Low inter-individual variation as a result of inbred animals

• Interspecies variability in placental structure and blood flood flow pattern

• Difficult to extrapolate findings to humans

Trophoblast Tissue

Preparations (Villous

Preparations, Membrane

Vesicles)

• Useful for studying transfer mechanisms (such as uptake, efflux and metabolism), even at earlier gestational ages

• Can study fetal and maternal plasma membrane fractions separately

• Not useful for studying trans-cellular transport.

• Transporter regulatory mechanisms may not be present in the preparation

• Does not incorporate the role of fetal endothelial cells

Trophoblast Cultures (primary cultures and

trophoblast cell lines)

• Useful for studying transfer mechanisms (such as uptake, efflux and metabolism)

• BeWo cell line available that forms cellular monolayers

• Transporter expression can vary between cultures

• Does not incorporate the role of fetal endothelial cells

Page 32: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

16

Figure 1-3. Schematic presentation of the placental perfusion experimental set-up.

Page 33: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

17

After tissue collection, a fetal vein-artery pair supplying a well-defined cotyledon is

identified. The corresponding maternal surface is also confirmed to have an intact basal plate and

no evident trauma. The fetal vein-artery pair is cannulated and flow of perfusate through the fetal

circulation is established. The lobule is clamped into a plexiglass chamber with the fetal side

downwards and the excess placental tissue is removed. Buffered saline in the chamber supports

the weight of the lobule and the chamber is placed in a water bath to maintain physiologic

temperature. Depending on the design of the perfusion chamber, saline solutions may not be

used and physiologic temperature may be maintained using a heated flowbench (Ala-Kokko et

al., 2000; Mathiesen et al., 2010).

Perfusion of the maternal side begins with insertion of blunt-tipped needles into the

intervillous space 2-3 mm below the decidual surface. Venous samples from the maternal

circulation are collected from multiple venous openings in the decidual plate. The fetal and

maternal circulations are independently controlled by roller pumps and are equilibrated with 95%

O2/5% CO2 in the maternal side and 95% N2/5% CO2 in the fetal side. Some groups utilize

atmospheric air instead of 95% O2/5% CO2 to avoid toxicity associated with hyperoxia

(Mathiesen et al., 2010). Flow rates are maintained in the fetal and maternal circulations at 3 and

12 mL/min, respectively. Flow rates used vary between laboratories (Mathiesen et al., 2010) and

can influence the time to achieve transplacental transfer and steady state between the fetal and

maternal circulations (Bassily et al., 1995; He et al., 2001).

The first portion of the experiment is usually a control period where the blood is washed

out and baseline parameters are established. After the control period, the compound of interest is

added either into the maternal or fetal circulation to investigate placental transfer in one

direction: maternal-to-fetal or fetal-to-maternal, respectively. The experiment is performed in

either a closed (recirculating) or open (single-pass or non-recirculating) configuration. In the

closed configuration, the perfusates are recycled to imitate physiological conditions. Drug

transfer, as well as the maternal-placental-fetal distribution, can be evaluated. Open

configuration allows for the calculation of drug clearance at steady-state concentrations. The

drug can also be added equally to both circulations in the closed configuration to assess any

accumulation in one direction (Gedeon et al., 2008; Kraemer et al., 2006; Pollex et al., 2008).

Page 34: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

18

Placental viability and integrity are monitored by several parameters both during and

after the perfusion. However, there is no standard set of parameters that are required and these

vary between laboratories (Mathiesen et al., 2010). Common parameters measured during the

perfusion include pressure in the fetal circulation, pH, oxygen consumption, net fetal oxygen

transfer, glucose consumption, lactate production, and synthesis and secretion of proteins

including human chorionic gonadotropin (hCG), human placental lactogen, and leptin (Hutson et

al., 2011). The optimal measure of tissue viability and integrity is suggested to be measuring

fetal volume loss (Pienimaki et al., 1997). There is no generally accepted maximum fetal volume

loss, however, a loss of 2 to 4 mL/hour from the fetal reservoir is typically allowed (Mathiesen et

al., 2010). The addition of a flow-limited marker that undergoes only passive diffusion, such as

antipyrine, into the maternal circulation is also be used to measure tissue integrity. Antipyrine or

other reference markers are also used to standardize results to account for variability in flow or

lobule size among experiments.

Results obtained from placental perfusion experiments are often compared to in vivo

results (if available) that have been collected by measuring drug concentrations in maternal

blood and fetal/umbilical cord blood at the time of delivery. The calculation of cord to maternal

blood concentration ratio (C:M) can provide information on how much drug crosses the placenta

in vivo. Drugs are either measured after the mother has been taking the medication chronically

during pregnancy and is likely in steady state or after single bolus administration before delivery.

Venous blood is usually collected from the mother, while umbilical cord blood can be arterial,

venous, or mixed cord blood. Ideally, drug levels should be measured from the umbilical vein,

when evaluating placental drug transfer, since this receives blood directly from the placenta.

Drug concentrations are likely to be higher in the umbilical vein before reaching steady state

compared to mixed or arterial cord blood. Measuring the concentration difference between the

umbilical vein and artery can also provide information on whether there is drug distribution or

metabolism by the fetus. However, it is not possible from cord blood measurements to

determine if there is accumulation of the drug in fetal tissues.

It is difficult to accurately determine the transplacental kinetics of a particular drug by the

C:M concentration ratios because it is measured at only one time point and typically in a small

number of mother-infant pairs. Often there is large variability between pairs due to sample

contamination, timing of samples, sample site, inter-individual differences, and timing of

Page 35: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

19

exposure. Maternal venous blood concentrations do not reflect those being presented to the

placenta and can depend on the tissues drained by the selected vein if an arteriovenous gradient

exists (Hermann et al., 1999). However, sampling at the time of delivery is easily accessible

using umbilical cord blood and is an ethically acceptable method. It is often the only source of in

vivo human data.

1.2 Therapeutic Drug Use during Pregnancy: Inflammatory Bowel Disease and Treatment with Thiopurines

Between 23 and 85% of women are estimated to require medication at some point during

their pregnancy (Gendron et al., 2009). Many maternal medical conditions such as diabetes,

epilepsy, and hypertension require drug therapy in order to ensure optimal health of both the

mother and fetus. A common medical condition encountered during pregnancy is inflammatory

bowel disease (IBD). The major types of IBD are Crohn's disease and ulcerative colitis and peak

incidence of IBD occurs in women of childbearing age (Cooper & Stroehla, 2003). During

pregnancy, maintaining disease remission is critical for positive fetal outcome (Coelho et al.,

2011; Kwan & Mahadevan, 2010) since disease activity is associated with adverse pregnancy

outcomes (Baiocco & Korelitz, 1984; Dominitz et al., 2002; Fedorkow et al., 1989; Woolfson et

al., 1990).

The thiopurines azathioprine (AZA) and 6-mercaptopurine (6-MP) are effective

treatments for IBD and are increasingly being used as treatment for this disease in pregnancy

(Goldstein et al., 2007). Thiopurines for the treatment of IBD can benefit patients by maintaining

remission during pregnancy, avoiding chronic use of steroids for disease control, and leading to

endoscopic remission (Ardizzone et al., 2006; Mowat et al., 2011; Timmer et al., 2007).

Avoiding the chronic use of oral corticosteroids in women of childbearing age is beneficial as it

has been associated with an increased risk of adverse perinatal outcomes including gestational

diabetes, preeclampsia, preterm delivery, oral cleft palate, premature rupture of membranes, and

impaired fetal growth (Jain & Gordon, 2011; Landy et al., 1988; Park-Wyllie et al., 2000; Schatz

et al., 2004).

Page 36: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

20

Animal studies have investigated the potential teratogenic effect of the thiopurines. In

mice given AZA intraperitoneally (i.p.) at doses 4-13 times human therapeutic doses, increased

rates of cleft palate and skeletal anomalies were observed (Githens et al., 1965; Rosenkrantz et

al., 1967). A similar study at these high doses, however, observed no fetal anomalies in mice

(Tuchmann-Duplessis & Mercier-Parot, 1964). No malformations were observed in mice born to

mothers injected i.p. with therapeutic levels or twice therapeutic levels of AZA during the time

of organogenesis, however, there was an increased frequency of fetal loss and growth retardation

(Tuchmann-Duplessis & Mercier-Parot, 1964; Githens et al., 1965; Rosenkrantz et al., 1967). In

pregnant rabbits injected i.p. with AZA at a dose two to six times higher than that used in

humans, increased frequencies of limb malformations, ocular anomalies, and cleft palate were

observed in the offspring (Tuchmann-Duplessis & Mercier-Parot, 1964). In pregnant rats given

a single injection of 6-MP equivalent to 37.5 to 156 times the maximum human therapeutic dose

on the 11th or 12th day of gestation, an increased incidence of cleft palate, skeletal and urogenital

anomalies, and diaphragmatic hernia were observed in fetuses (Murphy, 1960; Bragonier et al.,

1964; Chaub and Murphy, 1968; Kury et al., 1968; Puget et al., 1975). However, no

malformations were induced in fetuses of pregnant rats treated orally at various times during

organogenesis with 6-MP at the equivalent of one to 12 times the maximum human dose

(Thiersch, 1954).

Several studies have investigated the safety of thiopurines in human pregnancy. A

review of 27 case series and three large retrospective and prospective cohort studies did not find

an increased risk for major malformations associated with the use of AZA in pregnancy (Cleary

& Kallen, 2009; Coelho et al., 2011; Francella et al., 2003; Goldstein et al., 2007; Polifka &

Friedman, 2002). These studies are discussed further in detail in Chapter 4. However, one recent

study reported a trend for increased malformations and an increased risk specifically for

ventricular/atrial septal defects after prenatal AZA exposure (Cleary & Kallen, 2009). An

increased risk for prematurity and lower birth weight has also been reported, but this may be

related to the underlying maternal disease (Cleary & Kallen, 2009; Goldstein et al., 2007).

Although evidence supports the relative safety of AZA in pregnancy compared to the untreated

disease (Coelho et al., 2011), safety concerns still arise from its mechanism of action and its

toxicity during pregnancy in animal studies (Polifka & Friedman, 2002). Because of these

remaining safety concerns, over eighty percent of women with IBD have concerns about

Page 37: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

21

medications, including AZA, in pregnancy and fears surrounding medications are greater than

the negative effect of IBD exacerbation (Mountifield et al., 2010). Moreover, some physicians

still choose not to administer AZA during pregnancy because they consider it to be a potential

teratogen or it is discontinued in the third trimester (Mahey et al., 2013; Peyrin-Biroulet et al.,

2011).

After oral administration, AZA is extensively (>80%) cleaved into 6-MP in the blood

(Polifka & Friedman, 2002). 6-MP is a purine analogue, and after further conversion into 6-

thioguanine nucleotides (6-TGNs), the active metabolites (Figure 1-4), it can inhibit de novo

purine ribonucleotide synthesis leading to inhibition of cell proliferation. Furthermore, 6-MP

can be cytotoxic after incorporation of its metabolites into cellular DNA. Side effects include

bone marrow suppression, hepatotoxicity, and an increased risk of neoplasia (Nelson et al., 1975;

Van Scoik et al., 1985). Since 6-MP targets rapidly dividing cells, the developing fetus would

theoretically be expected to be sensitive to its cytotoxicity. However, even high dose 6-MP was

not found to be useful as a single-agent medical abortifacient in early pregnancy, leading the

authors to suggest that this drug does not alter cell division in trophoblastic tissues (Davis et al.,

1999).

The half-life of 6-MP is short at one hour (Chan et al., 1987; Ohlman et al., 1994). The

metabolite, 6-thioguanine (6-TG), is found intracellularly and the half-life is much longer at

three to 13 days (Lennard, 1992; Sandborn et al., 1995). The effect of pregnancy on the

pharmacokinetics of 6-MP has only recently been studied. Jharap et al. (2014) studied a small

cohort of 30 pregnant women with IBD receiving steady-state thiopurines. Median 6-TGN

concentrations decreased over the course of the pregnancy but returned to preconception baseline

after delivery. On the other hand, 6-methylmercaptopurine (6-MMP) concentrations increased

but returned to baseline after delivery. Although these changes were statistically significant, the

clinical significance is unknown as no woman in this study developed toxicity or therapeutic

failure. The authors also measured fetal concentrations of 6-TGN in umbilical cord blood

immediately after delivery and found fetal concentrations to correlate positively with maternal 6-

TGN levels.

Page 38: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

22

Figure 1-4 Metabolic pathway for 6-mercaptopurine and its prodrug azathioprine.

GMPS, guanine monophosphate synthetase; HPRT, hypoxanthine–guanine phosphoribosyltransferase; IMPDH,

inosine monophosphate dehydrogenase; meTGMP, methylated TGMP; meTIMP, methylated TIMP; TGMP,

6’thioguanosine 5’monophosphate; TGN, thioguanine nucleotides; TIMP, 6-thioinosine 5 ́ monophosphate; TPMT,

thiopurine S-methyltransferase; TXMP 6-thioxanthosine 5’monophosphate; XO, xanthine oxidase;

Page 39: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

23

1.3 Recreational Drug Use During Pregnancy: Alcohol and Fetal Alcohol Spectrum Disorder

In addition to therapeutic drugs, many women of childbearing age consume recreational

or illicit drugs. In a survey of Canadian women who had recently given birth, the self-reported

incidence of alcohol use in pregnancy was 5-15% and street drug use was 1-5% (Finnegan, 2013;

Zelner & Koren, 2013). However, because of the stigma associated with alcohol and recreational

drug use in pregnancy, self-reporting is likely to underestimate the true incidence (Ostrea et al.,

2001). In Canada, one standard drink contains 17.05 mL (13.45 g) of pure ethanol and in the US

it equals to 18 mL (14.00 g) of pure alcohol. This roughly corresponds to one 12 fl. oz. (341 mL)

bottle of beer, cider, or cooler (all 5% alcohol); one 5 fl. oz.(142 mL) glass of wine (12%

alcohol); or 1.5 fl.oz. (43 mL) shot of 80-proof spirits or liquor (40% alcohol) (Butt et al., 2011;

Centers for Disease Control and Prevention, 2012). Statistics Canada defines heavy drinking as

>5 drinks per occasion, >12 times over the past year (Dell & Roberts, 2006).

Prenatal alcohol exposure can lead to a wide range of deficits known as fetal alcohol

spectrum disorder (FASD). FASD is considered the leading preventable cause of intellectual

disability in the developed world (Abel & Sokol, 1986). The most severe form of FASD is fetal

alcohol syndrome (FAS) and is characterized by prenatal and postnatal growth restriction, facial

dysmorphology, and neurocognitive and behavioural dysfunction (Chudley et al., 2005). Fetal

alcohol syndrome is characteristic of heavy daily or weekend binge drinking (Abel, 1998). Also

under the umbrella term FASD are less severe outcomes such as partial FAS, alcohol-related

birth defects, and alcohol-related neurodevelopmental disorder. The less severe outcomes result

from differences in frequency and amount of alcohol consumed, timing of exposure during the

pregnancy, genetic factors, and maternal nutrition (Alati et al., 2006; Chudley et al., 2005;

Shankar et al., 2006; Warren & Li, 2005). Criteria for diagnosing FASD as set by the Institute of

Medicine (IOM) are detailed in Table 1-2.

Page 40: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

24

Table 1-2 The 1996 Institute of Medicine’s Diagnostic Criteria for FASD (Stratton et al.,

1996)

Category Description 1. FAS with confirmed maternal alcohol exposure

• Confirmed maternal alcohol exposure • Facial anomalies including thin upper

lip, flattened philtrum flat midface, small palpebral fissures

• Growth retardation • Central nervous system

neurodevelopmental abnormalities 2. FAS without confirmed maternal alcohol exposure

• Same as above but without confirmed maternal alcohol consumption

3. Partial FAS with confirmed maternal alcohol exposure

• Confirmed maternal alcohol exposure • Facial anomalies (less than seen in 1 &

2) • Growth retardation • Central nervous system

neurodevelopmental abnormalities • Behavioural/cognitive abnormalities

4. Alcohol-related birth defects • Confirmed maternal alcohol exposure • Congenital anomalies such as cardiac

defects, skeletal defects, among many others

5. Alcohol-related neurodevelopmental disorder

• Confirmed maternal alcohol exposure • Central nervous system

neurodevelopmental abnormalities OR Behavioural/cognitive abnormalities

Page 41: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

25

In addition to primary features associated with FASD, many children, adolescents, and

adults with FASD suffer from secondary disabilities (Table 1-3). The risk of developing

secondary disabilities associated with FASD is decreased 2- to 4-fold with an early diagnosis

and/or a stable, loving home environment (Streissguth et al., 2004). The large array of primary

and secondary deficits associated with FASD results in a large economic impact for affected

individuals, affected families, and society. In Canada, where the prevalence of FAS and FASD is

considered to be similar to USA general population estimates of 0.1% and 1.0% respectively

(May & Gossage, 2001), the associated annual economic loss is estimated at 4-6 billion dollars

(Stade et al., 2009). If extrapolated to include other countries, the global economic burden

would be enormous, exemplifying the need to establish FASD prevention strategies globally.

Table 1-3 Secondary Disabilities in Adolescents and Adults with FAS or FAE (O’Connor et

al., 2002; Streissguth et al., 2004)

Adverse Life Outcome Prevalence

Psychiatric Disorder 87% (n=23)

Disrupted School Experience 61% (n=415)

Trouble with the Law 60% (n=415)

Confinement 50% (n-415)

Inappropriate Sexual Behaviour 49% (n=415)

Alcohol/Drug Problems 35% (n=415)

Page 42: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

26

Ethanol can freely cross the placenta and transfer is bidirectional between the maternal

and fetal circulations (Brien et al., 1983; Nava-Ocampo et al., 2004). Ethanol concentrations are

similar in the maternal and fetal circulations; however, amniotic fluid may act as a reservoir for

ethanol and prolong fetal exposure (Brien et al., 1983; Nava-Ocampo et al., 2004). Numerous

mechanisms have been proposed as contributing to alcohol-induced fetal damage. Ethanol itself

is generally considered to be the teratogenic substance, however, metabolites such as

acetaldehyde also may play a role (reviewed in Goodlett et al., 2005). Suggested mechanisms

include oxidative stress to both the fetus and placenta leading to excessive cell death,

interference with developmental regulatory proteins, interference with neurotransmitters, and

interference with trophic factors that regulate neurogenesis and cell survival (Bosco & Diaz,

2012; Goodlett et al., 2005). However, no single mechanism can account for all the phenotypes

observed in children with FASD and we still do not have a good understanding of the mechanism

of alcohol teratogenesis (Goodlett et al., 2005).

In addition to ethanol, alcoholic beverages may contain a small amount of methanol as a

congener (Lachenmeier et al., 2011). Furthermore, methanol is produced endogenously in the

pituitary from S-adenosylmethionine (Axelrod & Daly, 1965; Sarkola & Eriksson, 2001).

Methanol shares a metabolic pathway with ethanol in that it is first metabolized by alcohol

dehydrogenase (ADH) into formaldehyde or acetaldehyde, respectively (Kalant & Khanna,

2007). Second, it is further metabolized by acetaldehyde dehdydrogenase (ALDH) into formic

acid or acetic acid, respectively. In heavy drinkers, methanol may accumulate since ethanol has a

higher affinity for alcohol dehydrogenase (ADH) and consequently methanol may reach plasma

concentrations above 2 mM (Kapur et al., 2007; Majchrowicz & Mendelson, 1971). Formic acid,

the toxic metabolite of methanol, has been detected in both sera and cerebrospinal fluid of

alcoholics in concentrations that are neurotoxic (Kapur et al., 2007). There is no current

regulation of methanol content in Canada, however, the European Union has set a limit for

naturally occurring methanol at 10g methanol/ litre of ethanol which equates to 0.4% (v/v)

methanol at 40% alcohol (Paine & Davan, 2001).

Formic acid has also been recently detected in maternal blood and umbilical cord blood

of infants born to heavy drinkers (Kapur, et al., 2009). Furthermore, preliminary studies have

shown a negative correlation between formic acid concentrations in umbilical cord blood at the

time of delivery and cognitive function as measured using Bayley scores (r= -0.6154, p=0.025,

Page 43: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

27

n=12 at 12 months and r=-0.6241, p=0.023, n=13 at 18 months) (Kapur, et al., 2009). Formic

acid has also been shown in animal studies to lead to growth restriction, physical malformations,

and depletion of glutathione in the embryo (Andrews et al., 1998; Brown-Woodman et al., 1995;

Hansen et al., 2005; Harris et al., 2004). Furthermore, human studies have reported fetal alcohol

syndrome-like facial features in infants born after solvent abuse (including methanol) by the

mother during pregnancy (Scheeres & Chudley, 2002). However, the role of methanol and

formic acid in the development of FASDs has not been investigated.

Formic acid is cleared by the body through metabolism into carbon dioxide and two non-

free radical pathways have been proposed for this step. First, oxidation through the catalase-

peroxidative system (Chance, 1950), and second by entering the one-carbon pool. Formate enters

the one carbon pool by combining with tetrahydrofolate (THF) to form 10-formyl-THF, a

reaction catalyzed by 10-formyl-THF synthetase (Johlin et al., 1987). This is followed by the

oxidation of 10-formyl-THF to carbon dioxide mediated by 10-formyl-THF dehydrogenase (10-

FTHFDH). Metabolism of formate through the one carbon pool has been shown to be the major

route (Johlin et al., 1987; Makar & Tephly, 1976; Chiao & Stokstad, 1977; Palese & Tephly,

1975) and also the predominant route in primates (McMartin et al., 1977). In humans, formate

oxidation to carbon dioxide is dependent upon folate (Johlin et al., 1987). Sokoro and

colleagues, using an animal model, showed that the formic acid half-life was significantly longer

in the folate deficient animals when compared with folate sufficient animals (Sokoro et al.,

2008). Studies on rat brain hippocampal slices showed that formic acid can cause neuronal cell

death at concentrations found in heavy alcohol users, and this toxicity can be mitigated by folate

(Kapur et al., 2007).

Folic acid deficiency is a common finding in chronic alcoholics (Herber et al., 1963;

Halsted et al., 2002a). Chronic alcohol ingestion reduces the intestinal absorption of dietary folic

acid leading to a decrease in the folate metabolic pool (Halsted et al., 2002b). A decrease in this

pool prolongs the formate blood levels by decreasing the rate at which formate combines with

tetrahydrofolate, the first step in its metabolism to carbon dioxide. Decreased detoxification of

formate may lead to formate-induced cytotoxicity mediated by oxidative stress (McMartin et al.,

1977). Since folic acid plays a major role as a coenzyme in one-carbon metabolism and is a key

participant in the biosynthesis of DNA, RNA and certain amino acids, it is vital to proper fetal

development. In an animal model, it has been suggested that ethanol can alter the normal transfer

Page 44: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

28

of folate across the placenta and is described further in Section 5.2. Furthermore, Lin & Lester

(1985) showed that the folate distribution in the rat placenta was altered after ethanol exposure as

accumulation of methyltetrahydrofolate was increased compared to non-methyltetrahydrofolates.

Ethanol has been shown to decrease the activity of methionine synthase, which has been

suggested to decrease the regeneration of non-methyltetrahydrofolates (Barak et al., 1985;

Finkelstein et al., 1974). To our knowledge, there are no studies in humans to determine the

effect of folic acid on the toxicity of formic acid in pregnancy as this is a novel hypothesis to

explain the deficits observed in the FASDs.

Page 45: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

29

1.4 Overall Rational and Primary Objectives

The human placenta performs many functions that support the maintenance of pregnancy

and the normal development of the fetus. This organ regulates the supply of oxygen and

nutrients to and removal of waste products from the fetus by bringing the maternal and fetal

circulations into close opposition. Because the human placenta separates the maternal and fetal

circulations, it can restrict fetal drug or toxin exposure from agents consumed by the pregnant

woman. Studies utilizing the human placenta are vital to understanding fetal drug exposure and

also drug safety in pregnancy. Dual perfusion of a single placental lobule is the only

experimental model to study human placental transfer of substances in organized placental

tissue. To date, there has not been any attempt at a systematic evaluation of this model. Careful

validation of the perfusion model is needed before it can be used routinely to predict placental

drug transfer during preclinical evaluation.

The first objective of this thesis was to systematically evaluate the placental perfusion

model in predicting placental drug transfer and to develop a pharmacokinetic model to

account for non-placental pharmacokinetic parameters in the perfusion results. Having a

validated and accurate model to predict drug transfer across the human placenta will be an

invaluable tool to help guide decisions regarding the benefits and risks of new medications

that may be required during pregnancy. We hypothesize that the placental perfusion model

will accurately predict in vivo placental drug transfer after adjusting results for non-placental

pharmacokinetic parameters.

The observed safety of AZA in pregnancy likely results, in part, from the placenta

limiting fetal exposure to the main metabolite, 6-MP. Indeed, only 1-2% of maternal 6-MP

concentrations are found in cord blood and the human placenta is considered to be a relative

barrier to 6-MP and its metabolites (de Boer et al., 2006; Saarikoski & Seppala, 1973).

However, the mechanism of how the placenta restricts 6-MP transfer to the fetus is unknown. 6-

MP is a substrate for several drug transport proteins, including breast cancer resistance protein

(BCRP) (de Wolf et al., 2008). BCRP is highly expressed in the placenta and is localized to the

brush border (maternal-facing) membrane of the syncytiotrophoblast (Maliepaard et al., 2001).

The importance of BCRP and other placental transporters in limiting fetal exposure to various

Page 46: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

30

drugs has been demonstrated in both human and animal models (Jonker et al., 2000; Pollex et al.,

2008).

The second objective of this thesis was to further evaluate the safety of AZA and 6-MP

in pregnancy using two approaches. First, to systematically review the literature and perform

a meta-analysis of available studies in order to further quantify the safety of thiopurines in

pregnancy. Second, to determine the mechanisms by which the placenta restricts 6-MP

transfer using the validated perfusion model from objective one. Understanding these

mechanisms will allow for prediction of fetuses that may be at risk of higher exposure to 6-MP

as a result of drug interactions with, or polymorphisms in, placental drug efflux proteins. We

hypothesize that 6-MP does not increase the risk for fetal malformations when taken during

pregnancy and that 6-MP is a substrate for placental drug efflux transporters.

Heavy alcohol consumption during pregnancy is associated with numerous placental

alterations including placental dysfunction, decreased placental size, gene expression changes

and endocrine disruptions (Bosco & Diaz, 2012; Burd et al., 2007; Rosenberg et al., 2010).

Methanol and its toxic metabolite, formic acid, have been detected in high concentrations in sera

of heavy alcohol users and may contribute to the teratogenic nature of alcohol beverages (Kapur

et al., 2007). Since formic acid can produce oxidative stress (Harris et al., 2004), it may be

placentotoxic; however, this has not been studied to date. Since the placenta is a reservoir for

folate (Henderson et al., 1995) and folate is required for detoxification of formic acid, there may

also be a potential role for detoxification of formic acid within the placenta itself. In the human

placenta, short-term exposure to ethanol during a human placental perfusion had no effect on

either binding or transfer of 5-methyltetrahydrofolate (Henderson et al., 1995). However, animal

studies have suggested that alcohol exposure can decrease transfer to the fetus (Fisher et al.,

1985; Keating et al., 2009; Keating et al., 2008). The chronic effect of alcohol during pregnancy

to folate status in the human placenta has not been investigated. Decreased folic acid transport

may limit the detoxification of formic acid in the fetus in addition to the other negative

consequences of folate deficiency in the fetus.

The third objective of this thesis was to evaluate the potential toxicity of formic acid to

the placenta and if folate may be protective to this toxicity. To determine this, the effect of

alcohol consumption in human pregnancy on placental transfer of folate will first be

Page 47: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

31

evaluated. We hypothesize that placental folate transfer to the fetus is decreased in

pregnancies with heavy alcohol exposure and that this is mediated by decreased expression of

placental folate transporters. Second, we will determine whether formic acid transfers across

the placenta and if it is toxic to the placenta. We hypothesize that formic acid can transfer

across the placenta because it is a small molecule and is toxic to the placenta but folate can

both decrease transplacental transfer of formic acid and mitigate toxicity.

Page 48: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

32

Chapter 2

The human placental perfusion model: a systematic review and development of a model to predict in vivo transfer of therapeutic drugs Janine R Hutson1,2, Facundo Garcia-Bournissen1,2, Amy Davis1, Gideon Koren1,2

1Division of Clinical Pharmacology and Toxicology, Hospital for Sick Children, 555 University

Ave, Toronto, Ontario, M5G 1X8, Canada

2Institute of Medical Science, University of Toronto, 1 King’s College Circle, Toronto, Ontario,

M5S 1A8, Canada

This work has been published and reproduced with permissions: Hutson JR, Garcia-Bournissen F, Davis A, Koren G. The human placental perfusion model: a

systematic review and development of a model to predict in vivo transfer of therapeutic drugs. Clin Pharmacol Ther 2011;90(1):67-76.

Page 49: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

33

The human placental perfusion model: a systematic 2review and development of a model to predict in vivo transfer of therapeutic drugs

2.1 Abstract

Dual perfusion of a single placental lobule is the only experimental model to study

human placental transfer of substances in organized placental tissue. To date, there has not been

any attempt at a systematic evaluation of this model. The aim of this study was to systematically

evaluate the perfusion model in predicting placental drug transfer and to develop a

pharmacokinetic-model to account for non-placental pharmacokinetic parameters in the

perfusion results. In general, the fetal-to-maternal drug concentration ratios matched well

between placental perfusion experiments and in vivo samples taken at the time of delivery. After

modeling for differences in maternal and fetal/neonatal protein binding and blood pH, the

perfusion results were able to accurately predict in vivo transfer at steady state (R2=0.85,

p<0.0001). Placental perfusion experiments can be used to predict placental drug transfer when

adjusting for extra parameters and can be useful for assessing drug therapy risks and benefits in

pregnancy.

Page 50: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

34

2.2 Introduction

Studies have estimated that between 23-85% of women require medication at some point

during their pregnancy (Gendron et al., 2009). There is often limited information on the fetal

safety of medications in pregnancy and the decision to continue treatment relies on weighing the

benefit of the drug against the fetal risks. An important determinant in this risk assessment is

estimation of fetal exposure, based on quantifying the amount of drug that crosses the placenta.

The rate-determining step in placental drug transfer is the crossing of drug molecules across a

single layer of fetal-derived syncytiotrophoblasts (Ceckova-Novotna et al., 2006). In vivo studies

evaluating the transplacental kinetics of drugs in humans have numerous ethical concerns

regarding fetal and maternal safety. Furthermore, the feto-placental unit in situ is not easily, or

ethically, accessible until delivery. Animal studies cannot always be extrapolated to humans

because the placenta is the most species-specific mammalian organ (Ala-Kokko et al., 2000). In

vitro models utilizing cell culture and membrane vesicles are mostly limited to the investigation

of specific mechanisms of transfer, such as active transport or passive diffusion, but they lack

anatomic integrity and blood flow.

Dual perfusion of a single placental lobule ex vivo is the only experimental model to

study human placental transfer of substances in organized placental tissue and theoretically may

be able to predict fetal exposure compared to other experimental methods (Table 1.1). Although

tedious, the placental perfusion model has proven useful in studying both endogenous and

exogenous substrates, such as amino acids, hormones, electrolytes, viruses, therapeutics, and

illicit drugs (Ala-Kokko et al., 2000; Malek et al., 2009; Muhlemann et al., 1995; Omarini et al.,

1992). This model allows to investigate simultaneously many properties that can influence

placental drug transfer, such as physiochemical properties (size, pKa, and lipophilicity) and

pharmacokinetic factors (active transport, placental binding, and metabolism). A limitation of

this model is that it cannot fully incorporate non-placental pharmacokinetic factors such as

protein binding, elimination, and distribution in the maternal and fetal compartments. However,

these factors can be determined in vitro. While protein can be added to the perfusion medium

(Gavard et al., 2006; Hemauer et al., 2009; Johnson et al., 1999; Schenker et al., 1999), it is

difficult to mimic physiological conditions since there are numerous proteins and endogenous

factors that could influence binding in vivo. A description of the model is given in section

1.1.4.1.

Page 51: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

35

Results obtained from placental perfusion experiments are often compared to in vivo

results (if available) that have been collected by measuring drug concentrations in maternal

blood and fetal/umbilical cord blood at the time of delivery. The calculation of cord to maternal

blood concentration ratio (C:M) can provide information on how much drug crosses the placenta

in vivo. Drugs are either measured after single bolus administration before delivery or after the

mother has been taking the medication chronically during pregnancy and is likely in steady state.

From the mother, venous blood is usually collected, while umbilical cord blood can be arterial,

venous, or mixed cord blood. Ideally, drug levels should be measured from the umbilical vein

for evaluating placental drug transfer, since this vessel receives blood directly from the placenta.

Drug concentrations are likely to be higher in the umbilical vein before reaching steady state

compared to mixed or arterial cord blood. Measuring the concentration difference between the

umbilical vein and artery can also provide information on whether there is drug distribution or

metabolism by the fetus. It is, however, not possible from cord blood measurements to

determine if there is accumulation of the drug in fetal tissues.

It is difficult to accurately determine the transplacental kinetics of a particular drug by

this in vivo method because it is measured at only one time point and typically in a small number

of mother-infant pairs. Often there is large variability between pairs due to sample

contamination, timing of samples, sample site, inter-individual differences, and timing of

exposure. Maternal venous blood concentrations do not reflect those being presented to the

placenta and can depend on the tissues drained by the selected vein if an arteriovenous gradient

exists (Hermann et al., 1999). However, sampling at the time of delivery is easily accessible

using umbilical cord blood and is an ethically acceptable method. It is often the only source of in

vivo human data.

Because in vivo data are not usually available, especially for new drugs, a method to

reliably estimate transfer across the placenta is needed. Careful validation of the perfusion

model is needed before it can be used routinely to predict placental drug transfer during

preclinical evaluation (Ala-Kokko et al., 2000). To date, no systematic evaluation of the

perfusion model has been conducted. This review aims to systematically evaluate the placental

perfusion model in predicting placental drug transfer by comparing it to in vivo data.

Subsequently, we constructed a pharmacokinetic model that best allows prediction of the in vivo

Page 52: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

36

maternal-fetal drug distribution at steady state. Finally, we also provide recommendations to

improve the reliability of the predictions provided by the perfusion method.

2.3 Methods

2.3.1 Literature Search

The in vivo data available in the literature are mostly limited to maternal blood and

umbilical cord blood drug concentrations measured at the time of delivery. Since only one time

point is available from these studies, the best measure to compare in vivo to perfusion studies

was the fetal to maternal drug concentration ratio. A systematic search for papers that evaluated

placental transfer of a therapeutic drug was performed by searching Medline, EMBASE, and

EMBASE Classic from inception to August 31, 2010 using the search strategy, placenta AND

perfusion and limited to human studies (Omarini et al., 1992). Perfusion studies were included if

they 1) were published in English 2) investigated a therapeutic drug 3) completed dual perfusion

of a term human placenta 4) involved placing a compound in the maternal circulation and

measuring its appearance in the fetal circulation 5) the fetal to maternal ratio (F:M) at the end of

the experiment or during steady state could be extracted. Papers were excluded if they 1)

investigated an endogenous compound, illicit drug, or a natural health product 2) did not present

F:M or did not provide sufficient information to calculate it 3) the length of the perfusion was

not sufficient to provide useful information (<20 minutes, since this is the time it would take for

the maternal reservoir to be completely cycled through once, if it is in closed-configuration). For

each drug studied using the perfusion model, a subsequent search was performed to retrieve

papers that measured cord blood and maternal blood drug levels.

2.3.2 Extraction of Data and Qualitative Comparisons

For each drug, the following data were extracted from the perfusion studies: F:M ratio,

initial maternal drug concentration, whether the experiment reached steady state, whether protein

was added to the perfusate, whether the system was in open- or closed-configuration, and the

number of completed perfusion experiments. From the in vivo studies, cord blood to maternal

Page 53: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

37

blood concentration ratio (C:M), time elapsed between drug administration and delivery (blood

sampling), and number of mother-child pairs included. The percentage of protein binding (in

non-pregnant adults) and clinical drug concentrations for each drug were obtained from

DRUGDEX® System [Internet database]. Greenwood Village, Colo: Thomson Reuters

(Healthcare) Inc. Updated periodically. Available www.micromedex.com). Data were

summarized and the placental transfer as determined by the perfusion and in vivo methods were

defined as limited transfer (F:M <0.1 ), transfer (F:M 0.1 to 1.0), or fetal accumulation (F:M >

1.0).

2.3.3 Quantitative Calculations

Quantitative comparisons between the perfusion and the in vivo data were performed on

data where a direct comparison was meaningful, ie. the perfusion and in vivo data were from

comparable conditions. The following inclusion criteria were developed to ensure comparable

conditions: 1) the perfusion was performed in closed-configuration 2) the perfusion and in vivo

experiments both reached steady state (the maternal and fetal compartments reached a fixed

proportionality resulting in a constant C:M or F:M; there was a constant ratio of at least three

data points and over at least 20 minutes) and 3) the placenta used in the perfusion and the cord

and maternal blood samples were both taken at the same gestational age (ie. term). The F:M and

C:M ratios were plotted and Pearson correlation was calculated using GraphPad Prism (version

4.0b for Macintosh, GraphPad Software, San Diego, California). Outliers were determined by

analysis of residuals.

2.3.4 Model Development

The perfusion experiments often did not take into account the physiological differences in

protein binding between fetal and maternal circulations. The physiological pH difference

between the circulations was also often not taken into consideration, where the fetal circulation is

slightly more acidic than maternal (7.35 vs 7.4 respectively) (Reynolds & Knott, 1989). To

account for these, the results obtained from the placental perfusion was adjusted using the

following F:M equation (Garland, 1998):

Page 54: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

38

equation (1)

where %unboundM and %unboundF are the % of the free drug in the maternal and fetal

circulations, respectively; the pHM and pHF are the pH in the maternal and fetal circulations,

respectively; and the CLMF, CLFM, and CLF are the maternal-to-fetal, fetal-to-maternal and fetal

clearances, respectively. The CLMF/CLFM term was taken from the perfusion experiments as the

F:M ratio at steady state from a closed-configuration study, or from the clearance rates obtained

from an open-configuration study. Furthermore, to adjust the results, information regarding fetal

and maternal protein binding had to be available. Only drugs from studies providing this

information could be adjusted. To apply the F:M equation (equation 1), the following

assumptions were made: 1) only the free drug can cross the placenta passively 2) the free fraction

is not concentration dependent 3) only the un-ionized form of the drug can cross the placenta

passively and 4) the fetal clearance rate is negligible. If studies measuring the percentage of

protein binding in fetal and maternal blood were not available, binding could be estimated from

non-pregnant adult protein binding using the method by Hill and Abramson (Hill & Abramson,

1988) (Equations 2 to 4). This method requires that the drug be bound to only albumin or α1-

acid glycoprotein (AAG).

equation (2)

equation (3)

equation (4)

maternal

fetal

maternalfetal roteinP

roteinP

FB

FB

][][

×"#

$%&

'="#

$%&

'

adultpregnantnon

maternalorfetal

adultpregnantnonmaternalorfetal roteinP

roteinP

FB

FB

−−

×#$

%&'

(=#$

%&'

(][][

%unbound=100−100×B

FBF+1

F :M =%unboundM%unboundF

×1+10pKa − pHF

1+10pKa − pHM×

CLMFCLFM +CLF

Page 55: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

39

where B/F is the ratio between bound and free drug in the fetal, maternal, or adult circulations;

[Protein] is the plasma protein concentration for either albumin or AAG. The following values

were used in the calculations (Hill & Abramson, 1988): for albumin, the plasma protein

concentration ratios between fetal:adult is 0.866, maternal:adult is 0.733, and fetal:maternal is

1.20; for AAG, the plasma protein concentration ratios between fetal:adult and fetal:maternal are

both 0.37 (Hill & Abramson, 1988).

2.4 Results

2.4.1 Search Results

A total of 1732 papers were retrieved in our systematic search for studies evaluating

placental drug transfer using the placental perfusion model. 1585 papers were excluded after

reviewing the title/abstract or by not being published in English. From the 147 full-text articles

that were assessed, 19 papers did not include sufficient information to extract the F:M ratio and

29 papers investigated a drug that had no corresponding in vivo data. Furthermore, 4 papers

were excluded since their perfusion lasted ≤ 5 minutes and 3 papers were excluded because the

experimental parameters fluctuated over the course of the experiment. Another 3 papers

described perfusion experiments after adding the drug to both maternal and fetal circulations

simultaneously. Of the 89 perfusion papers included, 70 drugs were compared. An additional 58

drugs have been investigated using the perfusion model, but were included in papers that were

excluded for the above reasons (Figure 2.1).

2.4.2 Qualitative Comparison of Perfusion Results to In Vivo Data

The agreement between perfusion and in vivo results for the drugs that met the inclusion

criteria are outlined in Table 2.1. The experimental parameters and results of the perfusion and

in vivo studies are detailed in Appendix B. Out of the 70 drugs compared, 49 showed placental

transfer (F:M 0.1 to 1.0) in both the placental perfusion model and in vivo. Another 9 drugs

showed limited transfer (F:M <0.1) in both models. The remaining 12 drugs had discrepancies

between the two methods (Table 2.2).

Page 56: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

40

Figure 2-1. Flow diagram of the search strategy and articles retrieved and included in the

review.

Page 57: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

41

Table 2-1. Drugs that had the maternal-to-fetal transfer evaluated using the placental

perfusion model and where the fetal to maternal drug concentration ratios (F:M) from the

perfusion agreed with in vivo data from mother-infant pairs.

Limited Transfer (F:M<0.1) Transfer (F:M 0.1 – 1.0) Azithromycin Enfuvirtide Fondaparinux Glyburide Heparin Indinavir Insulin Lispro Low Molecular Weight Heparins Saquinavir

Acetaminophen Fluoxetine Phenytoin Acyclovir Hydralazine Propofol Alfentanil Indomethacin Propranolol ASA Labetalol Pyrimethamine Atenolol Lamivudine Quetiapine Bupivacaine Lamotrigine Ranitidine Buprenorphine Lidocaine Ritonavir Carbamazepine Lopinavir Ritrodrine Celiprolol Methadone Ropivacaine Cimetidine Methimazole Salbutamol Citalopram Methohexital Salicylic Acid Clonidine Morphine Sufentanil Diclofenac Nelfinavir Sulindac Dideoxyinosine Nifedipine Theophylline Digoxin Nortriptyline Zidovudine Enalaprilat Olanzapine Erythromycin Phenobarbital

Table 2-2. Drugs that had the maternal-to-fetal transfer evaluated using the placental

perfusion model and where the fetal to maternal drug concentration ratios (F:M) from the

perfusion disagreed with in vivo data from mother-infant pairs.

Page 58: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

42

2.4.3 Quantitative Comparison of Perfusion Results to In Vivo Data

Twenty-six drugs met the inclusion criteria for a quantitative comparison. Furthermore,

data were available for both total and free concentrations of bupivacaine and ropivacaine, so both

measures were included. There was a significant correlation between the F:M and C:M ratios

obtained by the perfusion and the in vivo measurements, respectively (R2=0.28, p=0.004) (2-2A).

Twenty-four drugs had sufficient data available to be adjusted using our model as per the

F:M equation (Table 2-3). After adjustment, there was a stronger correlation between the

perfusion and in vivo results (R2=0.54, p<0.001) compared to the unadjusted (Figure 2-2B).

There were 4 clear outliers in the adjusted comparison from digoxin, glyburide, metformin, and

oxcarbazepine. For digoxin, there was wide variability among the F:M ratio obtained from 4

different studies and among the C:M ratio obtained from 6 different studies. Since digoxin is a

substrate for P-glycoprotein and this efflux transporter decreases towards term, the concentration

ratio may be dependent on the gestational age of the placenta. Errors in digoxin measurements

from digoxin-like substances cross-reacting may further contribute to the variability in the C:M

ratios. For oxcarbazepine, the fetal and maternal protein binding were estimated from adult

binding to albumin only. This may not be accurate since it does not take into consideration

binding to other proteins or displacement by maternal endogenous substances. For metformin,

the in vivo studies may over estimate the C:M since in the study by Vanky et al. (Vanky et al.,

2005), the maternal blood sample was taken within an hour after birth and not necessarily at the

same time as fetal sampling. Other studies on metformin suggest a C:M closer to 1, but the exact

ratio could not be calculated from the information provided in the papers (Charles et al., 2006;

Eyal et al., 2010). For glyburide, both perfusion papers and an in vivo paper showed limited

transfer of glyburide at term. However, a recent study by Hebert et al. (Hebert et al., 2009)

observed a mean C:M of 0.7. Since many samples were below the limit of quantification, it is not

clear how these samples were used in the C:M calculation. Also, the gestational age at delivery

(mean = 34.6 weeks) was much shorter than that of the placentas routinely used for perfusion

experiments. After excluding the 4 outliers, the perfusion results avidly predicted the in vivo

results (R2=0.85, p<0.0001) (Figure 2-2C).

Page 59: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

43

Table 2-3. The observed and calculated F:M ratios for data obtained from placental perfusion experiments

Drug In Vivo C:M*

Observed F:M*§

%unboundM %unboundF

pKa CLMF*** CLFM

Calculated F:M****

Protein Binding References

Alfentanil 0.30 - 0.36 6.5 1.02 0.37 (Meuldermans et al., 1986) Bupivacaine 0.30 0.69 0.20 8.1 - 0.15 (Mather et al. 1971) Carbamazepine 0.91 1.00 0.90 13.9 - 1.01 (Kuhnz et al., 1984) Diazepam 1.38 0.51 2.16 3.3 - 1.10 (Ridd et al., 1989) Dideoxyinosine 0.35 - - 9.13 0.27 0.30 - Digoxin 0.89 0.11 1.04** 13.5 - 0.13 (Gorodischer et al., 1974; Hill &

Abramson, 1988) Glyburide 0.59 0.03 - 5.3 - 0.03 - Insulin Lispro <LOD <LOD - - - <LOD - Lamotrigine 0.96 1.05 0.80 5.7 - 0.84 (Rambeck et al., 1997) Lidocaine 0.52 0.82 0.70 7.9 - 0.62 (Nau, 1985) Metformin 1.53 0.47 - 12.4 - 0.53 - Methadone 0.43 0.71 0.47** 8.25 - 0.37 (Hill & Abramson, 1988; Pond et al.,

1985) Methohexital 1.00 0.75 1.16** 7.9 - 1.27 (Hill & Abramson, 1988) Morphine 0.96 0.78 1.06** 8.0 - 0.91 (Hill & Abramson, 1988; Kurz et al., 1977) Oxcarbazepine 2.25 1.1 1.06** - - 1.17 (Hill & Abramson, 1988) Phenobarbital 0.96 0.99 1.02 7.4 - 1.07 (Nau et al., 1983) Phenytoin 0.95 0.77 0.99 8.3 - 0.84 (Hamar & Levy, 1980) Propranolol 0.26 1.00 0.54 9.0 - 0.60 (Belpaire et al., 1995) Propylthiouracil 1.51 1.00 1.15** 7.8 - 1.25 (Hill & Abramson, 1988) Ropivacaine 0.31 0.82 0.39 8.1 - 0.35 (Morton et al., 1997) Sufentanil 0.76 - 0.45 8.5 0.78 0.39 (Meuldermans et al., 1986) Theophylline 1.05 0.88 1.03** 8.7 - 1.01 (Frederiksen, et al., 1986; Hill &

Abramson, 1988) Valproic acid 1.51 0.88 1.65 4.8 - 1.48 (Froescher et al., 1984) Zidovudine 1.18 1.00 1.06** 9.68 - 1.12 (Hill & Abramson, 1988) § F:M ratio obtained from closed-configuration placental perfusions that obtained steady-state. *If more than one study available from (Appendix B), the weighted mean was calculated. **Fetal and/or maternal protein binding unavailable. Numbers were calculated using the method by Hill & Abramson (1988). ***Clearance ratios obtained from open-configuration placental perfusions. ****Calculated using equation

Page 60: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

44

Figure 2-2. Agreement between perfusion fetal to maternal (F:M) concentration ratios and

the in vivo umbilical cord to maternal blood (C:M) drug concentration ratios A) with data

obtained from publications B) after adjusting the F:M ratios using the equation 1 and C)

after adjusting the F:M ratios using equation 1 and removing the 4 outliers where

experimental limitations preclude a direct comparison.

A)

B)

C)

Page 61: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

45

2.5 Discussion

For drugs where there is limited fetal safety information from clinical studies, estimating

fetal drug exposure is critical in predicting whether a drug can be safely administered in

pregnancy. As studying human transplacental transfer in vivo is almost always unethical, other

experimental models are needed to estimate fetal exposure. Placental perfusion is the only

experimental model to study human placental transfer of substances in organized placental

tissue. In agreement with a previous study (Gavard et al., 2009), the placental perfusion model is

useful in predicting whether the drug will have limited placental transfer. When the drug does

transfer across the placenta, it has been suggested that in vivo data are required to adequately

quantify this transfer (Gavard et al., 2009).

By adjusting the perfusion results using the F:M equation, the transfer across the placenta

could be accurately predicted when the conditions from the placental perfusion experiment and

the in vivo data were comparable (ie. they were both from term placenta and they reached steady

state). Transfer was either underestimated or overestimated by the perfusion model mostly

because it did not mimic protein binding in the fetal circulations. Many groups added protein,

such as human or bovine serum albumin, to the perfusate in order to investigate the effects of

protein binding on transfer (Appendix B); however, addition of only albumin or AAG and in the

same concentration on both maternal and fetal sides, does not mimic physiological conditions.

Furthermore, use of human plasma in the perfusate is often not practical and also fails to mimic

physiological conditions since α-fetoprotein is present in fetal serum. Adding protein to the

perfusate is useful in evaluating the effects of protein on the rate of transfer, but theoretically

should not affect the F:M ratio of the free drug at steady state. By adjusting the perfusion results

using the F:M equation and addressing differences in protein binding, appreciably more accurate

predictions of in vivo placental transfer at steady state for was obtained (Figure 2-2). For

example, the adjusted perfusion data for bupivacaine, lidocaine and ropivacaine all resulted in

closer values to the in vivo situation compared to adding human plasma to the maternal circuit

during the perfusion (Table 2.3).

Page 62: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

46

Figure 2-3. Comparison of F:M ratios for morphine obtained from placental perfusion

experiments versus maternal blood and umbilical cord blood (Kopecky et al., 2000;

Kopecky, 1999). Each in vivo data point represents a single patient’s F:M Ratio (n=8).

Each perfusion data point represents the mean ± SD value of 4 placental perfusions with

addition of morphine into the maternal compartment only.

Maternal and fetal plasma differ in both albumin and AAG concentrations, the two most

important drug binding plasma proteins. Towards term, albumin is appreciable higher in the fetal

circulation compared to the maternal, with the fetal to maternal albumin ratio increasing from

0.28 in the first trimester to 1.20 at term (Krauer et al., 1984). Furthermore, a 3-fold higher

concentration of free fatty acids towards term in the maternal circulation can displace drugs from

binding to albumin in the maternal circulation (Nau et al., 1984; Ridd et al., 1983). This leads to

increased binding to albumin in the fetal circulation compared to the maternal circulation at term.

Increased fetal binding can lead to increased placental drug transfer since it is the free

concentration of the drug that will equilibrate. Fetal accumulation of drugs bound to albumin at

term is expected if the drug is passively diffused. The perfusion model was not able to predict

the in vivo fetal accumulation observed with oxcarbazepine, diazepam, valproic acid, metformin,

and propylthiouracil (Table 2.2). With the exception of metformin, the other four drugs are all

Page 63: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

47

highly bound to albumin and this is likely the reason that the perfusion model did not accurately

predict in vivo transfer. After adjusting for differences in fetal and maternal binding to albumin

using the F:M equation and the method by Hill & Abramson (Hill & Abramson, 1988), the fetal

accumulation observed in vivo could be predicted using the perfusion results (Table 2.3).

The other major plasma binding protein, AAG, gradually increases throughout gestation

in fetal plasma while maternal levels stay fairly constant (Krauer et al., 1984). The fetal to

maternal AAG ratio increases from 0.09 in the first trimester to 0.37 at term (Krauer et al., 1984).

At delivery, there is no apparent change in AAG levels compared to before induction of labour

(Porter et al., 2001). Compared to maternal, there is lower binding to AAG in the fetal

circulation throughout pregnancy. In contrast to drugs bound to albumin, the placental perfusion

model overestimated fetal exposure to drugs bound to AAG, including the anesthetics and

propranolol. After adjusting the perfusion results, there was a stronger correlation between the

perfusion F:M and the in vivo C:M for drugs bound to AAG (Table 2.3). Protein binding has

previously been suggested as the most significant factor in establishing the F:M ratio at steady

state (Hill & Abramson, 1988) and our results support this finding when it is the free

concentration of the drug that equilibrates across the placenta. Protein binding is not the most

significant factor if there is limited transfer of the free drug, as is the case with insulin lispro,

heparin, fondaparinux, and enoxaparin, where the large molecular size precludes placental

transfer. For drugs where efflux transporters limit transfer, such as indinavir, saquinavir,

enfuvirtide, azithromycin, and glyburide, protein binding is also not a significant factor. For

drugs that are effluxed back into the maternal circulation, the placental perfusion model was able

to show limited transfer. This suggests that the placental perfusion model is useful in identifying

drugs that are actively transported. The net transfer of these drugs differed from what would be

predicted based on the drug’s partition coefficient and passive diffusion alone. The perfusion

model can also be useful in identifying active transport when used under equilibrative conditions

(adding the drug to both the fetal and maternal circulations in equal concentrations). Drug

accumulation in one compartment represents drug transport against a concentration gradient.

The active transport of glyburide by placental breast cancer resistance protein (BCRP) was

established by this method (Kraemer et al., 2006; Pollex et al., 2008).

The placental perfusion model was able to accurately predict limited transfer of drugs

across the placenta at term. Thus, perfusion results that show no transfer to the fetal circulation

Page 64: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

48

do not need to be adjusted using the F:M equation. Since the drug originates in the maternal

circulation, it will not even reach the fetal circulation if transfer is restricted by the placenta.

Thus, the term CLFM would not be reliably assessed by the perfusion model if the drug is placed

on the fetal side at much higher concentrations than would be reached in vivo. For example, if

CLMF and CLFM for indinavir from open-configuration placental perfusion (Sudhakaran et al.,

2005), the differential protein binding for maternal and fetal circulations (Sudhakaran et al.,

2007), and the pKa were put into the F:M equation, a F:M of 0.26 would result. This result

clearly overestimates placental transfer since in vivo data show very limited transfer (Appendix

B). Therefore, although adjusting the results using the F:M equation works well for drugs that

cross the placenta, it is not relevant to drugs which have limited transfer across the placenta.

Interpretation of placental perfusion results must take into consideration how the

experimental conditions relate to in vivo physiological conditions. If perfusion conditions do not

include protein in the perfusate, then the drug added into the system will theoretically be free

(unless it binds to tubing, dextran in the perfusate, or residual blood). When no protein is added

to the perfusate, then comparison to the in vivo results should be comparable to the free fraction

of the drug. For example, during a perfusion, alfentanil was added into the maternal circulation

in open-configuration (Zakowski et al., 1994). At steady state, the F:M ratio was 0.22 and the

authors commented that this value was similar to those obtained in vivo. However, in vivo, the

C:M ratio for free alfentanil is 0.97 (Gepts et al., 1986). A direct comparison of the free drug

would suggest that the perfusion underestimates exposure. However, since the perfusion was

performed in open-configuration, it underestimates the F:M ratio because the maternal circuit

remains at the initial concentration. In order to predict fetal exposure to the bound and free drug,

the perfusion results (clearances from open-configuration) can be adjusted using the F:M

equation. After adjustment, the calculated F:M ratio was 0.37 (Table 2.3), which is very similar

to the in vivo results for the free plus bound drug concentration.

Since protein binding is a critical factor in predicting the extent of placental transfer (Hill

& Abramson, 1988), experimental conditions must properly control for inadvertent binding.

Without addition of exogenous protein into the perfusion medium, protein resulting from

washout and placental secretion has been measured in both the fetal and maternal circulations.

Protein in the maternal circulation was found to be around 10-fold higher than the fetal

circulation (Shah & Miller, 1985). This likely results from the fetal side having a better washout

Page 65: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

49

and the fact that some proteins (eg. hCG) are preferentially secreted into the maternal circulation

(Shah & Miller, 1985). Furthermore, haemoglobin content in the maternal circulation was found

to be 100-fold higher than in the fetal circulation (Mathiesen et al., 2010). If the drug is able to

bind to proteins that remain from washout or secreted placental proteins, then there may be an

experimental bias towards the drug being found in the maternal circulation.

The placental perfusion model offers a valid model to predict transplacental drug

passage. After careful interpretation of the results and adjustments to incorporate protein binding

and ion trapping, the perfusion model can predict the extent of placental transfer. A limitation of

the perfusion model is that it mostly uses term placentas and cannot necessarily be extrapolated

to earlier gestational ages. At term, the placental transfer layer is the thinnest and expression of

certain efflux transporters such as P-glycoprotein is substantially decreased (Gil et al., 2005;

Nanovskaya et al., 2008; Sun et al., 2006; Vahakangas & Myllynen, 2006). Therefore, drug

transfer at term may represent the highest exposure compared to earlier gestational ages

(Vahakangas & Myllynen, 2006). Other experimental methods may be useful to further

characterize transfer in first and second trimester tissue (Table 1.1). For example, knowing that

glyburide is a substrate for placental BCRP, kinetic parameters could be evaluated from

membrane vesicles prepared from early gestational tissue and compared to term tissue.

Comparing perfusion results to cord and maternal blood ratios at delivery also has

limitations. In vivo ratios obtained at delivery only represent one time point and it is not always

certain if the two circulations are at steady state (Chappuy et al., 2004). However, with a large

sample size and measurements at multiple times after the last dose, one can determine whether

the ratio is likely to be at steady state. A constant proportionality between the maternal and fetal

compartments occurs faster in vivo compared to the perfusion model. For example, for

bupivacaine, drug levels in maternal and cord blood equilibrate within minutes (Thomas et al.,

1976). In contrast, in the perfusion model, the F:M ratio equilibrates after 1 hour (Johnson et al.,

1995). For morphine, fetal and maternal concentrations in vivo were almost identical after 5 min

(Gerdin et al., 1990). In the perfusion model, equilibrium was not achieved until after 2 hours

(Figure 2-3) (Kopecky, 1999). The faster equilibrium obtained in vivo results from more rapid

distribution since the entire maternal blood volume is circulated by the heart in approximately

one minute. In the perfusion experiment, where the maternal reservoir is 250 mL, it would take

around 20 minutes to circulate this volume once. Furthermore, the villus surface area of the

Page 66: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

50

whole placenta in vivo at term (12 m2 for the whole placenta) is much larger than that of a single

cotyledon (20 cotyledons per placenta on average) (Cabezon et al., 1985; Vermeulen et al.,

1982).

Recommendations for Future Standardization of Placental Perfusion Experiments

Only twenty percent of therapeutic drugs that have been studied using the perfusion

model were available for a quantitative comparison to the in vivo data. A major factor was that

in vivo data were not available for almost thirty percent of the drugs. However, the majority of

drugs could not be compared because of factors associated with study design or insufficient data

presented in the publication. Many perfusion studies did not achieve steady state as stated by the

authors or determined by inspection of the figures. In some publications, it was not possible to

determine whether steady state was achieved. Since the perfusion model equilibrates slower than

the maternal-fetal compartments in vivo, it is critical to ensure that the perfusion experiment lasts

a sufficient length of time. Many studies also reported only data that were standardized to a

reference marker and only final calculations. These standardized calculations do not provide

insight into the drug distribution and thus absolute values should always be reported (Ala-Kokko

et al., 2000). Reporting the absolute concentrations and a corresponding graph would add

valuable information to the literature and enable more secondary analyses to be performed. The

pH difference between the maternal and fetal compartments can influence the extent of drug

transfer, especially for weak bases (Ueki et al., 2009). Many studies did not report the actual pH

of the perfusates during the experiment or whether they controlled the pH during the course of

the experiment. This information should also be reported in publications. Mathiesen et al.

(2010) have also proposed criteria to be included in perfusion publications and have

recommended that each laboratory publish a detailed methods paper so that all experimental

parameters can easily be obtained (Mathiesen et al., 2010).

Summary

After systematically reviewing all drugs that have been evaluated by both the placental

perfusion model and in vivo, it is evident that the dual perfusion of a single placental lobule is a

valid method to predict placental drug transfer at term. Data obtained from perfusion

experiments must take into consideration many factors that are known to influence drug transfer

including placental pharmacokinetic factors (active transport, passive diffusion, and metabolism)

Page 67: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

51

as well as physiochemical properties of the drug. Differential protein binding between the fetal

and maternal circulations is known to be a critical factor in the placental transfer of drugs (Hill &

Abramson, 1988). However, it is difficult and unpractical to exactly mimic physiological protein

binding during the perfusion experiment. Determining the protein binding in vitro using fetal

and maternal blood in parallel with perfusion experiments, and then modeling the results as

documented above would allow for more accurate predictions for placental drug transfer at

steady state. After modeling, results from placental perfusion experiments can be used to

accurately predict in vivo placental drug transfer and thus estimate fetal drug exposure. This will

allow conclusions to be reached regarding placental drug transfer for new drugs where there is

no in vivo fetal safety data available. When applied appropriately, the placental perfusion model

is an invaluable tool to help guide decisions regarding the benefits and risks of new medications

that may be required during pregnancy.

Page 68: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

52

Chapter 3

The transfer of 6-mercaptopurine in the dually perfused human placenta Hutson JR1,2, Lubetsky A1, Walfisch A1, Ballios BG2, Garcia-Bournissen F1,2, Koren G1,2

1Division of Clinical Pharmacology and Toxicology, Hospital for Sick Children, 555 University

Ave, Toronto, Ontario, M5G 1X8, Canada

2Institute of Medical Science, University of Toronto, 1 King’s College Circle, Toronto, Ontario,

M5S 1A8, Canada

This work has been published and reproduced with permissions: Hutson JR, Lubetsky A, Walfisch A, Ballios BG, Garcia-Bournissen F, Koren G. The transfer

of 6-mercaptopurine in the dually perfused human placenta. Reprod Toxicol 2011;32:349–353.

Page 69: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

53

The transfer of 6-mercaptopurine in the dually 3perfused human placenta

3.1 Abstract

The immunosuppressant azathioprine is increasingly being used in pregnancy. The

human placenta is considered a relative barrier to the major metabolite, 6-mercaptopurine (6-

MP), and likely explains the lack of proven teratogenicity in humans. The aim of this study was

to determine how the human placenta restricts 6-MP transfer using the human placental

perfusion model. After addition of 50ng/ml (n=4) and 500ng/ml (n=3) 6-MP into the maternal

circulation, there was a biphasic decline in its concentration and a delay in fetal circulation

appearance. Under equilibrative conditions, the fetal-to-maternal concentration ratio was >1.0 as

a result of ion trapping. Binding to placental tissue and maternal pharmacokinetic parameters are

the main factors that restrict placental transfer of 6-MP. Active transport is unlikely to play a

significant role and drug interactions involving, or polymorphisms in, placental drug efflux

transporters are not likely to put the fetus at risk of higher 6-MP exposure.

Page 70: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

54

3.2 Introduction

The immunosuppressant azathioprine (AZA) is increasingly being used in pregnancy for

the treatment of autoimmune diseases or for organ transplant patients (Goldstein et al., 2007).

Peak incidence of autoimmune diseases, including inflammatory bowel disease (IBD) and

systemic lupus erythematosus, occurs in women during the childbearing years (Cooper &

Stroehla, 2003) and successful pharmacologic therapy in these women has improved the

feasibility of pregnancy (Polifka & Friedman, 2002). Treatment with AZA is required into

pregnancy to prevent relapse of the disease or organ rejection; thus, treatment can also minimize

adverse fetal effects associated with the underlying maternal disease (Polifka & Friedman, 2002).

Several studies have investigated the safety of AZA in pregnancy. A review of 27 case

series and three large retrospective and prospective studies did not find an increased risk for

major malformations associated with the use of AZA in pregnancy (Cleary & Kallen, 2009;

Coelho et al., 2011; Francella et al., 2003; Goldstein et al., 2007; Polifka & Friedman, 2002).

However, one recent study reported a trend for increased malformations and an increased risk

specifically for ventricular/atrial septal defects after prenatal AZA exposure (Cleary & Kallen,

2009). An increased risk for prematurity and lower birth weight has also been reported, but this

may be related to the underlying maternal disease (Cleary & Kallen, 2009; Goldstein et al.,

2007). Although evidence supports the relative safety of AZA in pregnancy compared to the

untreated disease (Coelho et al., 2011), safety concerns still arise from its mechanism of action

and its toxicity during pregnancy in animal studies (Polifka & Friedman, 2002). Over eighty

percent of women with IBD have unwarranted concerns about medications, including AZA, in

pregnancy and fears surrounding medications are greater than the effect of IBD exacerbation

(Mountifield et al., 2010). Moreover, some physicians still choose not to administer AZA during

pregnancy or to discontinue treatment in the third trimester (Peyrin-Biroulet et al., 2011).

After oral administration, AZA is extensively (>80%) cleaved into 6-mercaptopurine (6-

MP) in the blood (Polifka & Friedman, 2002). 6-MP is a purine analogue, and after further

conversion into active metabolites, it can inhibit de novo purine ribonucleotide synthesis leading

to inhibition of cell proliferation. Furthermore, 6-MP can be cytotoxic after incorporation of its

metabolites into cellular DNA. Since 6-MP targets rapidly dividing cells, the developing fetus

would theoretically be expected to be sensitive to its cytotoxicity. However, even high dose 6-

Page 71: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

55

MP was not found to be useful as a single-agent medical abortifacient in early pregnancy,

leading the authors to suggest that this drug does not alter cell division in trophoblastic tissues

(Davis et al., 1999).

The observed safety of AZA in pregnancy likely results, in part, from the placenta

limiting fetal exposure to the main metabolite, 6-MP. Indeed, only 1-2% of maternal 6-MP

concentrations are found in cord blood and the human placenta is considered to be a relative

barrier to 6-MP and its metabolites (de Boer et al., 2006; Saarikoski & Seppala, 1973).

However, the mechanism of how the placenta restricts 6-MP transfer to the fetus is unknown. 6-

MP is a substrate for several drug transport proteins, including breast cancer resistance protein

(BCRP) (de Wolf et al., 2008). BCRP is highly expressed in the placenta and is localized to the

brush border (maternal-facing) membrane of the syncytiotrophoblast (Maliepaard et al., 2001).

The importance of BCRP and other placental transporters in limiting fetal exposure to various

drugs has been demonstrated in both human and animal models (Jonker et al., 2000; Pollex et al.,

2008). The objective of this study was to determine the mechanisms by which the placenta

restricts 6-MP transfer. Understanding these mechanisms will allow for prediction of fetuses that

may be at risk of higher exposure to 6-MP as a result of drug interactions with, or

polymorphisms in, placental drug efflux proteins.

3.3 Methods

Materials

6-methylmercaptopurine (6-MMP) and 6-thioxanthosine were obtained from MP

Biomedicals (Solon, Ohio). 6-thiouric acid was obtained from US Biological (Swampscott, MA)

and toluene from EMD Chemicals (Gibbstown, NJ). HPLC-grade methanol was ordered from

Fisher Scientific (Fair Lawn, NJ). All other chemicals were purchased from Sigma (St. Louis,

MO). The water used for all experimental procedures was obtained from a Milli-Q Advantage

A10 Ultrapure Water Purification System (Millipore, Billerica, MA).

Page 72: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

56

Placental perfusion

The dual perfusion of a placental lobule was previously described by Miller et al. (Miller

et al., 1985) and adapted in our laboratory (Derewlany et al., 1991; Pollex et al., 2008). Term

placentas were obtained immediately after elective caesarean sections from healthy mothers with

uncomplicated pregnancies from Mount Sinai or St. Michael’s Hospitals in Toronto, Ontario.

Research ethics board approval at each hospital was obtained and maternal consent was obtained

prior to the surgery. The placentas were transferred to the lab in heparinised ice-cold phosphate-

buffered saline where a vein/artery pair supplying a clearly identifiable cotyledon were chosen

for cannulation (Derewlany et al., 1991). Maternal and fetal circulations were established within

30 minutes of delivery.

The perfusate consisted of M199 tissue culture medium (Sigma, St. Louis, MO)

containing heparin (2000 U/L), glucose (1.0 g/L), kanamycin (100 mg/L), and 40,000 molecular-

weight dextran (maternal 7.5 g/L; fetal 30 mg/L). Antipyrine (1 mM) was added to the maternal

circulation as a marker of passive diffusion. Flow rates were maintained at 2 to 3 and 13 to 14

ml/min in the fetal and maternal circuits, respectively. The maternal perfusate was equilibrated

with 95% O2, 5% CO2 and the fetal with 95% N2, 5% CO2. Maternal and fetal circuits were

maintained at pH 7.4 and 7.35, respectively, by the addition of small volumes of sodium

bicarbonate and hydrochloric acid. This pH mimics the slightly more acidic fetal circulation

observed in vivo (Reynolds & Knott, 1989). The temperature of the circuits and the perfusion

chamber was kept at 37°C.

Each perfusion consisted of a 1 h closed pre-experimental control period, followed by a 3

h closed experimental period, and a final 1 h closed post-control period. The post-control period

was omitted from the perfusions where 6-MP was added to only the maternal circulation since

this allowed measurement of drug binding to placental tissue. The perfusates in both circulations

were replaced with fresh media prior to each of the 3 periods. During the pre- and post-control

periods, samples were taken every 15 min to analyze glucose and oxygen consumption, and

human chorionic gonadotropin (hCG) secretion as measures of tissue viability. These measures

were taken every 30 minutes during the experimental period. Tissue viability measures were

calculated as previously described (Pollex et al., 2008). Fetal reservoir volume and fetal

perfusion pressure were monitored as an indicator of tissue integrity. pH, pO2, and pCO2 were

Page 73: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

57

monitored using a blood gas analyzer (Radiometer ABL 725, Copenhagen, Denmark). During

the experimental period, 50 ng/ml or 500 ng/ml 6-MP was added to the maternal circulation only

or simultaneously to both the fetal and maternal circulations. 50 ng/ml is the maximum

concentration obtained after oral administration of AZA in adult renal transplant recipients

(Bergan et al., 1994; Zins et al., 1997) and similar doses are indicated for the treatment of IBD.

Samples were taken for measurement of 6-MP every 10 minutes for the first half-hour and every

half-hour following. The perfusion was terminated at any time if there was a >3 ml/hour loss in

fetal reservoir volume.

The expected fetal to maternal (F:M) ratio at steady state resulting from ion trapping –

given the change in pH between the fetal and maternal circulations - was calculated using a

modified form of the Henderson-Hasselbalch equation that incorporates the two pKa of 6-MP:

This equation was derived by assuming conservation of mass and rapid equilibration between the

neutral and protonated forms of 6-MP (6-MP, 6-MP+, and 6-MP++). Only the un-ionized form is

assumed to cross the membrane. Data are expressed as mean ± SEM unless otherwise indicated.

F:M ratios were compared using a two-tailed Student’s t-test.

Sample analysis

Perfusate samples were stored at -20°C until analysis. 6-MP and its metabolites (6-

MMP, 6-thioxanthosine, 6-thiouric acid, 6-thioguanine) were extracted using a method adapted

from Lennard and Singleton (Lennard & Singleton, 1992). 500 µl of perfusate was added to 500

µl of 6 mM DL-Dithiothreitol and 500 µl of 1.5 M H2SO4 in duplicate. Since 6-TG was not

detected in any of the perfusions, it was further used as an internal standard (100 ng/ml). One of

each sample was heated at 100°C for 1 hour for acid hydrolysis of the thionucleotide metabolites

back into the parent thiopurine and for extraction of 6-methylmercaptopurine (Lennard &

Singleton, 1992). Samples not being heated were kept on ice. After cooling on ice, 500 µl of 3.8

M NaOH was added followed by 6 ml of isoamyl alcohol (170 mM) and phenylmercury acetate

(1.3 mM) in toluene and placed in an automatic rotator for 15 min. The mixture was centrifuged

at 2500×g for 15 min at room temperature and the toluene layer was added to a new tube. 200 µl

FM

=1+10 pKa 2 − pHF( ) +10 pKa1 + pKa 2 −2pHF( )

1+10 pKa 2 − pHM( ) +10 pKa1 + pKa 2 −2pHM( )

Page 74: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

58

of 0.1 M HCl was added and vortexed for 2 × 1 minute. 70 µl of the aqueous phase was injected

onto a Atlantis T3 Column (3µm, 4.6×150mm) protected with a guard cartidge (3 µm, 4.6×20

mm)(Waters, Milford, MA). The mobile phase consisted of acetonitrile, methanol, and KH2PO4

(0.02 M; pH 2.25) (3:1:96, v/v/v) (Hawwa, Millership, Collier, & McElnay, 2009) and was

filtered through a 0.45 µm filter. The flow rate was 1.0 ml/min.

The HPLC system (Shimadzu, Columbia, MD) consisted of a solvent delivery pump (LC-

10AT), auto injector (SIL-10A), degasser (DGU-14A), controller (CBM-20), and dual

wavelength UV detector (SPD-20A). 6-MP was monitored at 313 nm; 6-MMP at 303 nm; and

6-thioguanine, 6-thioxanthosine, and 6-thiouric acid at 342 nm. The chromatograms were

acquired and analyzed using Shimadzu Class-VP Software, Version 7.4 SP2. The limit of

quantification was 2 ng/ml for 6-MP and 6-thioguanine and 5 ng/ml for 6-MMP, 6-thiouric acid,

and 6-thioxanthosine.

3.4 Results

A total of 16 cotyledons from different placentae were perfused with 6-MP and the

physical parameters for the perfusions are given in Table 3-1. The mass of the perfused

cotyledons ranged from 10.22 to 32.0 g. Throughout the experiments, measures of placental

viability, integrity, and function remained within normal ranges and were not significantly

different between the control and experimental phases (Table 3-1). There was no significant

difference between experimental and control periods in the fetal arterial pressures. The rate of

antipyrine appearance in the fetal circulation during the experimental period was equal to the rate

of disappearance from the maternal circulation with values of 0.030 ± 0.002 and 0.032 ± 0.003

µmol/g per minute, respectively.

Page 75: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

59

Table 3-1. Placental viability parameters and metabolic capacity throughout the perfusions

(n=16 cotyledons from independent placentae) (mean ± SEM)

Viability parameter Pre-control Experiment Post-control Fetal arterial inflow pressure (mm Hg) 40.36 ± 1.11 40.71 ± 1.08 42.91 ± 1.81 hCG production (mIU/g per min) 57.52 ± 10.12 39.62 ± 6.83 40.86 ± 13.07 Oxygen (µmol O2/g per min) Transfer 0.01 ± 0.00 0.01 ± 0.00 0.02 ± 0.00 Delivery 0.48 ± 0.05 0.48 ± 0.04 0.45 ± 0.07 Consumption 0.22 ± 0.02 0.22 ± 0.02 0.23 ± 0.04 Glucose consumption (µmol/g per min) 0.42 ± 0.05 0.36 ± 0.04 0.38 ± 0.06

After addition of 50 ng/ml 6-MP into the maternal circulation only, there was a biphasic

decline in its concentration from this compartment (Figure 3-1A); an initial rapid decline during

the first 30 minutes, followed by a slower decline in the final 150 minutes. There was a delay in

the transfer of 6-MP in the fetal circulation as measurable concentrations of 6-MP appeared only

after 30 minutes. This delay suggests uptake and retention of 6-MP by the placental tissue.

After 3 hours, the mean fetal concentration of 6-MP was 8.8 ± 1.71 ng/ml and the mean fetal to

maternal (F:M) concentration ratio was 0.44 ± 0.06. After addition of the higher 6-MP

concentration (500 ng/ml), there was a similar biphasic decline from the maternal circulation

(Figure 3-1B). 6-MP was detected in all three perfusions at 20 minutes, which is more rapid

compared to the perfusions at the lower concentration. After 3 hours, the mean fetal

concentration of 6-MP was 91.48 ± 7.50 ng/ml and the mean F:M ratio was 0.57 ± 0.04.

When equal concentrations (50 ng/ml or 500 ng/ml) were added to both fetal and

maternal circulations at the start of the experimental period, the fetal to maternal (F:M) ratio

increased over time (Figure 3-2). The F:M ratios of 6-MP at the end of the 180 minute

experimental period were 1.21 ± 0.08 and 1.19 ± 0.04 for 50 ng/ml and 500 ng/ml, respectively.

The F:M ratios increased significantly from time zero to 180 minutes at both the 50 and 500

ng/ml concentrations (p=0.007, p=0.02, respectively). 6-MP is a basic drug with two pKa’s

(pKa1=7.77, pKa2=11.17) (O'Neil et al., 2010), and basic drugs are more ionized in the fetal

circulation since it is slightly more acidic than the maternal (7.35 vs. 7.40, respectively)

(Reynolds & Knott, 1989). The expected F:M ratio at equilibrium based on ion trapping alone

was calculated to be 1.22.

Page 76: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

60

After perfusion with 50 ng/ml 6-MP, the metabolites including 6-MMP, 6-thioguanine, 6-

thioxanthosine, 6-thioinosine, and 6-thiouric acid (and the related thionucleotide metabolites)

were not detected in the fetal nor the maternal circulations. At the ten-fold higher 6-MP

concentration of 500 ng/ml, only 6-MMP was detected with a mean concentration of 11.75 ±

2.02 and 8.65 ± 5.97 ng/ml in the fetal and maternal circulations, respectively, at 180 minutes.

Two of the seven perfusions at 500 ng/ml had 6-MMP below the limit of detection. The tissue-

to-maternal perfusate ratio was 0.49 ± 0.08 at the end of 180 minutes of perfusion (n=7). The

tissue-to-maternal perfusate ratio was not significantly different between the 50 ng/ml and the

500 ng/ml perfusions.

Page 77: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

61

Figure 3-1. Maternal-to-fetal transfer of 6-mercaptopurine after dual perfusion of a single

placental lobule. 6-mercaptopurine was added to the maternal circulation at A) 50 ng/ml (n=4)

or B) 500 ng/ml (n=3) and transfer was determined for a period of 180 minutes. Data are shown

as mean values ± SEM at each time point.

0"50"100"150"200"250"300"350"400"450"

0" 50" 100" 150" 200"

6MP$(ng/ml)$

Time$(min)$

Fetal"Maternal"

0"5"10"15"20"25"30"35"40"45"50"

0" 50" 100" 150" 200"

6MP$(ng/ml)$

Time$(min)$

Fetal"Maternal"

A)

B)

Page 78: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

62

Figure 3-2 Concentrations of 6-MP in the fetal and maternal circulations after perfusion of a

single placental lobule under equilibrative concentrations of A) 50 ng/ml (n=5) and B) 500 ng/ml

(n=4). C) The fetal to maternal (F:M) concentration ratios for the perfusion under equilibrative

conditions. Data are shown as mean values ± SEM at each time point.

A)

B)

C)

Page 79: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

63

3.5 Discussion

The human placenta is considered a relative barrier to 6-MP and its metabolites (de Boer

et al., 2005; de Boer et al., 2006; Saarikoski & Seppala, 1973) and our study demonstrates that

placental binding and not active transport limits transfer. Limited fetal exposure to 6-MP may in

part explain why studies have not shown conclusive evidence that 6-MP is teratogenic (Cleary &

Kallen, 2009; Coelho et al., 2011; Goldstein et al., 2007). Understanding the mechanism of this

limited transfer will allow for prediction of fetuses that may be at risk for higher exposure or for

prediction of potential drug interactions.

After introduction of a clinically relevant concentration into the maternal circulation,

there was a rapid decline in maternal levels, however, there was a delay in appearance of 6-MP

in the fetal circulation. This delay can be attributed to the uptake and binding by the placental

tissue. After addition of a 10-fold higher concentration, there was also a delay in the appearance

of 6-MP in the fetal circulation. The appearance of 6-MP in the fetal circulation was more rapid

for 500 ng/ml, suggesting saturation of tissue binding sites. Although binding to the tissue

delayed placental transfer, the tissue to maternal perfusate ratio at the end of the perfusion was

low. This low ratio is similar to the placenta to maternal plasma ratio observed in vivo after

therapeutic abortions in the ninth to fifteenth week of gestation (Saarikoski & Seppala, 1973).

The human placenta does have a capacity to sequester 6-MP and thus decrease transfer to the

fetus (Saarikoski & Seppala, 1973).

To determine if there is active efflux of 6-MP into the maternal circulation, we

introduced the drug into both fetal and maternal circulations as this methodology has been used

previously to identify placental efflux (Kraemer et al., 2006; Pollex et al., 2008). After 180

minutes of perfusion, the F:M ratio was above 1.0. Under equilibrative conditions at 50 ng/ml

and 500 ng/ml, the term placenta was not able to concentrate 6-MP into the maternal circulation.

This suggests that efflux of 6-MP into the maternal circulation is not responsible for or plays a

very minor role in the limited placental transfer of 6-MP. The F:M ratio at the end of the 3 hour

experimental period at both 50 ng/ml and 500 ng/ml was not significantly different from the

calculated F:M ratio based on ion trapping. This supports ion trapping as the mechanism for the

observed F:M ratio at steady state.

Page 80: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

64

During pregnancy, nucleoside transporters transfer nucleotides from the mother to the

fetus. Equilibrative nucleoside transporters (ENT) 1, 2 and concentrative nucleoside transporter

2 (CNT2) have been localized to the brush-border membrane of the syncytiotrophoblast

(Govindarajan et al., 2007). 6-MP is a substrate for the ENTs (Nagai et al., 2007), however,

these are energy-independent transporters and are only capable of equilibrative (not

concentrative) transport (Ganapathy et al., 2000) and are not likely to play a role in the transfer

under equilibrative conditions. Only mRNA expression (not protein) of CNT2 was demonstrated

in the human placenta (Govindarajan et al., 2007), thus it is unlikely that CNTs would play a

major role in the F:M distribution of 6-MP. 6-MP is also a substrate for multidrug resistance

protein 5 (MRP5) (Wielinga et al., 2002; Wijnholds et al., 2000). MRP5 is localized to the

basolateral membrane of the syncytiotrophoblast and also on fetal endothelial cells (Meyer Zu

Schwabedissen et al., 2005) and could thus transport 6-MP to the fetal circulation. However, the

expression of MRP5 decreases in term compared to pre-term placentas (Meyer Zu

Schwabedissen et al., 2005), thus it is unlikely that MRP5 plays an important role in our

perfusion model.

Our results show that active efflux of 6-MP is not the mechanism responsible for limited

placental transfer. After introducing 6-MP in the maternal circulation only, the fetal 6-MP

concentration was approximately half of the maternal. This is higher than that observed in vivo

where 6-MP transfer to the fetus is limited (de Boer et al., 2006; Saarikoski & Seppala, 1973).

The placental perfusion model is a useful model for determining transport mechanisms, but does

not always correlate well with in vivo data if non-placental pharmacokinetic variables are

important factors in fetal exposure (Hutson et al., 2011; Myllynen et al., 2005). Two non-

placental pharmacokinetic factors for 6-MP, namely distribution and clearance, are likely to be

important in the limited transfer in vivo. First, distribution of 6-MP includes binding to placental

tissue as demonstrated in our experimental model, but also 6-MP is rapidly converted into

intracellular metabolites. Because of the intracellular location, the metabolites are not free to

cross the placenta in vivo (de Boer et al., 2005). Second, maternal metabolism of 6-MP and its

corresponding short t1/2 of 1 hour (Chan et al., 1987; Ohlman et al., 1994) will limit fetal

exposure by decreasing the maternal area under the curve (AUC). Saarikoski & Seppala (1973)

suggested that maternal metabolism of 6-MP was important in limiting exposure as they

observed that 6-thiouric acid, an inactive metabolite, was the main metabolite transferred to the

Page 81: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

65

fetus. Metabolism by the placenta does not likely play an important role in limited fetal 6-MP

exposure as in our perfusions 6-MMP was the only metabolite detected and only after perfusion

with 500 ng/ml of 6-MP. In vivo, 6-MMP was not detected in RBC from cord blood in 3 infants;

however, the limit of detection in this study was not provided (de Boer et al., 2006).

Furthermore, 6-MMP is not stable in blood samples and may have degraded before the samples

were analyzed (de Graaf et al., 2010).

A limitation of our placental perfusion model is that it uses term placentas.

Generalizations to earlier gestational ages are difficult, including extrapolations to the embryonic

period where the placental structure varies. Our results suggest that the distribution of 6-MP and

its short maternal t1/2 are the key factors limiting transfer. These factors are unlikely to change

in the different stages of pregnancy and monitoring maternal plasma for 6-MP and its

metabolites would prevent elevated drug concentrations during pregnancy (de Boer et al., 2006).

Another limitation to our study is that there is no information available regarding plasma

concentrations of 6-MP during pregnancy. The clinically relevant concentration used in our

study was determined in adults taking AZA at similar doses to those administered during

pregnancy. Although the Km for BCRP and 6-MP has not been published, a linear rate of

transfer was observed at higher concentrations than those utilized in our study (de Wolf et al.,

2008).

3.6 Conclusions

Pharmacological therapy with AZA during pregnancy is needed to prevent relapse of

autoimmune diseases or rejection of renal transplants. Studies looking at the safety of AZA and

its major metabolite, 6-MP, have demonstrated that the benefit of treating the disease outweighs

the risk of maintaining treatment during the pregnancy (Kwan & Mahadevan, 2010). Additional

large epidemiological studies and meta-analyses are needed to provide conclusive evidence

regarding the teratogenicity of 6-MP (Coelho et al., 2011), especially to account for the

pharmacogenetic variability in thiopurine S-methyltransferase (TPMT) (Sahasranaman, Howard,

& Roy, 2008). The placenta acts as a relative barrier to 6-MP and its metabolites (de Boer et al.,

2005; de Boer et al., 2006; Saarikoski & Seppala, 1973), and our results suggest that maternal

pharmacokinetic factors and placental binding limit transfer. These findings support monitoring

Page 82: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

66

maternal plasma for elevated levels of 6-MP and metabolites is useful in preventing potential

fetal toxicity (de Boer et al., 2006). Our results also suggest that active transport of 6-MP into

the maternal circulation is not an important mechanism. Therefore, polymorphisms or drug

interactions involving active drug transport proteins in the placenta are unlikely to leave a fetus

more vulnerable to 6-MP exposure.

Page 83: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

67

Chapter 4

The fetal safety of thiopurines for the treatment of inflammatory bowel disease in pregnancy: a meta-analysis

Janine R Hutson MSc1,2, Jeremy N Matlow BMSc1,3, Myla Moretti MSc1, Gideon Koren MD1,2,3

1Division of Clinical Pharmacology and Toxicology, Hospital for Sick Children, 555 University

Ave, Toronto, Ontario, M5G 1X8, Canada

2Institute of Medical Science, University of Toronto, 1 King’s College Circle, Toronto, Ontario,

M5S 1A8, Canada

3Department of Pharmacology, University of Toronto, 1 King’s College Circle, Toronto, Ontario,

M5S 1A8, Canada

This work has been published and reproduced with permissions: Hutson JR, Matlow JN, Moretti M, Koren G. The fetal safety of thiopurines for the treatment of

inflammatory bowel disease in pregnancy: a meta-analysis. J Obstet Gynecol 2013;33(1):1-8.

Page 84: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

68

The fetal safety of thiopurines for the treatment of 4inflammatory bowel disease in pregnancy: a meta-analysis

4.1 Abstract

Maintaining remission of inflammatory bowel disease (IBD) during pregnancy is critical

for positive pregnancy outcomes. Conflicting data exist regarding the association between

thiopurine use for IBD treatment in pregnancy and adverse pregnancy outcomes and this meta-

analysis aims to clarify this association. A meta-analysis was performed of all original human

studies reporting outcomes in pregnancy in patients receiving thiopurines. Nine studies satisfied

the inclusion criteria and a total of 494 patients with IBD and 2,782 disease-matched controls

were reported. When compared to healthy women, those receiving thiopurines had an increased

risk for congenital malformations (RR 1.45; 95% CI 1.07 to 1.96; p = 0.02); however, when

compared to IBD controls, there was no increased risk (RR 1.37; 95% CI 0.92 to 2.05; p = 0.1).

These data provide support for thiopurines having a minimal risk, if any, to the fetus.

Page 85: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

69

4.2 Introduction

The thiopurines azathioprine (AZA) and 6-mercaptopurine (6-MP) are effective

treatments for inflammatory bowel disease (IBD), leading to steroid-free, clinical and endoscopic

remission (Ardizzone et al., 2006; Mowat et al., 2011; Timmer et al., 2007). Peak incidence of

IBD occurs in women of childbearing age (Cooper & Stroehla, 2003) and thiopurines are

increasingly being used in pregnancy for their treatment (Goldstein et al., 2007). Maintaining

disease remission during pregnancy is critical for positive fetal outcome (Coelho et al., 2011;

Kwan & Mahadevan, 2010).

Animal studies have demonstrated that at high doses of thiopurines, there is an increased

risk of congenital malformations (reviewed in Polifka & Friedman, 2002). However, doses used

in human treatment regimens are below the no observable adverse effect level (NOAEL)

reported for mice/rats (Coelho et al., 2011; Platzek & Bochert, 1996). In humans, few studies

have evaluated the safety of thiopurines in pregnancy and are limited by their statistical power

(Coelho et al., 2011). Although there is evidence to support that the benefit of treatment with

thiopurines in IBD patients during pregnancy likely outweighs any fetal risk, safety concerns still

arise as a result of their cytotoxic mechanism of action and their toxicity during pregnancy in

animal studies. Over eighty percent of women with IBD have unwarranted concerns about

medications, including AZA, in pregnancy and fears surrounding medications are greater than

the effect of IBD exacerbation (Mountifield et al., 2010). Moreover, some physicians still

choose not to prescribe AZA during pregnancy or to discontinue treatment in the third trimester

(Peyrin-Biroulet et al., 2011). The present meta-analysis aims to clarify the potential effects of

thiopurines on fetal outcome in women with IBD and to compare outcomes from studies using a

general/healthy population compared to an IBD control group.

4.3 Methods

Study Selection

A literature search was performed using MEDLINE, EMBASE, and EMBASE Classic

for all articles from inception to March 2011 reporting pregnancy outcomes in women who took

thiopurines at any time during their pregnancy. The following search terms were used:

Page 86: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

70

“azathioprine OR 6-mercaptopurine OR inflammatory bowel disease” AND “pregnancy” and the

results were limited to human studies. Web of Science Database of Abstracts, Cochrane Central

Register of Controlled Trials, Reprotox (an information system from the Reproductive

Toxicology Centre), International Pharmaceutical Abstracts, and articles from the references of

selected reviews were also consulted for additional articles.

Data Extraction

Data extraction was done independently by JRH and JNM. The following information

was extracted from each of the eligible articles: author, year of publication, journal, study design,

timeline and location of study, overlap with other studies, patient characteristics, definition of

malformations used in study, and data for study outcomes. In the case of a disagreement, a third

party (MM) served as a tie-breaker.

Inclusion and Exclusion Criteria

Studies were included if they met the following criteria: human subjects; either case-

controlled or cohort design with a control group; exposure to either AZA or 6-MP during any

stage of pregnancy; and at least one outcome of interest reported. Outcomes of interest included

incidence of congenital abnormalities, prematurity (<37 weeks gestation), birth weight, low birth

weight (LBW) (<2,500 g), and spontaneous abortion (SA). There were no language restrictions.

We excluded reviews, letters to the editor, animal studies, studies with no or inappropriate

control group, and studies with fewer than 10 subjects in either the control or experimental

group. If more than one paper had overlap in the study cohort, the study with the larger sample

size was chosen for inclusion.

Statistical Analysis

Data were combined using a random effects model. Analysis of dichotomous variables

(malformations, prematurity, and LBW) was performed by estimating relative risk (RR) values

and the continuous variable (birth weight) was analyzed using the weighted mean difference

(WMD). Both RR values and WMDs were reported with 95% confidence intervals (CIs) and

were calculated using Review Manager (version 5.0, Copenhagen: Cochrane Collaboration).

Heterogeneity was assessed by the χ2 and I2 tests. I2 refers to the percentage of variability that is

Page 87: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

71

due to heterogeneity. While the Q test (χ2) determines the presence or absence of heterogeneity,

the I2 test allows for determination of the extent of heterogeneity.

4.4 Results

Eligible Studies

A total of 533 studies were identified by the initial search and another four unique articles

were identified using Reprotox, two abstracts through Web of Science Database of Abstracts,

and two articles from hand-searching reference lists. After inspection of the title and abstract, 45

papers were read in entirety. Eight peer-reviewed publications (Cleary & Kallen, 2009; Coelho

et al., 2011; Francella et al., 2003; Goldstein et al., 2007; Langagergaard et al., 2007; Norgard et

al., 2007a; Shim, Eslick, Simring, Murray, & Weltman, 2011; Tennenbaum et al., 1999) and one

abstract (Dejaco et al., 2005) met the inclusion criteria and were therefore included in the meta-

analysis (Table 4.1). Two additional studies that met the inclusion criteria were excluded since

their study cohort was believed to overlap with other included studies; Nørgård et al. (2003) had

overlap with both Nørgård et al. (2007) and Langagergaard et al. (2007); Angelberger et al.

(2011) had overlap with Dejaco et al. (2005). Five of the included studies compared women with

IBD receiving thiopurines to IBD controls not receiving thiopurines (Coelho et al., 2011;

Francella et al., 2003; Norgard et al., 2007a; Shim et al., 2011; Tennenbaum et al., 1999), where

IBD controls were those with any previous diagnosis of IBD. Two studies compared women

receiving thiopurines for any indication to healthy or general population controls (Goldstein et

al., 2007; Langagergaard et al., 2007). Two studies had both comparisons available (Cleary &

Kallen, 2009; Dejaco et al., 2005). Although Nørgård et al. (2007) and Langagergaard et al.

(2007) were believed to have overlap between their study cohorts, they were both included since

the former was used for the IBD group comparison and the latter was used for the any indication

group comparison. Independent analyses were performed comparing thiopurine-exposed to IBD

and healthy controls, thus both studies were included. Since disease-matched and healthy

controls were mostly from different studies and thus populations, there were not directly

compared.

Page 88: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

72

Table 4-1 Characteristics of studies included in the meta-analysis

First Author Year Study Design

Thiopurine Thiopurine Exposed Patients

(n)

Controls (n)

IBD Any Indication*

IBD Disease-Matched

Healthy/ General

Population

Tennenbaum

1999

rc

AZA

12

-

137

-

Francella 2003 rc 6-MP 32 - 94 - Dejaco 2005 pc AZA or 6-MP 26 26 27 310,323 Norgard 2007 rc AZA or 6-MP 26 - 628 - Langagergaard 2007 rc AZA or 6-MP - 65 - 1,274 Goldstein 2007 pc AZA - 189 - 230 Cleary 2009 rc AZA 324 481 1,739 1,180,969 Coelho 2011 p/rc AZA or 6-MP 55 - 83 - Shim 2011 rc AZA or 6-MP 19 - 74 - Total 494 761 2,782 1,492,796

*Exposed patients in the IBD group are also included in the any indication group where applicable. rc, retrospective cohort; pc, prospective cohort; IBD, inflammatory bowel disease; AZA, azathioprine; 6-MP, 6-mercaptopurine

Page 89: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

73

Malformation Rates

Six studies reported on the incidence of congenital malformations in 482 patients with

IBD taking a thiopurine and 2645 IBD controls (Cleary & Kallen, 2009; Coelho et al., 2011;

Dejaco et al., 2005; Francella et al., 2003; Norgard et al., 2007a; Shim et al., 2011) (Table 4.2).

No significant difference was observed (p = 0.12) (Table 4.2; Figure 4.1). A significant

difference was found in the incidence of congenital malformations in IBD patients receiving

thiopurines versus controls not receiving other medications for IBD (Coelho et al., 2011;

Norgard et al., 2007a) (RR 2.67; 95% CI 1.07 to 6.63; p = 0.03), but not versus controls

receiving other pharmacologic treatments for IBD (Cleary & Kallen, 2009; Coelho et al., 2011)

(p = 0.55). Other medications for treatment of IBD included systemic corticosteroids, 5-

aminosalicylic acid, antitumor necrosis factor alpha inhibitors, and other immunosuppressants

(Cleary & Kallen, 2009; Coelho et al., 2011). There was also a significant difference observed

between women receiving thiopurines for any indication versus healthy or general population

controls (Cleary & Kallen, 2009; Dejaco et al., 2005; Goldstein et al., 2007; Langagergaard et

al., 2007) (RR 1.45; 95% CI 1.07 to 1.96; p = 0.02). The specific malformations reported in the

exposed infants in seven of the nine included studies are reported in Table 4.3. The other two

studies did not report on specific malformations (Cleary & Kallen, 2009; Dejaco et al., 2005).

Page 90: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

74

Table 4-2 Pregnancy outcomes in thiopurine-exposed pregnancies versus controls

Outcome of Interest No of Studies

Exposed Patients (n)

Controls (n)

RR (95% CI) p Value HG I2 Value

IBD vs Disease-Matched Controls*

Congenital Abnormalities 6 482 2,645 1.37 (0.92, 2.05) 0.12 0% Prematurity 6 470 2,589 1.67 (1.17, 2.37) 0.004 28% LBW 6 467 2,582 1.30 (0.92, 1.83) 0.13 0% SA 2 51 269 0.48 (0.21, 1.10) 0.08 0% IBD vs Disease-Matched Controls ON other medications

Congenital Abnormalities 2 379 1,795 1.15 (0.73, 1.81) 0.55 0% Prematurity 2 373 1739 1.49 (1.13, 1.96) 0.005 0% LBW 2 370 1,732 1.36 (0.95, 1.94) 0.09 0% IBD vs Disease-Matched Controls NOT on other medications

Congenital Abnormalities 2 81 655 2.67 (1.07, 6.63) 0.03 0% Prematurity 2 75 655 2.49 (0.92, 6.75) 0.07 57% LBW 2 75 655 2.06 (0.56, 7.52) 0.28 0% Exposed vs Healthy or General Population Controls**

Congenital Abnormalities 4 743 1,492,735 1.45 (1.07, 1.96) 0.02 0% Prematurity 3 688 1,146,056 3.97 (2.65, 5.94) <0.001 62% LBW 3 667 1,142,385 3.13 (2.44, 4.02) <0.001 0% *Disease-matched controls were on either other pharmacologic treatments for IBD, on no medication, or medication use was not specified **Exposed are women taking thiopurines for any indication (including IBD) during pregnancy HG, heterogeneity; LBW, low birth weight (<2500 g); RR, relative risk ratio; SA, spontaneous abortion; IBD, inflammatory bowel disease Significant results (p < 0.05) are shown in bold

Page 91: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

75

Table 4-3 Specific malformations and other drug use reported in the included studies First Author Year Malformations from thiopurine-exposed

group Other drug use in thiopurine-exposed group

Tennenbaum 1999 Pyloric stenosis Hip dislocation

Not clearly reported

Francella 2003 Hydrocephalus

Not clearly reported

Dejaco 2005 Not specified Similar use of prednisolone /5-ASA in exposed and control groups

Norgard 2007 Cataracta congential Encephalocele occipitalis Malformation of sternocleidomastoideus Malformation congenitae aliae cutis

Other drugs used concurrently, but not specified

Langagergaard 2007 Stenosis of pulmonal valve Hydronephrosis Cyst at the ductus choledochus Occipital encephalocele Malformation of sternocleidomastoideus Asymmetrical face

All malformations in exposed-group were in women on multiple medications

Goldstein 2007 Hypospadias Port wine stain Cleft lip Dysmorphic features with ASD Hydrocephalus Hydrocephalus

39% of women taking AZA were also taking a potential teratogen

Cleary 2009 30 malformations, not all specified 9 VSD/ASD (all on many other meds) 2 hypospadias

Women in exposed group at higher risk for taking other medications including drugs for gastric ulcer, loperamide, antihypertensives, steroids,opioids, sedatives, antidepressants, immunosuppressants, heparins, antiplatelet agents

Coelho 2011 Bilateral cataract Cervical angioma

Exposed group also likely receiving 5-ASA, steroids, and anti-TNF.

Shim 2011 Congenital hyperplastic heart Exposed group more likely also receiving 5-ASA and corticosteroids

5-ASA, 5-aminosalicylic acid; anti-TNF, antitumor necrosis factor alpha

Page 92: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

76

A)

B)

C)

D)

Figure 4-1 Overall effect of the incidence of congenital abnormalities after in utero

exposure to thiopurines A) In patients with IBD receiving thiopurines compared to disease-

matched controls B) In patients with IBD receiving thiopurines compared to disease-

matched controls on other medications only C) In patients with IBD receiving thiopurines

compared to disease-matched controls not receiving other medications D) In patients

receiving thiopurines for any indications compared to healthy or general population

controls.

Page 93: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

77

Prematurity

Six studies reported on the incidence of premature birth (<37 weeks gestation) (Cleary &

Kallen, 2009; Coelho et al., 2011; Dejaco et al., 2005; Francella et al., 2003; Norgard et al.,

2007a; Shim et al., 2011). Patients with IBD receiving thiopurines were more likely to have

premature infants compared to IBD controls (RR 1.67; 95% CI 1.17 to 2.37; p = 0.004) and also

compared to IBD controls that received other IBD therapeutics (Cleary & Kallen, 2009; Coelho

et al., 2011) (RR 1.49; 95% CI 1.13 to 1.96; p = 0.005) (Figure 4-2). There was a trend towards

having an increased odds for prematurity in exposed IBD patients compared to IBD controls that

were not receiving other medications (Coelho et al., 2011; Norgard et al., 2007a) (RR 2.49; 95%

CI 0.92 to 6.75; p = 0.07). Women receiving thiopurines for any indication were also more

likely to have premature infants compared to healthy or general population controls (Cleary &

Kallen, 2009; Goldstein et al., 2007; Langagergaard et al., 2007) (RR 3.97; 95% CI 2.65 to 5.94;

p < 0.001).

Page 94: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

78

Birth weight

A)

B)

Figure 4-2 Overall effect of the incidence of prematurity (< 37 weeks gestation) after in utero

exposure to thiopurines A) In patients with IBD receiving thiopurines compared to disease-matched

controls B) In patients receiving thiopurines for any indications compared to healthy or general

population controls.

Page 95: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

79

The odds ratio for LBW (<2500 g) was not significant based on six studies (Cleary &

Kallen, 2009; Coelho et al., 2011; Dejaco et al., 2005; Francella et al., 2003; Norgard et al.,

2007a; Shim et al., 2011) comparing women with IBD receiving thiopurines and all IBD controls

(p = 0.13) (Figure 4-3). Likewise, the mean difference in birth weight was also not significant (p

= 0.07) based on three studies (Coelho et al., 2011; Norgard et al., 2007a; Shim et al., 2011) with

the same comparison group (Figure 4-4). Similar findings for LBW were observed when

comparing specifically to IBD controls either receiving (Cleary & Kallen, 2009; Coelho et al.,

2011) (p = 0.09) or not receiving (Coelho et al., 2011; Norgard et al., 2007a) (p = 0.28) other

therapeutics for IBD. There was, however, a lower mean birth weight in women with IBD

receiving thiopurines compared to IBD controls not receiving other medications (Coelho et al.,

2011; Norgard et al., 2007a) (-238.8 g; 95% CI -443.8 to -33.8; p = 0.02). A significant increase

was seen in the incidence of LBW in patients receiving thiopurines for any indication compared

to healthy or general population controls (Cleary & Kallen, 2009; Goldstein et al., 2007;

Langagergaard et al., 2007) (RR 3.13; 95% CI 2.44 to 4.02; p < 0.001). The mean difference in

birth weight for this comparison was significantly decreased (Goldstein et al., 2007;

Langagergaard et al., 2007) (-363.3 g; 95% CI -608.9 to -117.7; p = 0.004).

Page 96: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

80

Figure 4-3 Overall effect of the incidence of low birth weight (< 2500 g) after in utero exposure

to thiopurines A) In patients with IBD receiving thiopurines compared to disease-matched

controls B) In patients receiving thiopurines for any indications compared to healthy or general

population controls.

A)

B)

Page 97: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

81

Figure 4-4. Overall effect of in utero exposure to thiopurines on mean birth weight A) In

patients with IBD receiving thiopurines compared to disease-matched controls B) In patients

receiving thiopurines for any indications compared to healthy or general population controls.

Spontaneous Abortions

Only two studies reported on the incidence of SA in 51 women receiving thiopurines for

IBD compared to 269 IBD controls (Francella et al., 2003; Tennenbaum et al., 1999). No

significant difference was found, although there was a trend for a decreased rate of SA (p =

0.08). Shim et al. (2011) also reported on the incidence of SA in the same comparison group and

observed no SA in 19 thiopurine-exposed women or 74 controls.

B)

A)

Page 98: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

82

4.5 Discussion

Compared to IBD controls, there was no increased risk for congenital malformations in

women receiving thiopurines for IBD. Subgroup analysis of the IBD control groups revealed

that no significant increase was observed when the comparison group comprised of women with

IBD receiving other medications. On the other hand there was a significant increase in risk when

the comparison group was made of women not receiving any other medication for their IBD or

taken from the healthy, general population. However, only two studies were included in each

IBD subgroup analysis. Medication use during pregnancy has been suggested to be a surrogate

for disease activity during pregnancy (Dominitz et al., 2002) and Shim et al. (2011) reported a

significantly higher rate of hospitalization for an IBD flare during pregnancy compared to

controls. The only two comparisons that showed a significant increase for abnormalities used

comparison groups that were either likely to be in remission since they were not receiving any

therapeutics for IBD, or were healthy controls. This supports other reports of disease activity

being associated with adverse pregnancy outcomes (Baiocco & Korelitz, 1984; Dominitz et al.,

2002; Fedorkow et al., 1989; Woolfson et al., 1990). Proper comparison groups and disease

activity will need to be maintained in future studies. Since women with IBD are considered

high-risk (Cornish et al., 2007), some malformations may be more likely to be detected

prenatally due to more intensive monitoring. Furthermore, the largest study in the pooled

analysis by Cleary and Källén (2009) included both major and minor abnormalities and this may

result in overestimation of the true risk (Cornish et al., 2007).

In some cases, studies also reported that women receiving thiopurines were more likely to

be receiving a potential teratogen compared to the control group, including systemic

corticosteroids (Cleary & Kallen, 2009; Goldstein et al., 2007; Shim et al., 2011). Women

receiving thiopurines for IBD treatment were often on other pharmacologic treatment, including

5-aminosalicylic acid (5-ASA), corticosteroids, and antitumor necrosis factor alpha (anti-TNF).

A meta-analysis evaluating the safety of 5-ASA drugs in pregnancy did not find a significant

association with congenital abnormalities (Rahimi et al., 2008). Systemic corticosteroid use in

pregnancy is associated with a small increased risk for oral clefts, growth restriction, and

prematurity (Park-Wyllie et al., 2000). To date, limited evidence suggests a low risk, if any, with

the use of anti-TNF in pregnancy, but further studies are needed (Osting & Carter, 2010).

Page 99: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

83

Seven of the nine studies provided information on the specific malformations in the

thiopurine-exposed cohort (Table 4.3). Out of the 22 specific malformations reported in the

seven studies, there was no pattern observed. A previous review of 27 clinical series, mostly

with women receiving thiopurines for renal transplant, also did not identify a recurrent pattern of

congenital anomalies (Polifka & Friedman, 2002). Specifically, no pattern was identified from

six clinical series and reports in 40 IBD patients on thiopurines (Alstead et al., 1990; Campbell &

Ghosh, 2000; Hanauer, 1999; Khan et al., 2000; Levy et al., 1981; Present et al., 1989). Cleary

and Källén (2009) reported an increased risk for ASD/VSDs in women using AZA for any

indication compared to the general population (OR 3.18; 95% CI 1.45 to 6.04). However, this

association was not significant when comparing women using AZA for any indication to

unexposed women with IBD (OR 1.91; 95% 0.75 to 4.57). Three studies in addition to Cleary

and Källén (2009) reported the incidence of ASD/VSD malformations in exposed and unexposed

IBD patients (Coelho et al., 2011; Dejaco et al., 2005; Francella et al., 2003). However, these

three studies had no cases in both exposed and control groups and therefore the data could not be

pooled. In the other five studies (Goldstein et al., 2007; Langagergaard et al., 2007; Norgard et

al., 2007a; Shim et al., 2011; Tennenbaum et al., 1999), there were no cases of ASD/VSD in the

exposed group, however, no data were provided for the control group. Of the six clinical series

and reports reporting outcomes in 40 IBD patients, only one ASD/VSD was reported in the

exposed infants (Campbell & Ghosh, 2000).

The risk for preterm delivery was increased in women receiving thiopurines. Since use

of thiopurines in pregnancy may be related to active disease during conception or pregnancy

(Cleary & Kallen, 2009; Dominitz et al., 2002; Goldstein et al., 2007), this finding is not

surprising since active disease itself increases the risk for prematurity (Baird et al., 1990; Bortoli

et al., 2007; Ludvigsson & Ludvigsson, 2002). Furthermore, medical induction of preterm

delivery, rather than spontaneous, was observed to occur more in women receiving thiopurines

compared to healthy controls (Langagergaard et al., 2007). Compared to IBD controls, there was

no increased risk for LBW in women with IBD being treated with thiopurines. There was,

however, a decrease in the mean birth weight when comparing to IBD controls or the general

population. Similar to the risk of prematurity, active disease during pregnancy can also increase

the chances for LBW (Baird et al., 1990; Ludvigsson & Ludvigsson, 2002).

Page 100: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

84

Limited data were provided regarding the risk of SA in women receiving thiopurines. A

decreased trend for SA in women with IBD on thiopurines was observed, however more studies

are needed. Cleary & Kallen (2009) reported on the incidence of SA in women on AZA for any

indication compared to the general population and did not observe a significant difference. Even

at high doses, 6-MP was not found to be useful as a single-agent medical abortifacient in early

pregnancy (Davis et al., 1999).

There are several limitations to our study. First, all nine of the studies included in the

meta-analysis were observational cohort studies, and only three were prospective. Since patients

were not randomized, the studies may be subject to selection bias and confounding variables.

However, studies evaluating the fetal safety of therapeutics in pregnancy are almost always

observational since there are many ethical concerns regarding randomization in this population.

Second, overall there was a low frequency of adverse outcomes, making statistical precision

difficult. Furthermore, subgroup analysis for IBD controls was difficult since only two studies

with relatively small sample sizes were available. Third, all studies except for Tennenbaum et al.

(1999) and Shim et al. (2011) did not report on disease activity during pregnancy. Since disease

activity may be related to adverse fetal outcomes, future studies should comment on this. Fourth,

there was limited information available regarding the concurrent medications that the exposed

and control groups were taking. Fifth, we included patients that received treatment with

thiopurines at any time during gestation and therefore were not able to specifically evaluate any

association with timing of exposure in pregnancy.

In conclusion, there is no significant difference in the risk for malformations or LBW in

women with IBD being treated with thiopurines compared to IBD patients not receiving

thiopurines. A significant risk for malformations and LBW was only observed when comparing

women receiving thiopurines to healthy or general population controls. Women receiving

thiopurines for IBD have an increased risk of having a premature infant, but this may be related

to disease activity. Future studies evaluating the safety of thiopurines in pregnancy should

include data on disease activity during pregnancy and use appropriate control groups.

Page 101: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

85

Chapter 5

Folic acid transport to the human fetus is decreased in pregnancies with chronic alcohol exposure Janine R. Hutson1,2, Brenda Stade3, Denis C. Lehotay4, Christine P. Collier5,6, and Bhushan M.

Kapur1,7

1Division of Clinical Pharmacology and Toxicology, Hospital for Sick Children, Toronto,

Canada

2Institute of Medical Science, University of Toronto, Toronto, Canada

3Department of Paediatrics, Keenan Research Centre, St. Michael’s Hospital, Toronto, Canada

4College of Medicine, University of Saskatchewan, Regina, Canada

5Department of Pathology and Molecular Medicine, Queen's University, Kingston, Canada

6Kingston General Hospital, Kingston, Canada

7Department of Clinical Pathology, Sunnybrook Health Sciences Centre, Toronto, Canada

This work has been published and reproduced with permissions: Hutson JR, Stade B, Lehotay DC, Collier CP, Kapur B. Folic acid transport to the human fetus

is decreased in pregnancies with chronic alcohol exposure. PLoS ONE 2012:7(5);e38057.

Page 102: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

86

Folic acid transport to the human fetus is decreased in 5pregnancies with chronic alcohol exposure

5.1 Abstract

Background: During pregnancy, the demand for folic acid increases since the fetus requires this

nutrient for its rapid growth and cell proliferation. The placenta concentrates folic acid into the

fetal circulation; as a result the fetal levels are 2 to 4 fold higher than maternal. Animal and in

vitro studies have suggested that alcohol may impair transport of folic acid across the placenta by

decreasing expression of transport proteins. We aim to determine if folate transfer to the fetus is

altered in human pregnancies with chronic alcohol consumption. Methodology/Principal

Findings: Serum folate was measured in maternal blood and umbilical cord blood at the time of

delivery in pregnancies with chronic and heavy alcohol exposure (n=23) and in non-drinking

controls (n=24). In the alcohol-exposed pairs, the fetal:maternal serum folate ratio was ≤1.0 in

over half (n=14), whereas all but one of the controls were >1.0. Mean folate in cord samples was

lower in the alcohol-exposed group than in the controls (33.15 ± 19.89 vs 45.91 ± 20.73,

p=0.04). Conclusions/Significance: Our results demonstrate that chronic and heavy alcohol use

in pregnancy impairs folate transport to the fetus. Altered folate concentrations within the

placenta and in the fetus may in part contribute to the deficits observed in the fetal alcohol

spectrum disorders.

Page 103: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

87

5.2 Introduction

During pregnancy, the demand for folic acid increases since this nutrient is critically

important for DNA synthesis and cell proliferation. During pregnancy, the placenta concentrates

folic acid into the fetal circulation and as a result fetal levels are 2- to 4-fold higher than maternal

(Economides et al., 1992; Giugliani et al., 1985; Guerra-Shinohara et al., 2004; Molloy et al.,

2005; Navarro et al., 1984; Obeid et al., 2005; Relton et al., 2005; Stark et al., 2007; Thorand et

al., 1996). The fetal requirement for folate during pregnancy is paramount such that cord folate

status is maintained even when maternal status is low (Wallace et al., 2008). Transport of folates

across the placenta is mediated by placental folate receptors (PFRs) (Henderson et al., 1995),

namely folate receptor-α (FR-α) at the microvillous membrane of the syncytiotrophoblast

(Bisseling et al., 2004; Solanky et al., 2010). Folate in the maternal circulation binds to FR-α,

where it binds with high affinity and is internalized through receptor-mediated endocytosis

(Keating et al., 2009) (Figure 5-1). Other transporters including the reduced folate carrier (RFC)

and the proton-coupled folate transporter (PCFT) are also important for the placental uptake of

folate (Prasad et al., 1995; Yasuda et al., 2008). This results in a high intervillous blood

concentration of folate within the placenta where it can be transported to the fetus by passive

diffusion and by the RFC (Bisseling et al., 2004; Henderson et al., 1995). This mechanism of

transport is established early in pregnancy (within the first trimester (Solanky et al., 2010)) as

folate is vital to the proper development of the fetus.

Page 104: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

88

Figure 5-1. The transport proteins present on the syncytiotrophoblast that are involved in

the transfer of folate to the fetal circulation.

Animal and in vitro studies have suggested that alcohol may impair transport of folic acid

across the placenta by decreasing expression of folate transport proteins (Fisher et al., 1985;

Keating et al., 2009; Keating et al., 2008). Indeed, there are similarities between deficits

observed in the fetal alcohol spectrum disorders (FASDs) and folate status during pregnancy.

These similarities include common physical malformations such as neural tube defects,

congenital heart defects, and limb defects (Chudley et al., 2005; Goh & Koren, 2008; Jones,

2006). Furthermore, lower folate status has been associated with hyperactivity, peer problems,

and lower cognitive function (Schlotz et al., 2010; Veena et al., 2010), which are all neuro-

developmental consequences of alcohol exposure during pregnancy (Chudley et al., 2005).

Page 105: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

89

Folic acid is also important as an antioxidant during pregnancy. Alcohol consumption

during pregnancy creates oxidative stress to both the placenta and fetus and this stress can be

mitigated by folic acid (Cano et al., 2001; Gundogan et al., 2010). Formic acid, the toxic

metabolite of methanol, has been reported in the sera of alcohol abusing patients and has also

recently been detected in umbilical cord blood from pregnancies with heavy amounts of alcohol

consumption (Kapur, et al., 2009). Formic acid can lead to neurotoxicity and oxidative stress

(Kapur et al., 2007). Folic acid is required for the detoxification of formic acid. Folic acid is

critical to the rate of detoxification of formic acid (Sokoro et al., 2008). Furthermore, in folate

deficient animals, the adverse effects to the fetus after alcohol exposure are more severe

compared to controls (Gutierrez et al., 2007; Lin, 1991). Taken together, proper placental

transfer of folic acid is critical to proper fetal development and can influence the fetal effects of

alcohol.

To our knowledge, there are no studies that determine fetal folate levels from pregnancies

affected by heavy alcohol use. We hypothesize that folate transfer to the fetus is impaired in

alcohol-exposed pregnancies and that this may, in part, be responsible for the deficits associated

with the FASDs.

5.3 Methods

All procedures and protocols received prior approval from the Institutional Research

Ethics Board. The most severe diagnosis under the FASDs is fetal alcohol syndrome (FAS).

Children with FAS are usually born to mothers who consume large amounts of alcohol in the

pregnancy and for a long duration. However, there is large variation in the deficits produced with

alcohol consumption and likely results from differences in amount consumed, drinking patterns,

genetics, and nutrition. Only women consuming heavy amounts throughout the pregnancy were

included in this study and are thus the women most at risk of having a child affected by FAS.

This avoided any variability that may result from differences in amount of alcohol consumed and

timing of exposure. Healthy women with no known illicit drug or alcohol use during pregnancy

were also recruited to be representative of the general population to serve as our comparison

group. As part of this study, the infants were offered follow up care and diagnostic testing for

the FASDs.

Page 106: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

90

Written informed consent was obtained from the alcohol abusing mothers included in this

study. Maternal and cord blood samples were collected at the time of delivery. These women

were identified at the time of delivery and questionnaires regarding drug use were administered

after delivery. For controls maternal and cord blood samples were collected as part of routine

clinical workup of women delivering at the hospital. Plasma folic acid was measured in our

routine clinical laboratory by chemiluminescent immunoassay using the Beckman Coulter

UniCel DxI 800 AccessH Immunoassay System (Beckman Coulter Canada, Inc., Mississauga,

Ont. Canada). Manufacturer recommended assay protocol was used. Besides folic acid results no

other information on these women was available to us.

Normally distributed data were compared as follows: maternal and fetal folate levels

were compared using a paired t-test; folate levels in the control group compared to the alcohol

group were compared using an independent t-test. Pearson correlation was used to compare

maternal and fetal folate levels. Fetal to Maternal folate ratios were compared using the Mann

Whitney U Test because they were not normally distributed. Significance was obtained if p<0.05

and all performed tests were two-tailed (where applicable).

5.4 Results and Discussion

Twenty-three women consuming heavy amounts throughout the pregnancy were included

in this study. Twenty-four women were recruited from the general population to serve as the

comparison group with no known illicit drug or alcohol use during pregnancy. Maternal

demographics and fetal information for the women drinking alcohol during pregnancy is given in

Table 5.1. Maternal age for this group ranged from 16 to 44 years. All women reported regular

alcohol consumption throughout the pregnancy and were considered to be dependent users. The

women were not recruited until later in their pregnancy and self-reported alcohol consumption

ranged from daily use to > 8 drinks per week. The women included in this cohort were being

followed by an FASD clinic and were therefore considered at high risk of having an affected

child because of their alcohol consumption. In addition to alcohol, fourteen women reported

occasional to frequent cocaine use, eight reported marijuana use, and two reported opiate use

during the pregnancy. Cigarette use was also common to this population.

Page 107: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

91

Table 5-1. Demographic information on alcohol-using women included in this study and

fetal parameters. (n=23)

Mean (SD)

Maternal age 29.2 (6.9) years

Gravidity 4.8 (3.4)

Parity 2.4 (1.9)

Gestational age 36.6 (1.6) weeks

Length 42.3 (3.2) cm

Head Circumference 32.3 (1.4) cm

Birth weight 2830 (421) g

In order to determine whether the placenta was concentrating folate to the fetal

circulation, we calculated the fetal to maternal (F:M) folate ratio using the serum folate

measurements from the mother-cord pairs. The F:M folate ratio was ≤1.0 in over half (n=14) of

the alcohol-exposed pairs, whereas all but one of the controls were >1.0 (Figure 5.2). The F:M

folate ratios in the alcohol group were significantly lower than the control group (p=0.014). Also,

there was large variability observed in the F:M folate ratios in the alcohol group (range 0.24 to

7.65) and not in the control group (range 0.79 to 3.18), suggesting that the tight regulation of

folate transport to the fetus is deregulated (Figure 5.3).

As expected, the fetal folate levels in the control group were significantly higher than

maternal in controls (45.91nM ± 20.73nM vs 26.99nM ± 11.18nM, p<0.0001) (Figure 5.4). This

finding has been reported in all studies to date comparing maternal and fetal folate levels

(Economides et al., 1992; Giugliani et al., 1985; Guerra-Shinohara et al., 2004; Molloy et al.,

2005; Navarro et al., 1984; Obeid et al., 2005; Relton et al., 2005; Stark et al., 2007; Thorand et

al., 1996). However, this finding was not observed in the alcohol group and there was a trend for

lower fetal levels compared to maternal (33.15nM ± 19.89nM vs 38.13nM ± 29.55nM,

respectively, p=0.461). Furthermore, there was no correlation between the maternal and cord

folate levels in the alcohol group (R2=0.051, p=0.226), yet there was the expected correlation in

the control group (R2=0.550, p=0.018). Together, these data support folate transport being

impaired in pregnancies affected by chronic and heavy alcohol use. Because only half of mother-

fetal pairs had a F:M folate ratio of less than one, this suggests variability in the effect. This

Page 108: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

92

could result from differences in drinking patterns, genetic susceptibility, or other drug use in

addition to alcohol (to be discussed below).

Figure 5-2. Scatter-plot of the fetal to maternal (F:M) folate ratios as measured in cord

blood and maternal blood, respectively, at the time of delivery.

Page 109: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

93

Figure 5-3. Corresponding maternal and fetal folate concentrations at the time of delivery

in pregnancies with A) heavy alcohol exposure and B) in controls.

Page 110: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

94

Figure 5-4. Mean (± SEM) cord and maternal plasma folate concentrations at the time of

delivery in alcohol-abusing women and controls. *p<0.05 for a two-tailed paired t-test.

**p<0.05 for a two-tailed independent t-test.

After observing an impaired ability to concentrate folic acid to the fetus, we next wanted

to see if the actual folate concentration in the fetal circulation was lower in the alcohol-exposed

group. Indeed, folate levels were significantly lower in cord blood in alcohol-exposed fetuses

compared to controls (p=0.04). Interestingly, there was no difference in folate levels between

the alcohol-using and control women. There was a trend for increased folate levels in the alcohol

group. This was not expected since folate deficiency is common with chronic alcohol

consumption (Halsted et al., 2002). Two studies however, have reported higher folate levels in

alcohol-using pregnant women compared to controls and may reflect women drinking beer,

which contains folates (Larroque et al., 1992; Stark et al., 2005). Despite having higher maternal

folate levels, fetal folate levels were still lower in the alcohol group – emphasizing that alcohol

likely deregulates placental folate transport.

Page 111: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

95

It has been suggested that alcohol consumption during pregnancy may alter the placental

transport of folate in animal and in vitro studies; however, no conclusion has been reached. In

vitro studies using both a choriocarcinoma cell line (BeWo) and primary trophoblast cells from

human term placentas demonstrated decreased folic acid uptake after chronic exposure to ethanol

(Keating et al., 2008; Keating et al., 2009). Furthermore, after chronic exposure in rats, there

was a decrease in folic acid binding (Fisher et al., 1985). Conversely, other studies have shown

that alcohol may not alter folate transport. Dual-perfusion of a single lobule from a term human

placenta did not demonstrate altered folate binding or transfer after treatment with ethanol

(Henderson et al., 1995). Similarly, a single ethanol exposure to rats during pregnancy did not

alter transport of folate (Lin, 1991). The different findings likely result from the type of exposure

since only those studies where there was chronic alcohol exposure showed a decreased effect.

Our study population in the alcohol group was consuming alcohol continuously during

pregnancy and is in concordance with the chronic animal and in vitro studies.

Decreased folate transfer may be a result of altered expression of the folate binding and

transport proteins, including FR-α, RFC, and PCFT. After chronic alcohol exposure in vitro,

primary trophoblasts and BeWo cells had reduced mRNA expression of RFC and FR-α

respectively (Keating et al., 2008; Keating et al., 2009). However, there are no data available

looking at mRNA, protein, or folic acid uptake in human placenta from a population similar to

that in our study. Future studies should investigate this mechanism using human placentas from

alcohol-exposed pregnancies. Since this population is difficult and sensitive to recruit and has

many co-morbidities, animal studies may be crucial to a full mechanistic understanding.

A role for impaired folic acid transport in the development of FASDs is further supported

by animal studies. After alcohol exposure during pregnancy in mice and rats, more adverse fetal

effects were observed in those born to mothers receiving a folate free diet (Gutierrez et al., 2007;

Lin, 1988). Furthermore, several recent animal studies have reported that high dose folic acid

can mitigate the adverse fetal effects induced by alcohol, especially measures of oxidative stress

(Cano et al., 2001; Garcia-Rodriguez et al., 2003; Wang et al., 2009; Xu, Tang, & Li, 2008;

Yanaguita et al., 2008). A recent study has demonstrated that administration of folic acid by

injection decreases fetal resorption and malformations in mice consuming ethanol in pregnancy

(Wentzel & Eriksson, 2008). These findings suggests that it may be possible to overcome the

decreased placental transfer observed in this study if maternal levels are high enough.

Page 112: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

96

The fetotoxic effects resulting from folate deficiency may result from different possible

pathways. Folic acid is critically important for DNA synthesis and cell proliferation as well as in

cellular protection through its antioxidant properties (Antony, 2007). Folate deficiency within

the placenta has also been shown to alter placental DNA methylation (Kim et al., 2009). Fetal

growth and development are closely linked to DNA methylation (Kim et al., 2009). Disruption

in placental DNA methylation may have a role in the Developmental Origins of Adult Health

and Disease hypothesis and alterations in the placenta have already been associated with the

future risk of cardiovascular disease and cancer (Gallou-Kabani et al., 2010).

An additional complication that may arise from folate deficiency is related to formic acid,

the toxic metabolite of methanol. Formic acid has recently been detected in maternal and cord

blood in pregnancies with chronic ethanol exposure (Kapur, et al., 2009). This formic acid is

endogenously produced from methanol naturally found in alcoholic beverages as well as

methanol produced as a by-product in the pituitary (Axelrod & Daly, 1965). Formic acid

requires folic acid for detoxification. A recent study by our group observed a negative

correlation between formic acid in cord blood and cognitive function (r=-0.6154, p=0.025, n=12

at 12 months and r=-0.6241, p=0.023, n=13 at 18 months) using Bayley scores (Kapur, BM et

al., 2009). Formic acid has been shown to cause neuronal cell death in the rat hippocampal

explants and this could be mitigated by the folic acid (Kapur et al., 2007). Thus, lower levels of

folate may leave the fetus more vulnerable to the potential toxic effects of formic acid.

The women included in our study in the alcohol group also reported use of other drugs of

abuse, including cocaine, tetrahydrocannabinol (THC in marijuana), and cigarettes. The

influence of these drugs on folate transfer cannot be ruled out. However, in vitro studies do not

support a role for cocaine in altering folate transport. Studies using primary trophoblasts showed

that acute or chronic exposure to cocaine did not alter folate uptake (Keating et al., 2009).

However, chronic exposure (but not acute) to THC, did decrease folate uptake in primary

trophoblasts (Keating et al., 2009). Furthermore, amphetamine, and ectasy (MDMA) decreased

folate uptake after both acute and chronic exposure. No women in our study reported use of

amphetamine or ecstasy, however, self-report may not be 100% accurate due to the stigma

associated with illicit drug use. Also, it is unknown whether the reported occasional use of THC

by the three women in this study would be enough to cause an effect on folate transport since

chronic exposure was needed in the in vitro studies to have an effect.

Page 113: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

97

There are currently no in vivo data available that relate folate transport to drugs of abuse;

however, there are studies evaluating folate transport in pregnancy to cigarette smoking in

pregnancy. A small decrease in cord folate levels was determined in infants born to smoking

mothers compared to non-smoking mothers (Stark et al., 2007). However, the fetal folate levels

remained higher than maternal and the positive linear relationship was maintained in the

smoking group. The decrease in fetal folate levels and transfer was more profound in our study

population; thus the observed effect in our study is a result of more than simply maternal

cigarette use. An interaction or additive effect between smoking, alcohol, and other drugs of

abuse may be possible and should be further investigated.

To our knowledge, this is the first study to show that folic acid transport to the fetus is

compromised in pregnancies affected by heavy and chronic alcohol exposure. Decreased levels

of folate available to the fetus as well as bound to the placenta itself may in part be responsible

for the deficits observed in the FASDs. Although our study was focused on FASD, low fetal

folate may also be part of the aetiologies of diseases seen in children born to chronic alcohol

drinking women during pregnancy. Although future studies are needed to address the mechanism

for the decreased folate transfer, current practice should continue to properly counsel pregnant

women and women of childbearing age on proper folic acid supplementation as well as

abstinence from alcohol during pregnancy.

Page 114: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

98

Chapter 6

Adverse placental effect of formic acid on hCG secretion is mitigated by folic acid Janine R Hutson1,2, Lubetsky, A1, Eichhorst, J3, Hackmon, R1, Koren, G1,2, Kapur, BM1,4

1Division of Clinical Pharmacology and Toxicology, Hospital for Sick Children, Toronto, ON

2Insitute of Medical Sciences, University of Toronto, Toronto, ON

3Saskatchewan Disease Control Laboratory, Regina, SK

4Dept of Clinical Pathology, Sunnybrook Health Sciences Center, Toronto, ON.

This work has been published and reproduced with permissions: Hutson JR, Lubetsky A, Eichhorst J, Hackmon R, Koren G, Kapur BM. Adverse placental

effect of formic acid on hCG secretion is mitigated by folic acid. Alcohol Alcohol 2013;48(3):283-7.

Page 115: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

99

Adverse placental effect of formic acid on hCG 6secretion is mitigated by folic acid

6.1 Abstract

Formic acid has recently been detected in maternal blood and umbilical cord blood of

infants born to alcohol abusing mothers. This toxic metabolite of methanol requires folate for

detoxification. We hypothesized that formic acid produced in the maternal circulation will

transfer across the placenta and will be toxic to the placenta. Our objectives were first, to

determine whether formic acid transfers across the human placenta and whether it is toxic to the

placenta. Second, to determine whether folate can decrease transplacental transfer of formic acid

and mitigate toxicity. Dual perfusion of a single placental lobule ex vivo was used to

characterize the transfer of formic acid across the placenta. After a 1-hour control period, formic

acid (2mM) was introduced into the maternal circulation with (n=4) or without folate (1uM)

(n=4) and was allowed to equilibrate for 3-hours. Formic acid transferred rapidly from the

maternal to the fetal circulation and transfer was not altered with the addition of folate.

Compared to the control period, there was a significant decrease in hCG secretion (p=0.03) after

addition of formic acid. The addition of folic acid to the perfusate mitgiated the decrease in

hCG. Formic acid rapidly transfers across the placenta and thus has the potential to be toxic to

the developing fetus. Formic acid decreases hCG secretion in the placenta, which may alter

steroidogenesis and differentiation of the cytotrophoblasts, and this adverse effect can be

mitigated by folate.

Page 116: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

100

6.2 Introduction

Alcoholic beverages may contain a small amount of methanol as a congener

(Lachenmeier et al., 2011). Furthermore, methanol is produced endogenously in the pituitary

from S-adenosylmethionine (Axelrod & Daly, 1965; Sarkola & Eriksson, 2001). In heavy

drinkers, methanol may accumulate since ethanol has a higher affinity for alcohol dehydrogenase

(ADH) and may reach plasma concentrations above 2 mM (Kapur et al., 2007; Majchrowicz &

Mendelson, 1971). Formic acid, the toxic metabolite of methanol, has been detected in both sera

and cerebral spinal fluid of alcoholics in concentrations that are neurotoxic (Kapur et al., 2007).

Formic acid has also been recently detected in maternal blood and umbilical cord blood of

infants born to heavy drinkers (Kapur et al., 2009).

Formic acid requires folate for detoxification and folate status can influence the clearance

rate (Johlin et al., 1987; Sokoro et al., 2008). Studies on rat brain hippocampal slices showed

that formic acid can cause neuronal cell death, and this toxicity can be mitigated by folate (Kapur

et al., 2007). Formic acid has also been shown in animal studies to lead to growth restriction,

physical malformations, and depletion of glutathione in the embryo (Andrews et al., 1998;

Brown-Woodman et al., 1995; Hansen et al., 2005; Harris et al., 2004). Furthermore, human

studies have reported fetal alcohol syndrome-like facial features in infants born after solvent

abuse (including methanol) by the mother during pregnancy (Scheeres & Chudley, 2002).

The placenta separates the maternal and fetal circulations throughout human pregnancy.

Heavy alcohol consumption during pregnancy is associated with numerous placental alterations

including placental dysfunction, decreased placental size, gene expression changes and endocrine

disruptions (Burd et al., 2007; Rosenberg et al., 2010). Since formic acid can produce oxidative

stress (Harris et al., 2004), it may be placentotoxic. Compromised placental function is

associated with growth restriction, a deficit associated with the fetal alcohol spectrum disorders

(FASDs). Since the placenta is a reservoir for folate (Henderson et al., 1995), there may also be

a potential role for detoxification of formic acid within the placenta itself. The objectives of the

study were to first determine whether formic acid transfers across the placenta and is toxic to the

placenta. Second, to determine whether folate can decrease transplacental transfer of formic acid

and mitigate potential toxicity.

Page 117: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

101

6.3 Methods

6.3.1 Materials

Sodium formate was obtained from Sigma (St. Louis, MO). The water used for all

experimental procedures was obtained from a Milli-Q Advantage A10 Ultrapure Water

Purification System (Millipore, Billerica, MA).

6.3.2 Placental perfusion

The dual perfusion of a placental lobule was previously described by Miller et al. (1985)

and adapted in our laboratory (Derewlany et al., 1991; Pollex et al., 2008). Immediately after

elective caesarean sections, term placentas were obtained from healthy mothers with

uncomplicated pregnancies from St. Michael’s Hospital in Toronto, Ontario. Research ethics

board approval was obtained and maternal consent was attained prior to the surgery. For each

placenta, a vein/artery pair supplying a clearly identifiable cotyledon was chosen for cannulation

and maternal and fetal circulations were established within 30 minutes of delivery (Derewlany et

al., 1991).

The maternal perfusate was equilibrated with 95% O2, 5% CO2 and the fetal with 95%

N2, 5% CO2. The perfusate consisted of M199 tissue culture medium (Sigma, St. Louis, MO)

containing heparin (2000 U/L), glucose (1.0 g/L), kanamycin (100 mg/L), and 40,000 molecular-

weight dextran (maternal 7.5 g/L; fetal 30 mg/L). As a marker of passive diffusion, antipyrine (1

mM) was added to the maternal circulation. The fetal and maternal circuit flow rates were

maintained at 2 to 3 and 13 to 14 ml/min, respectively, and the temperature of both circuits and

the perfusion chamber was kept at 37°C. Small volumes of sodium bicarbonate and hydrochloric

acid were added to the perfusate to maintain the maternal and fetal circuits at pH 7.4 and 7.35,

respectively. This pH mimics the slightly more acidic fetal circulation observed in vivo

(Reynolds & Knott, 1989).

Each perfusion consisted of a 1 h pre-experimental control period, followed by a 3 h

experimental period, where both were in a recirculating (closed) configuration. The perfusates in

both circulations were replaced with fresh media prior to each of the 2 periods. During the pre-

Page 118: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

102

experimental period, samples were taken every 15 min to analyze glucose and oxygen

consumption, and human chorionic gonadotropin (hCG) secretion as measures of tissue viability.

During the experimental period, these measures were taken every 30 minutes. Tissue viability

measures were calculated as previously described (Pollex et al., 2008). Fetal perfusion pressure

and fetal reservoir volume were monitored as an indicator of tissue integrity. The perfusion was

terminated at any time if there was a >3 ml/hour loss in fetal reservoir volume. pH, pO2, and

pCO2 were monitored using a blood gas analyzer (Radiometer ABL 725, Copenhagen,

Denmark).

During the experimental period, 2 mM sodium formate was added to the maternal

circulation. 2 mM was chosen since this is a clinically relevant concentration that was

previously observed in heavy drinkers (Kapur et al., 2007). For perfusions that included folate, 1

µM was added into both the fetal and maternal circuits at the beginning of both the pre-

experimental and experimental periods. Samples were taken for measurement of formate and

folic acid every 10 minutes for the first half-hour and every half-hour following.

6.3.3 Sample analysis

At the end of each perfusion, the perfused lobule was isolated. Perfused and unperfused

tissue from the same placenta was homogenized (1:5 (w/v) in phosphate buffered saline, pH 7.4)

using a Polytron (Brinkmann Instruments, Inc., Westbury, NY) for 2 minutes. The homogenate

was centrifuged at 5000 ×g for 15 minutes at 4 °C. The supernatant was removed and analyzed

for both hCG and formic acid. hCG levels were measured by enzyme-linked immunosorbent

assay (Alpha Diagnostic Intl. Inc., San Antonio, Texas) after a 1:80 dilution. The rate of hCG

secretion was determined by calculating the slope of the hCG concentration vs time curve during

the initial linear portion of the curve.

Perfusate samples were stored at -20°C until analysis. Formic acid was detected by gas

chromatography-flame ionization detection using a previously published method (Sokoro et al.,

2008). Area under the curve (AUC) was calculated using the trapezoidal rule. Tissue hCG levels

of the perfused tissue are expressed as a percentage of the initial tissue concentration

(Linnemann et al., 2000). Differences between the pre-experimental control and experimental

Page 119: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

103

periods were compared using a paired t-test and differences between perfusion with and without

added folate were compared using an independent t-test. Differences were considered significant

if P < 0.05.

6.4 Results

A total of 8 lobules from different placentae were perfused with formic acid (4 with

folate added and 4 without folate added) and the physical parameters for the perfusions are given

in Table 6.1. The mass of the perfused cotyledons ranged from 8.53 to 17.12 g. Measures of

placental viability, integrity, and function throughout the experiments remained within normal

ranges and were not significantly different between the control and experimental phases (Table

6-1). Furthermore, the fetal arterial pressures remained similar between the control and

experimental periods. The rate of antipyrine disappearance from the maternal circulation was

equal to the rate of appearance in the fetal circulation during the experimental period with values

of 0.028 ± 0.004 and 0.029 ± 0.006 µmol/g per minute, respectively.

Formic acid transferred rapidly from the maternal to the fetal circulation (Figure 6-1). In

the presence or absence of folate to the perfusate, formic acid appeared in the fetal circulation

within 10 minutes in all eight perfusions. The addition of folate into the perfusate did not alter

the fetal AUC (1.30 ± 0.14 without folate; 1.23 ± 0.48 with folate; p = 0.79). Tissue

concentrations of formic acid measured in the perfused lobules at the completion of the

experiment were 425.83 ± 57.18 and 431.18 ± 133.07 nmol/g for perfusions without and with

folate added, respectively.

Page 120: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

104

Table 6-1 Placental viability parameters and metabolic capacity throughout the perfusions

(mean ± SD)

Viability Parameter

Formic Acid without Folate (n=4) Formic Acid with Folate (n=4)

Pre-Experimental Control

Experiment

Pre-Experimental Control

Experiment

Fetal arterial inflow pressure (mm Hg)

43.65 ± 6.02 41.24 ± 2.15 41.50 ± 12.24 38.62 ± 9.05

Oxygen (µmol O2/g per min)

Transfer 0.02 ± 0.01 0.02 ± 0.01 0.02 ± 0.00 0.02 ± 0.00

Delivery 0.75 ± 0.33 0.72 ± 0.29 0.62 ± 0.09 0.64 ± 0.05

Consumption 0.35 ± 0.16 0.32 ± 0.13 0.31 ± 0.02 0.34 ± 0.03

Glucose consumption (µmol/g per min)

0.65 ± 0.58 0.42 ± 0.11 0.41 ± 0.09 0.32 ± 0.07

Lactate Production (µmol/g per min)

0.38 ± 0.11 0.25 ± 0.06 0.28 ± 0.17 0.22 ± 0.06

Page 121: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

105

Figure 6-1 Maternal-to-fetal transfer of formic acid after dual perfusion of a single

placental lobule in the presence (n=4) or absence (n=4) of folate (1µM). Formic acid (2mM)

was added to the maternal circulation and transfer was determined for a period of 180 min.

Data are shown as mean values ± SEM at each time point.

Page 122: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

106

Figure 6-2 The rate of hCG secretion into the maternal circulation during the placental

perfusion was calculated in the pre-experimental control period (before addition of formic

acid) and during the experimental period in the A) presence (n=4) and B) absence (n=4) of

folate. C) The rate of hCG secretion was compared between the experimental period in the

presence and absence of folate.

0"

2000"

4000"

6000"

0" 50" 100" 150" 200"

hCG$(m

IU/g)$

Time$(min)$

A)$

Experimental"

"Control""

0"

2000"

4000"

6000"

0" 50" 100" 150" 200"

hCG$(m

IU/g)$

Time$(min)$

B)$

Experimental"

Control"

0"1000"2000"3000"4000"5000"6000"

0" 50" 100" 150" 200"

hCG$(m

IU/g)$

Time$(min)$

C)$

Without"Folic"Acid"

With"Folic"Acid"

Page 123: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

107

Figure 6-3 Tissue concentration of hCG in the perfused lobule at the end of the 180 minute

experimental period with formic acid expressed as a percentage of the initial hCG tissue

concentration in the same placenta in the presence (n=4) and absence (n=4) of folate.

Compared to the pre-experimental control period, there was a significant decrease the

rate of hCG secretion in the maternal circulation after addition of formic acid in the experimental

period (p = 0.03) (Figure 6-2). In contrast, there was no significant decrease when folate was

present in the perfusate. The percentage of initial placental tissue hCG was decreased in the

perfusions without folate compared to perfusions with folate (p = 0.04) (Figure 6-3).

Page 124: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

108

6.5 Discussion

The small molecule, formic acid, crosses the placenta rapidly. Since formic acid can

reach high concentrations in heavy drinkers (Kapur et al., 2007; Kapur et al., 2009), our results

suggest that any formic acid produced by the mother is likely transferred to the fetus in utero.

Although there are a lack of in vivo human studies regarding the developmental toxicity of

formic acid (Ema et al., 2010), this molecule has the ability to cause neurotoxicity, growth

restriction, physical malformations, and depletion of fetal glutathione in animal studies (Andrews

et al., 1998; Brown-Woodman et al., 1995; Hansen et al., 2005; Harris et al., 2004; Kapur et al.,

2007).

Our current study demonstrates that formic acid is placentotoxic in that placental hCG

secretion was decreased. Other toxic compounds have been shown to decrease or inhibit

placental hCG secretion including cocaine, 2,3,7,8-tetrachlorodibenzo-p-dioxin, a

polychlorinated dibenzo-p-dioxin/polychlorinated dibenzo-P-furan mixture,

difluoromethylornithine, bisphenol A, benzo[a]pyrene, and high levels of zidovudine

(Augustowska et al., 2007; Boal et al., 1997; Moore et al., 1988; Morck et al., 2010; Simone et

al., 1996; Zhang et al., 1995). Since formic acid may be present in the maternal circulation at the

same time as ethanol, further studies should evaluate how these two agents interact to alter hCG

secretion since in vitro studies in human trophoblasts demonstrated that ethanol can increase

hCG secretion (Karl & Fisher, 1993; Karl et al., 1994; Karl et al., 1998; Wimalasena et al.,

1994).

Alterations in hCG secretion by the placenta may have consequential implications for the

pregnancy since hCG is vitally important in maintaining the pregnancy, regulating blood flow,

and delivering nutrients to the conceptus (Boal et al., 1997). In early pregnancy, hCG maintains

the corpus luteum for progesterone production. Even in later stages of pregnancy, hCG

continues to have numerous functions that have been recently reviewed by Cole (2010). Briefly,

hCG promotes cellular differentiation in the placenta to make syncytiotrophoblast cells (Cronier

et al., 1994); blocks the maternal immunological response to invading placental cells (Akoum et

al., 2005); promotes uterine and umbilical cord growth (Reshef et al., 1990); and increases blood

flow to the fetoplacental unit by promoting angiogenesis in the uterine vasculature and

maintaining umbilical cord elasticity (Berndt et al., 2009; Herr et al., 2007; Toth et al., 1994).

Page 125: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

109

The addition of folate did not alter the transfer of formic acid, however, it did mitigate the

effects on hCG secretion. Since tissue concentrations of formic acid were similar in the presence

or absence of folate, this suggests that folate may mitigate toxicity to the placenta by acting as an

antioxidant to the oxidative stress caused from formic acid as opposed to increasing clearance of

formic acid. Our study shows that formic acid in the mother with heavy drinking during

pregnancy, at concentrations that are obtained through both endogenous production and

methanol found in alcoholic beverages, can be placentotoxic in addition to being neurotoxic as

shown in our previous study (Kapur et al., 2007). Using organotypic rat hippocampal brain slice

cultures, we previously showed that formic acid neurotoxicity is both dose and time-dependent:

cell death increased significantly between 1 to 5 mM formic acid and also between 24 and 72

hours (Kapur et al., 2007). Our current study evaluates the transfer and toxicity at 2 mM formic

acid as this is the maximal concentration observed in heavy drinkers (Kapur et al., 2007).

Studies evaluating fetal effects of formic acid using rodent whole embryo culture have

demonstrated concentration-dependent toxicity (Andrews et al., 1998; Brown-Woodman et al.,

1995; Hansen et al., 2005). After microinjection of sodium formate into the amniotic fluid of

pregnant mice and rats, incomplete fetal axial rotation was observed at concentrations as low as

0.029 mM (Hansen et al., 2005).

A limitation of our results is that we evaluated toxicity at one concentration of formic

acid, 2 mM. Since the placental perfusion model is very tedious, future in vitro studies are

needed to characterize the full dose-response relationship for placental toxicity. Formic acid in

heavy drinkers and in pregnant women has not been characterized in large populations. It is

possible that higher concentrations of formic acid may be obtained as methanol may accumulate

as long as ethanol is present in the blood (Kapur et al., 2007; Majchrowicz & Mendelson, 1971).

To properly characterize formic acid concentrations in pregnant alcohol-consuming women,

studies are needed that capture the full time-concentration curve and these patients will be

difficult to recruit. However, theoretically, we expect formic acid concentrations to reach similar

levels in pregnant women compared to non-pregnant adults. Preliminary results from our group

show formic acid concentrations as high as 140 and 533 µM in maternal blood and umbilical

cord blood, respectively, although these samples were not obtained at times where peak formic

acid concentrations are expected. In non-pregnant adults, a serum formic acid concentration of 2

mM was observed with a blood alcohol concentration of 150 mM (Kapur et al., 2007). This

Page 126: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

110

would correspond to consuming over 12 standard drinks on one occasion. Binge drinking at and

exceeding this level has been reported in pregnant women (Kuehn et al., 2012; Malone et al.,

2010), thus it is likely that similar concentrations of formic acid are achieved in this population.

We have recently shown that folate transport to the fetus is compromised in pregnancies

affected by heavy and chronic alcohol exposure (Hutson et al., 2012). Although we did not

focus on FASD, one could speculate that with decreased folate transport there will be an increase

in formic acid which may play a significant role in the etiology of FASD.

6.6 Conclusions

In summary, formic acid at peak concentrations observed in heavy drinkers can rapidly

transfer across the placenta. Formic acid may be harmful to the developing fetus and our results

demonstrate that it is also directly placentotoxic. Formic acid decreases hCG secretion in the

placenta, which may alter steroidogenesis and differentiation of the cytotrophoblasts, and this

can be mitigated by folate.

Page 127: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

111

Chapter 7

General Discussion 7Many women require pharmacologic treatment during pregnancy for chronic medical

conditions or illnesses that arise during pregnancy (Koren et al., 1998). Furthermore, many

women may be exposed to known teratogens inadvertently during pregnancy or because of

addiction issues. The benefit of drug treatment must be weighed against any potential risk to the

unborn child as well of risk of the untreated maternal disease. After the thalidomide disaster,

there is a current belief that every drug has the potential to be harmful to the fetus (Koren et al.,

1998) and thus every drug must be carefully evaluated for its safety in pregnancy before it can be

ethically administered to pregnant women. Clinical studies of drug safety in human pregnancy

are often limited because of ethical considerations, thus a validated model or framework for

predicting drug safety is needed. As the placenta is the interface between the maternal and fetal

circulations, a thorough understanding of the role of the placenta in maternal-fetal

pharmacokinetics is vital to understanding drug safety in pregnancy. Animal studies often do not

extrapolate well into humans because the placenta is the most species-specific mammalian organ

(Ala-Kokko et al., 2000). In vitro studies using trophoblast preparations are useful for studying

mechanisms of transfer but do not provide information on transfer in the intact organ, which

includes the often overlooked fetal endothelial cells that form part of the barrier (Elad et al.,

2014). Dual perfusion of a single placental lobule ex vivo is the only experimental model to

study human placental transfer of substances in organized placental tissue. Because of this

feature, it was thought that this experimental model was able to better predict fetal exposure

compared to other in vitro methods such as cell culture, placental explants, or membrane

vesicles. Validation of this placental perfusion model is needed before it can have a place in

clinical and regulatory pharmacology and toxicology. The studies conducted as part of this

doctoral thesis aimed to validate the placental perfusion model as an accurate source of

determining drug transfer across the human placenta. Furthermore, this thesis aimed to

subsequently use the perfusion model in conjunction with clinical studies to provide examples of

how the perfusion model can be used to draw conclusions that are clinically meaningful.

Page 128: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

112

The present chapter deals with the significance of the specific research findings outlined

in the preceding chapters, and summarizes these findings in terms of clinical significance,

limitations of the conclusions, and future studies needed in each area.

7.1 Summary of Research Findings and Future Directions

I. To systematically evaluate the placental perfusion model in predicting placental

drug transfer and to develop a pharmacokinetic-model to account for non-

placental pharmacokinetic parameters in the perfusion results (Chapter 2).

A large number of studies have evaluated the transplacental transfer of therapeutic drugs

using the placental perfusion model. However, the accuracy of how well the results from the

perfusion model can predict the in vivo placental transfer in humans was unknown. This was

important to determine in order to better generalize results from perfusion studies into data that

can be used clinically. Knowledge translation is an important part of completing research and

finding ways to use basic science data in a clinically meaningful way. The results of this study

showed that:

• A systematic review of the literature determined that the fetal-to-maternal ratio in both

the perfusion model and in vivo was determined for 70 different drugs. The fetal-to-

maternal drug concentration ratios matched well between placental perfusion experiments

and in vivo samples taken at the time of delivery of the infant.

• The placental perfusion model was best able to predict drugs that have limited transfer

(F:M <0.1) across the term human placenta.

• After modeling for differences in maternal and fetal/neonatal protein binding and blood

pH, the perfusion results were able to accurately predict in vivo transfer at steady state

(R2 = 0.85, P < 0.0001).

• Determining the protein binding in vitro using fetal and maternal blood in parallel with

perfusion experiments, and then modeling the results as documented above, would allow

Page 129: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

113

for more accurate predictions for placental drug transfer at steady state compared to

performing perfusion experiments alone.

• After modeling, the placental perfusion model can be used to accurately predict in vivo

placental drug transfer and thus estimate fetal drug exposure. This will allow conclusions

to be reached regarding placental drug transfer for new drugs for which no in vivo fetal

safety data are available.

Our systematic review and model development (Hutson et al., 2011, Chapter 2) determined

that after adjusting for maternal and fetal pharmacokinetic factors, the placental perfusion model

can be used to accurately predict in vivo placental drug transfer and thus estimate fetal drug

exposure. This will allow conclusions to be reached regarding placental drug transfer for new

drugs for which no in vivo fetal safety data are available. It is also the first study to show that

adjusting data from perfusion experiments mathematically for maternal and fetal protein binding

is more accurate than attempting to add protein into the maternal or fetal perfusate during the

experiment. This will prove to be cost effective in addition to improving the generalizability of

the results. Furthermore, drugs known to influence fetal growth or development that were studied

using our model showed good correlation including NSAIDs, antiepileptics, and endocrine

agents. In fact, Shintaku et al. were able to use data from perfusion experiments to predict fetal

toxicity of NSAIDs (detailed below) (Shintaku et al., 2012). Drugs where fetal exposure is

desired, for example antivirals, also showed good agreement between in vivo and perfusion

results.

A limitation of our study was that we compared a small proportion of drugs that have been

studied by the placental perfusion model. Out of 128 therapeutic drugs that have been

investigated by placental perfusion studies, only 26 drugs met inclusion criteria for a quantitative

comparison. Many of these drugs had no in vivo data available (there was no umbilical cord

blood to maternal blood drug concentration ratio available). Since many of these drugs are

considered safe in pregnancy and are commonly given to pregnant women, it may be useful to

measure these concentrations in order to increase the number of drugs compared quantitatively.

We also proposed that measuring protein binding in vitro in umbilical cord blood and maternal

blood would be a useful alternative to measuring in vivo protein binding for new drugs as new

medications can not be ethically administered to pregnant women without safety data. By

Page 130: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

114

determining protein binding in vitro together with placental perfusion experiments would be a

useful framework for evaluating placental transfer of new medications that could be utilized by

regulatory agencies.

To apply this framework for the evaluation of new medications, a similar protocol for the

perfusion experiment must be performed by different laboratories. The basic experimental

perfusion model set-up as described in this thesis is performed by most laboratories utilizing this

experimental model, although slight variations do exist (Hutson et al., 2011; Mose et al., 2012).

It is recommended to transfuse a reference compound in all placental perfusions, such as

antipyrine (Mose et al., 2012). Antipyrine is a neutral molecule which undergoes passive

diffusion across the placenta in a flow dependent manner without binding to protein at

physiological pH (Schneider et al., 1972). Normalizing perfusion data of a specific drug to

antipyrine shows good agreement between laboratories who may use slightly different variations

of the placental perfusion setup (Mose et al., 2012). Published results from perfusion

experiments should therefore include enough both original data on the drug being studied as well

as data on antipyrine transfer. Furthermore, calculated normalized data should also be provided.

This will allow for secondary analyses of perfusion data to be performed (Hutson et al., 2011).

Mathiesen et al. (2010) have also proposed criteria to be included in perfusion publications and

have recommended that each laboratory publish a detailed methods paper so that all

experimental parameters can easily be obtained (Mathiesen et al., 2010).

Adjusting results from placental perfusion data for non-placental pharmacokinetic factors,

including protein binding and pH, showed that this experimental model was able to accurately

predict in vivo placental transfer at term. The main limitation of using this experimental model to

assess fetal drug exposure in pregnancy is the conclusions may not be generalizable to earlier

gestational ages. However, at term, drug transfer at term may represent the highest exposure

compared to earlier gestational ages (Vahakangas & Myllynen, 2006) since the placental transfer

layer is the thinnest and expression of certain efflux transporters such as P-glycoprotein is

substantially decreased (Gil et al., 2005; Nanovskaya et al., 2008; Sun et al., 2006; Vahakangas

& Myllynen, 2006). Using data from the perfusion experiments together with other in vitro

experimental methods using human placental tissue from earlier gestational ages, such as

placental explants, membrane vesicles, and primary cell culture, may be useful. For example,

knowing that a specific drug is a substrate for placental BCRP, kinetic parameters could be

Page 131: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

115

evaluated from membrane vesicles prepared from early gestational tissue and compared to term

tissue to draw more generalizable conclusions.

Recently, a new in vitro model of placental transport was developed by Levkovitz et al.

(2013a). In customized wells, this model co-cultures human trophoblast cells with human

umbilical vein endolthelial cells on both sides of a denuded amniotic membrane resulting in a

monolayer mimicking the architecture of the human placenta. This is the first co-culture

technique reported to mimic the human placental barrier, however, other biological barriers have

been co-cultured successfully including the blood-brain barrier, the pulmonary blood-gas barrier,

and skin and eye (Dai et al., 2005; Hermanns et al., 2004; Levkovitz et al., 2013; Ma et al.,

2005). This model was shown to have good correlation with in vivo data for the transfer of

glucose across the placenta (Levkovitz et al. 2013b). This in vitro placental model more closely

resembles the in vivo architecture compared to other cellular monolayers that use only

trophoblast cells, however the amniotic membrane used to mimic the basement membrane is

thicker compared to in vivo which may influence transfer. Furthermore, it is unclear what

gestational age this placental model may best represent and whether drug transport proteins are

present consistently in these cellular preparations. Further studies are required to validate this

model, however, it may be an efficient model to for high throughput screening of drugs before

moving to the more tedious model of the placental perfusion. It may be yet another tool that

could be used in conjunction with the perfusion model to provide safety data for new

medications where no safety data in pregnancy is available.

One exciting new area of current research is the modeling of pharmacodynamic data together

with placental pharmacokinetic data. Shintaku et al. used transplacental pharmacokinetic

parameters obtained from placental perfusion studies and pharmacodynamic data obtained from

animal studies to quantitatively predict human fetal toxicity of NSAIDs (Shintaku et al., 2012).

Their model was able to predict the risk of constriction of the ductus arterioiusus and thus predict

fetal toxicity using maternal serum concentrations. Novel approaches, such as our model and the

integrated pharmacokinetic-pharmacodynamic modeling, are expected to be useful tools for

evaluation of drug safety in pregnancy. Developing models utilizing both pharmacodynamics

and pharmacokinetic data will allow better predictions of overall fetal drug safety and should be

validated in future studies. This will require the collaboration of basic science researchers,

clinical pharmacologists, and biomedical engineers to develop these complex models.

Page 132: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

116

II. To further evaluate the safety of AZA and 6-MP in pregnancy by systematically

reviewing the literature and by determining the mechanisms by which the

placenta restricts 6-MP transfer (Chapters 3 & 4).

In addition to new medications where no clinical safety data is available, the perfusion model

is also useful to evaluate mechanisms of placental transfer, or lack of, for drugs where there may

still be concern regarding fetal safety. Another objective of this thesis was to utilize the validated

perfusion model in assessing drug safety in pregnancy in conjunction with available clinical

studies. An example of this is the use of thiopurines for IBD in pregnant women. Although there

is evidence to support that the benefit of treatment with thiopurines in IBD patients during

pregnancy likely outweighs any fetal risk, safety concerns still arise as a result of their cytotoxic

mechanism of action and their toxicity during pregnancy in animal studies (Hutson et al., 2013).

Despite the risks of untreated disease in pregnancy, some physicians still choose not to prescribe

AZA during pregnancy or to discontinue treatment in the third trimester (Peyrin-Biroulet et al.,

2011). By using our results from the placental perfusion experiments together with the results of

other pharmacokinetic parameters known about 6-MP and available human studies in pregnancy,

clinically relevant conclusions can be drawn to reassure women about the safety of thiopurines in

pregnancy.

• A meta-analysis was performed of all original human studies reporting outcomes in

pregnancy in patients receiving thiopurines and included nine studies and a total of 494

patients with IBD and 2,782 IBD controls.

• When compared with healthy women, those receiving thiopurines had an increased risk

for congenital malformations (RR 1.45; 95% CI 1.07–1.96; p = 0.02); however, when

compared with IBD controls, there was no increased risk (RR 1.37; 95% CI 0.92–2.05; p

= 0.1). These data provide support for thiopurines having a minimal risk, if any, to the

fetus.

• Patients with IBD receiving thiopurines were more likely to have premature infants (<37

weeks GA) compared with IBD controls (RR 1.67; 95% CI 1.17–2.37; p = 0.004) but this

may be related to disease activity. The risk ratio for LBW (<2,500 g) was not significant

when comparing women with IBD receiving thiopurines and all IBD controls (p = 0.13).

Page 133: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

117

• In the dually perfused placental perfusion model, with 6MP being added to the maternal

circulation, there was a biphasic decline in its concentration and a delay in fetal

circulation appearance suggesting uptake and retention of 6MP by the placenta.

• In the dually perfused placental perfusion model, with equal concentrations of 6MP in

both the maternal and fetal circulations, the fetal-to-maternal concentration ratio was >1.0

as a result of ion trapping.

• In the dually perfused placental perfusion model, metabolites of 6MP were not found in

the fetal circulation at physiologic concentrations.

• Binding to placental tissue and maternal pharmacokinetic parameters are the main factors

that restrict placental transfer of 6-MP. Active transport is unlikely to play a significant

role and drug interactions involving, or polymorphisms in, placental drug efflux

transporters are not likely to put the fetus at risk of higher 6-MP exposure.

Our data from the meta-analysis and the perfusion experiments support the safety of using

azathioprine or 6-MP in pregnancy as the benefit of treating the disease likely outweighs the risk

of maintaining treatment during the pregnancy; however, this should be individualized to each

patient. Using the dually perfused placental perfusion model, we showed that binding to

placental tissue and maternal pharmacokinetic parameters are the main factors that restrict

placental transfer of 6-MP. Furthermore, we concluded that active transport is unlikely to play a

significant role and drug interactions involving, or polymorphisms in, placental drug efflux

transporters are not likely to put the fetus at risk of higher 6-MP exposure. This finding was not

in agreement with our hypothesis that BCRP would efflux 6-MP into the maternal circulation. It

is likely that placental expression of BCRP in normal physiological conditions does not

contribute significantly to the efflux of 6-MP and that other mechanisms of transfer are important

in achieving steady state concentrations. 6-MP was determined to be a substrate using an in vitro

model where MDCKII and HEK293 cells were transfected with human or mouse ABCG2 (de

Wolf et al., 2008). These cell lines consequently overexpressed BCRP and thus it is difficult to

extrapolate whether BCRP would play a rate-limiting role in placental drug transport in vivo.

Since expression of BCRP does not change throughout gestation (discussed in Appendix C), it is

unlikely that it would alter steady-state fetal to maternal concentrations in earlier gestations.

Page 134: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

118

In addition to placental perfusion data and in vivo measurements of 6-MP fetal transfer,

several observational studies have been published regarding the safety of thiopurines in

pregnancy. Randomized controlled drug studies are usually not available because they most

always exclude pregnant women owing to fear of teratogenic risk. As a result, epidemiological

studies based on observational data constitute the main clinical data source (Schaefer et al.,

2008). Often, reporting of observational research is not sufficiently detailed to accurately assess

the potential strengths and weaknesses of these investigations (Etwel et al., 2014). Multiple

deficiencies such as heterogeneity, methodological quality, insufficient statistical methods, and

control of potential sources of bias have been widely recognized (Maguire et al., 2008;

Simunovic et al., 2009). The synthesis of these observational cohort studies into a systematic

review and meta-analysis yields an important signal for the safety/risk of drug use in pregnancy.

In fact, a recent systematic review determined that meta-analyses of earlier, albeit smaller, cohort

studies tend to generate an accurate overall teratogenic signal in estimating human teratogenicity

years before large and methodologically superior cohort studies are published (Etwel et al.,

2014).

We completed a systematic review of all original human observational studies reporting

outcomes in pregnancy in patients receiving thiopurines and included nine studies and a total of

494 patients with IBD and 2,782 IBD controls. When comparing women receiving thiopurines

for IBD compared to women with IBD not receiving thiopurines, there was no increased risk for

congenital malformations or LBW (<2,500 g). However, patients with IBD receiving thiopurines

were more likely to have premature infants (<37 weeks GA) but this may be related to disease

activity. A similar meta-analysis reporting on the safety of azathioprine and 6-MP was recently

published alongside our results (Akbari et al., 2013). They included only five studies (Cleary &

Kallen, 2009; Coelho et al., 2011; Francella et al., 2003; Norgard et al., 2007a; Shim et al., 2011)

and found results similar to ours: women with IBD exposed to thiopurines compared to IBD-

controls not receiving thiopurines, the pooled ORs for LBW, preterm birth, and congenital

abnormalities were 1.01 (95% CI 0.96, 1.06), 1.67 (95% CI 1.26, 2.20), and 1.45 (95% CI 0.99,

2.13), respectively. Akbari et al. (2013) limited their analysis to include only studies with an IBD

control group. However, compared to our analysis of studies with IBD control group

comparison, they did not include an article in French (Tennenbaum et al., 1999) and an abstract

(Dejaco et al., 2005). Results from our meta-analysis and that by Akbari et al. (2013) emphasize

Page 135: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

119

the importance of study design when evaluating the safety of medications in pregnancy. An

increased risk for malformations was observed when thiopurine-exposed pregnancies were

compared to the general/healthy population, however, this association disappeared when a more

appropriate control group was used (IBD controls). As discussed in Chapter 4, disease activity

may be responsible for any increased fetal risk observed and thus it is important to have disease-

matched controls. Specifically for IBD, medication use during pregnancy has been suggested to

be a surrogate for disease activity during pregnancy and IBD flares during pregnancy are

associated with adverse pregnancy outcomes such as preterm birth and LBW (Norgard et al.,

2007b; Baiocco & Korelitz, 1984; Bush et al., 2004; Dominitz et al., 2002; Fedorkow et al.,

1989; Morales et al., 2000; Reddy et al., 2008; Woolfson et al., 1990).

A limitation of our meta-analysis was that there was not enough data provided by the studies

to subsequently look at the risk of cardiac malformations in thiopurine-exposed infants. Cleary

and Källén (2009) suggested an increased risk for ASD/VSD after fetal exposure to thiopurines;

however, three additional studies had no cases in both their exposed and control groups (Coelho

et al., 2011; Dejaco et al., 2005; Francella et al., 2003). Cleary and Källén (2009) used

retrospective study design using data from a national health registry. Since women with IBD are

considered high-risk (Cornish et al., 2007) and both women and physicians still perceive

thiopurines in pregnancy as unsafe (Mountifield et al., 2010; Peyrin-Biroulet et al., 2011), it is

likely that more postnatal tests and monitoring may be performed in order to ease any related

anxiety. Performing more postnatal tests may lead to an overestimate of the risk from

confounding by indication; a similar phenomena has been observed for the selective serotonin

reuptake inhibitors (SSRIs) in pregnancy (Jimenez-Solem et al., 2012). An increased risk for

VSD was originally associated with fetal SSRI exposure (Reis & Kallen, 2010); however, it was

also shown that in the first year of life, approximately twice as many echocardiograms were

performed in SSRI-exposed infants compared with infants of unexposed women (Bar-Oz et al.,

2007). This diagnostic bias may lead to more VSDs being detected since it is a congenital

malformation that often resolves on its own and may go undetected in otherwise healthy infants.

A large recent study was able to show that, in fact, the observed association with VSD and SSRIs

was a result of confounding by indication (Jimenez-Solem et al., 2012).

Since completion of our meta-analysis, three additional studies have recently been published.

First, a population-based study from the United Kingdom reviewed medical records of 1703

Page 136: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

120

children born to mothers with IBD and 384,811 children of mothers without IBD (Ban et al.,

2014). Maternal exposure to azathioprine or 6MP was identified if a woman received a

prescription for these medications during the first trimester. Using logistic regression, the

adjusted odds ratios of a major congenital anomaly associated with azathioprine or 6MP

exposure was 1.27 (95% CI, 0.48–3.39). In early pregnancy, 24.6% of women did not receive

further prescriptions in pregnancy and were considered to have discontinued azathioprine/6MP

in pregnancy. There was no evident association between discontinuing these medications and

flares later in pregnancy, however, this study had limited power. Second, a retrospective,

multicenter study in IBD patients compared 187 thiopurine-exposed pregnancies compared to

318 non-exposed pregnancies (Casanova et al., 2013). The rate of unfavourable global pregnancy

outcome (those who developed obstetric complications including LBW, preterm labour,

spontaneous abortion, congenital malformations, or required ICU admission) was lower in the

thiopurine exposed pregnancies compared to the non-exposed controls (21.9% vs 31.8%,

p=0.01). Third, a prospective, multicenter study of children born to mothers with IBD assessed

the long-term effects of in utero thiopurine exposure (de Meij et al., 2013). Compared to a

control group (n=340), those children exposed to thiopurines (n=30) had no increased

susceptibility to infection or immunodeficiency and no difference in global medical and

psychosocial health status up to six years of age. Together, these studies alongside our results

support continuing AZA/6-MP therapy during pregnancy for IBD as the benefit of the treatment

outweighs the risk of untreated maternal disease (Ban et al., 2014).

A treatment algorithm for IBD in pregnancy has been recently proposed and recommends

AZA/6-MP as a safe pharmacologic choice in pregnancy (Habal & Huang, 2012). Considering

that maternal pharmacokinetic factors have the largest influence on fetal exposure to thiopurines,

our findings support the recommendations by other groups that maternal drug monitoring of

thiopurine levels may lead to avoidance of high levels of fetal drug exposure (de Boer et al.,

2006; Jharap et al., 2014). The effect of pregnancy on the pharmacokinetics of 6-MP has only

recently been studied. In a small cohort of 30 pregnant women with IBD receiving AZA or 6-MP

determined that 6-TGN levels may decrease throughout pregnancy but return to baseline after

delivery (Jharap et al., 2014). Future studies should determine whether there is a role for

therapeutic drug monitoring of 6-TGN levels in pregnancy or the role of monitoring maternal

CBC and liver enzymes (Jharap et al., 2014; Katz, 2013). Detecting any changes in maternal

Page 137: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

121

CBC or liver enzymes may be a suitable surrogate marker for potential elevated 6-TGN levels as

toxic changes would start to be observed. Furthermore, monitoring maternal CBC and liver

enzymes would be available at any hospital laboratory and would not require specialized

equipment as would be required for therapeutic drug monitoring of 6-TGN.

III. To evaluate the effect of alcohol on folic acid transfer across the human placenta

and the potential toxicity of formic acid to the placenta (Chapters 5 & 6).

Compared to the study of therapeutic drug safety in pregnancy, similar experimental

approaches involving the placental perfusion model together with clinical studies can be utilized

to study known teratogens or toxins, such as alcohol. The placental perfusion model was

previously used to elucidate the mechanisms of folate transport across the placenta (Henderson et

al., 1995). Animal and in vitro studies were used to evaluate the mechanisms by which alcohol

can influence expression of folate transport proteins (Fisher et al., 1985; Keating et al., 2008;

Keating et al., 2009). Our cohort study comparing maternal blood and cord blood folate levels in

vivo demonstrated that chronic and heavy alcohol use in pregnancy impairs folate transport to

the fetus. When used together, these techniques are complimentary and allow us to better

understand how alcohol can lead to changes in folate transport across the placenta.

• Serum folate was measured in maternal blood and umbilical cord blood at the time of

delivery in pregnancies with chronic and heavy alcohol exposure (n=23) and in non-

drinking controls (n=24). In the alcohol-exposed pairs, the fetal:maternal serum folate

ratio was <1.0 in over half (n = 14), whereas all but one of the controls were >1.0. Mean

folate in cord samples was lower in the alcohol-exposed group than in the controls (33.15

± 19.89 vs 45.91 ± 20.73, p = 0.04) demonstrating that chronic and heavy alcohol use in

pregnancy impairs folate transport to the fetus.

• In the dually perfused placental perfusion model, formic acid at clinically relevant

concentrations transferred rapidly from the maternal to the fetal circulation, and transfer

was not altered with the addition of folate. The addition of folate did not alter fetal

exposure of formic acid originating from the maternal circulation, as measured by AUC

(1.30 ± 0.14 without folate; 1.23 ± 0.48 with folate; P = 0.79), or tissue concentration of

Page 138: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

122

formic acid as measured in the perfused lobules (425.83 ± 57.18 and 431.18 ± 133.07

nmol/g for perfusions without and with folate added, respectively).

• In the dually perfused placental perfusion model, there was a significant decrease in hCG

secretion (p = 0.03) after addition of formic acid and this decrease was mitigated by the

addition of folic acid

• Formic acid decreases hCG secretion in the placenta, which may alter steroidogenesis

and differentiation of the cytotrophoblasts, and this adverse effect can be mitigated by

folate.

Our results from a small cohort of heavy drinking women showed that the majority of infants

have decreased serum folate concentrations compared to the maternal concentrations. Since it is

well known that the placenta concentrates folate into the fetal circulation and fetal concentrations

are normally 2- to 4-fold higher than maternal (Economides et al., 1992; Giugliani et al., 1985;

Guerra-Shinohara et al., 2004; Malloy et al., 2005; Navarro et al., 1984; Obeid et al., 2005;

Relton et al., 2005; Stark et al., 2007; Thorand et al., 1996), our data demonstrates that alcohol

alters normal placental folate transfer to the fetus. We hypothesize that the mechanism is

decreased expression of folate transport proteins in the placenta as this is observed in animal and

in vitro models after alcohol exposure as discussed in Chapter 5. Our results were the first to

show this finding in human pregnancies and were in agreement with our original hypothesis that

heavy alcohol exposure will decrease folate transport to the fetus. However, we were unable to

obtain the placentas from the deliveries in our cohort to look at protein expression of PFRs in

order to investigate the mechanism. Since many women who drink heavily in pregnancy receive

less prenatal care and many women do not admit to drinking in pregnancy because of the

associated social stigma, this is a difficult population to recruit and this was our experience in

our attempt to further this study. Furthermore, women can deliver at any time of the day or night

and at any institution and the placenta is often sent to pathology and is thus unavailable for

research purposes. Therefore, it is logistically difficult to collect tissue samples from this patient

population. A large collaborative effort with multiple institutions will be needed to recruit an

adequate number of human placentae from pregnancies with heavy alcohol exposure. If placental

samples are available, folate binding to FR-α can be determined by preparation of microvillous

membranes and binding can be measured by incubating the membranes with [3H]-folic acid after

Page 139: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

123

an initial wash-out of endogenous folate using the method by Bisseling et al. (2004). Binding

characteristics should be compared to control placentae from births to women who do not use

alcohol during pregnancy. In addition to measuring binding, the amount of FR-α in the

microvillous membranes should be quantified using western blot. The amount of RFC in the

basolateral membrane vesicles should also be determined by western blot. Amounts of FR-α and

RFC should also be compared to control placentae.

The results of altering folic acid levels both within the placenta and to the fetus may have

many consequences since folic acid is critically important for DNA synthesis and cell

proliferation as well as in cellular protection through its antioxidant properties (Antony, 2007).

Our results showed that total folate concentrations are decreased in fetal umbilical cord blood in

those born to mothers drinking heavily compared to controls. The decrease in mean fetal folate

concentrations in the exposed group was 12 nM compared to controls. It is unknown whether this

decreased concentration is ultimately clinically significant on its effect on fetal development.

Lower folate status has been associated with hyperactivity, peer problems, and lower cognitive

function (Schlotz et al., 2010; Veena et al., 2010) but folate status in these studies was

determined by maternal folate concentrations and thus are not directly comparable.

Folate deficiency within the placenta has also been shown to alter placental DNA

methylation (Kim et al., 2009). Fetal growth and development are closely linked to DNA

methylation and disruption in placental and fetal DNA methylation may have a role in the

Developmental Origins of Adult Health and Disease hypothesis (Gallou-Kabani et al., 2010).

Alterations in the placenta have already been associated with the future risk of cardiovascular

disease and cancer (Gallou-Kabani et al., 2010). In fact, a recent study reported long-lasting

alterations in DNA methylation as a result of fetal alcohol exposure in a mouse model (Laufer et

al., 2013). Genomic imprinting by DNA methylation acts only on a select set of genes that are

important in early development, particularly neurodevelopment (Hager & Johnstone, 2003;

Laufer et al., 2013). Interestingly, in addition to folic acid supplementation in animal models,

treatments that contain other methyl group donors such as choline have been able to attenuate

some of the effects of alcohol on the fetus (Cano et al., 2001; Garcia-Rodriguez et al., 2003;

Wang et al., 2009; Wentzel & Eriksson, 2008; Xu et al., 2008; Yanaguita et al., 2008; Zeisel,

2011). Future studies should aim to elucidate the influence of altered folate transport both on

Page 140: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

124

placental and fetal epigenetic patterns as this may shed light on the mechanism of alcohol

teratogenicity.

Folate is routinely given to pregnant women during pregnancy. High dose folate

supplementation (5mg/day) has been recommended by the Society of Obstetricians and

Gynaecologists of Canada to women with recreational drug use (including alcohol) because of

concerns surrounding adequate diet and compliance with prenatal vitamins (Wilson et al., 2007).

Since our results demonstrate that folate transport is decreased to the fetus after heavy alcohol

exposure, it would be beneficial to determine the optimal dose of folate supplementation that

may prevent any associated fetal effects. Animal studies have suggested that high dose folate

supplementation may mitigate adverse fetal effects induced by alcohol as discussed in Chapter 5

and thus it may be possible to overcome the decreased placental transfer. This would require a

large-scale clinical study to evaluate with adequate power to control for the numerous

confounding variables such as diet, socioeconomic status, other drug use, etc. This study would

also only be useful if the fetal effects directly resulting from alcohol altering placental transport

and binding are known as discussed above.

In addition to the direct toxic effects of alcohol on the placenta and fetus, our studies also

evaluated the role that formic acid, the toxic metabolite of methanol, may have in the

development of FASDs. Formic acid has recently been detected in maternal and cord blood in

pregnancies with chronic ethanol exposure (Kapur et al., 2009). Using the dually perfused

placental perfusion model, we determined that formic acid transferred rapidly from the maternal

to the fetal circulation, and transfer was not altered with the addition of folate. This was in

disagreement with our hypothesis. We hypothesized that folate would decrease transfer of formic

acid as it is known that folate can influence the rate of clearance of formic acid (Johlin et al.,

1989; Sokoro et al., 2008). This could be a limitation of the perfusion model as it cannot account

for potential changes on formic acid clearance after chronic exposure and accumulation of folate

reserves in the placenta (Henderson et al., 1995). Because the transfer of formic acid across the

placenta at clinically relevant concentrations is rapid, the developing fetus will be exposed to

formic acid from the maternal circulation. Since the perfusion model accurately predicts drug

transfer in term human placentas as determined in Chapter 2, it is likely that the fetus will rapidly

be exposed to formic acid from a heavy drinking mother.

Page 141: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

125

We also determined the direct toxic effects of formic acid to the placenta during the placental

perfusion. Our results showed that there was a significant decrease in placental hCG secretion

after addition of formic acid and this decrease was mitigated by the addition of folic acid. This

was in agreement with our hypothesis. Alterations in hCG secretion by the placenta may have

consequential implications clinically since hCG is vitally important in maintaining the

pregnancy, regulating blood flow, and delivering nutrients to the conceptus (Boal et al., 1997).

The mechanism by which formic acid decreases hCG secretion should be determined in future

studies in addition to whether chronic exposure to methanol or formic acid has adverse placental

effects. The later will have to be completed in animal or in vitro studies since it is unethical to

study in humans.

Our research is the first to propose a role of methanol and formic acid in the development of

FASDs. The involvement of methanol in the toxic effect of alcoholic beverages is a novel

hypothesis proposed by Kapur et al. (2009). The proposed interaction of how alcohol and

methanol could lead to formic acid toxicity and also decreased placental folate is outlined in

Figure 7.1. Further studies are needed to fully understand how the toxicity of methanol/formic

acid plays a role together in the FASDs. Here, there is an important role for animal and in vitro

studies in addition to the validated placental perfusion model as it is difficult to recruit pregnant

women with chronic and heavy alcohol consumption.

Page 142: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

126

Figure 7-1 The interaction between alcohol use in pregnancy and the proposed mechanisms

of toxicity by reduced placental folate transfer to the fetus and by formic acid.

Page 143: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

127

7.2 Summary

In summary, our studies illustrate the benefits of using an integrative approach in the

desire to understand the safety of therapeutics or the toxicity of xenobiotics in pregnancy. When

used together, techniques including animal, in vitro, placental perfusion ex vivo, and clinical

studies can capture the dynamic nature of pregnancy and we can make generalized conclusions

and recommendations. Furthermore, meta-analyses offer clinicians, scientists, and regulators an

earlier signal for the presence or lack of teratogenic risk and hence can have an important impact

on clinical practice (Etwel et al., 2014). Since it is often unethical to study drug safety in

pregnant women, theoretical frameworks are needed on which to build the study of drugs in

human pregnancy. Our validation of the ex vivo perfusion model showed that it is effective in

predicting fetal exposure to drugs and should have a place in clinical and regulatory

pharmacology and toxicology. We then utilized the placental perfusion model in an integrative

approach to understand the safety of the therapeutic, 6-MP, and the toxicity of alcohol/methanol

in pregnancy. Our results from the placental perfusion model demonstrated that binding to

placental tissue and maternal pharmacokinetic parameters are the main factors that restrict

placental transfer of 6-MP. Active transport is unlikely to play a significant role and drug

interactions involving, or polymorphisms in, placental drug efflux transporters are not likely to

put the fetus at risk of higher 6-MP exposure. A meta-analysis showed that 6-MP does not

increase the risk of congential malformation when used during pregnancy and the benefit of

treating a woman with 6-MP and maintaining disease remission without the use of steroids

would outweigh any risk. Together, these findings support the safety of 6-MP in pregnancy and

woman and physicians can be reassured about its use in pregnancy. Using the validated perfusion

model, we also determined that formic acid transfers across the placenta rapidly and this transfer

is not influenced by folate. Pregnant women with heavy alcohol use are thus likely to exposure

the fetus to toxic levels of formic acid since we used clinically relevant concentrations of formic

acid that is found in heavy alcohol users. Formic acid was also found to be toxic to the placenta

at clinically relevant concentrations. Furthermore, we showed that placental transfer of folate is

decreased after heavy and chronic alcohol exposure. Since folate is required for formic

detoxification of formic acid, this may be a potential mechanism for the development of fetal

alcohol spectrum disorders.

Page 144: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

128

References 8

Abel EL. (1998). Fetal alcohol syndrome: the 'American Paradox'. Alcohol Alcohol 33 (3), 195-

201.

Abel EL, & Sokol RJ. (1986). Fetal alcohol syndrome is now leading cause of mental

retardation. Lancet 2 (8517), 1222.

Akbari M, Shah S, Velayos FS, Mahadevan U, & Cheifetz AS. (2013). Systematic review and

meta-analysis on the effects of thiopurines on birth outcomes from female and male

patients with inflammatory bowel disease. Inflamm Bowel Dis 19 (1), 15-22.

Akoum A, Metz CN, & Morin M. (2005). Marked increase in macrophage migration inhibitory

factor synthesis and secretion in human endometrial cells in response to human chorionic

gonadotropin hormone. J Clin Endocrinol Metab 90 (5), 2904-2910.

Ala-Kokko TI, Myllynen P, & Vahakangas K. (2000). Ex vivo perfusion of the human placental

cotyledon: Implications for anesthetic pharmacology. Int J Obstet Anesth 9 (1), 26-38.

Ala-Kokko TI, Alahuhta S, Jouppila P, Korpi K, Westerling P, & Vahakangas K. (1997). Feto-

maternal distribution of ropivacaine and bupivacaine after epidural administration for

cesarean section. Int J Obstet Anesth 6 (3), 147-152.

Ala-Kokko TI, Pienimaki P, Herva R, Hollmen AI, Pelkonen O, & Vahakangas K. (1995).

Transfer of lidocaine and bupivacaine across the isolated perfused human placenta.

Pharmacol Toxicol 77 (2), 142-148.

Alati R, Al Mamun A, Williams GM, O'Callaghan M, Najman JM, & Bor W. (2006). In utero

alcohol exposure and prediction of alcohol disorders in early adulthood: a birth cohort

study. Arch Gen Psychiatry 63 (9), 1009-1016.

Alstead EM, Ritchie JK, Lennard-Jones JE, Farthing MJ, & Clark ML. (1990). Safety of

azathioprine in pregnancy in inflammatory bowel disease. Gastroenterology 99 (2), 443-

446.

Page 145: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

129

Anderson GD. (2005). Pregnancy-induced changes in pharmacokinetics: a mechanistic-based

approach. Clin Pharmacokinet 44 (10), 989-1008.

Andrews JE, Ebron-Mccoy M, Schmid JE, & Svendsgaard D. (1998). Effects of combinations

of methanol and formic acid on rat embryos in culture. Teratology 58 (2), 54-61.

Angelberger S, Reinisch W, Messerschmidt A, Miehsler W, Novacek G, Vogelsang H et al.

(2011). Long-term follow-up of babies exposed to azathioprine in utero and via

breastfeeding. J Crohns Colitis 5 (2), 95-100.

Antony AC. (2007). In utero physiology: role of folic acid in nutrient delivery and fetal

development. Am J Clin Nutr 85 (2), 598S-603S.

Ardizzone S, Maconi G, Russo A, Imbesi V, Colombo E, & Bianchi Porro G. (2006).

Randomised controlled trial of azathioprine and 5-aminosalicylic acid for treatment of

steroid dependent ulcerative colitis. Gut 55 (1), 47-53.

Atkinson DE, Greenwood SL, Sibley CP, Glazier JD, & Fairbairn LJ. (2003). Role of MDR1

and MRP1 in trophoblast cells, elucidated using retroviral gene transfer. Am J Physiol Cell

Physiol 285 (3), C584-91.

Audus KL. (1999). Controlling drug delivery across the placenta. Eur J Pharm Sci 8 (3), 161-

165.

Augustowska K, Magnowska Z, Kapiszewska M, & Gregoraszczuk EL. (2007). Is the natural

PCDD/PCDF mixture toxic for human placental JEG-3 cell line? The action of the

toxicants on hormonal profile, CYP1A1 activity, DNA damage and cell apoptosis. Hum

Exp Toxicol 26 (5), 407-417.

Axelrod J, & Daly J. (1965). Pituitary gland: enzymic formation of methanol from S-

adenosylmethionine. Science 150 (698), 892-893.

Baiocco PJ, & Korelitz BI. (1984). The influence of inflammatory bowel disease and its

treatment on pregnancy and fetal outcome. J Clin Gastroenterol 6 (3), 211-216.

Page 146: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

130

Baird DD, Narendranathan M, & Sandler RS. (1990). Increased risk of preterm birth for women

with inflammatory bowel disease. Gastroenterology 99 (4), 987-994.

Ban L, Tata LJ, Fiaschi L, & Card T. (2014). Limited Risks of Major Congenital Anomalies in

Children of Mothers With IBD and Effects of Medications. Gastroenterology 146 (1), 76-

84.

Barak AJ, Beckenhauer HC, Tuma DJ. (1985) Ethanol feeding inhibits the activity of hepatic

N5-methyltetrahydrofolate-homocysteine methyltransferase in the rat. IRCS Medical

Science 13 (8), 760-761.

Bar-Oz B, Einarson T, Einarson A, Boskovic R, O'Brien L, Malm H et al. (2007). Paroxetine and

congenital malformations: meta-Analysis and consideration of potential confounding

factors. Clin Ther 29 (5), 918-926.

Bassily M, Ghabrial H, Smallwood RA, & Morgan DJ. (1995). Determinants of placental drug

transfer: Studies in the isolated perfused human placenta. J Pharm Sci 84 (9), 1054-1060.

Belpaire FM, Wynant P, Van Trappen P, Dhont M, Verstraete A, & Bogaert MG. (1995).

Protein binding of propranolol and verapamil enantiomers in maternal and foetal serum. Br

J Clin Pharmacol 39 (2), 190-193.

Bergan S, Rugstad HE, Bentdal O, Endresen L, & Stokke O. (1994). Kinetics of mercaptopurine

and thioguanine nucleotides in renal transplant recipients during azathioprine treatment.

Ther Drug Monit 16 (1), 13-20.

Berndt S, Blacher S, Perrier d'Hauterive S, Thiry M, Tsampalas M, Cruz A et al. (2009).

Chorionic gonadotropin stimulation of angiogenesis and pericyte recruitment. J Clin

Endocrinol Metab 94 (11), 4567-4574.

Biehl D, Shnider SM, Levinson G, & Callender K. (1978). Placental transfer of lidocaine:

effects of fetal acidosis. Anesthesiology 48 (6), 409-412.

Bisseling TM, Steegers EA, van den Heuvel JJ, Siero HL, van de Water FM, Walker AJ et al.

(2004). Placental folate transport and binding are not impaired in pregnancies complicated

by fetal growth restriction. Placenta 25 (6), 588-593.

Page 147: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

131

Boal JH, Plessinger MA, van den Reydt C, & Miller RK. (1997). Pharmacokinetic and toxicity

studies of AZT (zidovudine) following perfusion of human term placenta for 14 hours.

Toxicol Appl Pharmacol 143 (1), 13-21.

Bortoli A, Saibeni S, Tatarella M, Prada A, Beretta L, Rivolta R et al. (2007). Pregnancy before

and after the diagnosis of inflammatory bowel diseases: retrospective case-control study. J

Gastroenterol Hepatol 22 (4), 542-549.

Bosco C, & Diaz E. (2012). Placental hypoxia and foetal development versus alcohol exposure

in pregnancy. Alcohol Alcohol 47 (2), 109-117.

Bragonier JR, Roesky N, Carver MJ. (1964) Teratogenesis: Effects of substituted purines and the

influence of 4-hydroxypyrazolopyrimi- dine in the rat. Proc Soc Exp Biol Med 116, 685–

688.

Brien JF, Loomis CW, Tranmer J, & McGrath M. (1983). Disposition of ethanol in human

maternal venous blood and amniotic fluid. Am J Obstet Gynecol 146 (2), 181-186.

Brown-Woodman PD, Huq F, Hayes L, Herlihy C, Picker K, & Webster WS. (1995). In vitro

assessment of the effect of methanol and the metabolite, formic acid, on embryonic

development of the rat. Teratology 52 (4), 233-243.

Burd L, Roberts D, Olson M, & Odendaal H. (2007). Ethanol and the placenta: A review. J

Matern Fetal Neonatal Med 20 (5), 361-375.

Burney RG, DiFazio CA, & Foster JA. (1978). Effects of pH on protein binding of lidocaine.

Anesth Analg 57 (4), 478-480.

Bush MC, Patel S, Lapinski RH, & Stone JL. (2004). Perinatal outcomes in inflammatory bowel

disease. J Matern Fetal Neonatal Med 15 (4), 237-241.

Butt P, Beirness D, Gliksman L, Paradis C, and Stockwell T. (2011). Alcohol and health in

Canada: A summary of evidence and guidelines for low risk drinking (Ottawa, ON:

Canadian Centre on Substance Abuse).

Page 148: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

132

Cabezon C, De la Fuente F, Jurado M, & Lopez G. (1985). Histometry of the placental structures

involved in the respiratory interchange. Acta Obstet Gynecol Scand 64 (5), 411-416.

Campbell S, & Ghosh S. (2000). Safety of azathioprine use in pregnancy. Gut 46, A13.

Cano MJ, Ayala A, Murillo ML, & Carreras O. (2001). Protective effect of folic acid against

oxidative stress produced in 21-day postpartum rats by maternal-ethanol chronic

consumption during pregnancy and lactation period. Free Radic Res 34 (1), 1-8.

Casanova MJ, Chaparro M, Domenech E, Barreiro-de Acosta M, Bermejo F, Iglesias E et al.

(2013). Safety of thiopurines and anti-TNF-alpha drugs during pregnancy in patients with

inflammatory bowel disease. Am J Gastroenterol 108 (3), 433-440.

Ceckova-Novotna M, Pavek P, & Staud F. (2006). P-glycoprotein in the placenta: expression,

localization, regulation and function. Reprod Toxicol 22 (3), 400-410.

Centers for Disease Control and Prevention (2012). Frequently Asked Questions - Alcohol.

Available at: http://www.cdc.gov/alcohol/faqs.htm#standDrink (accessed 2013 Sept. 6)

Chan GL, Canafax DM, & Johnson CA. (1987). The therapeutic use of azathioprine in renal

transplantation. Pharmacotherapy 7 (5), 165-177.

Chance B. (1950) On the Reaction of Catalase Peroxide with Acceptors. J Biol Chem 182, 649-

658.

Chappuy H, Treluyer JM, Jullien V, Dimet J, Rey E, Fouche M et al. (2004). Maternal-fetal

transfer and amniotic fluid accumulation of nucleoside analogue reverse transcriptase

inhibitors in human immunodeficiency virus-infected pregnant women. Antimicrob Agents

Chemother 48 (11), 4332-4336.

Charles B, Norris R, Xiao X, & Hague W. (2006). Population pharmacokinetics of metformin in

late pregnancy. Ther Drug Monit 28 (1), 67-72.

Chaub S, Murphy ML. (1968) The teratogenic effects of the recent drugs active in cancer

chemotherapy. Adv Teratol 3, 181–237.

Page 149: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

133

Chiao F, Stokstad EL. (1977) Effect of methionine on the metabolism of formate and histidine

by rats fed folate/vitamin B-12-methionine-deficient diet. Biochimica et Biophysica Acta

497 (1), 225-33.

Chiloiro M, Darconza G, Piccioli E, De Carne M, Clemente C, & Riezzo G. (2001). Gastric

emptying and orocecal transit time in pregnancy. J Gastroenterol 36 (8), 538-543.

Chudley AE, Conry J, Cook JL, Loock C, Rosales T, & LeBlanc N. (2005). Fetal alcohol

spectrum disorder: Canadian guidelines for diagnosis. CMAJ 172 (5 Suppl), S1-S21.

Cleary BJ, & Kallen B. (2009). Early pregnancy azathioprine use and pregnancy outcomes.

Birth Defects Res A Clin Mol Teratol 85 (7), 647-654.

Coelho J, Beaugerie L, Colombel JF, Hebuterne X, Lerebours E, Lemann M et al. (2011).

Pregnancy outcome in patients with inflammatory bowel disease treated with thiopurines:

cohort from the CESAME Study. Gut 60 (2), 198-203.

Cole LA. (2010). Biological functions of hCG and hCG-related molecules. Reprod Biol

Endocrinol 8, 102.

Cooper GS, & Stroehla BC. (2003). The epidemiology of autoimmune diseases. Autoimmun Rev

2 (3), 119-125.

Cornish J, Tan E, Teare J, Teoh TG, Rai R, Clark SK et al. (2007). A meta-analysis on the

influence of inflammatory bowel disease on pregnancy. Gut 56 (6), 830-837.

Cronier L, Bastide B, Herve JC, Deleze J, & Malassine A. (1994). Gap junctional

communication during human trophoblast differentiation: influence of human chorionic

gonadotropin. Endocrinology 135 (1), 402-408.

Dai NT, Yeh MK, Liu DD, Adams EF, Chiang CH, Yen CY et al. (2005). A co-cultured skin

model based on cell support membranes. Biochem Biophys Res Commun 329 (3), 905-908.

Datta S, Camann W, Bader A, & VanderBurgh L. (1995). Clinical effects and maternal and fetal

plasma concentrations of epidural ropivacaine versus bupivacaine for cesarean section.

Anesthesiology 82 (6), 1346-1352.

Page 150: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

134

Davis AR, Miller L, Tamimi H, & Gown A. (1999). Methotrexate compared with

mercaptopurine for early induced abortion. Obstet Gynecol 93 (6), 904-909.

Davison JM, & Hytten FE. (1974). Glomerular filtration during and after pregnancy. J Obstet

Gynaecol Br Commonw 81 (8), 588-595.

De Barros Duarte L, Moises EC, Cavalli RC, Lanchote VL, Duarte G, & Da Cunha SP. (2010).

Distribution of Bupivacaine Enantiomers and Lidocaine and its Metabolite in the Placental

Intervillous Space and in the Different Maternal and Fetal Compartments in Term Pregnant

Women. J Clin Pharmacol 51 (2), 212-217.

de Boer NK, Jarbandhan SV, de Graaf P, Mulder CJ, van Elburg RM, & van Bodegraven AA.

(2006). Azathioprine use during pregnancy: unexpected intrauterine exposure to

metabolites. Am J Gastroenterol 101 (6), 1390-1392.

de Boer NK, Van Elburg RM, Wilhelm AJ, Remmink AJ, Van Vugt JM, Mulder CJ et al.

(2005). 6-Thioguanine for Crohn's disease during pregnancy: thiopurine metabolite

measurements in both mother and child. Scand J Gastroenterol 40 (11), 1374-1377.

de Graaf P, Vos RM, de Boer NH, Sinjewel A, Jharap B, Mulder CJ et al. (2010). Limited

stability of thiopurine metabolites in blood samples: Relevant in research and clinical

practise. J Chromatogr B Analyt Technol Biomed Life Sci 878 (19), 1437-1442.

de Meij TG, Jharap B, Kneepkens CM, van Bodegraven AA, & de Boer NK. (2013). Long-term

follow-up of children exposed intrauterine to maternal thiopurine therapy during pregnancy

in females with inflammatory bowel disease. Aliment Pharmacol Ther 38 (1), 38-43.

de Wolf C, Jansen R, Yamaguchi H, de Haas M, van de Wetering K, Wijnholds J et al. (2008).

Contribution of the drug transporter ABCG2 (breast cancer resistance protein) to resistance

against anticancer nucleosides. Mol Cancer Ther 7 (9), 3092-3102.

Dell C, and Roberts G. (2006). Alcohol use and pregnancy: An important Canadian public health

and social issue (Ottawa, ON: Public Health Agency of Canada). Available at:

http://www.phac- aspc.gc.ca/publicat/fasd-ru-ectaf-pr-06/pdf/fasd-ru-ectaf-pr-06_e.pdf)

(accessed 2013 Sep. 10)

Page 151: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

135

Dejaco C, Angelberger S, Waldhoer T, Meihsler W, Wenzl H, Haas T et al. (2005). Pregnancy

and birth outcome under thiopurine therapy for inflammatory bowel disease (IBD).

Gastroenterology 128 (4), A12.

Derewlany LO, Leeder JS, Kumar R, Radde IC, Knie B, & Koren G. (1991). The transport of

digoxin across the perfused human placental lobule. J Pharmacol Exp Ther 256 (3), 1107-

1111.

Dominitz JA, Young JC, & Boyko EJ. (2002). Outcomes of infants born to mothers with

inflammatory bowel disease: a population-based cohort study. Am J Gastroenterol 97 (3),

641-648.

Economides DL, Ferguson J, Mackenzie IZ, Darley J, Ware II, & Holmes-Siedle M. (1992).

Folate and vitamin B12 concentrations in maternal and fetal blood, and amniotic fluid in

second trimester pregnancies complicated by neural tube defects. Br J Obstet Gynaecol 99

(1), 23-25.

Elad D, Levkovitz R, Jaffa AJ, Desoye G, & Hod M. (2014). Have we neglected the role of fetal

endothelium in transplacental transport? Traffic 15 (1), 122-126.

Ema M, Naya M, Yoshida K, & Nagaosa R. (2010). Reproductive and developmental toxicity of

degradation products of refrigerants in experimental animals. Reprod Toxicol 29 (1), 1-9.

Enders AC. (1981). Anatomy of the placenta and its relationship to function. Mead Johnson

Symp Perinat Dev Med, 18, 3-7.

Etwel F, Hutson JR, Madadi P, Gareri J, & Koren G. (2014). Fetal and perinatal exposure to

drugs and chemicals: novel biomarkers of risk. Annu Rev Pharmacol Toxicol 54, 295-315.

Evseenko DA, Murthi P, Paxton JW, Reid G, Emerald BS, Mohankumar KM et al. (2007). The

ABC transporter BCRP/ABCG2 is a placental survival factor, and its expression is reduced

in idiopathic human fetal growth restriction. FASEB Journal 21 (13), 3592-3605.

Eyal S, Easterling TR, Carr D, Umans JG, Miodovnik M, Hankins GD et al. (2010).

Pharmacokinetics of metformin during pregnancy. Drug Metab Dispos 38 (5), 833-840.

Page 152: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

136

Fedorkow DM, Persaud D, & Nimrod CA. (1989). Inflammatory bowel disease: a controlled

study of late pregnancy outcome. Am J Obstet Gynecol 160 (4), 998-1001.

Finkelstein JD, Cello JP, Kyle WE. (1974) Ethanol induced changes in methionine metabolism

in rat liver. Biochem Biophys Res Commun 61 (2), 525-531.

Finnegan L. (2013). Licit and illicit drug use during pregnancy: maternal, neonatal and early

childhood consequences. Available: www.ccsa.ca/./CCSA-Drug-Use-during-Pregnancy-

Report-2013-en.pdf .

Finster M, Morishima HO, Boyes RN, & Covino BG. (1972a). The placental transfer of

lidocaine and its uptake by fetal tissues. Anesthesiology 36 (2), 159-163.

Finster M, Morishima HO, Mark LC, Perel JM, Dayton PG, & James LS. (1972b). Tissue

thiopental concentrations in the fetus and newborn. Anesthesiology 36 (2), 155-158.

Fisher SE, Inselman LS, Duffy L, Atkinson M, Spencer H, & Chang B. (1985). Ethanol and

fetal nutrition: effect of chronic ethanol exposure on rat placental growth and membrane-

associated folic acid receptor binding activity. J Pediatr Gastroenterol Nutr 4 (4), 645-649.

Francella A, Dyan A, Bodian C, Rubin P, Chapman M, & Present DH. (2003). The safety of 6-

mercaptopurine for childbearing patients with inflammatory bowel disease: a retrospective

cohort study. Gastroenterology 124 (1), 9-17.

Frederiksen MC, Ruo TI, Chow MJ, & Atkinson AJJ. (1986). Theophylline pharmacokinetics in

pregnancy. Clin Pharmacol Ther 40 (3), 321-328.

Froescher W, Gugler R, Niesen M, & Hoffmann F. (1984). Protein binding of valproic acid in

maternal and umbilical cord serum. Epilepsia 25 (2), 244-249.

Gallou-Kabani C, Gabory A, Tost J, Karimi M, Mayeur S, Lesage J et al. (2010). Sex- and diet-

specific changes of imprinted gene expression and DNA methylation in mouse placenta

under a high-fat diet. PLoS ONE 5 (12), e14398.

Page 153: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

137

Ganapathy V, Prasad PD, Ganapathy ME, & Leibach FH. (2000). Placental transporters relevant

to drug distribution across the maternal-fetal interface. J Pharmacol Exp Ther 294 (2), 413-

420.

Garcia-Rodriguez S, Arguelles S, Llopis R, Murillo ML, Machado A, Carreras O et al. (2003).

Effect of prenatal exposure to ethanol on hepatic elongation factor-2 and proteome in 21 d

old rats: protective effect of folic acid. Free Radic Biol Med 35 (4), 428-437.

Garland M. (1998). Pharmacology of drug transfer across the placenta. Obstet Gynecol Clin

North Am 25 (1), 21-42.

Garland M, Abildskov KM, Taylor S, Benzeroual K, Caspersen CS, Arroyo SE et al. (2006).

Fetal morphine metabolism and clearance are constant during late gestation. Drug Metab

Dispos 34 (4), 636-646.

Gavard L, Beghin D, Forestier F, Cayre Y, Peytavin G, Mandelbrot L et al. (2009). Contribution

and limit of the model of perfused cotyledon to the study of placental transfer of drugs.

Example of a protease inhibitor of HIV: nelfinavir. Eur J Obstet Gynecol Reprod Biol 147

(2), 157-160.

Gavard L, Gil S, Peytavin G, Ceccaldi PF, Ferreira C, Farinotti R et al. (2006). Placental transfer

of lopinavir/ritonavir in the ex vivo human cotyledon perfusion model. Am J Obstet

Gynecol 195 (1), 296-301.

Gedeon C, Anger G, Lubetsky A, Miller MP, & Koren G. (2008). Investigating the potential

role of multi-drug resistance protein (MRP) transporters in fetal to maternal glyburide

efflux in the human placenta. J Obstet Gynaecol 28 (5), 485-489.

Gendron MP, Martin B, Oraichi D, & Berard A. (2009). Health care providers' requests to

Teratogen Information Services on medication use during pregnancy and lactation. Eur J

Clin Pharmacol 65 (5), 523-531.

Gepts E, Heytens L, & Camu F. (1986). Pharmacokinetics and placental transfer of intravenous

and epidural alfentanil in parturient women. Anesth Analg 65 (11), 1155-1160.

Page 154: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

138

Gerdin E, Rane A, & Lindberg B. (1990). Transplacental transfer of morphine in man. J Perinat

Med 18 (4), 305-312.

Gil S, Saura R, Forestier F, & Farinotti R. (2005). P-glycoprotein expression of the human

placenta during pregnancy. Placenta 26 (2-3), 268-270.

Giroux M, Teixera MG, Dumas JC, Desprats R, Grandjean H, & Houin G. (1997). Influence of

maternal blood flow on the placental transfer of three opioids--fentanyl, alfentanil,

sufentanil. Biol Neonate 72 (3), 133-141.

Githens JH, Rosenkrantz JG, Tunnock SM. (1965) Teratogenic effects of azathioprine (Imuran).

J Pediatr 66, 959–961.

Giugliani ER, Jorge SM, & Goncalves AL. (1985). Serum and red blood cell folate levels in

parturients, in the intervillous space of the placenta and in full-term newborns. J Perinat

Med 13 (2), 55-59.

Goh YI, & Koren G. (2008). Folic acid in pregnancy and fetal outcomes. J Obstet Gynaecol 28

(1), 3-13.

Goldstein LH, Dolinsky G, Greenberg R, Schaefer C, Cohen-Kerem R, Diav-Citrin O et al.

(2007). Pregnancy outcome of women exposed to azathioprine during pregnancy. Birth

Defects Res A Clin Mol Teratol 79 (10), 696-701.

Goodlett CR, Horn KH, & Zhou FC. (2005). Alcohol teratogenesis: mechanisms of damage and

strategies for intervention. Exp Biol Med 230 (6), 394-406.

Gorodischer R, Krasner J, & Yaffe SJ. (1974). Serum protein binding of digoxin in newborn

infants. Res Commun Chem Pathol Pharmacol 9 (2), 387-390.

Govindarajan R, Bakken AH, Hudkins KL, Lai Y, Casado FJ, Pastor-Anglada M et al. (2007).

In situ hybridization and immunolocalization of concentrative and equilibrative nucleoside

transporters in the human intestine, liver, kidneys, and placenta. Am J Physiol Regul Integr

Comp Physiol 293 (5), 1809-1822.

Page 155: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

139

Guerra-Shinohara EM, Morita OE, Peres S, Pagliusi RA, Sampaio Neto LF, D'Almeida V et al.

(2004). Low ratio of S-adenosylmethionine to S-adenosylhomocysteine is associated with

vitamin deficiency in Brazilian pregnant women and newborns. Am J Clin Nutr 80 (5),

1312-1321.

Gundogan F, Elwood G, Mark P, Feijoo A, Longato L, Tong M et al. (2010). Ethanol-Induced

Oxidative Stress and Mitochondrial Dysfunction in Rat Placenta: Relevance to Pregnancy

Loss. Alcohol Clin Exp Res 34, 415-423.

Gutierrez CM, Ribeiro CN, de Lima GA, Yanaguita MY, & Peres LC. (2007). An experimental

study on the effects of ethanol and folic acid deficiency, alone or in combination, on

pregnant Swiss mice. Pathology 39 (5), 495-503.

Habal FM, & Huang VW. (2012). Review article: a decision-making algorithm for the

management of pregnancy in the inflammatory bowel disease patient. Aliment Pharmacol

Ther 35 (5), 501-515.

Hager R, & Johnstone RA. (2003). The genetic basis of family conflict resolution in mice.

Nature 421 (6922), 533-535.

Halsted CH, Villanueva JA, Devlin AM et al. (2002a) Folate deficiency disturbs hepatic

methionine metabolism and promotes liver injury in the ethanol-fed micropig. Proc Natl

Acad Sci U S A 99 (15), 10072-10077.

Halsted CH, Villanueva JA, Devlin AM, & Chandler CJ. (2002b). Metabolic interactions of

alcohol and folate. J Nutr 132 (8 Suppl), 2367S-2372S.

Hamar C, & Levy G. (1980). Factors affecting the serum protein binding of salicylic acid in

newborn infants and their mothers. Pediatr Pharmacol 1 (1), 31-43.

Hanauer SB. (1999). Updating the approach to Crohn's disease. Hosp Pract 34 (8), 77-78, 81-

83, 87-93.

Hansen JM, Contreras KM, & Harris C. (2005). Methanol, formaldehyde, and sodium formate

exposure in rat and mouse conceptuses: a potential role of the visceral yolk sac in

embryotoxicity. Birth Defects Res A Clin Mol Teratol 73 (2), 72-82.

Page 156: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

140

Haram K, Bakke OM, Johannessen KH, & Lund T. (1978). Transplacental passage of diazepam

during labor: influence of uterine contractions. Clin Pharmacol Ther 24 (5), 590-599.

Harris C, Dixon M, & Hansen JM. (2004). Glutathione depletion modulates methanol,

formaldehyde and formate toxicity in cultured rat conceptuses. Cell Biol Toxicol 20 (3),

133-145.

Hawwa AF, Millership JS, Collier PS, & McElnay JC. (2009). Development and validation of

an HPLC method for the rapid and simultaneous determination of 6-mercaptopurine and

four of its metabolites in plasma and red blood cells. J Pharm Biomed Anal 49 (2), 401-

409.

He Y-L, Seno H, Tsujimoto S, & Tashiro C. (2001). The effects of uterine and umbilical blood

flows on the transfer of propofol across the human placenta during in vitro perfusion.

Anesth Analg 93 (1), 151-156.

Hebert MF, Ma X, Naraharisetti SB, Krudys KM, Umans JG, Hankins GD et al. (2009). Are we

optimizing gestational diabetes treatment with glyburide? The pharmacologic basis for

better clinical practice. Clin Pharmacol Ther 85 (6), 607-614.

Hemauer SJ, Patrikeeva SL, Nanovskaya TN, Hankins GD, & Ahmed MS. (2009). Opiates

inhibit paclitaxel uptake by P-glycoprotein in preparations of human placental inside-out

vesicles. Biochem Pharmacol 78 (9), 1272-1278.

Henderson GI, Perez T, Schenker S, Mackins J, & Antony AC. (1995). Maternal-to-fetal

transfer of 5-methyltetrahydrofolate by the perfused human placental cotyledon: evidence

for a concentrative role by placental folate receptors in fetal folate delivery. J Lab Clin Med

126 (2), 184-203.

Herbert V, Zalusky R, Davidson CS. (1963) Correlation of folate deficiency with alcoholism

and associated macrocytosis, anemia, and liver disease. Ann Intern Med 58, 977-988.

Herman NL, Li A-T, Van D,T.K., Johnson RF, Bjoraker RW, Downing JW et al. (2000).

Transfer of methohexital across the perfused human placenta. J Clin Anesth 12 (1), 25-30.

Page 157: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

141

Hermann DJ, Egan TD, & Muir KT. (1999). Influence of arteriovenous sampling on

remifentanil pharmacokinetics and pharmacodynamics. Clin Pharmacol Ther 65 (5), 511-

518.

Hermanns MI, Unger RE, Kehe K, Peters K, & Kirkpatrick CJ. (2004). Lung epithelial cell lines

in coculture with human pulmonary microvascular endothelial cells: development of an

alveolo-capillary barrier in vitro. Lab Invest 84 (6), 736-752.

Herr F, Baal N, Reisinger K, Lorenz A, McKinnon T, Preissner KT et al. (2007). HCG in the

regulation of placental angiogenesis. Results of an in vitro study. Placenta 28 (Suppl A),

S85-S93.

Hill MD, & Abramson FP. (1988). The significance of plasma protein binding on the

fetal/maternal distribution of drugs at steady-state. Clin Pharmacokinet 14 (3), 156-170.

Hines RN. (2007). Ontogeny of human hepatic cytochromes P450. J Biochem Mol Toxicol 21

(4), 169-175.

Hutson JR, Klieger C, & Koren G. (2012). Pharmacokinetics in Pregnancy. In Ginosar Y,

Reynolds F, Halpern SH (Ed) Anesthesia and the Fetus (1st ed., p. 53-62). Oxford, UK:

Wiley-Blackwell.

Hutson JR, Garcia-Bournissen F, Davis A, & Koren G. (2011). The human placental perfusion

model: a systematic review and development of a model to predict in vivo transfer of

therapeutic drugs. Clin Pharmacol Ther 90 (1), 67-76.

Hutson JR, Koren G, & Matthews SG. (2010). Placental P-glycoprotein and breast cancer

resistance protein: influence of polymorphisms on fetal drug exposure and physiology.

Placenta 31 (5), 351-357.

Hutson JR, Matlow JN, Moretti ME, & Koren G. (2013). The fetal safety of thiopurines for the

treatment of inflammatory bowel disease in pregnancy. J Obstet Gynaecol 33 (1), 1-8.

Hutson JR, Stade B, Lehotay DC, Collier CP, & Kapur BM. (2012). Folic acid transport to the

human fetus is decreased in pregnancies with chronic alcohol exposure. PLoS ONE 7 (5),

e38057.

Page 158: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

142

Jain V, & Gordon C. (2011). Managing pregnancy in inflammatory rheumatological diseases.

Arthritis Res Ther 13 (1), 206.

Jharap B, de Boer NK, Stokkers P, Hommes DW, Oldenburg B, Dijkstra G et al. (2014).

Intrauterine exposure and pharmacology of conventional thiopurine therapy in pregnant

patients with inflammatory bowel disease. Gut 63 (3), 451-457.

Jimenez-Solem E, Andersen JT, Petersen M, Broedbaek K, Jensen JK, Afzal S et al. (2012).

Exposure to selective serotonin reuptake inhibitors and the risk of congenital

malformations: a nationwide cohort study. BMJ Open 2 (3), e001148.

Johlin FC, Fortman CS, Nghiem DD, & Tephly TR. (1987). Studies on the role of folic acid and

folate-dependent enzymes in human methanol poisoning. Mol Pharmacol 31 (5), 557-561.

Johnson RF, Herman N, Arney TL, Gonzalez H, Johnson HV, & Downing JW. (1995).

Bupivacaine transfer across the human term placenta: A study using the dual perfused

human placental model. Anesthesiology 82 (2), 459-468.

Johnson RF, Herman N, Arney TL, Olenick M, Paschall RL, Johnson HV et al. (1999). Transfer

of lidocaine across the dual perfused human placental cotyledon. Int J Obstet Anesth 8 (1),

17-23.

Johnson RF, Herman NL, Johnson HV, Arney TL, Paschall RL, & Downing JW. (1996). Effects

of fetal pH on local anesthetic transfer across the human placenta. Anesthesiology 85 (3),

608-615.

Jones KL. (2006). Smith's recognizable patterns of human malformation. (6 ed., p. 704-705).

Philadelphia: WB Saunders.

Jonker JW, Smit JW, Brinkhuis RF, Maliepaard M, Beijnen JH, Schellens JH et al. (2000). Role

of breast cancer resistance protein in the bioavailability and fetal penetration of topotecan.

J Natl Cancer Inst 92 (20), 1651-1656.

Kalant H, and Khanna J. (2007). The Alcohols. In Principles of Medical Pharmacology,7th

edition. H Kalant, D Grant, and J Mitchell, eds. (Toronto, ON: Elsevier Canada), pp. 275–

288.

Page 159: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

143

Kapur BM, Vandenbroucke AC, Adamchik Y, Lehotay DC, & Carlen PL. (2007). Formic acid, a

novel metabolite of chronic ethanol abuse, causes neurotoxicity, which is prevented by

folic acid. Alcohol Clin Exp Res 31 (12), 2114-2120.

Kapur, BM, Khuu M, Bennett D, Tran S, Lehotay D, Eichhorst JC et al. (2009). Endogenous

methanol derived formic acid correlates with cognitive dysfunction in children born to

drinking mothers. Alcohol Clin Exp Res 33 (s1), 134A.

Karl PI, Divald A, Diehl AM, & Fisher SE. (1998). Altered cyclic AMP-dependent human

chorionic gonadotropin production in cultured human placental trophoblasts exposed to

ethanol. Biochem Pharmacol 55 (1), 45-51.

Karl PI, Divald A, & Fisher SE. (1994). Ethanol enhancement of ligand-stimulated cAMP

production by cultured human placental trophoblasts. Biochem Pharmacol 48 (7), 1493-

1500.

Karl PI, & Fisher SE. (1993). Ethanol alters hormone production in cultured human placental

trophoblasts. Alcohol Clin Exp Res 17 (4), 816-821.

Katz S. (2013). My treatment approach to the management of ulcerative colitis. Mayo Clin Proc

88 (8), 841-853.

Keating E, Goncalves P, Campos I, Costa F, & Martel F. (2009). Folic acid uptake by the human

syncytiotrophoblast: interference by pharmacotherapy, drugs of abuse and pathological

conditions. Reprod Toxicol 28 (4), 511-520.

Keating E, Lemos C, Goncalves P, & Martel F. (2008). Acute and chronic effects of some

dietary bioactive compounds on folic acid uptake and on the expression of folic acid

transporters by the human trophoblast cell line BeWo. J Nutr Biochem 19 (2), 91-100.

Khan ZH, Mayberry JF, Spiers N, & Wicks AC. (2000). Retrospective case series analysis of

patients with inflammatory bowel disease on azathioprine. A district general hospital

experience. Digestion 62 (4), 249-254.

Kim JM, Hong K, Lee JH, Lee S, & Chang N. (2009). Effect of folate deficiency on placental

DNA methylation in hyperhomocysteinemic rats. J Nutr Biochem 20 (3), 172-176.

Page 160: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

144

Kopecky EA. (1999). Transplacental pharmacology and toxicology of morphine in humans: in

vitro and in vivo studies. Dissertation, University of Toronto.

Kopecky EA, Ryan ML, Barrett JF, Seaward PG, Ryan G, Koren G et al. (2000). Fetal response

to maternally administered morphine. Am J Obstet Gynecol 183 (2), 424-430.

Koren G, Pastuszak A, & Ito S. (1998). Drugs in pregnancy. N Engl J Med 338 (16), 1128-1137.

Koukouritaki SB, Manro JR, Marsh SA, Stevens JC, Rettie AE, McCarver DG et al. (2004).

Developmental expression of human hepatic CYP2C9 and CYP2C19. J Pharmacol Exp

Ther 308 (3), 965-974.

Kraemer J, Klein J, Lubetsky A, & Koren G. (2006). Perfusion studies of glyburide transfer

across the human placenta: implications for fetal safety. Am J Obstet Gynecol 195 (1), 270-

274.

Krauer B, Dayer P, & Anner R. (1984). Changes in serum albumin and alpha 1-acid

glycoprotein concentrations during pregnancy: an analysis of fetal-maternal pairs. Br J

Obstet Gynaecol 91 (9), 875-881.

Kuehn D, Aros S, Cassorla F, Avaria M, Unanue N, Henriquez C et al. (2012). A prospective

cohort study of the prevalence of growth, facial, and central nervous system abnormalities

in children with heavy prenatal alcohol exposure. Alcohol Clin Exp Res 36 (10), 1811-

1819.

Kuhnz W, Steldinger R, & Nau H. (1984). Protein binding of carbamazepine and its epoxide in

maternal and fetal plasma at delivery: comparison to other anticonvulsants. Dev Pharmacol

Ther 7 (1), 61-72.

Kury G, Chaube S, Murphy ML. (1968) Teratogenic effects of some purine analogues on fetal

rat. Arch Pathol 86, 395–402.

Kurz H, Mauser-Ganshorn A, & Stickel HH. (1977). Differences in the binding of drugs to

plasma proteins from newborn and adult man. I. Eur J Clin Pharmacol 11 (6), 463-467.

Page 161: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

145

Kwan LY, & Mahadevan U. (2010). Inflammatory bowel disease and pregnancy: An update.

Expert Rev Clin Immunol 6 (4), 643-657.

Lachenmeier DW, Schoeberl K, Kanteres F, Kuballa T, Sohnius EM, & Rehm J. (2011). Is

contaminated unrecorded alcohol a health problem in the European Union? A review of

existing and methodological outline for future studies. Addiction 106 (Suppl 1), 20-30.

Landy HJ, Isada NB, McGinnis J, Ratner R, & Grossman J. (1988). The effect of chronic steroid

therapy on glucose tolerance in pregnancy. Am J Obstet Gynecol 159 (3), 612-615.

Langagergaard V, Pedersen L, Gislum M, Norgard B, & Sorensen HT. (2007). Birth outcome in

women treated with azathioprine or mercaptopurine during pregnancy: A Danish

nationwide cohort study. Aliment Pharmacol Ther 25 (1), 73-81.

Larroque B, Kaminski M, Lelong N, d'Herbomez M, Dehaene P, Querleu D et al. (1992). Folate

status during pregnancy: relationship with alcohol consumption, other maternal risk factors

and pregnancy outcome. Eur J Obstet Gynecol Reprod Biol 43 (1), 19-27.

Laufer BI, Mantha K, Kleiber ML, Diehl EJ, Addison SM, & Singh SM. (2013). Long-lasting

alterations to DNA methylation and ncRNAs could underlie the effects of fetal alcohol

exposure in mice. Dis Model Mech 6 (4), 977-992.

Lennard L, & Singleton HJ. (1992). High-performance liquid chromatographic assay of the

methyl and nucleotide metabolites of 6-mercaptopurine: quantitation of red blood cell 6-

thioguanine nucleotide, 6-thioinosinic acid and 6-methylmercaptopurine metabolites in a

single sample. J Chromatogr 583 (1), 83-90.

Levkovitz R, Zaretsky U, Gordon Z, Jaffa AJ, & Elad D. (2013). In vitro simulation of placental

transport: part I. Biological model of the placental barrier. Placenta 34 (8), 699-707.

Levkovitz R, Zaretsky U, Jaffa AJ, Hod M, & Elad D. (2013). In vitro simulation of placental

transport: part II. Glucose transfer across the placental barrier model. Placenta 34 (8), 708-

715.

Levy N, Roisman I, & Teodor I. (1981). Ulcerative colitis in pregnancy in Israel. Dis Colon

Rectum 24 (5), 351-354.

Page 162: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

146

Lin GWJ, Lester D. (1985) Altered placental folate coenzyme distribution by ethanol

consumption during pregnancy. Nutrition Reports Int 31 (6), 1375-1383.

Lin GW-J. (1988). Folate deficiency and acute ethanol treatment on pregnancy outcome in the

rat. Nutrition Res 8 (10), 1151-1160.

Lin GW. (1991). Maternal-fetal folate transfer: effect of ethanol and dietary folate deficiency.

Alcohol 8 (3), 169-172.

Linnemann K, Malek A, Sager R, Blum WF, Schneider H, & Fusch C. (2000). Leptin

production and release in the dually in vitro perfused human placenta. J Clin Endocrinol

Metab 85 (11), 4298-4301.

Loebstein R, Lalkin A, & Koren G. (1997). Pharmacokinetic changes during pregnancy and

their clinical relevance. Clin Pharmacokinet 33 (5), 328-343.

Ludvigsson JF, & Ludvigsson J. (2002). Inflammatory bowel disease in mother or father and

neonatal outcome. Acta Paediatr 91 (2), 145-151.

Ma SH, Lepak LA, Hussain RJ, Shain W, & Shuler ML. (2005). An endothelial and astrocyte

co-culture model of the blood-brain barrier utilizing an ultra-thin, nanofabricated silicon

nitride membrane. Lab Chip 5 (1), 74-85.

MacFarland A, Abramovich DR, Ewen SW, & Pearson CK. (1994). Stage-specific distribution

of P-glycoprotein in first-trimester and full-term human placenta. Histochem J 26 (5), 417-

423.

Maguire MJ, Hemming K, Hutton JL, & Marson AG. (2008). Overwhelming heterogeneity in

systematic reviews of observational anti-epileptic studies. Epilepsy Res 80 (2-3), 201-212.

Mahey R, Kaur SD, Chumber S, Kriplani A, & Bhatla N. (2013). Splenectomy during

pregnancy: treatment of refractory immune thrombocytopenic purpura. BMJ Case Rep In

Press.

Majchrowicz E, & Mendelson JH. (1971). Blood methanol concentrations during experimentally

induced ethanol intoxication in alcoholics. J Pharmacol Exp Ther 179 (2), 293-300.

Page 163: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

147

Makar AB, Tephly TR. (1976) Methanol poisoning in the folate-deficient rat. Nature 261 (5562),

715-6.

Malek A, Obrist C, Wenzinger S, & von Mandach U. (2009). The impact of cocaine and heroin

on the placental transfer of methadone. Reprod Biol Endocrinol 7, 61.

Maliepaard M, Scheffer GL, Faneyte IF, van Gastelen MA, Pijnenborg AC, Schinkel AH et al.

(2001). Subcellular localization and distribution of the breast cancer resistance protein

transporter in normal human tissues. Cancer Res 61 (8), 3458-3464.

Malone SM, McGue M, & Iacono WG. (2010). Mothers' maximum drinks ever consumed in 24

hours predicts mental health problems in adolescent offspring. J Child Psychol Psychiatry

51 (9), 1067-1075.

Mather LE, Long GJ, & Thomas J. (1971). The binding of bupivacaine to maternal and foetal

plasma proteins. J Pharm Pharmacol 23 (5), 359-365.

Mathiesen L, Mose T, Morck TJ, Nielsen JK, Nielsen LK, Maroun LL et al. (2010). Quality

assessment of a placental perfusion protocol. Reprod Toxicol 30 (1), 138-146.

May PA, & Gossage JP. (2001). Estimating the prevalence of fetal alcohol syndrome. A

summary. Alcohol Res Health 25 (3), 159-167.

McBride WG. (1978). Teratogenic action of thalidomide. Lancet 1 (8078), 1362.

McMartin KE, Martin-Amat G, Makar AB, Tephly TR. (1977) Methanol poisoning. V. Role of

formate metabolism in the monkey. J Pharmacol Exp Ther 201 (3), 564-572.

Meuldermans W, Woestenborghs R, Noorduin H, Camu F, van Steenberge A, & Heykants J.

(1986). Protein binding of the analgesics alfentanil and sufentanil in maternal and neonatal

plasma. Eur J Clin Pharmacol 30 (2), 217-219.

Meyer zu Schwabedissen HE, Grube M, Dreisbach A, Jedlitschky G, Meissner K, Linnemann K

et al. (2006). Epidermal growth factor-mediated activation of the map kinase cascade

results in altered expression and function of ABCG2 (BCRP). Drug Metab Dispos 34 (4),

524-533.

Page 164: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

148

Meyer Zu Schwabedissen HE, Grube M, Heydrich B, Linnemann K, Fusch C, Kroemer HK et

al. (2005). Expression, localization, and function of MRP5 (ABCC5), a transporter for

cyclic nucleotides, in human placenta and cultured human trophoblasts: effects of

gestational age and cellular differentiation. Am J Pathol 166 (1), 39-48.

Miller RK, Wier PJ, Maulik D, & di Sant'Agnese PA. (1985). Human placenta in vitro:

characterization during 12 h of dual perfusion. Contrib Gynecol Obstet 13, 77-84.

Molloy AM, Mills JL, Cox C, Daly SF, Conley M, Brody LC et al. (2005). Choline and

homocysteine interrelations in umbilical cord and maternal plasma at delivery. Am J Clin

Nutr 82 (4), 836-842.

Moore JJ, Lundgren DW, Moore RM, & Andersen B. (1988). Polyamines control human

chorionic gonadotropin production in the JEG-3 choriocarcinoma cell. J Biol Chem 263

(25), 12765-12769.

Morales M, Berney T, Jenny A, Morel P, & Extermann P. (2000). Crohn's disease as a risk

factor for the outcome of pregnancy. Hepatogastroenterology 47 (36), 1595-1598.

Morck TJ, Sorda G, Bechi N, Rasmussen BS, Nielsen JB, Ietta F et al. (2010). Placental

transport and in vitro effects of Bisphenol A. Reprod Toxicol 30 (1), 131-137.

Morton CP, Bloomfield S, Magnusson A, Jozwiak H, & McClure JH. (1997). Ropivacaine

0.75% for extradural anaesthesia in elective caesarean section: an open clinical and

pharmacokinetic study in mother and neonate. Br J Anaesth 79 (1), 3-8.

Mose T, Mathiesen L, Karttunen V, Nielsen JK, Sieppi E, Kummu M et al. (2012). Meta-

analysis of data from human ex vivo placental perfusion studies on genotoxic and

immunotoxic agents within the integrated European project NewGeneris. Placenta 33 (5),

433-439.

Mountifield RE, Prosser R, Bampton P, Muller K, & Andrews JM. (2010). Pregnancy and IBD

treatment: This challenging interplay from a patients' perspective. J Crohn's Colitis 4 (2),

176-182.

Page 165: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

149

Mowat C, Cole A, Windsor A, Ahmad T, Arnott I, Driscoll R et al. (2011). Guidelines for the

management of inflammatory bowel disease in adults. Gut 60 (5), 571-607.

Muhlemann K, Menegus MA, & Miller RK. (1995). Cytomegalovirus in the perfused human

term placenta in vitro. Placenta 16 (4), 367-373.

Murphy ML. 1960. Teratogenic effects of tumor-inhibiting chemicals in the rat. In:

Wolstenholme GEW, O’Conner CM, editors. A CIBA Foundation Symposium on

congenital malformations. London: J & A Churchill Ltd. p 78.

Myllynen P, Immonen E, Kummu M, & Vahakangas K. (2009). Developmental expression of

drug metabolizing enzymes and transporter proteins in human placenta and fetal tissues.

Expert Opin Drug Metab Toxicol 5 (12), 1483-1499.

Myllynen P, Pienimaki P, & Vahakangas K. (2005). Human placental perfusion method in the

assessment of transplacental passage of antiepileptic drugs. Toxicol Appl Pharmacol 207 (2

Suppl), 489-494.

Nagai K, Nagasawa K, Kihara Y, Okuda H, & Fujimoto S. (2007). Anticancer nucleobase

analogues 6-mercaptopurine and 6-thioguanine are novel substrates for equilibrative

nucleoside transporter 2. Int J Pharm 333 (1-2), 56-61.

Nagashige M, Ushigome F, Koyabu N, Hirata K, Kawabuchi M, Hirakawa T et al. (2003). Basal

membrane localization of MRP1 in human placental trophoblast. Placenta 24 (10), 951-

958.

Nanovskaya TN, Nekhayeva IA, Hankins GDV, & Ahmed MS. (2008). Transfer of methadone

across the dually perfused preterm human placental lobule. Am J Obstet Gynecol 198 (1),

126e1-4.

Nau H. (1985). Clinical pharmacokinetics in pregnancy and perinatology. I. Placental transfer

and fetal side effects of local anaesthetic agents. Dev Pharmacol Ther 8 (3), 149-181.

Nau H, Helge H, & Luck W. (1984). Valproic acid in the perinatal period: decreased maternal

serum protein binding results in fetal accumulation and neonatal displacement of the drug

and some metabolites. J Pediatr 104 (4), 627-634.

Page 166: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

150

Nau H, Luck W, Kuhnz W, & Wegener S. (1983). Serum protein binding of diazepam,

desmethyldiazepam, furosemide, indomethacin, warfarin, and phenobarbital in human

fetus, mother, and newborn infant. Pediatr Pharmacol 3 (3-4), 219-227.

Nava-Ocampo AA, Velazquez-Armenta Y, Brien JF, & Koren G. (2004). Elimination kinetics

of ethanol in pregnant women. Reprod Toxicol 18 (4), 613-617.

Navarro J, Causse MB, Desquilbet N, Herve F, & Lallemand D. (1984). The vitamin status of

low birth weight infants and their mothers. J Pediatr Gastroenterol Nutr 3 (5), 744-748.

Nelson JA, Carpenter JW, Rose LM, Adamson DJ. (1975) Mechanisms of action of 6-

thioguanine, 6-mercaptopurine, and 8-azaguanine. Cancer Res 35, 2872–2878.

Norgard B, Pedersen L, Christensen LA, & Sorensen HT. (2007a). Therapeutic drug use in

women with crohn's disease and birth outcomes: A danish nationwide cohort study. Am J

Gastroenterol 102 (7), 1406-1413.

Norgard B, Hundborg HH, Jacobsen BA, Nielsen GL, & Fonager K. (2007b). Disease activity in

pregnant women with Crohn’s disease and birth outcomes: a regional Danish cohort study.

Am J Gastroenterol 102 (9), 1947-1954.

O'Connor MJ, Shah B, Whaley S, Cronin P, Gunderson B, & Graham J. (2002). Psychiatric

illness in a clinical sample of children with prenatal alcohol exposure. Am J Drug Alcohol

Abuse 28 (4), 743-754.

O’Neil M, Heckelman P, Koch C, Roman K, editors. The Merck Index: An Encyclopedia of

Chemicals, Drugs, and Biologicals. 6-Mercaptopurine (Monograph #05871). Whitehouse

Station, NJ, USA: Merck & Co., Inc.; 2010.

Obeid R, Munz W, Jager M, Schmidt W, & Herrmann W. (2005). Biochemical indexes of the B

vitamins in cord serum are predicted by maternal B vitamin status. Am J Clin Nutr 82 (1),

133-139.

Ohlman S, Albertioni F, & Peterson C. (1994). Day-to-day variability in azathioprine

pharmacokinetics in renal transplant recipients. Clin Transplant 8 (3 Pt 1), 217-223.

Page 167: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

151

Omarini D, Pistotti V, & Bonati M. (1992). Placental perfusion. An overview of the literature. J

Pharmacol Toxicol Methods 28 (2), 61-66.

Osting VC, & Carter JD. (2010). A safety assessment of tumor necrosis factor antagonists

during pregnancy. Expert Opin Drug Saf 9 (3), 421-429.

Ostrea EMJ, Knapp DK, Tannenbaum L, Ostrea AR, Romero A, Salari V et al. (2001).

Estimates of illicit drug use during pregnancy by maternal interview, hair analysis, and

meconium analysis. J Pediatr 138 (3), 344-348.

Paine A1, Davan AD. Defining a tolerable concentration of methanol in alcoholic drinks. (2001)

Hum Exp Toxicol 20 (11), 563-568.

Palese M, Tephly TR. (1975) Metabolism of formate in the rat. J Toxicol Environ Health 1 (1),

13-24.

Panigel M, Pascaud M, & Brun JL. (1967). [Radioangiographic study of circulation in the villi

and intervillous space of isolated human placental cotyledon kept viable by perfusion]. J

Physiol 59 (1 Suppl), 277.

Park-Wyllie L, Mazzotta P, Pastuszak A, Moretti ME, Beique L, Hunnisett L et al. (2000). Birth

defects after maternal exposure to corticosteroids: prospective cohort study and meta-

analysis of epidemiological studies. Teratology 62 (6), 385-392.

Peyrin-Biroulet L, Oussalah A, Roblin X, & Sparrow MP. (2011). The use of azathioprine in

Crohn's disease during pregnancy and in the post-operative setting: a worldwide survey of

experts. Aliment Pharmacol Ther 33 (6), 707-713.

Philipson EH, Kuhnert BR, & Syracuse CD. (1985). Fetal acidosis, 2-chloroprocaine, and

epidural anesthesia for cesarean section. Am J Obstet Gynecol 151 (3), 322-324.

Pickering B, Biehl D, & Meatherall R. (1981). The effect of foetal acidosis on bupivacaine

levels in utero. Can Anaesth Soc J 28 (6), 544-549.

Pienimaki P, Lampela E, Hakkola J, Arvela P, Raunio H, & kangas K. (1997). Pharmacokinetics

of oxcarbazepine and carbamazepine in human placenta. Epilepsia 38 (3), 309-316.

Page 168: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

152

Pirani BB, Campbell DM, & MacGillivray I. (1973). Plasma volume in normal first pregnancy.

J Obstet Gynaecol Br Commonw 80 (10), 884-887.

Platzek T, & Bochert G. (1996). Dose-response relationship of teratogenicity and prenatal-toxic

risk estimation of 6-mercaptopurine riboside in mice. Teratog Carcinog Mutagen 16 (3),

169-181.

Polifka JE, & Friedman JM. (2002). Teratogen update: azathioprine and 6-mercaptopurine.

Teratology 65 (5), 240-261.

Pollex E, Lubetsky A, & Koren G. (2008). The role of placental breast cancer resistance protein

in the efflux of glyburide across the human placenta. Placenta 29 (8), 743-747.

Pond SM, Kreek MJ, Tong TG, Raghunath J, & Benowitz NL. (1985). Altered methadone

pharmacokinetics in methadone-maintained pregnant women. J Pharmacol Exp Ther 233

(1), 1-6.

Porter JM, Kelleher N, Flynn R, & Shorten GD. (2001). Epidural ropivacaine hydrochloride

during labour: protein binding, placental transfer and neonatal outcome. Anaesthesia 56

(5), 418-423.

Porter JS, Bonello E, & Reynolds F. (1997). The influence of epidural administration of fentanyl

infusion on gastric emptying in labour. Anaesthesia 52 (12), 1151-1156.

Prasad PD, Ramamoorthy S, Leibach FH, & Ganapathy V. (1995). Molecular cloning of the

human placental folate transporter. Biochem Biophys Res Commun 206 (2), 681-687.

Present DH, Meltzer SJ, Krumholz MP, Wolke A, & Korelitz BI. (1989). 6-Mercaptopurine in

the management of inflammatory bowel disease: short- and long-term toxicity. Ann Intern

Med 111 (8), 641-649.

Puget A, Cros S, Oreglia J, Tollon Y. (1975) E ́ tude de la sensibilite ́ embryonnaire de

l’ochotone afghan (Ochotona rufescens rufescens) a deux agents te ́ratoge`nes,

l’azathioprine et la 6 mercaptopurine. Zentralbl Veterinarmed [A] 22, 38–56.

Page 169: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

153

Rahimi R, Nikfar S, Rezaie A, & Abdollahi M. (2008). Pregnancy outcome in women with

inflammatory bowel disease following exposure to 5-aminosalicylic acid drugs: a meta-

analysis. Reprod Toxicol 25 (2), 271-275.

Rambeck B, Kurlemann G, Stodieck SR, May TW, & Jurgens U. (1997). Concentrations of

lamotrigine in a mother on lamotrigine treatment and her newborn child. Eur J Clin

Pharmacol 51 (6), 481-484.

Reddy D, Murphy SJ, Kane SV, Present DH, & Kornbluth AA. (2008). Relapses of

inflammatory bowel disease during pregnancy: in-hospital management and birth

outcomes. Am J Gastroenterol 103 (5), 1203-1209.

Reis M, & Kallen B. (2010). Delivery outcome after maternal use of antidepressant drugs in

pregnancy: an update using Swedish data. Psychol Med 40 (10), 1723-1733.

Relton CL, Pearce MS, & Parker L. (2005). The influence of erythrocyte folate and serum

vitamin B12 status on birth weight. Br J Nutr 93 (5), 593-599.

Reshef E, Lei ZM, Rao CV, Pridham DD, Chegini N, & Luborsky JL. (1990). The presence of

gonadotropin receptors in nonpregnant human uterus, human placenta, fetal membranes,

and decidua. J Clin Endocrinol Metab 70 (2), 421-430.

Reynolds F, & Knott C. (1989). Pharmacokinetics in pregnancy and placental drug transfer. Oxf

Rev Reprod Biol 11, 389-449.

Reynolds F, & Seed PT. (2005). Anaesthesia for Caesarean section and neonatal acid-base

status: a meta-analysis. Anaesthesia 60 (7), 636-653.

Ridd MJ, Brown KF, Nation RL, & Collier CB. (1983). Differential transplacental binding of

diazepam: causes and implications. Eur J Clin Pharmacol 24 (5), 595-601.

Ridd MJ, Brown KF, Nation RL, & Collier CB. (1989). The disposition and placental transfer of

diazepam in cesarean section. Clin Pharmacol Ther 45 (5), 506-512.

Ring JA, Ghabrial H, Ching MS, Smallwood RA, & Morgan DJ. (1999). Fetal hepatic drug

elimination. Pharmacol Ther 84(3), 429-445.

Page 170: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

154

Rosenberg MJ, Wolff CR, El-Emawy A, Staples MC, Perrone-Bizzozero NI, & Savage DD.

(2010). Effects of moderate drinking during pregnancy on placental gene expression.

Alcohol 44 (7-8), 673-690.

Rosenkrantz JG, Githens JH, Cox SM, Kellum DL. (1967) Azathioprine (Imuran) and

pregnancy. Am J Obstet Gynecol 97, 387–394.

Saarikoski S, & Seppala M. (1973). Immunosuppression during pregnancy: transmission of

azathioprine and its metabolites from the mother to the fetus. Am J Obstet Gynecol 115 (8),

1100-1106.

Sahasranaman S, Howard D, & Roy S. (2008). Clinical pharmacology and pharmacogenetics of

thiopurines. Eur J Clin Pharmacol 64 (8), 753-767.

Sandborn WJ, Van Os EC, Zins BJ, Tremaine WJ, Mays DC, Lipsky JJ. (1995) An intravenous

loading dose of azathioprine decreases the time to response in patients with Crohn disease.

Gastroenterology 109, 1808 –1817.

Sarkola T, & Eriksson CJ. (2001). Effect of 4-methylpyrazole on endogenous plasma ethanol and

methanol levels in humans. Alcohol Clin Exp Res 25 (4), 513-516.

Schaefer C, Ornoy A, Clementi M, Meister R, & Weber-Schoendorfer C. (2008). Using

observational cohort data for studying drug effects on pregnancy outcome--methodological

considerations. Reprod Toxicol 26 (1), 36-41.

Schatz M, Dombrowski MP, Wise R, Momirova V, Landon M, Mabie W et al. (2004). The

relationship of asthma medication use to perinatal outcomes. J Allergy Clin Immunol 113

(6), 1040-1045.

Scheeres JJ, & Chudley AE. (2002). Solvent abuse in pregnancy: a perinatal perspective. J

Obstet Gynaecol Can 24 (1), 22-26.

Schenker S, Yang Y, Mattiuz E, Tatum D, & Lee M. (1999). Olanzapine transfer by human

placenta. Clin Exp Pharmacol Physiol 26 (9), 691-697.

Page 171: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

155

Schlotz W, Jones A, Phillips DI, Gale CR, Robinson SM, & Godfrey KM. (2010). Lower

maternal folate status in early pregnancy is associated with childhood hyperactivity and

peer problems in offspring. J Child Psychol Psychiatry 51 (5), 594-602.

Schmolling J, Renke K, Richter O, Pfeiffer K, Schlebusch H, & Holler T. (2000). Digoxin,

flecainide, and amiodarone transfer across the placenta and the effects of an elevated

umbilical venous pressure on the transfer rate. Ther Drug Monit 22 (5), 582-588.

Schneider H, Panigel M, & Dancis J. (1972). Transfer across the perfused human placenta of

antipyrine, sodium and leucine. Am J Obstet Gynecol 114 (6), 822-828.

Shah YG, & Miller RK. (1985). The pharmacokinetics of phenytoin in perfused human placenta.

Pediatr Pharmacol 5 (3), 165-179.

Shankar K, Hidestrand M, Liu X, Xiao R, Skinner CM, Simmen FA et al. (2006). Physiologic

and genomic analyses of nutrition-ethanol interactions during gestation: Implications for

fetal ethanol toxicity. Exp Biol Med 231 (8), 1379-1397.

Shim L, Eslick GD, Simring AA, Murray H, & Weltman MD. (2011). The effects of

azathioprine on birth outcomes in women with inflammatory bowel disease (IBD). J

Crohns Colitis 5 (3), 234-238.

Shintaku K, Hori S, Tsujimoto M, Nagata H, Satoh S, Tsukimori K et al. (2009). Transplacental

pharmacokinetics of diclofenac in perfused human placenta. Drug Metab Dispos 37 (5),

962-968.

Shintaku K, Hori S, Satoh H, Tsukimori K, Nakano H, Fujii T et al. (2012). Prediction and

evaluation of fetal toxicity induced by NSAIDs using transplacental kinetic parameters

obtained from human placental perfusion studies. Br J Clin Pharmacol 73 (2), 248-256.

Shintaku K, Arima Y, Dan Y, Takeda T, Kogushi K, Tsujimoto M et al. (2007). Kinetic analysis

of the transport of salicylic acid, a nonsteroidal anti-inflammatory drug, across human

placenta. Drug Metab Dispos 35 (5), 772-778.

Simone C, Byrne BM, Derewlany LO, & Koren G. (1996). Cocaine inhibits hCG secretion by

the human term placental cotyledon perfused in vitro. Life Sciences 58 (5), PL 63-66.

Page 172: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

156

Simunovic N, Sprague S, & Bhandari M. (2009). Methodological issues in systematic reviews

and meta-analyses of observational studies in orthopaedic research. J Bone Joint Surg Am

91(Suppl 3), 87-94.

Sokoro AA, Zhang Z, Eichhorst JC, Zello GA, House JD, Alcorn J et al. (2008). Formate

pharmacokinetics during formate administration in folate-deficient young swine.

Metabolism 57 (7), 920-926.

Solanky N, Requena Jimenez A, D'Souza SW, Sibley CP, & Glazier JD. (2010). Expression of

folate transporters in human placenta and implications for homocysteine metabolism.

Placenta 31 (2), 134-143.

Sonnier M, & Cresteil T. (1998). Delayed ontogenesis of CYP1A2 in the human liver. Eur J

Biochem 251 (3), 893-898.

St-Pierre MV, Serrano MA, Macias RI, Dubs U, Hoechli M, Lauper U et al. (2000). Expression

of members of the multidrug resistance protein family in human term placenta. Am J

Physiol Regul Integr Comp Physiol 279 (4), R1495-1503.

Stade B, Ali A, Bennett D, Campbell D, Johnston M, Lens C et al. (2009). The burden of

prenatal exposure to alcohol: revised measurement of cost. Can J Clin Pharmacol 16 (1),

e91-102.

Stark KD, Pawlosky RJ, Beblo S, Murthy M, Flanagan VP, Janisse J et al. (2005). Status of

plasma folate after folic acid fortification of the food supply in pregnant African American

women and the influences of diet, smoking, and alcohol consumption. Am J Clin Nutr 81

(3), 669-677.

Stark KD, Pawlosky RJ, Sokol RJ, Hannigan JH, & Salem NJ. (2007). Maternal smoking is

associated with decreased 5-methyltetrahydrofolate in cord plasma. Am J Clin Nutr 85 (3),

796-802.

Stratton K, Howe C, Battaglia FC. (1996) Fetal alcohol syndrome: diagnosis, epidemiology,

prevention, and treatment. Institute of Medicine and National Academy Press, Washington,

USA.

Page 173: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

157

Streissguth AP, Bookstein FL, Barr HM, Sampson PD, O'Malley K, & Young JK. (2004). Risk

factors for adverse life outcomes in fetal alcohol syndrome and fetal alcohol effects. J Dev

Behav Pediatr 25 (4), 228-238.

Sudhakaran S, Rayner CR, Li J, Kong DC, Gude NM, & Nation RL. (2007). Differential protein

binding of indinavir and saquinavir in matched maternal and umbilical cord plasma. Br J

Clin Pharmacol 63 (3), 315-321.

Sudhakaran S, Ghabrial H, Nation RL, Kong DCM, Gude NM, Angus PW et al. (2005).

Differential bidirectional transfer of indinavir in the isolated perfused human placenta.

Antimicrob Agents Chemother 49 (3), 1023-1028.

Sugawara I, Akiyama S, Scheper RJ, & Itoyama S. (1997). Lung resistance protein (LRP)

expression in human normal tissues in comparison with that of MDR1 and MRP. Cancer

Letters 112 (1), 23-31.

Sun M, Kingdom J, Baczyk D, Lye SJ, Matthews SG, & Gibb W. (2006). Expression of the

multidrug resistance P-glycoprotein, (ABCB1 glycoprotein) in the human placenta

decreases with advancing gestation. Placenta 27 (6-7), 602-609.

Syme MR, Paxton JW, & Keelan JA. (2004). Drug transfer and metabolism by the human

placenta. Clin Pharmacokinet 43 (8), 487-514.

Tennenbaum R, Marteau P, Elefant E, Rambaud JC, Modigliani R, Gendre JP et al. (1999).

[Pregnancy outcome in inflammatory bowel diseases]. e Gastroenterol Clin Biol 23 (5),

464-469.

Thiersch JB. (1954) The effect of 6-mercaptopurine on the rat fetus and on reproduction of the

rat. Ann NY Acad Sci 60, 220–227.

Thomas J, Long G, Moore G, & Morgan D. (1976). Plasma protein binding and placental

transfer of bupivacaine. Clin Pharmacol Ther 19 (4), 426-434.

Thorand B, Pietrzik K, Prinz-Langenohl R, Hages M, & Holzgreve W. (1996). Maternal and

fetal serum and red blood cell folate and vitamin B12 concentrations in pregnancies

affected by neural tube defects. Z Geburtshilfe Neonatol 200 (5), 176-180.

Page 174: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

158

Timmer A, McDonald JW, & Macdonald JK. (2007). Azathioprine and 6-mercaptopurine for

maintenance of remission in ulcerative colitis. Cochrane Database Syst Rev 1, CD000478.

Toth P, Li X, Rao CV, Lincoln SR, Sanfilippo JS, Spinnato JAn et al. (1994). Expression of

functional human chorionic gonadotropin/human luteinizing hormone receptor gene in

human uterine arteries. J Clin Endocrinol Metab 79 (1), 307-315.

Tsen LC, Arthur GR, Datta S, Hornstein MD, & Bader AM. (1997). Estrogen-induced changes

in protein binding of bupivacaine during in vitro fertilization. Anesthesiology 87 (4), 879-

883.

Tuchmann-Duplessis H, Mercier-Parot L. (1964) Considerations sur les tests teratogenes.

Differences de reaction de trois especes animales a Iegard un antitumoral. CRSBA 158,

1984–1990.

Ueki R, Tatara T, Kariya N, Shimode N, & Tashiro C. (2009). Comparison of placental transfer

of local anesthetics in perfusates with different pH values in a human cotyledon model. J

Anesth 23 (4), 526-529.

Vahakangas K, & Myllynen P. (2006). Experimental methods to study human transplacental

exposure to genotoxic agents. Mutat Res 608 (2), 129-135.

Van Scoik KG, Johnson CA, Porter WR. 1985. The pharmacology and metabolism of the

thiopurine drugs 6-mercaptopurine and azathioprine. Drug Metab Rev 16, 157–174.

Vanky E, Zahlsen K, Spigset O, & Carlsen SM. (2005). Placental passage of metformin in

women with polycystic ovary syndrome. Fertil Steril 83 (5), 1575-1578.

Veena SR, Krishnaveni GV, Srinivasan K, Wills AK, Muthayya S, Kurpad AV et al. (2010).

Higher maternal plasma folate but not vitamin B-12 concentrations during pregnancy are

associated with better cognitive function scores in 9- to 10- year-old children in South

India. J Nutr 140 (5), 1014-1022.

Vermeulen RC, Kurver PH, Arts NF, Van Kessel H, Wilson GR, & Klopper A. (1982). The

relationship between the surface area of the trophoblast and some placental products.

Placenta 3 (4), 359-366.

Page 175: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

159

Wallace JM, Bonham MP, Strain J, Duffy EM, Robson PJ, Ward M et al. (2008). Homocysteine

concentration, related B vitamins, and betaine in pregnant women recruited to the

Seychelles Child Development Study. Am J Clin Nutr 87 (2), 391-397.

Wang LL, Zhang Z, Li Q, Yang R, Pei X, Xu Y et al. (2009). Ethanol exposure induces

differential microRNA and target gene expression and teratogenic effects which can be

suppressed by folic acid supplementation. Hum Reprod 24 (3), 562-579.

Warren KR, & Li TK. (2005). Genetic polymorphisms: impact on the risk of fetal alcohol

spectrum disorders. Birth Defects Res A Clin Mol Teratol 73 (4), 195-203.

Wentzel P, & Eriksson UJ. (2008). Folic acid prevention of ethanol damage in rat offspring.

Alcohol Clin Exp Res 32 (s1), 48A.

Wielinga PR, Reid G, Challa EE, van der Heijden I, van Deemter L, de Haas M et al. (2002).

Thiopurine metabolism and identification of the thiopurine metabolites transported by

MRP4 and MRP5 overexpressed in human embryonic kidney cells. Mol Pharmacol 62 (6),

1321-1331.

Wijnholds J, Mol CA, van Deemter L, de Haas M, Scheffer GL, Baas F et al. (2000). Multidrug-

resistance protein 5 is a multispecific organic anion transporter able to transport nucleotide

analogs. Proc Natl Acad Sci U S A 97 (13), 7476-7481.

Wilson RD1, Johnson JA, Wyatt P, Allen V, Gagnon A, Langlois S, Blight C, Audibert F,

Désilets V, Brock JA, Koren G, Goh YI, Nguyen P, Kapur B; Genetics Committee of the

Society of Obstetricians and Gynaecologists of Canada and The Motherisk Program.

(2007) Pre-conceptional vitamin/folic acid supplementation 2007: the use of folic acid in

combination with a multivitamin supplement for the prevention of neural tube defects and

other congenital anomalies. J Obstet Gynaecol Can. 29 (12), 1003-1026.

Wimalasena J, Beams F, & Caudle MR. (1994). Ethanol modulates the hormone secretory

responses induced by epidermal growth factor in choriocarcinoma cells. Alcohol Clin Exp

Res 18 (6), 1448-1455.

Page 176: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

160

Woolfson K, Cohen Z, & McLeod RS. (1990). Crohn's disease and pregnancy. Dis Colon

Rectum 33 (10), 869-873.

Xu Y, Tang Y, & Li Y. (2008). Effect of folic acid on prenatal alcohol-induced modification of

brain proteome in mice. Br J Nutr 99 (3), 455-461.

Yanaguita MY, Gutierrez CM, Ribeiro CN, Lima GA, Machado HR, & Peres LC. (2008).

Pregnancy outcome in ethanol-treated mice with folic acid supplementation in saccharose.

Childs Nerv Syst 24 (1), 99-104.

Yasuda S, Hasui S, Kobayashi M, Itagaki S, Hirano T, & Iseki K. (2008). The mechanism of

carrier-mediated transport of folates in BeWo cells: the involvement of heme carrier

protein 1 in placental folate transport. Biosci Biotechnol Biochem 72 (2), 329-334.

Yeboah D, Sun M, Kingdom J, Baczyk D, Lye SJ, Matthews SG et al. (2006). Expression of

breast cancer resistance protein (BCRP/ABCG2) in human placenta throughout gestation

and at term before and after labor. Can J Physiol Pharmacol 84 (12), 1251-1258.

Zakowski MI, Ham AA, & Grant GJ. (1994). Transfer and uptake of alfentanil in the human

placenta during in vitro perfusion. Anesth Analg 79 (6), 1089-1093.

Zeisel SH. (2011). What choline metabolism can tell us about the underlying mechanisms of

fetal alcohol spectrum disorders. Mol Neurobiol 44 (2), 185-191.

Zelner I, & Koren G. (2013). Alcohol consumption among women. J Popul Ther Clin

Pharmacol 20 (2), e201-206.

Zhang L, Connor EE, Chegini N, & Shiverick KT. (1995). Modulation by benzo[a]pyrene of

epidermal growth factor receptors, cell proliferation, and secretion of human chorionic

gonadotropin in human placental cell lines. Biochem Pharmacol 50 (8), 1171-1180.

Zins BJ, Sandborn WJ, McKinney JA, Mays DC, van Os EC, Tremaine WJ et al. (1997). A

dose-ranging study of azathioprine pharmacokinetics after single-dose administration of a

delayed-release oral formulation. J Clin Pharmacol 37 (1), 38-46.

Page 177: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

161

Appendices

Page 178: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

162

Appendix A: Placental Perfusion System Materials and Set-up

1. Materials 2. Maternal Ismatec Pump (Ismatec Instruments, Zurich Switzerland) 3. Fetal polystaltic SA 8031 pump (Buchler Instruments, (VWR Canlab) Missisauga

Canada) 4. Maternal Reservoir- 500 ml Flask 5. Fetal reservoir-chromatography Ampholine® column (type 8101) glass water jacket

(LKB Sweden) 6. Model 86 hot water bath (Precision Scientific, Chicago, USA) 7. Plexiglass Perfusion Chamber (Figure 5) 8. Buchi rotary evaporator- Brinkmann CH 9230 Rotavapor R (Laboratorium Technik AG,

Switzerland) 9. ABL 725 acid-base analysis laboratory (Radiometer A/S, Copenhagen Denmark) 10. Model 759 sphygmomanometer (Sunbeam Corp. (Canada) Ltd, Toronto, Canada) 11. Surgical instruments: fine and coarse scissors, 5.5" straight needle holder, 5.5"straight

Kelly clamps x 2, fine and coarse straight edge splinter forceps, German surgical 3.75" curved iris Castroviejo scissors (Irex, Toronto, Canada)

12. Three-way stop cocks x 13 13. Three hole rubber stopper x 1 14. # 5 French Catheter 15. # 8 French feeding tube 16. Blunt ended catheter 17. Narrow bore tubing (PE 280) 18. 19 gauge metal bulbous-tip cannula 19. 15 gauge metal bulbous-tip cannula 20. 1 cc and 3 cc syringes 21. 20 ml syringe 22. Oxygen- D tank (Praxair, Mississauga, Canada) with a pressure gauge, flow meter

regulator Model No ME1-540-FG (Medigas, Mississauga, Canada) 23. Carbon Dioxide- D tank (Praxair, Mississauga, Canada) with a pressure gauge, flow

meter regulator Model No ME1-540-FG (Medigas, Mississauga, Canada) 24. Nitrogen- C tank (Praxair, Mississauga, Canada) with a pressure gauge, flow meter

regulator Model No Ml-940 PG (Western Enterprises, Westlake, USA) 25. 4-0 black braided silk sutures with a 17 mm T-31 needles (Daves & Geek, Wayne, USA) 26. Catheter introducer (Becton-Dickinson, Rutherford, USA) 27. Gas dispersion tube 28. T-connector x 3 29. Y connector x 1 30. Ice-chest or cooler

Page 179: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

163

2. Fetal Circuit Hardware Set-up

The fetal reservoir is a chromatography column glass water jacket, fitted with a three-

hole rubber stopper on the top and a three-way stopcock at the base as illustrated below in figure

1. A gas dispersion tube is inserted into the reservoir, through one of the three stopper holes, and

functions to both gas and mix the fetal perfusate. The fetal venous line is connected to a tube that

is inserted in a second of the three stopper holes. It returns perfusate to the reservoir for re-

circulation. Sampling the fetal perfusate from the reservoir is done through an umbilical catheter

(# 5 French) which is inserted in the third hole of the rubber stopper located at the top of the fetal

reservoir.

Tubing extending from the base of the reservoir leads to the fetal pump. From the pump,

the fetal circuit follows the following sequence: a flowmeter, then to a sampling port (three way

stopcock) from which fetal arterial (FA) samples are drawn. Distal to the sampling port, a

sphygmomanometer is connected to the fetal circuit tubing by a T-connector. From the

sphygmomanometer, the tubing leads to a connector located on the outside of the fetal perfusion

chamber. The FA cannula is connected to the inside of the chamber by a #5 French umbilical

vessel catheter with a short 19 gauge metal bulbous tip. The fetal vein (FV) cannula is also

connected to the inside of the chamber; it is a #8 French feeding tube with a short 15 gauge metal

bulbous tip. Tubing connected to the outside of the perfusion chamber at the venous port leads to

the FV sampling port (another three-way stopcock). From the FV sampling port, tubing extends

back to the fetal reservoir enabling the perfusate to be re-circulated or sampled (Figure 2).

Page 180: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

164

Figure 1. Fetal reservoir setup

Figure 2. Configuration of fetal perfusion circuit.

Page 181: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

165

3. Maternal Circuit Hardware Setup

The maternal reservoir is composed of a 500 ml flask attached to the steam duct of the

rotary evaporator (Figure 3). The maternal flask is rotated over the progress of the experiment

since this aids in the mixing and oxygenation of the maternal perfusate. There are three lines

inserted into the maternal reservoir: one line delivers oxygen into the maternal reservoir while

the other two lines deliver perfusate to and from the placenta. Perfusate samples from the

reservoir are withdrawn through a #5 French umbilical catheter that inserts into the maternal

reservoir. The maternal pump is a peristaltic pump that delivers perfusate to the placenta and

pumps maternal venous return from the placenta.

Prior to entering the placenta, the maternal perfusate enters a bubble trap consisting of a

20 ml syringe barrel fitted with a rubber stopper at the top (inflow). A small piece of gauze is put

inside the bubble trap to filter any tissue debris that could clog the maternal arterial cannulas.

From the bubble trap, the maternal artery (MA) sampling port (three way stopcock) branches off

the main line tubing by way of a T connector. Before reaching the perfusion chamber, the arterial

line branches at a Y-connector thereby creating two arterial inflows that enter the maternal

surface of the placenta. The MA cannulas are two #5 French umbilical catheters inserted into the

maternal surface of the placenta. The maternal venous (MV) return collects on the surface of the

lobule. The MV is suctioned off and returned to the maternal reservoir for re-circulation (Figure

4). The MV sampling port (three-way stopcock) branches off the main circuit tubing by way of

another T-connector.

Page 182: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

166

Figure 3. Maternal reservoir Setup.

Figure 4. Configuration of maternal perfusion circuit.

Page 183: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

167

4. Perfusion Chamber The plexiglass perfusion chamber consists of 3 individual sections (Figure 30):

1. The base of the perfusion chamber consists of three ports: FA port which leads perfusate to the reservoir and the FV leading perfusate away from the chamber. There is a third port used to fill the bottom section of the perfusion chamber with warmed phosphate buffered saline (amniotic fluid) once the placental tissue has been mounted.

2. The middle section of the chamber is a plexiglass collar that is used to firmly clamp the placental tissue to the bottom portion of the chamber. This causes the tissue surrounding the cotyledon of interest to be clamped into position within the chamber.

3. The top section of the perfusion chamber has three ports. Two of the ports support MA cannulas, which are attached to the inside of the chamber. The third port supports tubing that is attached on the inside of the chamber suctioning off maternal venous return to the maternal reservoir The top section of the perfusion chamber is securely clamped to the middle section enabling the maternal venous return to collect around the cotyledon without leaking out of the perfusion chamber.

Figure 5. Configuration of the perfusion chamber.

Page 184: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

168

5. Composition of Medium 199 Culture Medium in Fetal and Maternal Perfusate

Compound mg/L Compound mg/L DL-Alanine 50.0 Thiamine HCl 0.01 L-Arginine ! HCl 70.0 Thymine 0.30 DL-Aspartic acid 60.0 Vitamin A acetate 0.14 L-Cysteine HCl ! H20 0.11 Xanthine ! Na 0.34 L-Cysteine ! 2HCl 26.0 Calcium chloride ! 2H2O 185.0 DL-Glutamic Acid 133.6 Ferric nitrate ! 9H2O 0.72 Glycine 50.0 Magnesium sulfate (anhydrous) 97.7 L-Histidine ! HCl ! H20 21.9 Potassium chloride 400.0 L-Hydroxyproline 10.0 Potassium phosphate 60.0 DL-Isoleucine 40.0 Sodium acetate (anhydrous) 50.0 DL-Leucine 120.0 Sodium chloride 8000.0 L-Lysine ! HCl 70.0 Sodium phosphate dibasic (anhydrous) 47.9 DL-Methionine 30.0 D-Glucose 1000.0 DL-Phylalanine 50.0 L-Proline 40.0 DL-Serine 50.0 DL-Threonine 60.0 DL-Tryptophan 20.0 L-Tyrosine ! 2Na 57.7 DL-Valine 50.0 Adenine sulfate 10.0 Adenosine triphosphate ! 2Na 1.0 Adenylic acid 0.20 α-Tocopherol phosphate ! 2Na 0.01 Ascorbic acid 0.05 Biotin 0.01 Calciferol 0.10 Cholestesterol 0.20 Choline chloride 0.50 Deoxyribose 0.50 Folic Acid 0.01 Glutathione (reduced) 0.50 Guanine ! HCl 0.30 Niacinamide 0.03 Nicotinic acid 0.03 PABA 0.05 D-Pantothenic acid ! Ca 0.01 Polyoxyethylene sorbitan 20.0 monooleate Pyridoxal HCl 0.03 Pyridoxine HCl 0.03 Riboflavin 0.01 Ribose 0.50

Page 185: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

169

Appendix B: Systematic Review of Perfusion Data - Supplementary Table (Chapter 2) Supplementary Table. Therapeutics evaluated using the placental perfusion model and in vivo as identified through a systematic search. Data is presented as the mean fetal to maternal concentration ratio (F:M) in the placental perfusion model ex vivo and at the time of delivery in vivo as the mean umbilical cord blood to maternal blood concentration ratio (C:M). For the perfusion experiments, the amount of protein added to the perfusate is described. Analgesics

Perfusion In vivo Drug PB F:M [ ] on M SS Albumin Sys n Ref C:M Time after n Ref

last dose _____________________________________________________________________________________________________________________________________ ASA 50-80% 0.38 100µM Y 0.1% BSA C 11 (1) 0.39 12-123 min 9 (2) (SA) 50-80% 0.21 8µg/ml Y 2 g/l HSA O 3 (3) 0.75 12-123 min 9 (2)

1.6 8 hr 1 (4) >1 <20 hr 4 (5) >1 1-4 hr 115 (6)

~1.0*** 96 min 18 (7) _____________________________________________________________________________________________________________________________________ Diclofenac >99% ~0.6 20µg/ml Y 1 g/l BSA O 4 (8) 0.95*** 2 hr 30 (9)

0.12 100ng/ml Y 2 g/l HSA O 3 (10) _____________________________________________________________________________________________________________________________________ Buprenorphine 96% 0.12 10ng/ml Y No O 5 (11) 0.5 16.5 hr 1 (12)

0.48 10ng/ml Y No C 9 (11) <LOD 24 hr 1 (12) 0.59 10ng/ml Y 30g/l HSA C 5 (13) 0.47 10ng/ml Y M>F AAG C 4 (13) 0.43 10ng/ml Y HSA + AAG C 7 (13) 0.14 10ng/ml NR No O 4 (14)

_____________________________________________________________________________________________________________________________________ Methadone 71-88% 0.27 200ng/ml Y No O 8 (15) 0.21 SS 19 (16)

0.71 100ng/ml Y No C ≥5 (17) 0.58 SS 27 (18) 0.19*** 200ng/ml Y No O 8 (19)

<LOD 500ng/ml No 4% BSA C 6 (20) _____________________________________________________________________________________________________________________________________ Sulindac 93% 0.42 3.6µg/ml No 2-5g/dl HSA C 4 (21) 0.99 4.4-6.7 hr 9 (22)

0.34 64µg/ml No 2mg/ml HSA C 4 (23)

_____________________________________________________________________________________________________________________________________

Page 186: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

170

Indomethacin 99% 0.45 32µg/ml No 2mg/ml HSA C 4 (23) 0.97 NR 26 (24) _____________________________________________________________________________________________________________________________________ Morphine 20-36% ~0.78 50ng/ml Y No C 4 (25) 0.96 5-74 min 5 (26)

~0.50 50ng/ml Y No O 2 (25) _____________________________________________________________________________________________________________________________________ Sufentanil 93% ~0.12 1ng/ml Y No O 10 (27) 0.38 ~45min 1 (28)

0.81 NR 8 (29)

_____________________________________________________________________________________________________________________________________ Alfentanil 92% 0.22 10ng/ml Y No O 10 (30) 0.29(0.97free) SS 10 (31) 0.29 6-10min PD 8 (28)

0.33 16-67 PI 6 (32) 0.35 9-25minPD 21 (33)

_____________________________________________________________________________________________________________________________________ Acetaminophen 10-30% 0.8 5-30µg/ml Y No C 4 (34) 0.78 overdose 1 (35)

>1 overdose 1 (36) _____________________________________________________________________________________________________________________________________ Anaesthetic

Perfusion In vivo Drug PB F:M [ ] on M SS Albumin Sys n Ref C:M Time after n Ref

last dose _____________________________________________________________________________________________________________________________________ Bupivacaine 95% 0.81 4µg/ml Y 2% HSA C 4 (37) 0.29 5-175 min 19 (38) 0.51 1µg/ml Y HP(M) / 4%HSA(F) C 6 (37) 0.32 28.5 min 10 (39) 0.73 4µg/ml Y 0.2% HSA O 4 (37) 0.35 0.58-3.5 hr 9 (40)

0.74 1µg/ml Y 2% HSA C 4 (41) 0.34 31 min 6 (42) 0.40 1µg/ml Y HP(M) / 4%HSA(F) C 4 (41) 0.24 3-135 min 31 (43)

0.56 2µg/ml Y 200mg/l HSA C 5 (44) 0.44 40-500 min 12 (45) 0.27 71 min 20 (46)

0.41 50 min 14 (47) 0.25 42 min 10 (48) 0.14 5-30 min 8 (49) 0.31 62 min 23 (50) 0.26(0.79free) 1-1.5 hr 10 (51) 0.25 NR 20 (52) 0.69(free) NR 29 (53) 0.29 38 min 28 (54) 0.37(0.77free) SS 12 (55)

Page 187: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

171

_____________________________________________________________________________________________________________________________________ Lidocaine 33-80% 0.74 5µg/ml Y 2%HSA C 5 (56) 0.52 SS 1 (57)

1.10 5µg/ml Y HP(M) / 4%HSA(F) C 5 (56) 0.46 11 min 23 (58) 0.90 2µg/ml Y 200mg/l HSA C 5 (44) 0.48 10-71 min 18 (59) 0.55 7-39 min 9 (60) 0.54 40 min 11 (61)

0.69 NR 9 (62) 0.66 NR 19 (63) 0.48 30 min 29 (54) 0.37 15-18 min 10 (64)

_____________________________________________________________________________________________________________________________________ Ropivacaine 94% 0.82 1µg/ml Y 2%HSA C 4 (41) 0.28(0.81free) 77-103 min 11 (51)

0.42 1µg/ml Y HP(M) / 4%HSA(F) C 4 (41) 0.72(free) NR 31 (53) 0.30 1µg/ml Y 4%BSA O 6 (65) 0.31(0.74free) NR 9 (66)

0.33(0.66free) SS 12 (55) _____________________________________________________________________________________________________________________________________ Propofol 97-99% 0.51 15µg/ml Y 22g/l_M-44g/l_F O 6 (67) 0.65 5-14 min 20 (68)

0.13 15µg/ml Y 44g/l O 6 (67) 0.70 13-45 min 10 (69) 0.74 8.7-12.3min 8 (70) 0.85 4.0-7.4 min 10 (71) 0.72(0.2-1.7) 5-23 min 21 (72) 0.68 2-10min 13 (73) 0.22 4-10min 10 (74)

_____________________________________________________________________________________________________________________________________ Methohexital 80-85% 1.0 25µg/ml Y 2%HSA C 3 (75) -rapid transfer and approaches 1 1 (76)

0.66 25µg/ml Y M>F HSA C 9 (75) _____________________________________________________________________________________________________________________________________ Antidepressants

Perfusion In vivo Drug PB F:M [ ] on M SS Albumin Sys n Ref C:M Time after n Ref

last dose _____________________________________________________________________________________________________________________________________ Nortriptyline 86-95% 0.22 150 ng/ml Y 30 g/l BSA O 9 (77) 0.43 SS 1 (78)

0.68 SS 10 (79) _____________________________________________________________________________________________________________________________________ Citalopram 80% 0.30 1.23 µM Y 30g/l BSA O 8 (80) 0.71 SS 4 (81) 0.16 1.23 µM Y No O 5 (80) 0.83 SS 9 (82)

Page 188: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

172

(Desmethylcitalopram) 0.19 0.60 µM Y 30g/l BSA O 8 (80) 0.54 SS 2 (81)

0.86 SS 9 (82) _____________________________________________________________________________________________________________________________________ Fluoxetine 95% 0.29 1.45 µM Y 30g/l BSA O 7 (80) 0.64 SS 15 (81) 0.04 1.45 µM Y No O 4 (80) 0.67 SS 4 (83) 0.91 SS 6 (84)

0.72 SS 2 (82) (Norfluoxetine) 0.30 1.23 µM Y 30g/l BSA O 7 (80) 0.65 SS 15 (81) 0.71 SS 4 (83)

1.04 SS 6 (84) 0.78 SS 2 (82)

_____________________________________________________________________________________________________________________________________ Antiepileptic

Perfusion In vivo Drug PB F:M [ ] on M SS Albumin Sys n Ref C:M Time after n Ref

last dose _____________________________________________________________________________________________________________________________________ Carbamazepine 76% 1.0 50µg/ml Y No C 6 (85) ~1 SS 16 (85)

~1 SS 5 (86) 0.73(1.42free) SS 7 (87) 0.78 SS 5 (88) 0.63*** 0.5->24hr 18 (89)

_____________________________________________________________________________________________________________________________________ Oxcarbazepine 40-60% ~1.1 50µg/ml Y No C 3 (90) ~1 SS 1 (91)

2.35 SS 3 (90) 2.40 SS 7 (92)

_____________________________________________________________________________________________________________________________________ Phenytoin 88-93% 0.73 0.16-0.5µg/ml Y No C 10 (93) 0.97 SS 3 (94)

0.94 0.7µg/ml Y No C 2 (95) 0.91 SS 3 (96) ~0.58 20µg/ml Y No O 4 (97) 0.91(1.1free) SS 7 (87)

~1 SS 8 (86) _____________________________________________________________________________________________________________________________________

Page 189: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

173

Phenobarbital 20-60% 0.99 0.5µg/ml Y No C 2 (95) 1.05 SS 31 (98) ~1 SS 25 (99) 0.87 SS 23 (100) 0.95 SS 5 (94) 1.05 SS 25 (101) 0.86(1.13free) SS 13 (87) 0.84 SS 27 (102)

_____________________________________________________________________________________________________________________________________ Diazepam 94-99% 0.48 2µg/ml Y No C 4 (103) 1.31 0.1-9.5 hr 16 (104)

0.55 200ng/ml Y No C 3 (103) 1.73(0.92free) 1.5-3 hr 5 (105) 1.32 0.5-3.5 hr 7 (106)

1.46 ~15 hr 26 (107) 1.3 30-140 min 16 (108)

1.9 17-265 min 6 (108) 1.44 12 min 6 (109)

0.57 3-13 min 30 (110) 1.8 5-401min 37 (111) 1.69 NR 10 (112) >1*** 1-6 hr 8 (113) 0.84 3-13.5 min 30 (114) 2.4 0.6-10.9 hr 15 (115) 0.82 1-5 min 33 (116) 1.01 4-514 min 18 (117) >1 NR 20 (118)

_____________________________________________________________________________________________________________________________________ Valproic Acid 90% 0.90 80mg/l Y 0.2%BSA C 6 (119) 1.43 SS 1 (120) ~0.85 67µg/ml Y No C 4 (121) 1.7 SS 6 (122) ~0.50 20µg/ml Y No O 4 (121) 1.59 SS 8 (87)

1.71 SS 4 (94) 1.38 SS 18 (123)

_____________________________________________________________________________________________________________________________________ Lamotrigine 55% 0.83 2.5µg/ml Y No C 4 (124) 0.91 SS 51 (125)

1.26 10µg/ml Y No C 4 (124) 1.29 SS 2 (124) 0.9 SS 9 (126) 1.2 SS 1 (127) 0.90 SS 4 (128) 1.01 SS 6 (129)

_____________________________________________________________________________________________________________________________________

Page 190: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

174

Antipsychotic Perfusion In vivo

Drug PB F:M [ ] on M SS Albumin Sys n Ref C:M Time after n Ref last dose

_____________________________________________________________________________________________________________________________________ Olanzapine 93% 0.20 10 ng/ml Y No C 3 (130) 0.32 SS 1 (131)

0.72 SS 13 (132) 0.3 SS 1 (133)

_____________________________________________________________________________________________________________________________________ Quetiapine 83% 0.12 75ng/ml Y 30g/l BSA O 6 (134) 0.24 SS 20 (132) _____________________________________________________________________________________________________________________________________ Antivirals

Perfusion In vivo Drug PB F:M [ ] on M SS Albumin Sys n Ref C:M Time after n Ref

last dose _____________________________________________________________________________________________________________________________________ Protease Inhibitors Indinavir 60% 0.06 7.6 mg/l Y No O 5 (135) <LOD 12h 21 (136)

0.04 7.6 mg/l Y No O 12 (137) 0.01 SS 4 (138) _____________________________________________________________________________________________________________________________________ Lopinavir 98-99% 0.30 1.1-10.6 mg/l NR 2g/l BSA O 6 (139) <250ng/ml SS 1 (140)

0.24 1.1-10.6 mg/L NR 10 g/l BSA O 3 (139) 0.22 SS 11 (141) 0.03 1.1-10.6 mg/L NR 40 g/l BSA O 3 (139) <LOD SS 11 (142)

0.23 SS 23 (143) 0.57 SS 6 (144) 0.2 SS 10 (145)

_____________________________________________________________________________________________________________________________________ Nelfinavir >98% 0.001 ~430 µg/l Y 2 g/l BSA O 3 (146) 0-0.3 SS 9 (140)

0.14 ~1.7 mg/l Y 2 g/l BSA O 6 (146) 0.24 SS 8 (141) 0.14 ~4.4 mg/ l Y 2 g/l BSA O 4 (146) 0.37 SS 7 (142)

0.49 SS 3 (144) 0.25 SS 73 (147) 0.24 SS 39 (138)

_____________________________________________________________________________________________________________________________________ Ritonavir 98-99% <LOD 1-2 µg/ml NR No C 2 (148) <LOD SS 2 (140)

0.12 0.3-1.1mg/l NR 2 g/l BSA O 3 (139) 0.31 SS 1 (141) 0.08 0.3-1.1mg/l NR 10 g/l BSA O 3 (139) 0.11 SS 15 (141) 0.00 0.3-1.1mg/l NR 40 g/l BSA O 3 (139) 0.55 SS 6 (144)

0.00 SS 10 (138)

Page 191: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

175

_____________________________________________________________________________________________________________________________________ Saquinavir 97% 0.02 322-2197 ng/ml Y 2 g/l HSA O 10 (149) <LOD SS 1 (140)

0.04 SS 4 (141) <LOD SS 6 (138)

_____________________________________________________________________________________________________________________________________ Nucleoside Analogue Acyclovir 9-33% ~0.43 1µg/ml No 0.2g/L C 4 (150) 0.77 SS 10 (151)

0.16 1µg/ml NR 3%BSA C 4 (152) 0.74 SS 12 (153) 0.09 10µg/ml NR 3%BSA C 4 (152) 0.92 SS 5 (154) 0.89 100µg/ml NR 3%BSA C 2 (152) _____________________________________________________________________________________________________________________________________ Dideoxyinosine <5% ~0.35 1µM NR 0.1g/dl BSA O 5 (155) 0.38 SS 10 (156) (DDI) ~0.39 500µM NR 0.1g/dl BSA O 5 (155) 0.17 1 hr 2 (157)

~0.56 30µM No 2 g/l BSA C 5 (158)

_____________________________________________________________________________________________________________________________________ Lamivudine <36% 0.21 1.39µg/l NR 3% BSA C 3 (159) 1.12 SS 16 (160)

0.11 14.68µg/l NR 3% BSA C 3 (159) 0.96 SS 59 (156) 1.00 SS 10 (144) 1.31 2.3-35 hr 50 (161)

_____________________________________________________________________________________________________________________________________ Zidovudine <38% 0.90 1.0mg/ml No NR C 2 (162) 1.13-1.27 SS 2 (163)

~1.0 3.8mM Y N C 4 (164) 1.22 SS 75 (156) ~0.75 3µM No 0.2g/L HSA C 3 (165) 0.86 SS 10 (144) ~0.45 3µM No 0.1g/dl C 5 (166) 1.89*** 60 min 7 (167)

~1.0 10µM Y N C 5 (168) 3.55*** 2.5&3hr 2 (169) 1.25 0.25-4.5hr 7 (170)

_____________________________________________________________________________________________________________________________________ Antiretroviral Fusion Inhibitor Enfuvirtide 92% <LOD 12.4µg/ml** NR 2g/L HSA O 3 (171) <LOD SS 2 (172) _____________________________________________________________________________________________________________________________________

Page 192: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

176

Asthma Perfusion In vivo

Drug PB F:M [ ] on M SS Albumin Sys n Ref C:M Time after n Ref last dose

_____________________________________________________________________________________________________________________________________ Theophylline 40% 0.33 15mg/l Y 0.2%BSA O 8 (173) 1.06 SS 9 (174)

0.88 15mg/l Y 0.2%BSA C 6 (175) 0.96 5&9 hr 2 (176) 1.10 SS 10 (177) 0.91 <6hr 3 (178)

_____________________________________________________________________________________________________________________________________ Salbutamol 10% 0.12 2µg/ml Y No O 7 (179) 0.70 27-105min 5 (180) _____________________________________________________________________________________________________________________________________ H-blockers

Perfusion In vivo Drug PB F:M [ ] on M SS Albumin Sys n Ref C:M Time after n Ref

last dose _____________________________________________________________________________________________________________________________________ Cimetidine 13-26% ~0.15 8µg/ml No 1 g/L BSA O 4 (8) 0.39 5-780min 17 (181) 0.46 4µg/ml No 0.1% BSA C 4 (182) 0.36 1-2hr 8 (183)

~0.5 1µg/ml No 2%HSA+13%HP C 6 (184) 0.6 1-1.5hr 16 (185) 0.8 1.5-2hr 16 (185)

_____________________________________________________________________________________________________________________________________ Ranitidine 15% ~0.4 60µg/ml No 2%HSA+13%HP C 3 (184) 0.9 0.5-8 hr (IV) 20 (186)

0.38 1-9 hr (oral) 45 (186) 0.95 5 hr 1 (187)

_____________________________________________________________________________________________________________________________________ Antimicrobials

Perfusion In vivo Drug PB F:M [ ] on M SS Albumin Sys n Ref C:M Time after n Ref

last dose _____________________________________________________________________________________________________________________________________ Pyrimethamine 87% 0.30 0.8mg/l NS 0-20g/l O 2 (188) 0.66 0.5-13 hr 10 (189)

0.30 5mg/l NS 0-20g/l O 11 (188) 0.5-1.0 SS 4 (190) 0.32 10mg/l NS 0-20g/l O 2 (188)

_____________________________________________________________________________________________________________________________________

Page 193: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

177

Erythromycin 75-90% 0.10 2µg/ml Y N O 7 (191) 0.22 SS 12 (192) 0.06 NR 4 (193)

_____________________________________________________________________________________________________________________________________ Azithromycin 7-50% 0.09 0.3µg/ml Y N O 7 (191) limited transfer 6-168 hr 20 (194) _____________________________________________________________________________________________________________________________________ Ofloxacin 20-32% 0.07 10µg/ml No 3g/l C 6 (195) 0.8-0.9 3-4 hr 11 (196) 0.6-0.7 NR 8 (197) _____________________________________________________________________________________________________________________________________ Ticarcillin 45% 0.02 10-94µg/ml No 3g/dl BSA O 5 (198) 0.6 NR 5 (199) Clavulanic acid 25% <LOD 2-6µg/ml No 3g/dl BSA O 4 (198) 0.8 NR 5 (199)

0.05 10µg/ml No 3g/dl BSA O 4 (198) _____________________________________________________________________________________________________________________________________ Ceftizoxime 28-50% 0.03 56-166µg/ml No 3% BSA O 6 (200) 1.6 1-4 hr 20 (201) _____________________________________________________________________________________________________________________________________ Sulbactam 38% 0.03 40µg/ml N 3g/dl BSA O 4 (202) 1.3 NR 5 (199)

0.04 27-166µg/ml N 3g/dl BSA O 4 (202) _____________________________________________________________________________________________________________________________________ Diabetic Agents

Perfusion In vivo Drug PB F:M [ ] on M SS Albumin Sys n Ref C:M Time after n Ref

last dose _____________________________________________________________________________________________________________________________________ Insulin Lispro <LOD 100µU/ml N/A No O 4 (203) <LOD NR 4 (204)

<LOD 200µU/ml N/A No O 1 (203) <LOD 100µU/ml N/A NR C 4 (205) _____________________________________________________________________________________________________________________________________ Metformin 0.55 5µg/ml Y No C 12 (206) ~1.0?? 3.5-336 hr 7 (207)

0.16 5µg/ml Y No O 6 (206) 2.0 4-32 hr 8 (208) ~0.30 1µg/ml Y 3g/l C 6 (209) 0.11 1µg/ml NR 3g/l C 7 (210) 0.17 10mg/ml NR 3g/l C 2 (210)

_____________________________________________________________________________________________________________________________________

Page 194: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

178

Glyburide 99% 0.006 1µg/ml NR 2g/dl C 3 (211) <LOD 8 hr 12 (212) 0.02 20µg/ml NR 2g/dl C 1 (211) 0.7 NR 66 (213)

0.26 150ng/ml NR No O 6 (214) 0.29 150ng/ml NR 42µg/mlHSA O 2 (214) 0.13 150ng/ml NR 210µg/mlHSA O 2 (214) 0.11 150ng/ml NR 630µg/mlHSA O 2 (214) 0.03 150ng/ml NR 2.1mg/mlHSA O 2 (214) 0.09 150ng/ml N No C 4 (215) 0.03 150ng/ml Y 30mg/mlHSA C 3 (215)

_____________________________________________________________________________________________________________________________________ Cardiac Drugs

Perfusion In vivo Drug PB F:M [ ] on M SS Albumin Sys n Ref C:M Time after n Ref

last dose _____________________________________________________________________________________________________________________________________ Digoxin 20-25% 0.39 5ng/ml No No C 4 (216) 0.26 SS 1 (217)

0.11 5.5ng/ml Y 3g/L BSA C 6 (218) 0.97 NR 3 (219) <LOD 5.5ng/ml Y 1mg/ml BSA C 5 (220) ~1 5-7 hr 7 (221) 0.11 5.5ng/ml Y 3mg/ml BSA C 4 (220) 0.27 SS/C 1 (222) 0.12 5.5ng/ml Y 5mg/ml BSA C 4 (220) 0.97*** 0.25-24 hr 6 (223) 0.70 5.7ng/ml No 21g/l BSA C 5 (224) 0.62 SS 1 (225) 0.61 5.7ng/ml No M>F BSA C 5 (224)

0.73 5.7ng/ml No 3g/l BSA C 5 (224) _____________________________________________________________________________________________________________________________________ Enalaprilat 50-60% ~0.7 150ng/ml No No C 5 (226) 0.59 NS 1 (226) _____________________________________________________________________________________________________________________________________ Clonidine 20-40% 0.85 4.4nM No 200mg/l HSA C 4 (227) ~1 SS 10 (228)

0.87 1-13.5 hr 5 (229) 1.0 SS 6 (230) 0.92 117 min 8 (231)

_____________________________________________________________________________________________________________________________________ Fondaparinux 94% <LOD 1.75 mg/l N/A 10g/l HSA O 6 (232) 0.1 22hrPD-SS 5 (233)

<LOD >24hr 1 (234) <0.07 >24hr 1 (235)

_____________________________________________________________________________________________________________________________________

Page 195: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

179

Heparin 0.004 650IU/ml No No C 12 (236) NC 15-170 min 7 (237) _____________________________________________________________________________________________________________________________________ Low Molecular Weight Heparins Enoxaparin 80% <LOD 1 IU/mL N/A 10g/l HSA O 6 (232) NC*** 3 hr 14 (238) Other LMWHs NC*** 3 hr 5 (239)

NC*** 3 hr 7 (240) <LOD, NC*** 3&15 hr 15 (241) NC*** 3-7 hr 9 (242)

_____________________________________________________________________________________________________________________________________ Nitroglycerin 11-60% 0.19 100 nmol/l Y No O 6 (243) <0.01 1 min 32 (244) _____________________________________________________________________________________________________________________________________ Hydralazine 88-90% 0.58 15µg/ml NR No O 4 (245) ~1 SS 4 (246) 0.30 115µg/ml NR No O 4 (245) 0.40 750µg/ml NR No O 4 (245) _____________________________________________________________________________________________________________________________________ Atenolol <5% 0.09 5µg/ml Y 0.02&4g/dlHSA O 6 (247) 0.88 SS 6 (248)

~0.85 5µg/ml No 2g/dlHSA C 3 (247) _____________________________________________________________________________________________________________________________________ Celiprolol 22-24% 0.05 5µg/ml Y 0.02&4g/dlHSA O 2 (247) 0.28 SS 4 (249)

0.11 5µM NR No O 5 (250) _____________________________________________________________________________________________________________________________________ Labetalol 50% 0.27 10µg/ml Y 0.02g/dlHSA O 3 (247) ~0.5 NR 25 (251)

0.22 1µg/ml Y 4g/dl HSA O 2 (247) 0.10 6nmol/ml Y No NR 6 (252)

_____________________________________________________________________________________________________________________________________ Propranolol 93% 0.21 5µg/ml Y 0.02g/dl HSA O 5 (247) 0.26 3 hr 8 (253)

0.32 5µg/ml Y 4g/dl HSA O 5 (247) ~1 5µg/ml Y 2g/dl HSA C 3 (247) ~1.38 5µg/ml Y M<F HSA C 3 (247)

_____________________________________________________________________________________________________________________________________ Nifedipine 90-96% 0.18 0.2µM NS No O 5 (250) 0.76 2.5 hr 10 (254) _____________________________________________________________________________________________________________________________________

Page 196: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

180

Immunological Agents Perfusion In vivo

Drug PB F:M [ ] on M SS Albumin Sys n Ref C:M Time after n Ref last dose

_____________________________________________________________________________________________________________________________________ Cyclosporin 90% <0.05 1.66µM NR No NR 6 (255) 0.56 8&10 hr 2 (256)

0.2 6 hr 1 (257) ~1 20 hr 1 (258)

Endocrine Agents

Perfusion In vivo Drug PB F:M [ ] on M SS Albumin Sys n Ref C:M Time after n Ref

last dose _____________________________________________________________________________________________________________________________________ Methimazole ~1 1.5µg/ml Y No C 3 (259) 0.77*** 2 hr 2 (260)

~1 15µg/ml Y No C 3 (259) ~1 1.5µg/ml Y 40g/l BSA C 3 (259)

_____________________________________________________________________________________________________________________________________ Propylthiouracil 80% ~1 4µg/ml Y No C 3 (259) 0.31*** 2 hr 2 (260)

~1 40µg/ml Y No C 3 (259) 1.99 1-9 hr 5 (261) ~1 4µg/ml Y 40g/l BSA C 3 (259)

_____________________________________________________________________________________________________________________________________ Tocolytic

Perfusion In vivo Drug PB F:M [ ] on M SS Albumin Sys n Ref C:M Time after n Ref

last dose _____________________________________________________________________________________________________________________________________ Ritodrine 32-38% 0.16 2µg/ml Y No O 4 (262) 0.58 NR 9 (263)

0.40 1.4-2.8µg/ml No 20g/L HSA C 8 (264) 1.17 10-990 min 28 (265) 0.30 120 min 8 (266) 0.70 SS 5 (267) 0.26 45-190 min 7 (268) 1.24 SS 19 (269)

_____________________________________________________________________________________________________________________________________ PB=protein binding [ ] on M= concentration of drug added into the maternal perfusate Sys=System configuration for perfusion experiment

Page 197: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

181

BSA=bovine serum albumin HSA=human serum albumin HP=human plasma O=open C=closed NC=no change in anticoagulation activity compared to a control group No change in anti-Xa or anti-IIa activity ~ =calculated/estimated from data provided AAG=α1-acid glycoprotein NR=not reported SS=steady state ***placenta or in vivo results are from preterm pregnancies

Page 198: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

182

Supplementary Table References

1. Jacobson, R.L., Brewer, A., Eis, A., Siddiqi, T.A., Myatt, L. Transfer of aspirin across the perfused human placental cotyledon. Am J Obstet Gynecol 165, 939-944 (1991).

2. Wolff, F., Berg, R., Bolte, A., Putter, J. Perinatal pharmacokinetics of acetylsalicylic acid. Arch Gynecol 233, 15-22 (1982).

3. Shintaku, K. et al. Kinetic analysis of the transport of salicylic acid, a nonsteroidal anti-inflammatory drug, across human placenta. Drug Metab Dispos 35, 772-778 (2007).

4. Garrettson, L.K., Procknal, J.A., Levy, G. Fetal acquisition and neonatal elimination of a large amount of salicylate. Study of a neonate whose mother regularly took therapeutic doses of aspirin during pregnancy. Clin Pharmacol Ther 17, 98-103 (1975).

5. Levy, G., Garrettson, L.K. Kinetics of salicylate elimination by newborn infants of mothers who ingested aspirin before delivery. Pediatrics 53, 201-210 (1974).

6. Noschel, H., Bonow, A., Moller, R., Estel, C., Muller, B. [Placental passage of sodium salicylate]. Zentralbl Gynakol 94, 437-442 (1972).

7. Elis, J., Sechserova, M., Stribrny, J., Drabkova, J. The distribution of sodium salicylate in the human fetus. Int J Clin Pharmacol Biopharm 16, 365-367 (1978).

8. Bassily, M., Ghabrial, H., Smallwood, R.A., Morgan, D.J. Determinants of placental drug transfer: Studies in the isolated perfused human placenta. J Pharm Sci 84, 1054-1060 (1995).

9. Siu, S.S., Yeung, J.H., Lau, T.K. A study on placental transfer of diclofenac in first trimester of human pregnancy. Hum Reprod 15, 2423-2425 (2000).

10. Shintaku, K. et al. Transplacental pharmacokinetics of diclofenac in perfused human placenta. Drug Metab Dispos 37, 962-968 (2009).

11. Nanovskaya, T., Deshmukh, S., Brooks, M., Ahmed, M.S. Transplacental transfer and metabolism of buprenorphine. J Pharmacol Exp Ther 300, 26-33 (2002).

12. Concheiro, M., Jones, H.E., Johnson, R.E., Choo, R., Shakleya, D.M., Huestis, M.A. Maternal buprenorphine dose, placenta buprenorphine, and metabolite concentrations and neonatal outcomes. Ther Drug Monit 32, 206-215 (2010).

13. Nanovskaya, T.N., Bowen, R.S., Patrikeeva, S.L., Hankins, G.D., Ahmed, M.S. Effect of plasma proteins on buprenorphine transfer across dually perfused placental lobule. J Matern Fetal Neonatal Med 22, 646-653 (2009).

14. Nekhayeva, I.A., Nanovskaya, T.N., Hankins, G.D.V., Ahmed, M.S. Role of human placental efflux transporter P-glycoprotein in the transfer of buprenorphine, levo-alpha-acetylmethadol, and paclitaxel. Am J Perinatol 23, 423-430 (2006).

Page 199: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

183

15. Nanovskaya, T., Nekhayeva, I., Karunaratne, N., Audus, K., Hankins, G.D.V., Ahmed, M.S. Role of P-glycoprotein in transplacental transfer of methadone. Biochem Pharmacol 69, 1869-1878 (2005).

16. Harper, R.G., Solish, G., Feingold, E., Gersten-Woolf, N.B., Sokal, M.M. Maternal ingested methadone, body fluid methadone, and the neonatal withdrawal syndrome. Am J Obstet Gynecol 129, 417-424 (1977).

17. Nekhayeva, I.A., Nanovskaya, T.N., Deshmukh, S.V., Zharikova, O.L., Hankins, G.D., Ahmed, M.S. Bidirectional transfer of methadone across human placenta. Biochem Pharmacol 69, 187-197 (2005).

18. Rosen, T.S., Pippenger, C.E. Disposition of methadone and its relationship to severity of withdrawal in the newborn. Addict Dis 2, 169-178 (1975).

19. Nanovskaya, T.N., Nekhayeva, I.A., Hankins, G.D.V., Ahmed, M.S. Transfer of methadone across the dually perfused preterm human placental lobule. Am J Obstet Gynecol 198, (2008).

20. Malek, A., Obrist, C., Wenzinger, S., von Mandach, U. The impact of cocaine and heroin on the placental transfer of methadone. Reprod Biol Endocrinol 7, 61 (2009).

21. Higby, K., Elliott, B., King, T.S., Frasier, D., Langer, O. Human placental transfer of the prostaglandin inhibitor sulindac using an in vitro model. J Soc Gynecol Investig 2, 526-530 (1995).

22. Kramer, W.B. et al. Placental transfer of sulindac and its active sulfide metabolite in humans. Am J Obstet Gynecol 172, 886-890 (1995).

23. Lampela, E.S. et al. Placental transfer of sulindac, sulindac sulfide, and indomethacin in a human placental perfusion model. Am J Obstet Gynecol 180, 174-180 (1999).

24. Moise, K.J.J., Ou, C.N., Kirshon, B., Cano, L.E., Rognerud, C., Carpenter, R.J.J. Placental transfer of indomethacin in the human pregnancy. Am J Obstet Gynecol 162, 549-554 (1990).

25. Kopecky, E.A., Simone, C., Knie, B., Koren, G. Transfer of morphine across the human placenta and its interaction with naloxone. Life Sci 65, 2359-2371 (1999).

26. Gerdin, E., Rane, A., Lindberg, B. Transplacental transfer of morphine in man. J. Perinat. Med. 18, 305-312 (1990).

27. Krishna, B.R., Zakowski, M.I., Grant, G.J. Sufentanil transfer in the human placenta during in vitro perfusion. Can J Anaesth 44, 996-1001 (1997).

28. Meuldermans, W., Woestenborghs, R., Noorduin, H., Camu, F., van Steenberge, A., Heykants, J. Protein binding of the analgesics alfentanil and sufentanil in maternal and neonatal plasma. Eur J Clin Pharmacol 30, 217-219 (1986).

Page 200: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

184

29. Loftus, J.R., Hill, H., Cohen, S.E. Placental transfer and neonatal effects of epidural sufentanil and fentanyl administered with bupivacaine during labor. Anesthesiology 83, 300-308 (1995).

30. Zakowski, M.I., Ham, A.A., Grant, G.J. Transfer and uptake of alfentanil in the human placenta during in vitro perfusion. Anesth Analg 79, 1089-1093 (1994).

31. Gepts, E., Heytens, L., Camu, F. Pharmacokinetics and placental transfer of intravenous and epidural alfentanil in parturient women. Anesth Analg 65, 1155-1160 (1986).

32. Heytens, L., Cammu, H., Camu, F. Extradural analgesia during labour using alfentanil. Br J Anaesth 59, 331-337 (1987).

33. Cartwright, D.P., Dann, W.L., Hutchinson, A. Placental transfer of alfentanil at caesarean section. Eur J Anaesthesiol 6, 103-109 (1989).

34. Weigand, U.W., Chou, R.C., Maulik, D., Levy, G. Assessment of biotransformation during transfer of propoxyphene and acetaminophen across the isolated perfused human placenta. Pediatr Pharmacol 4, 145-153 (1984).

35. Roberts, I., Robinson, M.J., Mughal, M.Z., Ratcliffe, J.G., Prescott, L.F. Paracetamol metabolites in the neonate following maternal overdose. Br J Clin Pharmacol 18, 201-206 (1984).

36. Sancewicz-Pach, K., Chmiest, W., Lichota, E. Suicidal paracetamol poisoning of a pregnant woman just before a delivery. Przegl Lek 56, 459-462 (1999).

37. Johnson, R.F., Herman, N., Arney, T.L., Gonzalez, H., Johnson, H.V., Downing, J.W. Bupivacaine transfer across the human term placenta: A study using the dual perfused human placental model. Anesthesiology 82, 459-468 (1995).

38. Decocq, G. et al. Serum bupivacaine concentrations and transplacental transfer following repeated epidural administrations in term parturients during labour. Fundam Clin Pharmacol 11, 365-370 (1997).

39. de Barros Duarte, L., Moises, E.C., Carvalho Cavalli, R., Lanchote, V.L., Duarte, G., Pereira da Cunha, S. Placental transfer of bupivacaine enantiomers in normal pregnant women receiving epidural anesthesia for cesarean section. Eur J Clin Pharmacol 63, 523-526 (2007).

40. Papini, O., Mathes, A.C., Cunha, S.P., Lanchote, V.L. Stereoselectivity in the placental transfer and kinetic disposition of racemic bupivacaine administered to parturients with or without a vasoconstrictor. Chirality 16, 65-71 (2004).

41. Johnson, R.F. et al. A comparison of the placental transfer of ropivacaine versus bupivacaine. Anesth Analg 89, 703-708 (1999).

Page 201: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

185

42. Kuhnert, B.R., Zuspan, K.J., Kuhnert, P.M., Syracuse, C.D., Brown, D.E. Bupivacaine disposition in mother, fetus, and neonate after spinal anesthesia for cesarean section. Anesth Analg 66, 407-412 (1987).

43. Thomas, J., Long, G., Moore, G., Morgan, D. Plasma protein binding and placental transfer of bupivacaine. Clin Pharmacol Ther 19, 426-434 (1976).

44. Ala-Kokko, T.I., Pienimaki, P., Herva, R., Hollmen, A.I., Pelkonen, O., Vahakangas, K. Transfer of lidocaine and bupivacaine across the isolated perfused human placenta. Pharmacol Toxicol 77, 142-148 (1995).

45. Thomas, J., Climie, C.R., Mather, L.E. The maternal plasma levels and placental transfer of bupivacaine following epidural analgesia. Br J Anaesth 41, 1035-1040 (1969).

46. Scanlon, J.W., Ostheimer, G.W., Lurie, A.O., Brown wu, J.R., Weiss, J.B., Alper, M.H. Neurobehavioral responses and drug concentrations in newborns after maternal epidural anesthesia with bupivacaine. Anesthesiology 45, 400-405 (1976).

47. Magno, R., Berlin, A., Karlsson, K., Kjellmer, I. Anesthesia for cesarean section IV: placental transfer and neonatal elimination of bupivacaine following epidural analgesia for elective cesarean section. Acta Anaesthesiol Scand 20, 141-146 (1976).

48. McGuinness, G.A., Merkow, A.J., Kennedy, R.L., Erenberg, A. Epidural anesthesia with bupivacaine for Cesarean section: neonatal blood levels and neurobehavioral responses. Anesthesiology 49, 270-273 (1978).

49. Beazley, J.M., Taylor, G., Reynolds, F. Placental transfer of bupivacaine after paracervical block. Obstet. Gynecol. 39, 2-6 (1972).

50. Reynolds, F., Taylor, G. Maternal and neonatal blood concentrations of bupivacaine: a comparison with lignocaine during continuous extradural analgesia. Anaesthesia 25, 14-23 (1970).

51. Ala-Kokko, T.I., Alahuhta, S., Jouppila, P., Korpi, K., Westerling, P., Vahakangas, K. Feto-maternal distribution of ropivacaine and bupivacaine after epidural administration for cesarean section. Int J Obstet Anesth 6, 147-152 (1997).

52. Bader, A.M., Tsen, L.C., Camann, W.R., Nephew, E., Datta, S. Clinical effects and maternal and fetal plasma concentrations of 0.5% epidural levobupivacaine versus bupivacaine for cesarean delivery. Anesthesiology 90, 1596-1601 (1999).

53. Datta, S., Camann, W., Bader, A., VanderBurgh, L. Clinical effects and maternal and fetal plasma concentrations of epidural ropivacaine versus bupivacaine for cesarean section. Anesthesiology 82, 1346-1352 (1995).

54. Loftus, J.R., Holbrook, R.H., Cohen, S.E. Fetal heart rate after epidural lidocaine and bupivacaine for elective cesarean section. Anesthesiology 75, 406-412 (1991).

Page 202: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

186

55. Irestedt, L., Ekblom, A., Olofsson, C., Dahlstrom, A.C., Emanuelsson, B.M. Pharmacokinetics and clinical effect during continuous epidural infusion with ropivacaine 2.5 mg/ml or bupivacaine 2.5 mg/ml for labour pain relief. Acta Anaesthesiol Scand 42, 890-896 (1998).

56. Johnson, R.F. et al. Transfer of lidocaine across the dual perfused human placental cotyledon. Int J Obstet Anesth 8, 17-23 (1999).

57. Banzai, M., Sato, S., Tezuka, N., Komiya, H., Chimura, T., Hiroi, M. Placental transfer of lidocaine hydrochloride after prolonged continuous maternal intravenous administration. Can J Anaesth 42, 338-340 (1995).

58. de Carvalho Cavalli, R. et al. Pharmacokinetics and transplacental transfer of lidocaine and its metabolite for perineal analgesic assistance to pregnant women. Eur J Clin Pharmacol 60, 569-574 (2004).

59. Thomas, J., Climie, C.R., Long, G., Nighjoy, L.E. The influence of adrenaline on the maternal plasma levels and placental transfer of lignocaine following lumbar epidural administration. Br J Anaesth 41, 1029-1034 (1969).

60. Shnider, S.M., Way, E.L. The kinetics of transfer of lidocaine (Xylocaine) across the human placenta. Anesthesiology 29, 944-950 (1968).

61. Guay, J., Gaudreault, P., Boulanger, A., Tang, A., Lortie, L., Dupuis, C. Lidocaine hydrocarbonate and lidocaine hydrochloride for cesarean section: transplacental passage and neonatal effects. Acta Anaesthesiol Scand 36, 722-727 (1992).

62. Kileff, M.E., James, F.M.r., Dewan, D.M., Floyd, H.M. Neonatal neurobehavioral responses after epidural anesthesia for cesarean section using lidocaine and bupivacaine. Anesth Analg 63, 413-417 (1984).

63. Abboud, T.K., Sarkis, F., Blikian, A., Varakian, L., Earl, S., Henriksen, E. Lack of adverse neonatal neurobehavioral effects of lidocaine. Anesth Analg 62, 473-475 (1983).

64. Kuhnert, B.R., Philipson, E.H., Pimental, R., Kuhnert, P.M., Zuspan, K.J., Syracuse, C.D. Lidocaine disposition in mother, fetus, and neonate after spinal anesthesia. Anesth Analg 65, 139-144 (1986).

65. Zuo, M., Duan, G.-L., Ge, Z.-G. Simultaneous determination of ropivacaine and antipyrine by high performance liquid chromatography and its application to the in vitro transplacental study. Biomed Chromatogr 18, 752-755 (2004).

66. Morton, C.P., Bloomfield, S., Magnusson, A., Jozwiak, H., McClure, J.H. Ropivacaine 0.75% for extradural anaesthesia in elective caesarean section: an open clinical and pharmacokinetic study in mother and neonate. Br J Anaesth 79, 3-8 (1997).

67. He, Y.L., Seno, H., Sasaki, K., Tashiro, C. The influences of maternal albumin concentrations on the placental transfer of propofol in human dually perfused cotyledon in vitro. Anesth Analg 94, 1312-1314 (2002).

Page 203: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

187

68. Gin, T., Gregory, M.A., Chan, K., Oh, T.E. Maternal and fetal levels of propofol at caesarean section. Anaesth Intensive Care 18, 180-184 (1990).

69. Dailland, P. et al. Intravenous propofol during cesarean section: placental transfer, concentrations in breast milk, and neonatal effects. A preliminary study. Anesthesiology 71, 827-834 (1989).

70. Gin, T., Yau, G., Jong, W., Tan, P., Leung, R.K., Chan, K. Disposition of propofol at caesarean section and in the postpartum period. Br J Anaesth 67, 49-53 (1991).

71. Valtonen, M., Kanto, J., Rosenberg, P. Comparison of propofol and thiopentone for induction of anaesthesia for elective caesarean section. Anaesthesia 44, 758-762 (1989).

72. Moore, J., Bill, K.M., Flynn, R.J., McKeating, K.T., Howard, P.J. A comparison between propofol and thiopentone as induction agents in obstetric anaesthesia. Anaesthesia 44, 753-757 (1989).

73. Kanto, J., Rosenberg, P. Propofol in cesarean section. A pharmacokinetic and pharmacodynamic study. Methods Find Exp Clin Pharmacol 12, 707-711 (1990).

74. Sanchez-Alcaraz, A., Quintana, M.B., Laguarda, M. Placental transfer and neonatal effects of propofol in caesarean section. J Clin Pharm Ther 23, 19-23 (1998).

75. Herman, N.L. et al. Transfer of methohexital across the perfused human placenta. J Clin Anesth 12, 25-30 (2000).

76. Marshall, J.R. Human antepartum placental passage of methohexital sodium. Obtet Gynecol 23, 589-592 (1964).

77. Heikkinen, T., Ekblad, U., Laine, K. Transplacental transfer of amitriptyline and nortriptyline in isolated perfused human placenta. Psychopharmacology 153, 450-454 (2001).

78. Sjoqvist, F., Bergfors, P.G., Borga, O., Lind, M., Ygge, H. Plasma disappearance of nortriptyline in a newborn infant following placental transfer from an intoxicated mother: evidence for drug metabolism. J Pediatr 80, 496-500 (1972).

79. Loughhead, A.M., Stowe, Z.N., Newport, D.J., Ritchie, J.C., DeVane, C.L., Owens, M.J. Placental passage of tricyclic antidepressants. Biol Psychiatry 59, 287-290 (2006).

80. Heikkinen, T., Ekblad, U., Laine, K. Transplacental transfer of citalopram, fluoxetine and their primary demethylated metabolites in isolated perfused human placenta. BJOG 109, 1003-1008 (2002).

81. Hendrick, V., Stowe, Z.N., Altshuler, L.L., Hwang, S., Lee, E., Haynes, D. Placental passage of antidepressant medications. Am J Psychiatry 160, 993-996 (2003).

82. Rampono, J. et al. Placental transfer of SSRI and SNRI antidepressants and effects on the neonate. Pharmacopsychiatry 42, 95-100 (2009).

Page 204: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

188

83. Rampono, J., Proud, S., Hackett, L.P., Kristensen, J.H., Ilett, K.F. A pilot study of newer antidepressant concentrations in cord and maternal serum and possible effects in the neonate. Int J Neuropsychopharmacol 7, 329-334 (2004).

84. Kim, J. et al. Stereoselective disposition of fluoxetine and norfluoxetine during pregnancy and breast-feeding. Br J Clin Pharmacol 61, 155-163 (2006).

85. Pienimaki, P. et al. Carbamazepine and its metabolites in human perfused placenta and in maternal and cord blood. Epilepsia 36, 241-248 (1995).

86. Bardy, A.H., Hiilesmaa, V.K., Teramo, K., Neuvonen, P.J. Protein binding of antiepileptic drugs during pregnancy, labor, and puerperium. Ther Drug Monit 12, 40-46 (1990).

87. Takeda, A., Okada, H., Tanaka, H., Izumi, M., Ishikawa, S., Noro, T. Protein binding of four antiepileptic drugs in maternal and umbilical cord serum. Epilepsy Res. 13, 147-151 (1992).

88. Kuhnz, W. et al. Carbamazepine and carbamazepine-10,11- epoxide during pregnancy and postnatal period in epileptic mother and their nursed infants: pharmacokinetics and clinical effects. Pediatr Pharmacol (New York) 3, 199-208 (1983).

89. Pynnonen, S., Kanto, J., Sillanpaa, M., Erkkola, R. Carbamazepine: placental transport, tissue concentrations in foetus and newborn, and level in milk. Acta Pharmacol Toxicol (Copenh) 41, 244-253 (1977).

90. Pienimaki, P., Lampela, E., Hakkola, J., Arvela, P., Raunio, H., kangas, K. Pharmacokinetics of oxcarbazepine and carbamazepine in human placenta. Epilepsia 38, 309-316 (1997).

91. Bulau, P., Paar, W.D., von Unruh, G.E. Pharmacokinetics of oxcarbazepine and 10-hydroxy-carbazepine in the newborn child of an oxcarbazepine-treated mother. Eur J Clin Pharmacol 34, 311-313 (1988).

92. Myllynen, P., ki, P., Jouppila, P., kangas, K. Transplacental passage of oxcarbazepine and its metabolites in vivo. Epilepsia 42, 1482-1485 (2001).

93. Shah, Y.G., Miller, R.K. The pharmacokinetics of phenytoin in perfused human placenta. Pediatr Pharmacol 5, 165-179 (1985).

94. Ishizaki, T., Yokochi, K., Chiba, K., Tabuchi, T., Wagatsuma, T. Placental transfer of anticonvulsants (phenobarbital, phenytoin, valproic acid) and the elimination from neonates. Pediatr Pharmacol (New York) 1, 291-303 (1981).

95. Kluck, R.M., Cannell, G.R., Hooper, W.D., Eadie, M.J., Dickinson, R.G. Disposition of phenytoin and phenobarbitone in the isolated perfused human placenta. Clin Exp Pharmacol Physiol 15, 827-836 (1988).

96. Mirkin, B.L. Diphenylhydantoin: placental transport, fetal localization, neonatal metabolism, and possible teratogenic effects. J Pediatr 78, 329-337 (1971).

Page 205: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

189

97. Dickinson, R.G., Fowler, D.W., Kluck, R.M. Maternofetal transfer of phenytoin, p-hydroxy-phenytoin and p-hydroxy-phenytoin-glucuronide in the perfused human placenta. Clin Exp Pharmacol Physiol 16, 789-797 (1989).

98. Rayburn, W., Donn, S., Compton, A., Piehl, E. Oral phenobarbital given antenatally to reduce neonatal intraventricular hemorrhage. A comparison between maternal and umbilical cord serum levels at delivery. J Perinatol 9, 268-270 (1989).

99. Melchior, J.C., Svensmark, O., Trolle, D. Placental transfer of phenobarbitone in epileptic women, and elimination in newborns. Lancet 2, 860-861 (1967).

100. De Carolis, M.P. et al. Placental transfer of phenobarbital: what is new? Dev Pharmacol Ther 19, 19-26 (1992).

101. Shankaran, S., Cepeda, E.E., Ilagan, N., Kauffman, R.E. Pharmacokinetic basis for antenatal dosing of phenobarbital for the prevention of neonatal intracerebral hemorrhage. Dev Pharmacol Ther 9, 171-177 (1986).

102. Kuhnz, W., Koch, S., Helge, H., Nau, H. Primidone and phenobarbital during lactation period in epileptic women: total and free drug serum levels in the nursed infants and their effects on neonatal behavior. Dev Pharmacol Ther 11, 147-154 (1988).

103. Myllynen, P., Vahakangas, K. An examination of whether human placental perfusion allows accurate prediction of placental drug transport: studies with diazepam. J Pharmacol Toxicol Methods 48, 131-138 (2002).

104. Moore, R.G., McBride, W.G. The disposition kinetics of diazepam in pregnant women at parturition. Eur J Clin Pharmacol 13, 275-284 (1978).

105. Ridd, M.J., Brown, K.F., Nation, R.L., Collier, C.B. The disposition and placental transfer of diazepam in cesarean section. Clin Pharmacol Ther 45, 506-512 (1989).

106. Cavanah, D., Condo, C.S. Diazepam-- A pilot study of drug concentrations in maternal blood, amniotic fluid and cord blood. Curr Ther Res Clin Exp 6, 122-126 (1964).

107. Cree, J.E., Meyer, J., Hailey, D.M. Diazepam in labour: its metabolism and effect on the clinical condition and thermogenesis of the newborn. Br Med J 4, 251-255 (1973).

108. Kanto, J., Erkkola, R., Sellman, R. Accumulation of diazepam and N-demethyldiazepam in the fetal blood during the labour. Ann Clin Res 5, 375-379 (1973).

109. Mandelli, M. et al. Placental transfer to diazepam and its disposition in the newborn. Clin Pharmacol Ther 17, 564-572 (1975).

110. Bakke, O.M., Haram, K., Lygre, T., Wallem, G. Comparison of the placental transfer of thiopental and diazepam in caesarean section. Eur J Clin Pharmacol 21, 221-227 (1981).

111. Erkkola, R., Kangas, L., Pekkarinen, A. The transfer of diazepam across the placenta during labour. Acta Obstet Gynecol Scand 52, 167-170 (1973).

Page 206: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

190

112. Owen, J.R., Irani, S.F., Blair, A.W. Effect of diazepam administered to mothers during labour on temperature regulation of neonate. Arch Dis Child 47, 107-110 (1972).

113. Idanpaan-Heikkila, J.E., Jouppila, P.I., Puolakka, J.O., Vorne, M.S. Placental transer and fetal metabolism of diazepam in early human pregnancy. Am J Obstet Gynecol 109, 1011-1016 (1971).

114. Haram, K., Bakke, O.M. Diazepam as an induction agent for caesarean section: a clinical and pharmacokinetic study of fetal drug exposure. Br J Obstet Gynaecol 87, 506-512 (1980).

115. Kanto, J., Scheinin, M. Placental and blood-CSF transfer of orally administered diazepam in the same person. Pharmacol Toxicol 61, 72-74 (1987).

116. Haram, K., Bakke, O.M., Johannessen, K.H., Lund, T. Transplacental passage of diazepam during labor: influence of uterine contractions. Clin Pharmacol Ther 24, 590-599 (1978).

117. Desilva, J.A., D'Arconte, L., Kaplan, J. The determination of blood levels and the placental transfer of diazepam in humans. Curr Ther Res Clin Exp 6, 115-121 (1964).

118. Eliot, B.W., Hill, J.G., Cole, A.P., Hailey, D.M. Continuous pethidine/diazepam infunsion during labour and its effects on the newborn. Br J Obstet Gynaecol 82, 126-131 (1975).

119. Barzago, M.M. et al. Placental transfer of valproic acid after liposome encapsulation during in vitro human placenta perfusion. J Pharmacol Exp Ther 277, 79-86 (1996).

120. Dickinson, R.G., Harland, R.C., Lynn, R.K., Smith, W.B., Gerber, N. Transmission of valproic acid (Depakene) across the placenta: half-life of the drug in mother and baby. J Pediatr 94, 832-835 (1979).

121. Fowler, D.W., Eadie, M.J., Dickinson, R.G. Transplacental transfer and biotransformation studies of valproic acid and its glucuronide(s) in the perfused human placenta. J Pharmacol Exp Ther 249, 318-323 (1989).

122. Nau, H., Rating, D., Koch, S., Hauser, I., Helge, H. Valproic acid and its metabolites: placental transfer, neonatal pharmacokinetics, transfer via mother's milk and clinical status in neonates of epileptic mothers. J Pharmacol Exp Ther 219, 768-777 (1981).

123. Froescher, W., Gugler, R., Niesen, M., Hoffmann, F. Protein binding of valproic acid in maternal and umbilical cord serum. Epilepsia 25, 244-249 (1984).

124. Myllynen, P.K., ki, P.K., kangas, K.H. Transplacental passage of lamotrigine in a human placental perfusion system in vitro and in maternal and cord blood in vivo. Eur J Clin Pharmacol 58, 677-682 (2003).

125. Kacirova, I., Grundmann, M., Brozmanova, H. Serum levels of lamotrigine during delivery in mothers and their infants. Epilepsy Res. 91, 161-165 (2010).

Page 207: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

191

126. Ohman, I., Vitols, S., Tomson, T. Lamotrigine in pregnancy: pharmacokinetics during delivery, in the neonate, and during lactation. Epilepsia 41, 709-713 (2000).

127. Tomson, T., Ohman, I., Vitols, S. Lamotrigine in pregnancy and lactation: a case report. Epilepsia 38, 1039-1041 (1997).

128. de Haan, G.J. et al. Gestation-induced changes in lamotrigine pharmacokinetics: a monotherapy study. Neurology 63, 571-573 (2004).

129. Fotopoulou, C. et al. Prospectively assessed changes in lamotrigine-concentration in women with epilepsy during pregnancy, lactation and the neonatal period. Epilepsy Res. 85, 60-64 (2009).

130. Schenker, S., Yang, Y., Mattiuz, E., Tatum, D., Lee, M. Olanzapine transfer by human placenta. Clin Exp Pharmacol Physiol 26, 691-697 (1999).

131. Aichhorn, W., Yazdi, K., Kralovec, K., Steiner, H., Whitworth, S., Stuppaeck, C. Olanzapine plasma concentration in a newborn. J Psychopharmacol 22, 923-924 (2008).

132. Newport, D.J. et al. Atypical antipsychotic administration during late pregnancy: placental passage and obstetrical outcomes. Am J Psychiatry 164, 1214-1220 (2007).

133. Kirchheiner, J., Berghofer, A., Bolk-Weischedel, D. Healthy outcome under olanzapine treatment in a pregnant woman. Pharmacopsychiatry 33, 78-80 (2000).

134. Rahi, M. et al. Placental transfer of quetiapine in relation to P-glycoprotein activity. J Psychopharmacol 21, 751-756 (2007).

135. Sudhakaran, S. et al. Differential bidirectional transfer of indinavir in the isolated perfused human placenta. Antimicrob Agents Chemother 49, 1023-1028 (2005).

136. Mirochnick, M. et al. Concentrations of protease inhibitors in cord blood after in utero exposure. Pediatr Infect Dis J 21, 835-838 (2002).

137. Sudhakaran, S., Rayner, C.R., Li, J., Kong, D.C.M., Gude, N.M., Nation, R.L. Inhibition of placental P-glycoprotein: Impact on indinavir transfer to the foetus. Br J Clin Pharmacol 65, 667-673 (2008).

138. Chappuy, H. et al. Maternal-fetal transfer and amniotic fluid accumulation of protease inhibitors in pregnant women who are infected with human immunodeficiency virus. Am J Obstet Gynecol 191, 558-562 (2004).

139. Gavard, L. et al. Placental transfer of lopinavir/ritonavir in the ex vivo human cotyledon perfusion model. Am J Obstet Gynecol 195, 296-301 (2006).

140. Marzolini, C. et al. Transplacental passage of protease inhibitors at delivery. AIDS 16, 889-893 (2002).

Page 208: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

192

141. Gingelmaier, A. et al. Placental transfer and pharmacokinetics of lopinavir and other protease inhibitors in combination with nevirapine at delivery. AIDS 20, 1737-1743 (2006).

142. Ivanovic, J. et al. Transplacental transfer of antiretroviral drugs and newborn birth weight in HIV-infected pregnant women. Curr HIV Res 7, 620-625 (2009).

143. Mirochnick, M. et al. Lopinavir exposure with an increased dose during pregnancy. J Acquir Immune Defic Syndr 49, 485-491 (2008).

144. Yeh, R.F. et al. Genital tract, cord blood, and amniotic fluid exposures of seven antiretroviral drugs during and after pregnancy in human immunodeficiency virus type 1-infected women. Antimicrob Agents Chemother 53, 2367-2374 (2009).

145. Stek, A.M. et al. Reduced lopinavir exposure during pregnancy. AIDS 20, 1931-1939 (2006).

146. Gavard, L. et al. Contribution and limit of the model of perfused cotyledon to the study of placental transfer of drugs. Example of a protease inhibitor of HIV: nelfinavir. Eur J Obstet Gynecol Reprod Biol 147, 157-160 (2009).

147. Hirt, D. et al. Pharmacokinetic modelling of the placental transfer of nelfinavir and its M8 metabolite: a population study using 75 maternal-cord plasma samples. Br J Clin Pharmacol 64, 634-644 (2007).

148. Casey, B.M., Bawdon, R.E. Placental transfer of ritonavir with zidovudine in the ex vivo placental perfusion model. Am J Obstet Gynecol 179, 758-761 (1998).

149. Forestier, F., de, R.,P., Peytavin, G., Dohin, E., Farinotti, R., Mandelbrot, L. Maternal-fetal transfer of saquinavir studied in the ex vivo placental perfusion model. Am J Obstet Gynecol 185, 178-181 (2001).

150. Henderson, G.I., Hu, Z.Q., Johnson, R.F., Perez, A.B., Yang, Y., Schenker, S. Acyclovir transport by the human placenta. J Lab Clin Med 120, 885-892 (1992).

151. Kimberlin, D.F. et al. Pharmacokinetics of oral valacyclovir and acyclovir in late pregnancy. Am J Obstet Gynecol 179, 846-851 (1998).

152. Gilstrap, L.C., Bawdon, R.E., Roberts, S.W., Sobhi, S. The transfer of the nucleoside analog ganciclovir across the perfused human placenta. Am J Obstet Gynecol 170, 967-972 (1994).

153. Frenkel, L.M. et al. Pharmacokinetics of acyclovir in the term human pregnancy and neonate. Am J Obstet Gynecol 164, 569-576 (1991).

154. Haddad, J., Langer, B., Astruc, D., Messer, J., Lokiec, F. Oral acyclovir and recurrent genital herpes during late pregnancy. Obstet Gynecol 82, 102-104 (1993).

Page 209: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

193

155. Dancis, J., Lee, J.D., Mendoza, S., Liebes, L. Transfer and metabolism of dideoxyinosine by the perfused human placenta. J Acquir Immune Defic Syndr 6, 2-6 (1993).

156. Chappuy, H. et al. Maternal-fetal transfer and amniotic fluid accumulation of nucleoside analogue reverse transcriptase inhibitors in human immunodeficiency virus-infected pregnant women. Antimicrob Agents Chemother 48, 4332-4336 (2004).

157. Pons, J.C. et al. Fetoplacental passage of 2',3'-dideoxyinosine. Lancet 337, 732 (1991).

158. Henderson, G.I., Perez, A.B., Yang, Y., Hamby, R.L., Schenken, R.S., Schenker, S. Transfer of dideoxyinosine across the human isolated placenta. Br J Clin Pharmacol 38, 237-242 (1994).

159. Bloom, S.L., Dias, K.M., Bawdon, R.E., Gilstrap, I.I.I.,L.C. The maternal-fetal transfer of lamivudine in the ex vivo human placenta. Am J Obstet Gynecol 176, 291-293 (1997).

160. Moodley, D. et al. Pharmacokinetics of zidovudine and lamivudine in neonates following coadministration of oral doses every 12 hours. J Clin Pharmacol 41, 732-741 (2001).

161. Mandelbrot, L., Peytavin, G., Firtion, G., Farinotti, R. Maternal-fetal transfer and amniotic fluid accumulation of lamivudine in human immunodeficiency virus-infected pregnant women. Am J Obstet Gynecol 184, 153-158 (2001).

162. Olivero, O.A., Parikka, R., Poirier, M.C., Vahakangas, K. 3'-azido-3'-deoxythymidine (AZT) transplacental perfusion kinetics and DNA incorporation in normal human placentas perfused with AZT. Mutat Res 428, 41-47 (1999).

163. Watts, D.H. et al. Pharmacokinetic disposition of zidovudine during pregnancy. J Infect Dis 163, 226-232 (1991).

164. Boal, J.H., Plessinger, M.A., van, d.,Reydt C., Miller, R.K. Pharmacokinetic and toxicity studies of AZT (zidovudine) following perfusion of human term placenta for 14 hours. Toxicol Appl Pharmacol 143, 13-21 (1997).

165. Schenker, S., Johnson, R.F., King, T.S., Schenken, R.S., Henderson, G.I. Azidothymidine (Zidovudine) transport by the human placenta. Am J Med Sci 299, 16-20 (1990).

166. Liebes, L., Mendoza, S., Wilson, D., Dancis, J. Transfer of zidovudine (AZT) by human placenta. J Infect Dis 161, 203-207 (1990).

167. Gillet, J.Y., Garraffo, R., Abrar, D., Bongain, A., Lapalus, P., Dellamonica, P. Fetoplacental passage of zidovudine.[letter]. Lancet 1989;2(8657):269-270.

168. Collier, A.C., Keelan, J.A., Van, Z.,P.E., Paxton, J.W., Mitchell, M.D., Tingle, M.D. Human placental glucuronidation and transport of 3'azido-3'- deoxythymidine and uridine diphosphate glucuronic acid. Drug Metab Dispos 32, 813-820 (2004).

Page 210: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

194

169. Pons, J.C., Taburet, A.M., Singlas, E., Delfraissy, J.F., Papiernik, E. Placental passage of azathiothymidine (AZT) during the second trimester of pregnancy: study by direct fetal blood sampling under ultrasound. Eur J Obstet Gynecol Reprod Biol 40, 229-231 (1991).

170. O'Sullivan, M.J. et al. The pharmacokinetics and safety of zidovudine in the third trimester of pregnancy for women infected with human immunodeficiency virus and their infants: phase I acquired immunodeficiency syndrome clinical trials group study (protocol 082). Zidovudine Collaborative Working Group. Am J Obstet Gynecol 168, 1510-1516 (1993).

171. Ceccaldi, P.-F., Ferreira, C., Gavard, L., Gil, S., Peytavin, G., Mandelbrot, L. Placental transfer of enfuvirtide in the ex vivo human placenta perfusion model. Am J Obstet Gynecol 198, 433.e1-433.e2 (2008).

172. Brennan-Benson, P. et al. Enfurvitide prevents vertical transmission of multidrug-resistant HIV-1 in pregnancy but does not cross the placenta. AIDS 20, 297-299 (2006).

173. Omarini, D., Barzago, M.M., Aramayona, J., Bortolotti, A., Lucchini, G., Bonati, M. Theophylline transfer across human placental cotyledon during in vitro dual perfusion. J Med 23, 101-116 (1992).

174. Labovitz, E., Spector, S. Placental theophylline transfer in pregnant asthmatics. JAMA 247, 786-788 (1982).

175. Omarini, D. et al. Placental transfer of theophylline in an in vitro closed perfusion system of human placenta isolated lobule. Eur J Drug Metab Pharmacokinet 18, 369-374 (1993).

176. Arwood, L.L., Dasta, J.F., Friedman, C. Placental transfer of theophylline: two case reports. Pediatrics 63, 844-846 (1979).

177. Ron, M., Hochner-Celnikier, D., Menczel, J., Palti, Z., Kidroni, G. Maternal-fetal transfer of aminophylline. Acta Obstet Gynecol Scand 63, 217-218 (1984).

178. Romero, R., Kadar, N., Gonzales Govea, F., Hobbins, J.C. Pharmacokinetics of intravenous theophylline in pregnant patients at term. Am J Perinatol 1, 31-35 (1983).

179. Nandakumaran, M., Gardey, C.L., Challier, J.C. Transfer of salbutamol in the human placenta in vitro. Dev Pharmacol Ther 3, 88-98 (1981).

180. Boulton, D.W., Fawcett, J.P., Fiddes, T.M. Transplacental distribution of salbutamol enantiomers at Caesarian section. Br J Clin Pharmacol 44, 587-590 (1997).

181. McGowan, W.A. Safety of cimetidine in obstetric patients. J R Soc Med 72, 902-907 (1979).

182. Ching, M.S., Mihaly, G.W., Morgan, D.J., Date, N.M., Hardy, K.J., Smallwood, R.A. Low clearance of cimetidine across the human placenta. J Pharmacol Exp Ther 241, 1006-1009 (1987).

Page 211: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

195

183. Qvist, N., Storm, K., Holmskov, A. Cimetidine as pre-anesthetic agent for cesarean section: perinatal effects on the infant, the placental transfer of cimetidine and its elimination in the infants. J Perinat Med 13, 179-183 (1985).

184. Dicke, J.M., Johnson, R.F., Henderson, G.I., Kuehl, T.J., Schenker, S. A comparative evaluation of the transport of H2-receptor antagonists by the human and baboon placenta. Am J Med Sci 295, 198-206 (1988).

185. Howe, J.P., McGowan, W.A., Moore, J., McCaughey, W., Dundee, J.W. The placental transfer of cimetidine. Anaesthesia 36, 371-375 (1981).

186. McAuley, D.M., Moore, J., McCaughey, W., Donnelly, B.D., Dundee, J.W. Ranitidine as an antacid before elective Caesarean section. Anaesthesia 38, 108-114 (1983).

187. Armentano, G., Bracco, P.L., Di Silverio, C. Ranitidine in the treatment of reflux oesophagitis in pregnancy. Clin Exp Obstet Gynecol 16, 130-133 (1989).

188. Peytavin, G., Leng, J.J., Forestier, F., Saux, M., Hohlfeld, P., Farinotti, R. Placental transfer of pyrimethamine studied in an ex vivo placental perfusion model. Biol Neonate 78, 83-85 (2000).

189. Trenque, T. et al. Human maternofoetal distribution of pyrimethamine-sulphadoxine. Br J Clin Pharmacol 45, 179-180 (1998).

190. Dorangeon, P.H. et al. Transplacental passage of the pyrimethamine-sulfadoxine combination in the prenatal treatment of congenital toxoplasmosis. Presse Medicale 19, 2036 (1990).

191. Heikkinen, T., Laine, K., Neuvonen, P.J., Ekblad, U. The transplacental transfer of the macrolide antibiotics erythromycin, roxithromycin and azithromycin. BJOG 107, 770-775 (2000).

192. Kiefer, L., Rubin, A., McCoy, J.B., Foltz, E.L. The placental transfer of erythomycin. Am J Obstet Gynecol 69, 174-177 (1955).

193. Heilman, F.R., Herrell, W.E., Wellman, W.E., Geraci, J.E. Some laboratory and clinical observations on a new antibiotic, erythromycin (ilotycin). Proc Staff Meet Mayo Clin 27, 285-304 (1952).

194. Ramsey, P.S., Vaules, M.B., Vasdev, G.M., Andrews, W.W., Ramin, K.D. Maternal and transplacental pharmacokinetics of azithromycin. Am J Obstet Gynecol 188, 714-718 (2003).

195. Polachek, H. et al. Transfer of ciprofloxacin, ofloxacin and levofloxacin across the perfused human placenta in vitro. Eur J Obstet Gynecol Reprod Biol 122, 61-65 (2005).

196. Matsuda, S., Kashiwagura, T., Kokuho, K. Experimental and clinical studies of DL-8280 in the field of obstetrics and gynecology. Chemotherapy 32, 900-907 (1984).

Page 212: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

196

197. Yamamoto, T., Yasuda, J., Kanao, M., Okada, H. Fundamental and clinical studies on DL-8280 in the field of obstetrics and gynecology. Chemotherapy 32, 944-949 (1984).

198. Fortunato, S.J., Bawdon, R.E., Swan, K.F., Bryant, E.C., Sobhi, S. Transfer of Timentin (ticarcillin and clavulanic acid) across the in vitro perfused human placenta: comparison with other agents. Am J Obstet Gynecol 167, 1595-1599 (1992).

199. Maberry, M.C., Trimmer, K.J., Bawdon, R.E., Sobhi, S., Dax, J.B., Gilstrap, L.C.r. Antibiotic concentration in maternal blood, cord blood and placental tissue in women with chorioamnionitis. Gynecol Obstet Invest 33, 185-186 (1992).

200. Fortunato, S.J., Bawdon, R.E., Maberry, M.C., Swan, K.F. Transfer of ceftizoxime surpasses that of cefoperazone by the isolated human placenta perfused in vitro. Obstet Gynecol 75, 830-833 (1990).

201. Fortunato, S.J., Bawdon, R.E., Welt, S.I., Swan, K.F. Steady-state cord and amniotic fluid ceftizoxime levels continuously surpass maternal levels. Am J Obstet Gynecol 159, 570-573 (1988).

202. Fortunato, S.J., Bawdon, R.E., Baum, M. Placental transfer of cefoperazone and sulbactam in the isolated in vitro perfused human placenta. Am J Obstet Gynecol 159, 1002-1006 (1988).

203. Boskovic, R., Feig, D.S., Derewlany, L., Knie, B., Portnoi, G., Koren, G. Transfer of insulin lispro across the human placenta: in vitro perfusion studies. Diabetes Care 26, 1390-1394 (2003).

204. Jovanovic, L. et al. Metabolic and immunologic effects of insulin lispro in gestational diabetes. Diabetes Care 22, 1422-1427 (1999).

205. Holcberg, G. et al. Transfer of insulin lispro across the human placenta. Eur J Obstet Gynecol Reprod Biol 115, 117-118 (2004).

206. Nanovskaya, T.N., Nekhayeva, I.A., Patrikeeva, S.L., Hankins, G.D.V., Ahmed, M.S. Transfer of metformin across the dually perfused human placental lobule. Am J Obstet Gynecol 195, 1081-1085 (2006).

207. Hague, W., Davoren, P., McIntyre, D., Norris, R., Xiaonian, X. Metformin crosses the placenta: A modulator for fetal insulin resistance? BMJ 327, 880-881 (2003).

208. Vanky, E., Zahlsen, K., Spigset, O., Carlsen, S.M. Placental passage of metformin in women with polycystic ovary syndrome. Fertil Steril 83, 1575-1578 (2005).

209. Kovo, M., Haroutiunian, S., Feldman, N., Hoffman, A., Glezerman, M. Determination of metformin transfer across the human placenta using a dually perfused ex vivo placental cotyledon model. Eur J Obstet Gynecol Reprod Biol 136, 29-33 (2008).

Page 213: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

197

210. Kovo, M., Kogman, N., Ovadia, O., Nakash, I., Golan, A., Hoffman, A. Carrier-mediated transport of metformin across the human placenta determined by using the ex vivo perfusion of the placental cotyledon model. Prenat Diagn 28, 544-548 (2008).

211. Elliott, B.D., Langer, O., Schenker, S., Johnson, R.F. Insignificant transfer of glyburide occurs across the human placenta. Am J Obstet Gynecol 165, 807-812 (1991).

212. Langer, O., Conway, D.L., Berkus, M.D., Xenakis, E.M., Gonzales, O. A comparison of glyburide and insulin in women with gestational diabetes mellitus. N Engl J Med 343, 1134-1138 (2000).

213. Hebert, M.F. et al. Are we optimizing gestational diabetes treatment with glyburide? The pharmacologic basis for better clinical practice. Clin Pharmacol Ther 85, 607-614 (2009).

214. Nanovskaya, T.N., Nekhayeva, I., Hankins, G.D.V., Ahmed, M.S. Effect of human serum albumin on transplacental transfer of glyburide. Biochem Pharmacol 72, 632-639 (2006).

215. Nanovskaya, T.N. et al. Effect of albumin on transplacental transfer and distribution of rosiglitazone and glyburide. J Matern Fetal Neonatal Med 21, 197-207 (2008).

216. Derewlany, L.O., Leeder, J.S., Kumar, R., Radde, I.C., Knie, B., Koren, G. The transport of digoxin across the perfused human placental lobule. J Pharmacol Exp Ther 256, 1107-1111 (1991).

217. Kerenyi, T.D. et al. Transplancental cardioversion of intrauterine supraventricular tachycardia with digitalis. Lancet 2, 393-394 (1980).

218. Holcberg, G. et al. Lack of interaction of digoxin and P-glycoprotein inhibitors, quinidine and verapamil in human placenta in vitro. Eur J Obstet Gynecol Reprod Biol 109, 133-137 (2003).

219. Nagashima, M., Asai, T., Suzuki, C., Matsushima, M., Ogawa, A. Intrauterine supraventricular tachyarrhythmias and transplacental digitalisation. Arch Dis Child 61, 996-1000 (1986).

220. Tsadkin, M. et al. Albumin-dependent digoxin transfer in isolated perfused human placenta. Int J Clin Pharmacol Ther 39, 158-161 (2001).

221. Rogers, M.C., Willerson, J.T., Goldblatt, A., Smith, T.W. Serum digoxin concentrations in the human fetus, neonate and infant. N Engl J Med 287, 1010-1013 (1972).

222. Younis, J.S., Granat, M. Insufficient transplacental digoxin transfer in severe hydrops fetalis. Am J Obstet Gynecol 157, 1268-1269 (1987).

223. Saarikoski, S. Placental transfer and fetal uptake of 3H-digoxin in humans. Br J Obstet Gynaecol 83, 879-884 (1976).

224. Schmolling, J. Digoxin transfer across the isolated placenta is influenced by maternal and fetal albumin concentrations. Reprod Fertil Dev 8, 969-974 (1996).

Page 214: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

198

225. Mimura, S., Suzuki, C., Yamazaki, T. Transplacental passage of digoxin in the case of nonimmune hydrops fetalis. Clin Cardiol 10, 63-65 (1987).

226. Miller, R.K., Jessee, L., Barrish, A., Gilbert, J., Manson, J.M. Pharmacokinetic studies of enalaprilat in the in vitro perfused human placental lobule system. Teratology 58, 76-81 (1998).

227. Ala-Kokko, T.I., Pienimaki, P., Lampela, E., Hollmen, A.I., Pelkonen, O., Vahakangas, K. Transfer of clonidine and dexmedetomidine across the isolated perfused human placenta. Acta Anaesthesiol Scand 41, 313-319 (1997).

228. Hartikainen-Sorri, A.L., Heikkinen, J.E., Koivisto, M. Pharmacokinetics of clonidine during pregnancy and nursing. Obstet Gynecol 69, 598-600 (1987).

229. Boutroy, M.J., Gisonna, C.R., Legagneur, M. Clonidine: placental transfer and neonatal adaption. Early Hum Dev 17, 275-286 (1988).

230. Buchanan, M.L. et al. Clonidine pharmacokinetics in pregnancy. Drug Metab Dispos 37, 702-705 (2009).

231. Cigarini, I. et al. Epidural clonidine combined with bupivacaine for analgesia in labor. Effects on mother and neonate. Reg Anesth 20, 113-120 (1995).

232. Lagrange, F. et al. Absence of placental transfer of pentasaccharide (fondaparinux, arixtra) in the dually perfused human cotyledon in vitro. Thromb. Haemost. 87, 831-835 (2002).

233. Dempfle, C.E. Minor transplacental passage of fondaparinux in vivo.[letter]. N Engl J Med 2004;350(18):1914-1915.

234. Mazzolai, L., Hohlfeld, P., Spertini, F., Hayoz, D., Schapira, M., Duchosal, M.A. Fondaparinux is a safe alternative in case of heparin intolerance during pregnancy. Blood 108, 1569-1570 (2006).

235. Harenberg, J. Treatment of a woman with lupus and thromboembolism and cutaneous intolerance to heparins using fondaparinux during pregnancy. Thromb Res 119, 385-388 (2007).

236. Bajoria, R., Contractor, S.F. Transfer of heparin across the human perfused placental lobule. J Pharm Pharmacol 44, 952-959 (1992).

237. Flessa, H.C., Kapstrom, A.B., Glueck, H.I., Will, J.J. Placental transport of heparin. Am J Obstet Gynecol 93, 570-573 (1965).

238. Dimitrakakis, C., Papageorgiou, P., Papageorgiou, I., Antzaklis, A., Sakarelou, N., Michalas, S. Absence of transplacental passage of the low molecular weight heparin enoxaparin. Haemostasis 30, 243-248 (2000).

Page 215: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

199

239. Forestier, F., Daffos, F., Capella-Pavlovsky, M. Low molecular weight heparin (PK 10169) does not cross the placenta during the second trimester of pregnancy study by direct fetal blood sampling under ultrasound. Thromb Res 34, 557-560 (1984).

240. Forestier, F., Daffos, F., Rainaut, M., Toulemonde, F. Low molecular weight heparin (CY 216) does not cross the placenta during the third trimester of pregnancy.[letter]. Thromb Haemost 1987;57(2):234.

241. Melissari, E. et al. Use of low molecular weight heparin in pregnancy. Thromb Haemost 68, 652-656 (1992).

242. Omri, A., Delaloye, J.F., Andersen, H., Bachmann, F. Low molecular weight heparin Novo (LHN-1) does not cross the placenta during the second trimester of pregnancy. Thromb. Haemost. 61, 55-56 (1989).

243. Bustard, M.A., Farley, A.E., Smith, G.N. The pharmacokinetics of glyceryl trinitrate with the use of the in vitro term human placental perfusion setup. Am J Obstet Gynecol 187, 187-190 (2002).

244. David, M., Walka, M.M., Schmid, B., Sinha, P., Veit, S., Lichtenegger, W. Nitroglycerin application during cesarean delivery: plasma levels, fetal/maternal ratio of nitroglycerin, and effects in newborns. Am J Obstet Gynecol 182, 955-961 (2000).

245. Magee, K.P., Bawdon, R.E. Ex vivo human placental transfer and the vasoactive properties of hydralazine. Am J Obstet Gynecol 182, 167-169 (2000).

246. Liedholm, H., Wahlin-Boll, E., Hanson, A., Ingemarsson, I., Melander, A. Transplacental passage and breast milk concentrations of hydralazine. Eur J Clin Pharmacol 21, 417-419 (1982).

247. Schneider, H., Proegler, M. Placental transfer of beta-adrenergic antagonists studied in an in vitro perfusion system of human placental tissue. Am J Obstet Gynecol 159, 42-47 (1988).

248. Melander, A., Niklasson, B., Ingemarsson, I., Liedholm, H., Schersten, B., Sjoberg, N.O. Transplacental passage of atenolol in man. Eur J Clin Pharmacol 14, 93-94 (1978).

249. Kofahl, B., Henke, D., Hettenbach, A., Mutschler, E. Studies on placental transfer of celiprolol. Eur J Clin Pharmacol 44, 381-382 (1993).

250. Poranen, A.-K., Nurmi, H., Malminiemi, K., Ekblad, U. Vasoactive effects and placental transfer of nifedipine, celiprolol, and magnesium sulfate in the placenta perfused in vitro. Hypertension in Pregnancy 17, 93-102 (1998).

251. Michael, C.A. Use of labetalol in the treatment of severe hypertension during pregnancy. Br J Clin Pharmacol 8, 211S-215S (1979).

252. Nandakumaran, M., Angelo-Khattar, M., Ibrahim, M.E., Hathout, H. Transfer of labetalol in the perfused human placenta: An in vitro study. Med Princ Pract 1, 81-85 (1989).

Page 216: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

200

253. Erkkola, R., Lammintausta, R., Liukko, P., Anttila, M. Transfer of propranolol and sotalol across the human placenta. Their effect on maternal and fetal plasma renin activity. Acta Obstet Gynecol Scand 61, 31-34 (1982).

254. Pirhonen, J.P., Erkkola, R.U., Ekblad, U.U., Nyman, L. Single dose of nifedipine in normotensive pregnancy: nifedipine concentrations, hemodynamic responses, and uterine and fetal flow velocity waveforms. Obstet Gynecol 76, 807-811 (1990).

255. Nandakumaran, M., Eldeen, A.S. Transfer of cyclosporine in the perfused human placenta. Dev Pharmacol Ther 15, 101-105 (1990).

256. Venkataramanan, R., Koneru, B., Wang, C.C., Burckart, G.J., Caritis, S.N., Starzl, T.E. Cyclosporine and its metabolites in mother and baby. Transplantation 46, 468-469 (1988).

257. Bourget, P., Fernandez, H., Bismuth, H., Papiernik, E. Transplacental passage of cyclosporine after liver transplantation. Transplantation 49, 663 (1990).

258. Flechner, S.M., Katz, A.R., Rogers, A.J., Van Buren, C., Kahan, B.D. The presence of cyclosporine in body tissues and fluids during pregnancy. Am J Kidney Dis 5, 60-63 (1985).

259. Mortimer, R.H., Cannell, G.R., Addison, R.S., Johnson, L.P., Roberts, M.S., Bernus, I. Methimazole and propylthiouracil equally cross the perfused human term placental lobule. J Clin Endocrinol Metab 82, 3099-3102 (1997).

260. Marchant, B., Brownlie, B.E., Hart, D.M., Horton, P.W., Alexander, W.D. The placental transfer of propylthiouracil, methimazole and carbimazole. J Clin Endocrinol Metab 45, 1187-1193 (1977).

261. Gardner, D.F., Cruikshank, D.P., Hays, P.M., Cooper, D.S. Pharmacology of propylthiouracil (PTU) in pregnant hyperthyroid women: correlation of maternal PTU concentrations with cord serum thyroid function tests. J Clin Endocrinol Metab 62, 217-220 (1986).

262. Nandakumaran, M., Gardey, C., Rey, E. Transfer of ritodrine and norepinephrine in human placenta: In vitro study. Dev Pharmacol Ther 4, 71-80 (1982).

263. Gandar, R., de Zoeten, L.W., van der Schoot, J.B. Serum level of ritodrine in man. Eur J Clin Pharmacol 17, 117-122 (1980).

264. Urbach, J., Mor, L., Fuchs, S., Brandes, J.M. Transplacental transfer of ritodrine and its effect on placental glucose and oxygen consumption in an in vitro human placental cotyledon perfusion. Gynecol Obstet Invest 32, 10-14 (1991).

265. Gross, T.L., Kuhnert, B.R., Kuhnert, P.M., Rosen, M.G., Kazzi, N.J. Maternal and fetal plasma concentrations of ritodrine. Obstet Gynecol 65, 793-797 (1985).

266. van Lierde, M., Thomas, K. Ritodrine concentrations in maternal and fetal serum and amniotic fluid. J Perinat Med 10, 119-124 (1982).

Page 217: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

201

267. Fujimoto, S., Tanaka, T., Akahane, M. Levels of ritodrine hydrochloride in fetal blood and amniotic fluid following long-term continuous administration in late pregnancy. Eur J Obstet Gynecol Reprod Biol 38, 15-18 (1991).

268. Fujimoto, S., Akahane, M., Sakai, A. Concentrations of ritodrine hydrochloride in maternal and fetal serum and amniotic fluid following intravenous administration in late pregnancy. Eur J Obstet Gynecol Reprod Biol 23, 145-152 (1986).

269. Kuhnert, B.R., Gross, T.L., Kuhnert, P.M., Erhard, P., Brashar, W.T. Ritodrine pharmacokinetics. Clin Pharmacol Ther 40, 656-664 (1986).

Page 218: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

202

Appendix C: Placental P-glycoprotein and breast cancer resistance protein: influence of

polymorphisms on fetal drug exposure and physiology

Janine R Hutsona,b, Gideon Korena,b, Stephen G Matthewsc

aInstitute of Medical Sciences, University of Toronto, 1 King’s College Circle, Toronto, Ontario M5S 1A8, Canada

bDivision of Clinical Pharmacology and Toxicology, Hospital for Sick Children, 555 University Ave, Toronto, Ontario M5G 1X8, Canada

cDepartments of Physiology, Obstetrics & Gynaecology and Medicine, University of Toronto, 1 King’s College Circle, Toronto, Ontario M5S 1A8, Canada

This work has been published and reproduced with permissions:

Hutson JR, Koren G, Matthews SG. Placental P-glycoprotein and breast cancer resistance protein: influence of polymorphisms on fetal drug exposure and physiology. Placenta 2010;31(5):351-357.

Page 219: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

203

Abstract

Recent studies have illustrated the importance of placental drug transport proteins, such

as P-glycoprotein (Pgp) and breast cancer resistance protein (BCRP) in limiting fetal exposure to

drugs and toxins. Moreover, increasing evidence supports a role for Pgp and BCRP in the

normal development and physiological function of the placenta. Several single nucleotide

polymorphisms (SNPs) in the genes encoding Pgp and BCRP have been described and are

associated with altered protein expression, transporter activity, and clinical outcome in studies

focusing on tissues other than the placenta. This review aims to summarize current research

regarding the association between these polymorphisms and expression and function in the

placenta. The influence of these genotypes on fetal drug exposure and altered placental

physiology or development is also presented. To date, evidence suggests that SNPs in both

ABCB1 and ABCG1 can alter expression of their respective protein; however, the functional

significance of these polymorphisms is less clear. An understanding of this genotype-phenotype

relationship will allow for prediction of susceptible or favorable genotypes in order to

personalize medication choices to minimize fetal exposure to teratogens, or to maximize

pharmacological therapy to the fetus.

Page 220: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

204

Introduction

Numerous factors are known to mediate transplacental transfer of substances, including

physiochemical and pharmacokinetic properties. Furthermore, transport proteins present on both

the microvillous brush-border and basal membranes (maternal blood-facing and fetal endothelial

cell-facing, respectively) of the syncytiotrophoblast cells have recently been identified as being

important for facilitated and active transport. Specifically, several members of the ATP-binding

cassette (ABC) drug efflux proteins are expressed, including P-glycoprotein (Pgp), breast cancer-

resistance protein (BCRP), and multidrug resistance-associated proteins (MRPs)(Table 1). Over

the last decade, the importance of Pgp and BCRP in mediating drug transfer across the placenta

and in the normal development and function of the placenta has been investigated.

Numerous polymorphisms have been reported in the genes encoding Pgp and BCRP

(ABCB1 and ABCG2, respectively). An understanding of these polymorphisms will allow

prediction of when normal placental physiology could be altered and lead to the development of

appropriate preventative therapies. Furthermore, knowledge of susceptible genotypes to fetal

drug exposure will influence medication choices during pregnancy, allowing minimized drug

exposure to potential teratogens or maximized drug exposure when pharmacological therapy to

the fetus is desired. This review summarizes current research regarding the association between

ABCB1 and ABCG2 polymorphisms and expression and function in the placenta and how this

knowledge will translate into personalized medicine.

Page 221: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

205

Table 1. Selected substrates for Pgp and BCRP (Marzolini et al., 2004; Mao & Unadkat, 2005). Pgp Substrates BCRP Substrates Cimetidine Bisantrene Cyclosporin Diflomotecan Dexamethasone Cimetidine Digoxin Doxorubicin Doxorubicin Etoposide Erythromycin Flavonoids Indinavir Glyburide Levofloxacin Imatinib Morphine Lamivudine Phenobarbital Methotrexate Phenytoin Nitrofurantoin Quinidine Porphyrins Rifampin Sulfated estrogens Ritonavir Topotecan Saquinavir Zidovudine Talinolol Topotecan Verapamil Vinblastine

P-glycoprotein

Expression and function in the placenta

The expression of Pgp in the placenta throughout gestation has been localized to the

microvillous border of the syncytiotrophoblast[1-3]. Expression of Pgp is relatively high in the

placenta as ABCB1 mRNA levels are comparable to the intestine and liver[4]. However,

placental expression decreases throughout gestation[3;5]. This decrease in late gestation may be

important in allowing increased fetal exposure to endogenous factors in the maternal circulation;

including cortisol[3]. The decrease in Pgp may also increase susceptibility of the fetus to toxins

present in the maternal circulation. Many studies have shown close correlation between P-gp

expression and activity[6;7]. However, when P-gp levels are very low at term, this relationship is

not quite so clear[8].

Mechanisms regulating Pgp expression in the placenta are not fully understood as

regulation of Pgp appears tissue-specific[7;9]. In the trophoblast, ABCB1 transcription is

Page 222: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

206

regulated by the Ap-2 and Sp families of transcription factors[10]. Hormonal regulation of Pgp

by progesterone has been proposed, however, administration of progesterone to pregnant mice

did not alter ABCB1 expression[7]. Placental expression of Pgp correlates with that of human

chorionic gonadotropin-β (hCG-β), suggesting that hCG-β may regulate Pgp expression or they

are both under the same control mechanisms[11]. Evidence suggests that cytokines may also be

important regulators of placental Pgp; in vitro incubation of primary term trophoblasts with

tumor necrosis factor-α (TNF-α) and interleukin-1β (IL-1β) decreased ABCB1 mRNA and

protein expression[12]. Furthermore, in rats, endotoxin-induced inflammation decreased

expression of placental Pgp[13].

Placental Pgp confers fetal protection to toxins in the maternal circulation. Deficiency of

mdr1a (knockout mice) or pharmacological inhibition of Pgp increased fetal exposure by up to

16-fold to digoxin, saquinavir, and paclitaxel after administration to pregnant dams[14]. Studies

using human placental tissue have also demonstrated altered drug transfer after inhibition of Pgp.

Using the dually perfused placenta model, increased drug transfer was seen for the protease

inhibitors saquinavir[15] and indinavir[16], and for methadone[17] after inhibition of Pgp.

Increased fetal exposure to protease inhibitors is warranted to prevent vertical transmission of

HIV, thus co-administration with other Pgp substrates or inhibitors will lead to the desired higher

fetal levels.

Polymorphisms in Pgp and effect on placental Pgp expression

The ABCB1 gene is over 100 kb in length and consists of a core promoter region and 28

exons located on the long arm of chromosome 7 at q21.1[18]. There is currently no evidence for

a spontaneous mutation in humans that would result in complete Pgp deficiency and may suggest

that mutations in ABCB1 may be embryolethal in humans[19]. There are, on the other hand,

over 50 SNPs and 3 insertion/deletion polymorphisms described[20]. Of these, specific

polymorphisms in an exon 26 wobble position (C3435T), an exon 21 coding region (G2677T/A),

an exon 12 coding region (C1236T), and the promoter region (T-129C) have been the focus of

studies in placental tissue. The C3435T SNP is synonymous, however, the G2677T/A SNP

results in an amino acid conversion (Ala to Thr/Ser) and the C1236T may influence RNA

stability[21;22].

Page 223: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

207

Five studies have examined the relationship between placental Pgp expression and

ABCB1 polymorphisms and are summarized in Table 2. The first examined placentae from

Japanese women and detected nine SNPs. Compared to the -129TT genotype, -129CT exhibited

significantly reduced Pgp protein expression[21]. Interestingly, there were no subjects that were

homozygous with the mutation (-129CC). In a subsequent study, no placentae from Caucasian

subjects carried the T-129C SNP[23]. The maternal haplotype 2677T/3435T exhibited lower

mRNA and Pgp protein levels compared to other pooled haplotypes. Mothers that were

homozygous for both 2677T and 3435T (TT/TT) exhibited placental Pgp (protein) that was more

than 2-fold lower compared to GG/CC mothers. Interestingly, there was no significant

association between fetal genotype or haplotype and Pgp protein expression. When genotype of

both the mother and fetus was considered, there was a >30% reduction in Pgp protein expression

when both the mother and fetus were homozygous or heterozygous for the C3435T SNP (TT/tt

and CT/ct) compared to CC/cc. Limited sample size precluded full haplotype analysis. A

challenge with these correlation studies is that ABCB1 mRNA likely degrades rapidly after

delivery; rapid collection is critical. A more recent study, using brush border membrane

preparations, identified decreased protein expression in C1236T (11% decrease), C3435T (16%),

and G2677T/A (16%) variants compared to their homozygous wild-type control group[24]. A

fourth study contrasts to the previous three reports: a 29% higher Pgp protein expression for the

3435TT genotype was determined compared to 3435CC[25]. A major limitation of the above

studies is that most performed genotyping[21;24;25] and quantification of Pgp[23;25] using

whole placental tissue, which contains both maternal and fetal tissue. Since Pgp localized to the

syncytiotrophoblast will be of fetal origin, care must be taken to genotype only fetal tissue[23].

Furthermore, placental blood vessels of maternal origin have been shown to express Pgp, thus

fetal and maternal-derived Pgp should be separated in order to produce a clear genotype�

phenotype association[23].

Although most studies linking ABCB1 polymorphisms to protein expression have focused

on the coding region, Takane et al. focused on promoter variants[26]. Haplotypes that

simultaneously included T-1517aC, T-1017aC, and T-129C mutations in Japanese and Caucasian

subjects were associated with increased mRNA expression and altered protein-DNA binding,

independent of the C3435T mutation. These promoter polymorphisms also altered gene

Page 224: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

208

transcription. In Japanese and Caucasian populations, the aforementioned haplotypes are found

in low frequency and thus their clinical relevance remains to be determined[26].

In addition to mutations in the ABCB1 gene, epigenetic regulation may control

transcription. Despite a CpG-rich promoter region found in the human ABCB1 gene, potential

transcription factor binding sites were found to be hypomethylated in full term-placenta

tissue[26]. On the other hand, hypermethylation of the ABCB1 promoter has been shown to

decrease Pgp expression in different types of cancer cells[27-29]. The small study in term

placenta (n=7), failed to identify an association between methylation status and mRNA

expression[26]. Further studies utilizing a larger sample size, expanded promoter mapping and

other promoter genotypes are clearly required.

In each study that examined placental ABCB1 polymorphisms, there is large variation in

Pgp expression found for each genotype. Given this large interindividual variation in expression,

the sample sizes of the five studies may be too small to provide adequate statistical power.

Furthermore, the large variability may result from other environmental factors that could also

influence expression. For example, common medications in pregnancy may influence Pgp

expression (such as morphine[30]). Only the study by Rahi et al.[25] required mothers to be

healthy and not taking any medications throughout the pregnancy. Once environmental factors

are controlled for, full haplotype analysis may result in less variability between genotypes and

stronger associations and should be included in future studies. Another limitation from the

studies linking ABCB1 genotype to Pgp expression is that only full term placentae were

analyzed. Since placental Pgp expression is at its lowest at term[3;5], any association with

genotype may be harder to establish and require a much larger sample size to obtain significance.

The influence of ABCB1 SNPs and Pgp expression outside of the placenta has also led to

conflicting results. A meta-analysis of three studies (total n=261) found no effect of the C3435T

SNP on intestinal ABCB1 mRNA expression[31]. On the other hand, a large systematic study

not included in the meta-analysis found that the 3435TT genotype had a 2-fold reduction in

intestinal Pgp protein expression compared to 3435CC[32]. Similar findings were observed in

lymphocytes[33;34] and renal epithelial tissue[35].

Page 225: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

209

Table 2. Summary of the associations between A) ABCB1 genotype and placental Pgp expression B) ABCG2 genotype and placental BCRP expression. A) Study Population SNP Finding Placental Preparation Tanabe et al.(2001) Japanese (n=100) T-129C -129TC < -129TT protein from enriched trophoblasts G2677A/T NS C3435T NS Hitzl et al. (2004) Caucasian (n=73) T-129C N/A protein/mRNA from whole placental tissue G2677A/T NS C3435T NS Hemauer et al. (2009) Caucasian, C1236T 1236T < 1236CC protein from brush border membranes African-American G2677A/T 2677A/T < 2677GG Hispanic (n=199) C3435T 3435T < 3435CC Rahi et al. (2008) Caucasian (n=44) G2677A/T NS protein from whole placental tissue C3435T 3435TT > 3435CC Takane et al. (2004) Japanese (n=89) -1517aC, -1017aC, -129C mRNA from enriched trophoblasts

haplotype has > mRNA

B) Study Population SNP Finding Placental Preparation Kobayashi et al. (2005) Japanese (n=100) C376T Creates a stop codon protein from enriched trophoblasts

C421A 421AA < 421CC G34A NS

Page 226: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

210

Pgp polymorphisms and altered placental transfer of drugs

The effect of ABCB1 polymorphisms on transplacental drug transfer has been

investigated for a limited number of substrates. The first series of studies investigated the

influence of 2 SNPs (C3435T and G2677A/T) on the transfer of the Pgp substrate saquinavir

using the dual perfusion of a single lobule in a human term placenta. There was no effect of the

C3435T or G2677A/T genotype on saquinavir transport[15;27]. On the basis of these studies,

the authors suggested that SNPs in ABCB1 do not have a significant impact on transfer.

However, these studies had small sample size and restricted the analysis to only two SNPs. A

more practical and rapid method for determining Pgp function is through preparation of inside

out vesicles from the brush border membrane[24]. Using vesicles from 105 term placentae,

Hemauer et al. identified a >40% increase in Pgp activity in the 1236TT genotype compared to

1236CC and the 3435TT genotype compared to 3435CC[24]. This increase in Pgp activity was

observed despite the decrease in protein expression for the 1236TT and 3435TT genotypes.

Further studies are clearly needed to elucidate the clinical relevance of these findings.

Pgp polymorphisms and influence on normal physiologic function of the placenta and

development

The importance of Pgp in normal physiology and development is still unclear. Mice

completely deficient in Pgp were initially thought to have little change in various physiological

parameters, including viability, fertility, multiple measures of serum clinical chemistry and

hematological parameters, and lymphocyte differentiation[36]. These initial physiological

observations led to the speculation that the role of Pgp was compensated by other

mechanisms[36]. We have shown that this is likely the case. Mice deficient in abcb1a and

abcb1b exhibit a substantial elevation of placental levels of BCRP protein (unpublished

observation).

Recently, an age-related phenotype has been identified in Pgp-deficient mice; mdr1a null

mice develop colitis[37]. This is likely due to inflammation caused by toxins that are normally

excluded from the intestinal epithelial cells by Pgp efflux[37] or from abnormalities in the

Page 227: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

211

epithelial lining of the gut[38]. Pgp deficient mice also exhibit abnormal renal tubular function

as a result of altered intracellular ATP levels and altered mitochondrial morphology[39].

P-glycoprotein likely affects normal development and function of the placenta through

the apoptotic pathways. Pgp may counteract apoptosis by regulating intracellular concentrations

of intermediates in the apoptotic pathway, such as sphingomyelin[40]. Formation of the

syncytiotrophoblast layer involves apoptotic processes and increases in apoptosis are seen in

complications during pregnancy, such as pre-eclampsia and intra-uterine growth restriction (see

Heazell & Crocker[41] for a review). Interestingly, decreased expression of ABCB1 mRNA was

observed in placentas complicated by idiopathic fetal growth restriction (FGR) compared to

controls[42]. Any role of Pgp and its polymorphisms in these pathologies has yet to be

determined.

Breast Cancer Resistance Protein

Expression and function in the placenta

Similar to Pgp, placental BCRP expression has been localized to the microvillous border

of the syncytiotrophoblast[42-45]. BCRP is also expressed on the luminal membrane of fetal

endothelial cells in terminal and intermediate villi[42]. The number of fetal capillaries

expressing BCRP was higher in term tissue compared to that at 6-13 weeks of gestation[45].

Unlike Pgp, there is no clear consensus on how BCRP expression relates to gestational age since

three studies have obtained conflicting results.

The first study reported that ABCG2 mRNA levels and BCRP expression in microvillous

plasma membranes did not differ with gestational age[11]. The second study found no change in

ABCG2 mRNA levels with advancing gestational age, but did identify a significant increase in

whole placental BCRP protein levels at term[45]. The increase in protein but not mRNA levels

suggests that posttranscriptional regulation may be altered with advancing gestation.

Alternatively, the increase in BCRP protein levels may result from the increased BCRP

expression in fetal capillaries and not necessarily in the syncytiotrophoblast[45]. A third study

reported a 2-fold decrease in ABCG2 mRNA as well as BCRP protein from crude trophoblast

Page 228: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

212

membranes at term compared to the preterm group (28 ± 1 week)[44]. All three studies

demonstrated large variability in BCRP expression within gestational age group. These studies

suggest that BCRP expression may be more dependent on genetic and environmental influences

than gestational age.

The regulation of human placental BCRP remains unclear; however, in vitro and animal

studies have brought forward possible mechanisms. The ABCG2 promoter region contains

estrogen and progesterone response elements[46;47]. Similar to Pgp, cytokines are also likely

important regulators of placental BCRP. Decreased expression of BCRP in trophoblasts isolated

from term placentae was observed after incubation with TNF-α and IL-1β[12]. Other factors

shown to influence BCRP expression include estriol, human placental lactogen, and prolactin in

BeWo cells[48]; and insulin-like growth factor II and epidermal growth factor in trophoblasts

isolated from term placentae[12;44].

Like Pgp, placental BCRP also confers fetal protection to toxins in the maternal

circulation. In Pgp-deficient mice, chemical inhibition of BCRP resulted in a 2-fold increase in

fetal exposure to maternally administered topotecan (a BCRP substrate) relative to wild-type[49].

Likewise, using a BCRP knockout model, fetal exposure to topotecan[50], genistein[51],

nitrofurantoin[52], and glyburide[53] was increased 2 to 5-fold compared to wild-type controls.

BCRP has also been demonstrated to limit fetal exposure to glyburide using the dually perfused

human placenta model: inhibition of BCRP resulted in an increase in the fetal-maternal ratio of

glyburide[54]. A recent meta-analysis examining the safety of the oral hypoglycemic, glyburide

in pregnancy compared to insulin treatment reported that there was no increased risk for fetal

hypoglycemia, macrosomia, and other adverse fetal effects[55]. Understanding the role of

BCRP in regulating glyburide transfer across the placenta will allow for identification of women

at risk of altered placental transfer due to drug interactions or polymorphisms in BCRP.

Polymorphisms in BCRP and effect on placental BCRP expression

The ABCG2 gene is located on 4q22 and contains 16 exons and 15 introns[56]. The

encoded protein, BCRP, has only one ATP-binding region and one transmembrane domain[56].

Unlike Pgp, BCRP is considered a half transporter and must homodimerize to become

Page 229: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

213

functional[57]. In most ethnicities, the G34A and C421A SNPs occur at a relatively high

frequency[58-62], however, there are ethnic differences between SNP and haplotype

distributions[62;63]. The G34A and C421 SNPs result in amino acid changes (V12M and

Q141K respectively).

The relationship between ABCG2 polymorphisms and BCRP expression has only been

examined in one study (Table 2). In a Japanese cohort, 20 polymorphisms were detected in the

16 exons and 5’-flanking region: five SNPs resulted in amino acid substitutions[61]. A SNP

encoding a premature stop codon (C376T) was detected with an allelic frequency of 1%

(heterozygosity only). Due to the low frequency of the 376T allele, a statistical analysis could

not be undertaken. However, the two individuals with the 376T allele, exhibited low levels of

BCRP protein. There was also no association between ABCG2 mRNA and the C421A and G34A

SNPs, despite a difference in protein expression for the C421A SNP. The two SNPs, C421A and

G34A, did not result in an allelic expression imbalance. Interestingly, Kobayashi et al. attempted

to perform a haplotype analysis; however, the C421A variant did not coexist with G34A and

C376T in the Japanese population. Furthermore, the authors did not find any significant

haplotype-dependent changes in mRNA or protein expression when considering polymorphisms

in the 5’-flanking and 3’untranslated regions.

In tissues other than the placenta, the C421A SNP has been shown to decrease or not alter

BCRP expression[60]. There was no significant association between the C421A variant and

intestinal expression of BCRP protein or mRNA[62]. Functional changes in BCRP have also

been demonstrated in cancer patients, where patients with the 421A allele had 3-fold higher

plasma concentrations of diflomotecan compared to wild type controls[64]. The C421A SNP

results in an amino acid change that is differentially charged and is also located in the ATP-

binding region[61]. It has also been suggested that this C421A variant protein has altered

tertiary structure and is more easily degraded[60]. Thus, it is very likely that this SNP has

functional significance, even in the placenta.

Epigenetic modulation has also been shown to be an important regulator of BCRP

expression, particularly in cancer cells. The promoter of the ABCG2 gene contains a CpG island

and aberrant methylation has been shown to suppress transcription in renal carcinoma[65].

Moreover, in human multiple myeloma cell lines and patient plasma cells, expression of BCRP

Page 230: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

214

was associated with promoter methylation[66]. Demethylation of the ABCG2 promoter

increased protein expression in lung cell carcinoma[67]. The role of epigenetic regulation of

BCRP has not been studied in the placenta.

BCRP polymorphisms and influence on normal physiologic function of the placenta,

development, and altered placental transfer of drugs

There are currently no studies available that have determined the effect of ABCG2

genotype on BCRP function in the placenta. As discussed previously, BCRP is important in

limiting fetal exposure to various drugs, therefore, altered function could lead to modified fetal

exposure to potential teratogens. Furthermore, new physiological roles for BCRP are currently

being identified and recent associations between BCRP and pregnancy complications involving

the placenta have been made.

Using a BeWo cell model, BCRP was shown to be important in trophoblast

differentiation. Silencing ABCG2 resulted in increases in phosphatidylserine externalization,

intracellular concentrations of ceramide, and apoptosis[68]. Additionally, there was a decrease

in markers of syncytial formation. Similar to Pgp, decreased placental expression of BCRP

(mRNA) was observed in pregnancies with idiopathic FGR[42]. It was suggested that BCRP is a

survival factor for trophoblasts, protecting them from cytokine- or ceramine-induced apoptosis.

Moreover, if reduced BCRP is associated with FGR or other placental complications, it could

render this vulnerable fetus more susceptible to exposure to drugs or toxins normally effluxed by

BCRP.

Another BCRP survival role has been indicated in myeloid progenitor cells from mice in

protecting against hypoxia. Hypoxia results in upregulation of BCRP by HIF-1 and in hypoxic

conditions, BCRP effluxes heme and porphyrins to protect against their toxic effects[69]. In fact,

mice deficient in BCRP develop a protoporphyria[50]. BCRP may also be important in

maintenance of amniotic stem cell differentiation, viability and pluripotency[70]. However, in

BeWo cells, silencing BCRP did not alter intracellular levels of protoporphyrin IX after exposure

to TNF-α and IFN-γ[42]. Thus, the significance of this protective role in the placenta requires

further investigation.

Page 231: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

215

BCRP may also play a role in regulating placental estrogen synthesis by modulating

levels of precursor molecules, such as dehydroepiandrosterone-sulfate (DHEAS), or of metabolic

products, such as estrone-3-sulfate[71]. Regulation of transport of essential nutrients to the fetus,

such as folate and porphyrins also involves BCRP[50]. The localization of BCRP to the junction

between amnion epithelial cells and chorion nuclei suggest that the protein may also play a role

in intercellular communication and regulating access of substrates to the nuclear

compartment[72].

Summary and Implications

To date, there is evidence that the T-129C, C1236T, G2677T/A and C3435T SNPs in

ABCB1 alter expression of Pgp, and that the C376T and C421A SNPs in ABCG1 alter expression

of BCRP in the placenta. However, the functional significance of these polymorphisms in the

placenta is less clear. This lack of clarity likely partially results from methodological limitations

resulting in reduced statistical power. In addition, studies have focused on use of term placental

tissue (when P-gp levels are very low), thus the impact of polymorphisms in function of the

preterm tissue remain unknown. Animal studies will be crucial to investigate the functional

relevance of altered Pgp and BCRP expression earlier in gestation since normal human preterm

tissue will not be obtainable. Many of the human SNP studies have not included haplotype

analysis; investigation of one individual SNP may have poor predictive value[31]. More studies

are needed to determine the impact of polymorphisms in placental Pgp and BCRP on fetal drug

exposure and normal placental function and development.

To date, the role of polymorphisms in Pgp and BCRP in modulating drug-induced

malformations in humans has not been addressed[23]. However, there are many important

examples of how a better understanding of polymorphisms in Pgp and BCRP may lead to

personalized pharmacotherapy in pregnancy. First, the pharmacologic treatment of epilepsy is

often unavoidable in pregnancy because uncontrolled epilepsy carries risks for complications and

adverse fetal outcomes[73]. The antiepileptic drug phenytoin taken during pregnancy is

associated with a definite fetal malformation phenotype[74]. Phenytoin is a substrate for

Pgp[75] and variability in Pgp function due to genetic polymorphisms may explain why most

infants born to epileptic mothers are protected from the congenital malformations and cognitive

Page 232: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

216

impairment while others are not[76]. Establishing a hierarchy of haplotypes could lead to genetic

counseling to better assess fetal risk to phenytoin exposure. This will influence drug choice and

give a better prediction of teratogenicity[76].

Other important examples include the use of glyburide and protease inhibitors during

pregnancy. Placental BCRP is important in limiting fetal exposure to glyburide. However, the

influence of BCRP genotype has not been specifically studied. Susceptible genotypes should be

investigated to avoid any potential increased fetal exposure to glyburide since it could lead to

fetal hypoglycemia. A safe orally administered hypoglycemic is desired for the treatment of

gestational diabetes since current treatment involves daily insulin injections. Daily injections

may lead to impaired adherence and the cost of insulin injections may not be affordable in many

parts of the world[55].

In contrast to glyburide where limited fetal exposure is desired, it may be favorable to

transfer protease inhibitors to the fetus to prevent vertical HIV transmission. Favorable

genotypes could be identified that will allow for optimal transfer of these drugs to the fetus and

for personalized medication selection. Furthermore, a complete understanding of the

relationship between placental drug transporter genotype and phenotype could result in better

drug design for safe and effective drugs in pregnancy.

In summary, a full understanding of which SNPs in Pgp and BCRP affect placental

expression and function would be useful for predicting altered fetal exposure to drugs. This is

clinically relevant when limiting fetal exposure to potential teratogens is desired or when drug

delivery to the fetus is warranted for therapeutic effects. The importance of Pgp and BCRP in

normal placental physiology and development is emerging and susceptible genotypes to

complications such as idiopathic FGR may be identified. Future studies must give consideration

to haplotypes, environmental factors, sample size, gestational age of the placenta, comorbidities

and concurrent medications in order to avoid spurious associations.

Page 233: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

217

Reference List

[1] Macfarland A, Abramovich DR, Ewen SWB, Pearson CK. Stage-specific distribution

of P-glycoprotein in first-trimester and full-term human placenta. Histochem J 1994;

26:417-23.

[2] Sugawara I, Akiyama S, Scheper RJ, Itoyama S. Lung resistance protein (LRP)

expression in human normal tissues in comparison with that of MDR1 and MRP.

Cancer Lett 1997; 112:23-31.

[3] Sun M, Kingdom J, Baczyk D, Lye SJ, Matthews SG, Gibb W. Expression of the

Multidrug Resistance P-Glycoprotein, (ABCB1 glycoprotein) in the Human Placenta

Decreases with Advancing Gestation. Placenta 2006; 27:602-9.

[4] Bremer S, Hoof T, Wilke M, Busche R, Scholte B, Riordan JR et al. Quantitative

expression patterns of multidrug-resistance P-glycoprotein (MDR1) and differentially

spliced cystic-fibrosis transmembrane-conductance regulatory mRNA transcripts in

human epithelia. Eur J Biochem 1992; 206:137-49.

[5] Gil S, Saura R, Forestier F, Farinotti R. P-glycoprotein expression of the human

placenta during pregnancy. Placenta 2005; 26:268-70.

[6] Atkinson DE, Greenwood SL, Sibley CP, Glazier JD, Fairbairn LJ. Role of MDR1 and

MRP1 in trophoblast cells, elucidated using retroviral gene transfer. Am J Physiol Cell

Physiol 2003; 285:C584-91.

[7] Petropoulos S, Kalabis GM, Gibb W, Matthews SG. Functional changes of mouse

placental multidrug resistance phosphoglycoprotein (ABCB1) with advancing gestation

and regulation by progesterone. Reprod Sci 2007; 14:321-8.

[8] Hemauer SJ, Patrikeeva SL, Nanovskaya TN, Hankins GDV, Ahmed MS. Opiates

inhibit paclitaxel uptake by P-glycoprotein in preparations of human placental inside-

out vesicles. Biochem Pharmacol 2009; 78:1272-8.

[9] Kohno K, Sato S, Uchiumi T, Takano H, Kato S, Kuwano M. Tissue-specific enhancer

of the human multidrug-resistance (MDR1) gene. J Biol Chem 1990; 265:19690-6.

Page 234: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

218

[10] Loregger T, Pollheimer J, Knofler M. Regulatory transcription factors controlling

function and differentiation of human trophoblast - A review. Placenta 2003; 24:S104-

S110.

[11] Mathias AA, Hitti J, Unadkat JD. P-glycoprotein and breast cancer resistance protein

expression in human placentae of various gestational ages. Am J Physiol Regul Integr

Comp Physiol 2005; 289:R963-R969.

[12] Evseenko DA, Paxton JW, Keelan JA. Independent regulation of apical and basolateral

drug transporter expression and function in placental trophoblasts by cytokines,

steroids, and growth factors. Drug Metab Dispos 2007; 35:595-601.

[13] Petrovic V, Wang JH, Piquette-Miller M. Effect of endotoxin on the expression of

placental drug transporters and glyburide disposition in pregnant rats. Drug Metab

Dispos 2008; 36:1944-50.

[14] Smit JW, Huisman MT, Van Tellingen O, Wiltshire HR, Schinkel AH. Absence or

pharmacological blocking of placental P-glycoprotein profoundly increases fetal drug

exposure. J Clin Invest 1999; 104:1441-7.

[15] Molsa M, Heikkinen T, Hakkola J, Hakala K, Wallerman O, Wadelius M et al.

Functional role of P-glycoprotein in the human blood-placental barrier. Clin Pharmacol

Ther 2005; 78:123-31.

[16] Sudhakaran S, Rayner CR, Li J, Kong DCM, Gude NM, Nation RL. Inhibition of

placental P-glycoprotein: Impact on indinavir transfer to the foetus. Br J Clin

Pharmacol 2008; 65:667-73.

[17] Nanovskaya T, Nekhayeva I, Karunaratne N, Audus K, Hankins GDV, Ahmed MS.

Role of P-glycoprotein in transplacental transfer of methadone. Biochem Pharmacol

2005; 69:1869-78.

[18] Callen DF, Baker E, Simmers RN, Seshadri R, Roninson IB. Localization of the human

multiple drug resistance gene, MDR1, to 7q21.1. Hum Genet 1987; 77:142-4.

Page 235: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

219

[19] Lankas GR, Wise LD, Cartwright ME, Pippert T, Umbenhauer DR. Placental P-

glycoprotein deficiency enhances susceptibility to chemically induced birth defects in

mice. Reprod Toxicol 1998; 12:457-63.

[20] Li YH, Wang YH, Li Y, Yang L. MDR1 gene polymorphisms and clinical relevance.

Acta Genetica Sinica 2006; 33:93-104.

[21] Tanabe M, Ieiri I, Nagata N, Inoue K, Ito S, Kanamori Y et al. Expression of P-

glycoprotein in human placenta: Relation to genetic polymorphism of the multidrug

resistance (MDR)-1 gene. J Pharmacol Exp Ther 2001; 297:1137-43.

[22] Mathijssen RHJ, Marsh S, Karlsson MO, Xie R, Baker SD, Verweij J et al. Irinotecan

pathway genotype analysis to predict pharmacokinetics. Clin Cancer Res 2003; 9:3246-

53.

[23] Hitzl M, Schaeffeler E, Hocher B, Slowinski T, Halle H, Eichelbaum M et al. Variable

expression of P-glycoprotein in the human placenta and its association with mutations

of the multidrug resistance 1 gene (MDR1, ABCB1). Pharmacogenetics 2004; 14:309-

18.

[24] Hemauer SJ, Nanovskaya TN, Abdel-Rahman SZ, Patrikeeva SL, Hankins GDV,

Ahmed MS. Modulation of human placental P-glycoprotein expression and activity by

MDR1 gene polymorphisms. Biochem Pharmacol 2009; In Press.

[25] Rahi M, Heikkinen T, Hakkola J, Hakala K, Wallerman O, Wadelius M et al. Influence

of adenosine triphosphate and ABCB1 (MDR1) genotype on the P-glycoprotein-

dependent transfer of saquinavir in the dually perfused human placenta. Hum Exp

Toxicol 2008; 27:65-71.

[26] Takane H, Kobayashi D, Hirota T, Kigawa J, Terakawa N, Otsubo K et al. Haplotype-

oriented genetic analysis and functional assessment of promoter variants in the MDR1

(ABCB1) gene. J Pharmacol Exp Ther 2004; 311:1179-87.

[27] Baker EK, Johnstone RW, Zalcberg JR, El Osta A. Epigenetic changes to the MDR1

locus in response to chemotherapeutic drugs. Oncogene 2005; 24:8061-75.

Page 236: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

220

[28] Fryxell KB, McGee SB, Simoneaux DK, Willman CL, Cornwell MM. Methylation

analysis of the human multidrug resistance 1 gene in normal and leukemic

hematopoietic cells. Leukemia 1999; 13:910-7.

[29] Kusaba H, Nakayama M, Harada T, Nomoto M, Kohno K, Kuwano M et al.

Association of 5' CpG demethylation and altered chromatin structure in the promoter

region with transcriptional activation of the multidrug resistance gene in human cancer

cells. Eur J Biochem 1999; 262:924-32.

[30] Aquilante CL, Letrent SP, Pollack GM, Brouwer KLR. Increased brain P-glycoprotein

in morphine tolerant rats. Life Sci 1999; 66:PL47-51.

[31] Chowbay B, Li H, David M, Cheung YB, Lee EJD. Meta-analysis of the influence of

MDR1 C3435T polymorphism on digoxin pharmacokinetics and MDR1 gene

expression. Br J Clin Pharmacol 2005; 60:159-71.

[32] Hoffmeyer S, Burk O, Von Richter O, Arnold HP, Brockm+¦ller J, Johne A et al.

Functional polymorphisms of the human multidrug-resistance gene: Multiple sequence

variations and correlation of one allele with P-glycoprotein expression and activity in

vivo. Proc Natl Acad Sci USA 2000; 97:3473-8.

[33] Fellay J, Marzolini C, Meaden ER, Back DJ, Buclin T, Chave JP et al. Response to

antiretroviral treatment in HIV-1-infected individuals with allelic variants of the

multidrug resistance transporter 1: A pharmacogenetics study. Lancet 2002; 359:30-6.

[34] Hitzl M, Drescher S, Van der Kuip H, Scha¦êffeler E, Fischer J, Schwab M et al. The

C3435T mutation in the human MDR1 gene is associated with altered efflux of the P-

glycoprotein substrate rhodamine 123 from CD56+ natural killer cells.

Pharmacogenetics 2001; 11:293-8.

[35] Siegsmund M, Brinkmann U, Scha¦êffeler E, Weirich G, Schwab M, Eichelbaum M et

al. Association of the P-glycoprotein transporter MDR1C3435T polymorphism with the

susceptibility to renal epithelial tumors. J Am Soc Nephrol 2002; 13:1847-54.

Page 237: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

221

[36] Schinkel AH, Mayer U, Wagenaar E, Mol CAAM, Van Deemter L, Smit JJM et al.

Normal viability and altered pharmacokinetics in mice lacking mdr1-type (drug-

transporting) P-glycoproteins. Proc Natl Acad Sci USA 1997; 94:4028-33.

[37] Panwala CM, Jones JC, Viney JL. A novel model of inflammatory bowel disease: Mice

deficient for the multiple drug resistance gene, mdr1a, spontaneously develop colitis. J

Immunol 1998; 161:5733-44.

[38] Resta-Lenert S, Smitham J, Barrett KE. Epithelial dysfunction associated with the

development of colitis in conventionally housed mdr1a-/- mice. Am J Physiol

Gastrointest Liver Physiol 2005; 289:G153-G162.

[39] Huls M, Kramers C, Levtchenko EN, Wilmer MJG, Dijkman HBPM, Kluijtmans LAJ

et al. P-glycoprotein-deficient mice have proximal tubule dysfunction but are protected

against ischemic renal injury. Kidney Int 2007; 72:1233-41.

[40] Smyth MJ, Krasovskis E, Sutton VR, Johnstone RW. The drug efflux protein, P-

glycoprotein, additionally protects drug-resistant tumor cells from multiple forms of

caspase-dependent apoptosis. Proc Natl Acad Sci USA 1998; 95:7024-9.

[41] Heazell AEP, Crocker IP. Live and Let Die - Regulation of Villous Trophoblast

Apoptosis in Normal and Abnormal Pregnancies. Placenta 2008; 29:772-83.

[42] Evseenko DA, Murthi P, Paxton JW, Reid G, Emerald BS, Mohankumar KM et al. The

ABC transporter BCRP/ABCG2 is a placental survival factor, and its expression is

reduced in idiopathic human fetal growth restriction. FASEB J 2007; 21:3592-605.

[43] Maliepaard M, Scheffer GL, Faneyte IF, Van Gastelen MA, Pijnenborg ACLM,

Schinkel AH et al. Subcellular localization and distribution of the Breast Resistance

Protein Transporter in normal human tissues. Cancer Res 2001; 61:3458-64.

[44] Meyer Zu Schwabedissen HE, Grube M, Dreisbach A, Jedlitschky G, Meissner K,

Linnemann K et al. Epidermal growth factor-mediated activation of the MAP kinase

cascade results in altered expression and function of ABCG2 (BCRP). Drug Metab

Dispos 2006; 34:524-33.

Page 238: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

222

[45] Yeboah D, Sun M, Kingdom J, Baczyk D, Lye SJ, Matthews SG et al. Expression of

breast cancer resistance protein (BCRP/ABCG2) in human placenta throughout

gestation and at term before and after labor. Can J Physiol Pharmacol 2006; 84:1251-8.

[46] Ee PLR, Kamalakaran S, Tonetti D, He X, Ross DD, Beck WT. Identification of a

Novel Estrogen Response Element in the Breast Cancer Resistance Protein (ABCG2)

Gene. Cancer Res 2004; 64:1247-51.

[47] Wang H, Lee EW, Zhou L, Leung PCK, Ross DD, Unadkat JD et al. Progesterone

receptor (PR) isoforms PRA and PRB differentially regulate expression of the breast

cancer resistance protein in human placental choriocarcinoma BeWo cells. Mol

Pharmacol 2008; 73:845-54.

[48] Wang H, Unadkat JD, Mao Q. Hormonal regulation of BCRP expression in human

placental BeWo cells. Pharm Res 2008; 25:444-52.

[49] Jonker JW, Smit JW, Brinkhuis RF, Maliepaard M, Beijnen JH, Schellens JHM et al.

Role of breast cancer resistance protein in the bioavailability and fetal penetration of

topotecan. J Natl Cancer Inst 2000; 92:1651-6.

[50] Jonker JW, Buitelaar M, Wagenaar E, Van der Valk MA, Scheffer GL, Scheper RJ et

al. The breast cancer resistance protein protects against a major chlorophyll-derived

dietary phototoxin and protoporphyria. Proc Natl Acad Sci USA 2002; 99:15649-54.

[51] Enokizono J, Kusuhara H, Sugiyama Y. Effect of breast cancer resistance protein

(Bcrp/Abcg2) on the disposition of phytoestrogens. Mol Pharmacol 2007; 72:967-75.

[52] Zhang Y, Wang H, Unadkat JD, Mao Q. Breast cancer resistance protein 1 limits fetal

distribution of nitrofurantoin in the pregnant mouse. Drug Metab Dispos 2007;

35:2154-8.

[53] Zhou L, Naraharisetti SB, Wang H, Unadkat JD, Hebert MF, Mao Q. The breast cancer

resistance protein (Bcrp1/Abcg2) limits fetal distribution of glyburide in the pregnant

mouse: An Obstetric-Fetal Pharmacology Research Unit Network and University of

Washington Specialized Center of Research Study. Mol Pharmacol 2008; 73:949-59.

Page 239: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

223

[54] Pollex E, Lubetsky A, Koren G. The Role of Placental Breast Cancer Resistance

Protein in the Efflux of Glyburide across the Human Placenta. Placenta 2008; 29:743-7.

[55] Moretti ME, Rezvani M, Koren G. Safety of glyburide for gestational diabetes: A meta-

analysis of pregnancy outcomes. Ann Pharmacother 2008; 42:483-90.

[56] Allikmets R, Schriml LM, Hutchinson A, Romano-Spica V, Dean M. A human

placenta-specific ATP-binding cassette gene (ABCP) on chromosome 4q22 that is

involved in multidrug resistance. Cancer Res 1998; 58:5337-9.

[57] Kage K, Tsukahara S, Sugiyama T, Asada S, Ishikawa E, Tsuruo T et al. Dominant-

negative inhibition of breast cancer resistance protein as drug efflux pump through the

inhibition of S-S dependent homodimerization. Int J Cancer 2002; 97:626-30.

[58] Backstrom G, Taipalensuu J, Melhus H, Brandstrom H, Svensson AC, Artursson P et

al. Genetic variation in the ATP-binding Cassette Transporter gene ABCG2 (BCRP) in

a Swedish population. Eur J Pharm Sci 2003; 18:359-64.

[59] Bosch TM, Kjellberg LM, Bouwers A, Koeleman BPC, Schellens JHM, Beijnen JH et

al. Detection of single nucleotide polymorphisms in the ABCG2 gene in a Dutch

population. Am J Pharmacogenomics 2005; 5:123-31.

[60] Imai Y, Nakane M, Kage K, Tsukahara S, Ishikawa E, Tsuruo T et al. C421A

polymorphism in the human breast cancer resistance protein gene is associated with

low expression of Q141K protein and low-level drug resistance. Mol Cancer Ther

2002; 1:611-6.

[61] Kobayashi D, Ieiri I, Hirota T, Takane H, Maegawa S, Kigawa J et al. Functional

assessment of ABCG2 (BRCP) gene polymorphisms to protein expression in human

placenta. Drug Metab Dispos 2005; 33:94-101.

[62] Zamber CP, Lamba JK, Yasuda K, Farnum J, Thummel K, Schuetz JD et al. Natural

allelic variants of breast cancer resistance protein (BCRP) and their relationship to

BCRP expression in human intestine. Pharmacogenetics 2003; 13:19-28.

Page 240: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

224

[63] Maekawa K, Itoda M, Sai K, Saito Y, Kaniwa N, Shirao K et al. Genetic variation and

haplotype structure of the ABC transporter gene ABCG2 in a Japanese population.

Drug metab pharmacokinet 2006; 21:109-21.

[64] Sparreboom A, Gelderblom H, Marsh S, Ahluwalia R, Obach R, Principe P et al.

Diflomotecan pharmacokinetics in relation to ABCG2 421C>A genotype. Clin

Pharmacol Ther 2004; 76:38-44.

[65] To KKW, Zhan Z, Bates SE. Aberrant promoter methylation of the ABCG2 gene in

renal carcinoma. Mol Cell Biol 2006; 26:8572-85.

[66] Turner JG, Gump JL, Zhang C, Cook JM, Marchion D, Hazlehurst L et al. ABCG2

expression, function, and promoter methylation in human multiple myeloma. Blood

2006; 108:3881-9.

[67] Nakano H, Nakamura Y, Soda H, Kamikatahira M, Uchida K, Takasu M et al.

Methylation status of breast cancer resistance protein detected by methylation-specific

polymerase chain reaction analysis is correlated inversely with its expression in drug-

resistant lung cancer cells. Cancer 2008; 112:1122-30.

[68] Evseenko DA, Paxton JW, Keelan JA. The Xenobiotic Transporter ABCG2 Plays a

Novel Role in Differentiation of Trophoblast-like BeWo Cells. Placenta 2007; 28:S116-

20.

[69] Krishnamurthy P, Ross DD, Nakanishi T, Bailey-Dell K, Zhou S, Mercer KE et al. The

stem cell marker Bcrp/ABCG2 enhances hypoxic cell survival through interactions

with heme. J Biol Chem 2004; 279:24218-25.

[70] Aye IL, Paxton JW, Evseenko DA, Keelan JA. Expression, Localisation and Activity of

ATP Binding Cassette (ABC) Family of Drug Transporters in Human Amnion

Membranes. Placenta 2007; 28:868-77.

[71] Grube M, Reuther S, Meyer Zu Schwabedissen H, Köck K, Draber K, Ritter CA et al.

Organic anion transporting polypeptide 2B1 and breast cancer resistance protein

Page 241: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

225

interact in the transepithelial transport of steroid sulfates in human placenta. Drug

Metab Dispos 2007; 35:30-5.

[72] Yeboah D, Kalabis GM, Sun M, Ou RC, Matthews SG, Gibb W. Expression and

localisation of breast cancer resistance protein (BCRP) in human fetal membranes and

decidua and the influence of labour at term. Reprod Fertil Dev 2008; 20:328-34.

[73] Viinikainen K, Heinonen S, Eriksson K, Kälviäinen R. Community-based, prospective,

controlled study of obstetric and neonatal outcome of 179 pregnancies in women with

epilepsy. Epilepsia 2006; 47:186-92.

[74] Nulman I, Laslo D, Koren G. Treatment of epilepsy in pregnancy. Drugs 1999; 57:535-

44.

[75] Luna-Tortos C, Fedrowitz M, Loscher W. Several major antiepileptic drugs are

substrates for human P-glycoprotein. Neuropharmacology 2008; 55:1364-75.

[76] Atkinson DE, Brice-Bennett S, D'Souza SW. Antiepileptic medication during

pregnancy: Does fetal genotype affect outcome? Pediatr Res 2007; 62:120-7.

[77]Marzolini C, Paus E, Buclin T, Kim RB. Polymorphisms in human MDR1 (P-

glycoprotein): recent advances and clinical relevance. Clin Pharmacol Ther 2004

;75:13-33

[78] Mao Q, Unadkat JD. Role of the breast cancer resistance protein (ABCG2) in drug

transport. AAPS J 2005 ;7:E118-33.

[79] St-Pierre MV, Serrano MA, Macias RI, Dubs U, Hoechli M, Lauper U et al.

Expression of members of the multidrug resistance protein family in human term

placenta. Am J Physiol Regul Integr Comp Physiol 2000; 279:R1495-503.

[80] Meyer Zu Schwabedissen HE, Grube M, Heydrich B, Linnemann K, Fusch C, Kroemer

HK, Jedlitschky G. Expression, localization, and function of MRP5 (ABCC5), a

transporter for cyclic nucleotides, in human placenta and cultured human trophoblasts:

effects of gestational age and cellular differentiation. Am J Pathol 2005;166:39-48.

Page 242: Evaluation of transplacental pharmacology and toxicology from … · 2015-06-17 · ii Evaluation of transplacental pharmacology and toxicology from bench to bedside Janine Rose Hutson

226

[81] Nagashige M, Ushigome F, Koyabu N, Hirata K, Kawabuchi M, Hirakawa T et al.

Basal membrane localization of MRP1 in human placental trophoblast. Placenta.

2003;24:951-8.