copyright 2018 seán martin mythen

158
Copyright 2018 Seán Martin Mythen

Upload: others

Post on 04-Jun-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Copyright 2018 Seán Martin Mythen

Copyright 2018 Seán Martin Mythen

Page 2: Copyright 2018 Seán Martin Mythen

CHARACTERIZATION OF ENZYMATIC PATHWAYS INVOLVED IN CORTISOL AND BILE ACID METABOLISM BY THE GUT MICROBIOTA

BY

SEÁN MARTIN MYTHEN

THESIS

Submitted in partial fulfillment of the requirements for the degree of Master of Science in Animal Sciences

in the Graduate College of the University of Illinois at Urbana-Champaign, 2018

Urbana, Illinois

Advisor:

Assistant Professor Jason M. Ridlon

Page 3: Copyright 2018 Seán Martin Mythen

ii

ABSTRACT

The gut microbiota consists of a complex network of distinct bacterial taxa that

together affect the physiology of the human host. Of the many facets affected by gut

microbiota, their impact on the endocrine system is both significant and understudied.

The diversity observed in microbial isolates and operational taxonomic units from one

human gut to another is due largely in part by the strain specific proteins expressed by

the microbes that inhabit them. Some of these unique proteins have evolved to

biotransform host sterols which can be reabsorbed by the body and potentially affect

endocrine functions.

Bacterial hydroxysteroid dehydrogenases (HSDHs) have the potential to

significantly alter the physicochemical properties of bile acids with implications for

increased/decreased toxicity for gut bacteria and the host. We located a gene-cluster in

Eggerthella CAG:298 predicted to encode three HSDHs (CDD59473, CDD59474,

CDD59475), synthesized the genes for heterologous expression in E. coli and then

screened bile acid substrates against the purified recombinant enzymes, revealing

novel 3α-, 3β-, and 12α-HSDHs.

We also developed an enzyme-linked continuous spectrophotometric assay to

quantify steroid-17,20-desmolase activity from recombinant enzymes encoded by the

desA and desB genes from Clostridium scindens ATCC 35704. Steroid-17,20-

desmolase is responsible for the side chain cleavage that biotransforms cortisol into

11β-hydroxyandrostenedione (11β-OHAD).

The reaction performed by the steroid-17,20-desmolase appears to be regulated

via pyridine nucleotide-dependent HSDHs, which are capable of converting cortisol into

Page 4: Copyright 2018 Seán Martin Mythen

iii

a form unable to be utilized by steroid-17,20-desmolase. A recently identified 20β-

HSDH from Bifidobacterium adolescentis strain L2-32 has been biochemically

characterized and crystallized. This 20β-HSDH has the potential to directly affect

androgen formation by functioning as a metabolic rheostat controlling the steroid-17,20-

desmolase activity of Clostridium scindens ATCC 35704.

Page 5: Copyright 2018 Seán Martin Mythen

iv

ACKNOWLEDGMENTS

This research would not have been possible without the support of my adviser,

Dr. Jason Ridlon, who took me under his wing when I barely knew how to use a pipette.

I am also thankful for the guidance provided by Dr. Roderick Mackie and Dr. Isaac

Cann. I would like to thank Dr. Saravanan Devendran for his teachings over the past

few years. I would also like to thank each of the students and post-docs in the Ridlon,

Mackie, and Cann labs, with whom I have shared many great moments. I would also

like to thank my family and friends for their endless encouragement. Finally, I would like

to thank my girlfriend, Samantha, for her patience, love, and support.

Page 6: Copyright 2018 Seán Martin Mythen

v

TABLE OF CONTENTS

LIST OF FIGURES ......................................................................................................... vii

LIST OF TABLES ............................................................................................................ix

CHAPTER 1: Literature Review ...................................................................................... 1

INTRODUCTION .................................................................................................... 1

STEROIDS ............................................................................................................. 1

STEROID HORMONES .......................................................................................... 7

BILE ACIDS .......................................................................................................... 12

THE GUT MICROBIOTA ...................................................................................... 17

Clostridium scindens ............................................................................................. 17

Eggerthella lenta ................................................................................................... 19

Butyricicoccus desmolans .................................................................................... 20

Bifidobacterium adolescentis ................................................................................ 21

OBJECTIVES ....................................................................................................... 23

CHAPTER 2: Targeted synthesis and characterization of a gene-cluster encoding NAD(P)H-dependent 3α-, 3β-, and 12α-hydroxysteroid dehydrogenases from Eggerthella CAG:298, a gut metagenomic sequence ................................................... 25

ABSTRACT .......................................................................................................... 25

INTRODUCTION .................................................................................................. 26

MATERIALS AND METHODS .............................................................................. 29

RESULTS ............................................................................................................. 33

DISCUSSION ....................................................................................................... 54

CHAPTER 3: The desA and desB genes from Clostridium scindens ATCC 35704 encode steroid-17,20-desmolase .................................................................................. 60

ABSTRACT .......................................................................................................... 60

INTRODUCTION .................................................................................................. 61

MATERIALS AND METHODS .............................................................................. 62

RESULTS ............................................................................................................. 69

DISCUSSION ....................................................................................................... 79

CHAPTER 4: Structural and biochemical characterization of 20β-hydroxysteroid dehydrogenase from Bifidobacterium adolescentis strain L2-32 ................................... 95

ABSTRACT .......................................................................................................... 95

INTRODUCTION .................................................................................................. 95

Page 7: Copyright 2018 Seán Martin Mythen

vi

MATERIALS AND METHODS .............................................................................. 97

RESULTS ........................................................................................................... 104

DISCUSSION ..................................................................................................... 114

SUMMARY OF RESEARCH ....................................................................................... 118

FUTURE RESEARCH ................................................................................................. 124

REFERENCES ............................................................................................................ 126

Page 8: Copyright 2018 Seán Martin Mythen

vii

LIST OF FIGURES

Figure 1.1 Steroid classifications .................................................................................... 3

Figure 1.2 Steroid numbering ......................................................................................... 4

Figure 1.3 Alpha and beta isomeric variation of the C-7 hydroxyl group ........................ 6

Figure 1.4 Phylogenetic analysis of identified 20β-hydroxysteroid dehydrogenases .... 22

Figure 2.1 SDS-PAGE of purified recombinant 3β-, 3α-, and 12α-hydroxysteroid dehydrogenases from a gene cluster in Eggerthella CAG:298 ...................................... 38

Figure 2.2 Representative TLC of bile acid reaction products from recombinant CDD59473 (A), CDD59474 (B), and CDD59475 (C) .................................................... 39

Figure 2.3 Mass spectrometry of enzymatic reaction products ..................................... 42

Figure 2.4 pH optima for purified recombinant CDD59473 (A), CDD59474 (B), and CDD59475 (C) in the oxidative and reductive directions. .............................................. 44

Figure 2.5 Kinetic analysis of recombinant CDD59473, CDD59474, and CDD59475 from metagenomic sequence of Eggerthella CAG:298 ................................................. 46

Figure 2.6 Schematic representation of bile acid metabolism by Eggerthella lenta. ..... 50

Figure 2.7 iTOl representation of FastTree maximum likelihood phylogeny from MUSCLE alignment of bacterial hydroxysteroid dehydrogenases and the placement of CDD59473, CDD59474, and CDD59475. ..................................................................... 51

Figure 2.8 CLUSTAL alignment of biochemically characterized gut bacterial hydroxysteroid dehydrogenases ................................................................................... 57

Figure 3.1 Steroid-17,20-desmolase (DesAB) shares conserved catalytic and co-factor binding residues with transketolases ............................................................................. 70

Figure 3.2 SDS-PAGE of affinity purified rDesA and rDesB and native molecular mass estimate of rDesAB. ...................................................................................................... 72

Figure 3.3 Liquid chromatography-mass spectrometry of rDesAB reaction products ... 73

Figure 3.4 SDS-PAGE of purified recombinant 17β-HSDH from Cochliobolus lunatus (rCL17HSDH) and schematic representation of enzyme-coupled continuous spectrophotometric steroid-17,20-desmolase activity assay. ........................................ 74

Figure 3.5 Biochemical characterization of rDesAB by enzyme-coupled continuous spectrophotometric assay. ............................................................................................ 76

Page 9: Copyright 2018 Seán Martin Mythen

viii

Figure 3.6 Determination of steroid-17,20-desmolase activity against substrates lacking a 17-hydroxy, and tetrahydro(deoxy)cortisol by LC/MS................................................. 80

Figure 3.7 Functional groups important for rDesAB specificity ..................................... 82

Figure 3.8 Mass spectra of rDesAB catalyzed reactions. ............................................. 83

Figure 3.9 Gene organization and domain architecture of steroid-17,20-desmolase ... 91

Figure 3.10 Schematic representation of side-chain cleavage reactions for steroid-17,20-desmolase from C. scindens ATCC 35704, steroid-transaldolase from Mycobacterium tuberculosis, and Cytochrome P450 monooxygenase (CYP11A1) in Eukaryotes. ................................................................................................................... 93

Figure 4.1 Reaction mechanism of 20β-HSDH ........................................................... 105

Figure 4.2 Whole cell assays of B. adolescentis strain L2-32, B. desmolans ATCC 43058, and C. scindens ATCC 35704 ......................................................................... 106

Figure 4.3 Purified 20β-hydroxysteroid dehydrogenase (20β-HSDH) from Bifidobacterium adolescentis strain L2-32 ................................................................... 108

Figure 4.4 pH optimum and Michaelis-Menten saturation curve of wild-type 20β-HSDH .................................................................................................................................... 109

Figure 4.5 Circular dichroism spectroscopy ................................................................ 112

Figure 4.6 20β-HSDH crystals .................................................................................... 115

Figure 4.7 Competing cortisol metabolizing reactions by B. adolescentis, E. lenta, B. desmolans, and C. scindens ....................................................................................... 123

Page 10: Copyright 2018 Seán Martin Mythen

ix

LIST OF TABLES

Table 1.1 Concentrations of major steroids found in humans ......................................... 9

Table 2.1 Oligonucleotides used for amplification and directional cloning .................... 30

Table 2.2 Sequences utilized in phylogenetic analysis ................................................. 34

Table 2.3 Kinetic parameters of recombinant CDD59473, CDD59474, and CDD59475 ...................................................................................................................................... 45

Table 2.4 Substrate-specificity of Eggerthella CAG:298 HSDHs .................................. 48

Table 3.1 Oligonucleotides and synthetic gene sequences used in this study .............. 64

Table 3.2 Substrate specificity for rDesAB from C. scindens ATCC 35704 .................. 78

Table 4.1 Kinetic parameters of cortisol reduction by wild-type and mutant 20β-HSDH .................................................................................................................................... 111

Table 4.2 Substrate specificity of 20β-HSDH .............................................................. 113

Page 11: Copyright 2018 Seán Martin Mythen

1

CHAPTER 1: Literature Review

INTRODUCTION

It is now widely recognized that the community of microbes that inhabit the guts

of higher eukaryotes function as a “virtual organ”, with many metabolic, immunologic,

and endocrine influences (1). Many microbes in the gut are known to benefit the host

by fermenting indigestible food, synthesizing vitamins, and protecting against

pathogens. It is known that steroids circulating throughout the body alter metabolism,

salt and water balance, immune functions, digestion, and sexual characteristics. The

effect of steroid biotransformations performed by microbes within the gut is significantly

understudied. We have proposed the term “sterolbiome” to describe the repertoire of

human gut microbiome genes involved in the uptake and metabolism of host,

pharmaceutical, and diet derived steroids (2). Small changes in steroid structure can

have significant consequences on both host physiology and disease, as well as

microbial diversity, spore germination, and microbial growth (3). Therefore, there is a

pressing need to understand the nature of steroid biotransformation in the gut and the

biological significance it has on the host. Here, we characterize important gut microbe

enzymatic pathways involved in the biotransformation of steroids.

STEROIDS

Steroids are biologically important molecules utilized by eukaryotic organisms as

essential signaling molecules (4). They exert their effects on the host by binding to

intracellular nuclear receptors which can either induce or repress gene transcription (5).

Small changes in the structure of a steroid determines its overall biological effect and

functional class to which it belongs. The functional classes discussed here are

Page 12: Copyright 2018 Seán Martin Mythen

2

glucocorticoids, mineralocorticoids, androgens, estrogens, progestogens, and bile acids

(Figure 1.1) (6).

Steroid Structure and Host Steroid Diversity- Each of these steroid groups is

a biological derivative of cholesterol, and because of that, every steroid shares a

common cyclopentanophenanthrene parent ring structure of three six-carbon

cyclohexane rings and one five-carbon cyclopentane ring, denoted as the A, B, C, and

D rings, respectively (Figure 1.2). Host enzymes create steroid diversity via

modifications to the eight-carbon aliphatic side-chain (carbon positions 20-27), and the

hydroxylation of carbons on the parent ring structure resulting in variations in steroid

length, branching, and oxidation. The steroids secreted into the human gut are uniquely

modified by anaerobic microbes that are capable of producing steroids that the host is

incapable of synthesizing. This explains the differences in gut steroid composition

observed in germ-free versus conventional animals (7).

Steroid Biotransformation by Gut Microbes- A multitude of microbial enzymes

exist that can perform various transformations to the ring system and the side chain of

steroid molecules (8). The main enzymatic reactions catalyzed by gut microbial

enzymes tend to be net reductive and hydrolytic, whereas host enzymatic reactions

tend to be oxidative and conjugative. It is possible to form more than 230 steroids If you

take into account the interplay between host and microbial enzymes (6).

Hydroxysteroid dehydrogenases (HSDHs) are a class of pyridine nucleotide-dependent

oxidoreductases that are involved in the reversible oxidation and epimerization of host

steroids and are the most common type of steroid metabolizing enzymes.

Page 13: Copyright 2018 Seán Martin Mythen

3

Figu

re 1

.1 S

tero

id c

lass

ifica

tions

Al

l ste

roid

s ar

e bi

olog

ical

ly d

eriv

ed fr

om c

hole

ster

ol.

The

mai

n ac

tive

ster

oids

pro

duce

d in

hum

ans

are

liste

d in

th

e bo

ttom

row

. Th

ey a

re o

rgan

ized

into

gro

ups

base

d on

the

rece

ptor

to w

hich

they

bin

d an

d w

here

they

are

pr

oduc

ed in

the

body

.

Page 14: Copyright 2018 Seán Martin Mythen

4

Figure 1.2 Steroid numbering At its core, a steroid is composed of 17 carbons organized into 4 rings (A, B, C, and D). Rings A, B, and C are cyclohexanes. Ring D is a cyclopentane. Steroids are made unique by the addition or removal of functional groups, by reduction or oxidation of carbons on the 4 ring structure, and many other modifications.

Page 15: Copyright 2018 Seán Martin Mythen

5

Hydroxysteroid dehydrogenases- HSDHs expressed by the host are a means

of rapidly regulating the intracellular concentration of ligands for various host nuclear

receptors by functioning as enzymatic “switches” that shuttle steroids between active

and inactive forms. The best example is the interconversion of cortisol (11β-hydroxyl) to

cortisone (11-keto) by the human 11β-HSD1 and 2 isoforms (9). Cortisol is capable of

binding to and activating both the glucocorticoid and the mineralocorticoid receptor, but

cortisone is not a ligand nor an activator of these receptors. In this way, reversible

oxidation and epimerization by HSDHs produce slight, readily reversible, structural

modifications that determine the biological activity of a population of steroid molecules

in host tissues and cells (Figure 1.3). By contrast, HSDHs expressed by microbes

function as redox sinks and have the potential to indirectly affect host physiology and

the surrounding microbial community.

Most HSDHs belong to the superfamily of proteins known as the short-chain

dehydrogenases/reductases (SDRs) (10). Proteins in the SDR superfamily are typically

composed of 250-350 residues and present highly similar tertiary structure folding

patterns consisting of a characteristic Rossmann-fold; which consists of a central β-

sheet (composed of 6-7 β-strands), flanked by α-helices (11). HSDHs are important

SDR type enzymes with a role in various processes such as host growth, energy

homeostasis, reproduction, and cell signaling (12). HSDHs are also widespread in the

gut microbiome, with the appearance of having been randomly scattered between the

various genomes of gut species. This random appearance is largely due to an

inadequate understanding of the function(s) of HSDHs for particular isolates and the

Page 16: Copyright 2018 Seán Martin Mythen

6

Figure 1.3 Alpha and beta isomeric variation of the C-7 hydroxyl group The hydroxyl (-OH) group bound to carbon 7 (red) of ring B can be in one of two isomeric orientations. Alpha (α) is represented by dashes, signifying the bond going into the plane of reference or away from the reader. Beta (β) is represented by a wedge, signifying the bond coming out of the plane of reference or toward the reader. In this example, distinct 7α- or 7β-HSDHs perform the biotransformations.

Page 17: Copyright 2018 Seán Martin Mythen

7

effect of bile acid chemistry on gut bacterial fitness (13). Many of the genes encoding

HSDHs have yet to be identified. It is imperative that we characterize these enzymes if

we are to understand the full extent to which they modify steroids and affect the host

and surrounding microbial community.

STEROID HORMONES

Many steroids act as hormone signaling molecules that bind to receptors and

cause physiological changes to the host. These “steroid hormones” are produced in

numerous host tissues, including the adrenal cortex, gonads, and placenta, and are all

synthesized from cholesterol (5). Steroids can then be transported through the blood

via plasma transport proteins to target tissues that have receptors for that specific

steroid hormone (6). Steroids exert their regulatory effects by binding to cell surface

receptors or entering target cells by passive diffusion and associating with and

activating the nuclear receptors therein (5, 14, 15). Once bound to the receptor, the

steroid induces a conformational change in the receptor which then generates a specific

biological response based on the target tissue, receptor, and steroid in question (6).

Steroid hormones are categorized into two classes and five types according to

the receptors to which they bind. The two classes are corticosteroids and sex steroids.

The 5 types are: glucocorticoids, mineralocorticoids, androgens, progestogens, and

estrogens (Figure 1.1).

Corticosteroids- The corticosteroids consist of glucocorticoids and

mineralocorticoids. Corticosteroids are synthesized in the adrenal cortex in mammals

(16). Corticosteroids contain 21 carbons and are characterized by an oxo group at C-3

and 20, a Δ4,5 double bond, a two-carbon side chain on C-17, and a hydroxyl on C-21.

Page 18: Copyright 2018 Seán Martin Mythen

8

Glucocorticoids- Glucocorticoids are categorized by the characteristics

described in corticosteroids, and by the presence or absence of hydroxyls at C-11 and

C-17. Production of glucocorticoids is a classic endocrine response to stress.

Glucocorticoids synthesized in the adrenal cortex stimulate gluconeogenesis to provide

energy for the “flight or fight” response (17). In humans and animals, it has been shown

that excess stress or excess exogenous glucocorticoids cause hyperglycemia,

hypertension, and increased anxiety-related behavior (18). Glucocorticoids act by

binding to glucocorticoid receptors, which subsequently act as transcription factors to

alter gene expression. Various stressors, including infection, malnutrition, anxiety, and

depression, trigger a rise in glucocorticoids (19, 20).

Cortisol- In tetrapods, the active glucocorticoid can be either cortisol or

corticosterone (16). But rodents, birds, lizard, and other animals only secrete

corticosterone (21). In species that secrete both steroids, cortisol is often regarded as

the main glucocorticoid, while corticosterone is just considered the second

glucocorticoid with similar properties to cortisol (22). But, corticosterone actually has

different properties than cortisol. Cortisol and its metabolites cannot enter

enterohepatic circulation, whereas corticosterone can; meaning that it can be recycled

throughout the body instead of being excreted. Overall though, it is still not known why

humans and many other mammals produce both glucocorticoids. Many more studies

have been done on the physiological effects of cortisol relating to obesity and stress,

(23, 24). The amount of cortisol synthesized daily in humans is listed in Table 1.1.

Mineralocorticoids- Mineralocorticoids are a class of corticosteroids, with a

hydroxyl group at carbon 11 and an aldehyde on carbon 18. The primary active

Page 19: Copyright 2018 Seán Martin Mythen

9

Table 1.1 Concentrations of major steroids found in humans

Concentration Produced Per Day

(nM)

Steroid Hormone Male Female Major Route of Excretion

Cortisol 55.18 - 772.52 Urine Aldosterone 0.0277 - 0.582.5 Urine Testosterone 10.4 - 41.6 0.3 - 2.1 Urine Estradiol 0.0367 - 0.146 0.09912 - 1.6007 Urine Progesterone ≤ 3.18 ≤ 3.18 - 954 Urine

Bile Acid Concentration Measured in Human Feces (µM)

Major Route of Excretion

Cholic Acid 0 - 97 (25) (26) Feces Deoxycholic Acid 30 - 700 (25) (26) Feces Chenodeoxycholic Acid 0 - 94 (25) (26) Feces Lithocholic Acid 1 - 450 (25) (26) Feces Iso Deoxycholic Acid 0 - 390 (26) Feces Iso Lithocholic Acid 0 - 260 (26) Feces

Page 20: Copyright 2018 Seán Martin Mythen

10

mineralocorticoid is aldosterone (16). The amount of aldosterone synthesized daily in

humans is listed in Table 1.1. Mineralocorticoids contribute to electrolyte homeostasis

(4). Uniquely, the mineralocorticoid receptor responds to both aldosterone and cortisol

(corticosterone in rodents), a glucocorticoid (27).

Sex Steroids- The sex steroids are comprised of androgens, estrogens, and

progestogens. Typically, estrogens and progestogens are thought of as the female sex

steroids, and androgens are thought of as the male sex steroids. Sex steroids are

essential for reproductive function but can act on other organ systems as well. Each

sex steroid is produced by both sexes, though the quantity differs by age and sex (28).

Sex steroids are not only produced by the gonads, but by other organs as well,

including the central nervous system (29, 30).

Androgens- The androgens are steroids that contain 19 carbons. Androgens

are characterized by the absence of a carbon side chain on C-17, and by the presence

of a ketone functional group at C-3 (6). Androgens are important steroid hormones that

control normal male development and reproductive function (31). Androgens are

produced in males in the testes and in females by the ovaries and placenta. The most

important androgens are 5α-dihydrotestosterone (DHT) and testosterone. The amount

of testosterone synthesized daily in humans is listed in Table 1.1. The biological actions

of DHT and testosterone are facilitated by androgen receptors, which upon activation,

regulate gene expression in target tissues. Although testosterone and DHT have affinity

for the androgen receptor, they interact differently. Testosterone has a two-fold lower

affinity than DHT, while the dissociation rate of testosterone from the receptor is five-

Page 21: Copyright 2018 Seán Martin Mythen

11

fold faster than DHT (32). Overall, androgens are important for normal development

and function of male reproductive tissues and for development of muscles and bone.

11β-hydroxyandrostenedione- An important androstane synthesized in the

adrenal glands that is often metabolized by steroidogenic enzymes is 11β-

hydroxyandrostenedione (11β-OHAD). The biological role of 11β-OHAD specifically is

still being elucidated, but it is known that 11β-OHAD is a precursor for many recently

recognized novel androgens (33). These androgens bind to and activate the human

androgen receptors similarly to testosterone and DHT (34). This is significant when

considering androgen dependent diseases such as prostate cancer and other diseases

associated with androgen excess such as polycystic ovary syndrome (35). Bacterial

enzymes have been identified that cleave the side chain of cortisol, converting cortisol

into 11β-OHAD in the gut. It is important to further understand 11β-OHAD and the

enzyme responsible for its synthesis from cortisol. The synthesis of 11β-OHAD in the

gut highlights an alternative pathway for the formation of potent androgens, the majority

of which, were classically thought to be synthesized in the adrenal glands.

Estrogens- The estrogens are steroids that contain 18 carbons. They are

produced in females in the ovaries and the placenta, and in the testes in males.

Estrogens are also produced in the adrenal cortex. Estrogens are characterized by a

loss of C-19, the presence of an aromatic A ring, the absence of a carbon side chain on

C-17, and the presence of a hydroxyl group at both C-3 and C-17 (6). One of the

principal regulators of circulating estrogens is the gut microbiome (36). The three major

estrogens produced are, estrone, estradiol, and estriol, the most common of which is

estradiol. The amount of estradiol synthesized daily in humans is listed in Table 1.1.

Page 22: Copyright 2018 Seán Martin Mythen

12

Estradiol is required for breast development as well as body fat and skin changes seen

in females. Pregnancy is characterized in humans by a massive increase in the

production of estradiol (6). Overall, the ovaries are the principle source of estradiol,

which functions as a circulating hormone to act on near or distant target tissues (29).

Progestogens- Progestogens bind and activate the progesterone receptor.

Progesterone, the primary progestogen, is produced in the corpus luteum of the ovaries

and the placenta, as well as the adrenal cortex (37). The amount of progesterone

synthesized daily in humans is listed in Table 1.1. Prevotella intermedius has been

shown to take up progesterone, which is then metabolized to enhance the growth of the

microbe (38). The synthesis of every other steroid hormone starts with the production

of progesterone from cholesterol; progesterone leads to the production of

mineralocorticoids (aldosterone), glucocorticoids (cortisol and corticosterone), other

progestogens, estrogens (estradiol), and androgens (testosterone). Similarly to

estradiol, pregnancy is characterized in humans by a massive increase in the

production of progesterone (6). Progesterone has also been shown to be protective

and preventative of breast cancer (39).

BILE ACIDS

Historically, bile acids were viewed solely as detergents, but it is now widely

recognized that bile acids can regulate various metabolic processes, by acting as

nutrient signaling hormones (40). Bile acids, are synthesized from cholesterol in the

liver (41). From there they are stored in the gallbladder and released upon consumption

of a meal to aid in digestion and the absorption of lipids (6, 42).

Page 23: Copyright 2018 Seán Martin Mythen

13

Bile acid synthesis and enterohepatic circulation- In the liver, bile acids are

conjugated with either taurine or glycine, which increases their hydrophilicity, improving

their detergent properties, and preventing them from penetrating cell membranes.

Individuals that consume a “western diet” have predominantly taurine conjugated bile

acids, while individuals who consume a vegetarian diet have predominantly glycine

conjugated bile acids (43). Bile salts are then secreted into bile and stored in the

gallbladder (41). Upon consumption of a meal, the gallbladder contracts and bile salts

enter the duodenum of the small intestine where they aid in the absorption of fat-soluble

vitamins and cholesterol (8, 40, 44). Upon reaching the distal ileum, more than 95% of

the bile salts are reabsorbed and returned to the liver via portal veins (8, 45). The bile

salts are transported across the sinusoidal membrane and re-secreted into the

gallbladder. This cyclical process is known as enterohepatic circulation (EHC) and

allows for the recycling of bile salts multiple times throughout the day (8, 46). The

process of EHC occurs between four and twelve times per day and helps to maintains a

bile acid pool of about 2 to 4 grams in humans (45).

EHC is about 95% efficient with most bile acids reabsorbed in the terminal ileum,

however, about 400-800 mg of bile salts enter the large intestine per day (26). In the

large intestine, bile salts come into contact with a large community of microbes capable

of metabolizing the bile salts (42). In the gut, the bile salts are deconjugated back into

bile acids by microbial enzymes called bile salt hydrolases (BSH) (47). The bile acids

are then further metabolized by a small group of gut bacteria within the genus

Clostridium via 7α-dehydroxylating enzymes, creating toxic secondary bile acids.

Additionally, microbial HSDHs can create toxic secondary bile acids. Secondary bile

Page 24: Copyright 2018 Seán Martin Mythen

14

acids are then mostly excreted in the feces, but a certain proportion is also absorbed in

the colon and sent back to the liver via EHC. In the liver, bile acids are re-conjugated

and secreted into the biliary pool along with the primary bile acids (48).

Primary bile acids- The major primary bile acids synthesized in humans, are

cholic acid (CA) and chenodeoxycholic acid (CDCA). When produced, they are

conjugated with either taurine or glycine and then referred to as bile salts. Classically,

these bile salts are known to function as detergents that aid in the absorption of

nutrients (49). The ability of bile salts to aid in digestion is due to their amphipathic

structure, characterized by the orientation of the hydroxyl groups toward one side of the

steroid structure and the methyl groups toward the other when viewed edge-on. More

importantly, however, is the recent recognition that the various bile acids have their own

unique set of biological activities which are able to regulate gene expression by serving

as ligands for various nuclear receptors.

Secondary bile acids- Secondary bile acids are derivatives of primary bile acids

resulting from bacterial metabolism in the colon. The major secondary bile acids in

humans are deoxycholic acid (DCA), produced from CA; and lithocholic acid (LCA) and

ursodeoxycholic acid (UDCA), produced from CDCA. There are additional important

secondary bile acids called iso bile acids which are 3β-hydroxylated. In fact, after DCA

and LCA, isoDCA and isoLCA are the most abundant secondary bile acids. For

secondary bile acids to be produced; however, primary bile salts have to first be

deconjugated by an enzyme called bile salt hydrolase (BSH). The primary bile acids

can then be converted into more than twenty different secondary bile acid metabolites

by various gut microbes (45). The enzymes responsible for the multitude of secondary

Page 25: Copyright 2018 Seán Martin Mythen

15

bile acids include those that function to dehydroxylate the 7 position, to epimerize the C-

5 hydrogen, and to oxidize or epimerize hydroxyl groups at position 3, 7, or 12 (50).

Some of these modifications make secondary bile acids more hydrophobic, thereby

increasing their toxicity. Hydrophobic secondary bile acids such as DCA and LCA can

lead to the development of apoptosis resistance and, therefore, colon cancer (42).

Other secondary bile acids appear to bind and activate various host receptors to an

even greater extent than primary bile acids (47). Thus, it is necessary to study these

secondary bile acids if we want to understand their full effect on the host. Overall, of the

main human bile acids, the circulating bile acid pool is comprised of about 35% CA and

CDCA each, about 25% of DCA, and less than 5% of LCA (51). The concentration of

bile acids in the large intestine of humans is listed in Table 1.1.

Iso Bile Acids- Following DCA and LCA, isoDCA and isoLCA are the most

abundant secondary bile acids in humans (26). It is hypothesized that the formation of

iso bile acids is a pathway evolved by gut bacteria to detoxify DCA or LCA (13). DCA

and LCA act as detergents, which can solubilize membrane lipids causing cells to lyse

and die (52). By flipping the C3-hydroxyl stereocenter to form iso bile acids, the

amphipathic nature of the bile acids is altered, decreasing the molecule’s detergent

properties, which in turn decreases toxicity to bacteria (26). It has been shown in

Ruminococcus gnavus that isoDCA results in less membrane damage than DCA,

meaning isoDCA is less toxic (26). The formation of an iso bile acid requires two

concerted enzymatic reactions performed by 3α- and 3β- HSDHs, respectively. In order

for DCA or LCA to be converted into isoDCA or isoLCA, the C3-hydroxyl group, which is

in an α orientation in DCA and LCA, has to first be oxidized by a 3α-HSDH. Then, a 3β-

Page 26: Copyright 2018 Seán Martin Mythen

16

HSDH can reduce the C3-ketone into β oriented hydroxyl. Iso bile acid pathways have

been found in Eggerthella lenta (previously Eubacterium lentum) and R. gnavus.

Bile acid signaling- Many of the genes induced by bile acids encode enzymes

involved in their own regulation, transport, conjugation, and metabolism (53). Bile acids

can function as important regulatory molecules in the small intestine by inducing genes

that are involved in resistance to bacterial growth (54). Bile acids have been found to

have affinity for the Farnesoid X Receptor (FXR) and G-protein coupled receptors such

as TGR5, and the sphingosine phosphate receptor (S1PR2). Further, the bile acid

structure determines its affinity for a certain receptor and therefore extent of agonism or

antagonism.

FXR is a member of the nuclear receptor superfamily and controls multiple

metabolic pathways. FXRs are mostly found in the liver and intestine. Upon activation

of the FXR by bile acids, FXR can regulate bile acid synthesis and conjugation in the

liver and transport, as well as aspects of lipid and glucose metabolism (40, 53, 55)

Activation of TGR5 has been shown to have beneficial effects on body weight as well as

glucose homeostasis (47, 56). Lithocholic acid (LCA) is the most potent natural agonist

of TGR5. Conjugated and unconjugated forms of DCA, CDCA, and CA have 53%,

12%, and 7% potency, respectively, compared to LCA (57). S1PR2 is only activated by

taurine or glycine conjugated bile acids. Activation of S1PR2 may help these

conjugated bile acids control hepatic glucose levels, reduce bile acid levels overall, and

increase lipid metabolism. These bile acid signaling pathways elicit hormone-like

regulating mechanisms, highlighting their importance in the sterolbiome.

Page 27: Copyright 2018 Seán Martin Mythen

17

THE GUT MICROBIOTA

The intestinal tract of human adults contains about 1000 g of bacteria, which

equates to about 1014 organisms from more than 400 distinct species (58). More than

99% of these organisms are obligate anaerobes (58). Adult colons are predominantly

colonized with members of two phyla; the Bacteroidetes, comprising of mainly gram-

negative bacteria, and the Firmicutes, comprising of mainly gram-positive clostridia (59).

The gut microbial community encodes about 150 times more genes than are present in

the human genome (60). Steroids in the gut can undergo drastic transformative

changes caused by the strain specific enzymes certain microbes express. If we are to

understand the impact bacteria and their metabolites have on the host, characterization

of bacteria is needed at the strain level. Both dietary and endogenous steroids come

into contact with the microbial population in the gut (8). The resulting metabolites affect

the overall microbial community and, through passive diffusion, can also circulate

throughout the body and alter the host endocrine system. One of the most important

goals in the field of gut microbiology is to gain a better understanding of microbial

interactions with the host in the hopes of modulating or reengineering the microbiome to

improve health states in the host.

Clostridium scindens

Early studies during the 1950’s revealed that rectal infusion of hydrocortisone

(cortisol), a C-21 steroid, resulted in a 100-fold increase in urinary excretion of C-19 17-

ketosteroids (61, 62). Rectal infusion of cortisol and concomitant oral antibiotic

treatment with neomycin, however, ablated the formation of 17-ketosteroids, suggesting

that gut bacteria were responsible for the biotransformation (62). Twenty years later,

Page 28: Copyright 2018 Seán Martin Mythen

18

similar findings were reported when cortisol was incubated with human and rat fecal

samples, resulting in a notable conversion of cortisol to various other C-19 metabolites

(63, 64). These studies led to the hypothesis that the side chain of cortisol was being

enzymatically removed by intestinal bacteria. It was postulated that intestinal bacteria

encode an enzyme known as steroid-17,20-desmolase, which is responsible for the

conversion of cortisol into the 17-ketosteroid metabolite known as 11β-OHAD. In 1984,

the first gut bacterium that expresses steroid-17,20-desmolase activity was isolated and

was named Clostridium scindens (type strain ATCC 35705), which means “to cut” (65,

66). C. scindens is a gram-positive, spore forming, motile, obligate anaerobe. C.

scindens is a known bile acid 7α-dehydroxylating species. In addition to steroid-17,20-

desmolase activity, C. scindens expressed cortisol-inducible 20α-hydroxysteroid

dehydrogenase (20α-HSDH) activity. The two enzymes were originally characterized

from whole cell extracts; substrate-specificity and cofactor requirements were reported

(67). Native 20α-HSDH was thereafter purified to apparent electrophoretic homogeneity

(although 2D-gel revealed a major contaminant of identical molecular weight) and

characterized from cell extracts of C. scindens ATCC 35704 (68). Up until recently, the

genes encoding these enzymes and the mechanism by which steroid-17,20-desmolase

proceeded was unclear. Using RNA sequencing, Ridlon et al (2013), identified a

cortisol-inducible operon (desABCD) hypothesized to encode the steroid-17,20-

desmolase and 20α-HSDH (69). Genes involved in the biosynthesis of thiamine

pyrophosphate (TPP) and the expression of pentose-phosphate pathway enzymes were

notably upregulated. The desA and desB genes are annotated as N-terminal and C-

terminal transketolases, respectively, and it is hypothesized that DesA and DesB form a

Page 29: Copyright 2018 Seán Martin Mythen

19

complex that has steroid-17,20-desmolase activity that functions similarly to the

transketolases found in the pentose-phosphate pathway. The desC gene was cloned,

overexpressed, purified, and shown to encode the cortisol 20α-HSDH (69). The desD

gene is hypothesized to encode a cortisol transporter.

Eggerthella lenta

Eggerthella lenta (formerly Eubacterium lentum) is a gram-positive, non-spore

forming, non-motile, obligate anaerobic bacterium whose growth is stimulated by

arginine (70, 71). E. lenta is a particularly important participant in the gut sterolbiome,

various strains of which have the potential to encode enzymes that metabolize cardiac

glycosides (72, 73), 21-dehydroxylate corticosteroids (74), hydrolyze conjugated bile

acids (75), and oxidize bile acid hydroxyl groups (76, 77). E. lenta strains have

previously been grouped according to their pattern of hydroxysteroid dehydrogenase

(HSDH) activities; some strains having no HSDH activity (Group V), with others

expressing 3α-HSDH, 3β-HSDH, 7α-HSDH, and 12α-HSDH (Group III) (77). E. lenta

and other members of the Coriobacteriaceae family have been correlated with changes

in host liver metabolome, specifically through reduction in serum triglycerides,

cholesterol excretion, and metabolism of xenobiotics (78).

Of the mentioned enzymatic activities expressed by E. lenta, 21-dehydroxylation

is of notable importance. 21-dehydroxylation of the glucocorticoid cortisol forms 11β-

hydroxyprogesterone. Metabolites of 21-dehydroxylation have been shown to be potent

inhibitors of host 11β-HSD2 (79, 80). 11β-HSD2 is responsible for the conversion of

cortisol to an inactive form known as cortisone. Inhibition of 11β-HSD2 allows cortisol to

Page 30: Copyright 2018 Seán Martin Mythen

20

accumulate and continuously occupy glucocorticoid and mineralocorticoid receptors,

potentially causing hypertension (81).

The E. lenta sterolbiome offers a potential link between associations with strains

of this species and changes in host liver metabolome and physiology. Alteration of

steroid structure, and absorption of E. lenta steroid metabolites (and derivatives from

additional microbial biotransformations) from the colon and transport to the liver via

portal circulation may alter nuclear receptor and G-protein coupled receptor activation,

as well as affect liver enzyme activities, leading to altered lipid and steroid metabolism.

Currently, the sterolbiome of E. lenta is poorly characterized. Genes encoding HSDHs,

particularly 7α-HSDH and 12α-HSDH, have yet to be identified. It may be expected that

diverse isoforms of currently identified 3α-HSDH and 3β-HSDH may exist, requiring

functional characterization of genes whose protein products are predicted to encode

these enzymes, namely those within the short chain dehydrogenases/reductases (SDR)

family (82). A major barrier to such mechanistic study is the need to identify and

characterize much of the E. lenta sterolbiome.

Butyricicoccus desmolans

B. desmolans is a gram-positive, non-spore forming, non-motile obligate

anaerobic and butyrate-producing bacterium recovered from a healthy human’s feces

(83). Previously, the bacterium was named Eubacterium desmolans (83). Steroid-

17,20-desmoalse activity had been reported in B. desmolans, but unlike C. scindens

ATCC 35704, which expresses cortisol 20α-HSDH activity, B. desmolans was

demonstrated to express cortisol 20β-HSDH activity (84). Soon after the reporting of

the desABCD operon from C. scindens ATCC 35704, the genome sequence for B.

Page 31: Copyright 2018 Seán Martin Mythen

21

desmolans ATCC 43058 became available and the desAB gene in B. desmolans ATCC

43058 was shown to encode steroid-17,20-desmolase (85). The desE gene, not

present in C. scindens ATCC 35704, but present in B. desmolans ATCC 43058 was

shown to encode cortisol 20β-HSDH (85). The 20β-HSDH from B. desmolans 43058

was overexpressed and purified from E. coli. The enzyme was determined to be a

homotetramer with cortisol as the preferred substrate. Phylogenetic analysis of 70

amino acid sequences with at least 30% identity to the gene encoding cortisol 20β-

HSDH from B. desmolans ATCC 43058 was performed. A phylogenetic tree was

constructed with cortisol 20β-HSDH from B. desmolans and closest neighboring

sequences, forming three major clusters (Figure 1.4). The cluster containing the most

distantly related proteins relative to cortisol 20β-HSDH from B. desmolans was

composed exclusively of polypeptides from Mycobacterium spp., followed by a cluster of

proteins from anaerobic gut bacteria that inhabit ruminants and monogastrics. The final

cluster, populated by B. desmolans ATCC 43058 and other steroid-17,20-desmolase

expressing strains, also harbored polypeptides from a number of species of

Bifidobacterium. The desA and desB genes, hypothesized to encode steroid-17,20-

desmolase were not located in the available genomes from Bifidobacterium represented

in the tree. B. scardovii DSM 13734 (ATCC BAA-773) was tested for cortisol

metabolism and after 48 hours of incubation in BHI medium, cortisol was converted to

20β -dihydrocortisol, but did not form 11β- OHAD (85).

Bifidobacterium adolescentis

In 1982, a strain of B. adolescentis was isolated from human feces that

Page 32: Copyright 2018 Seán Martin Mythen

22

Figure 1.4 Phylogenetic analysis of identified 20β-hydroxysteroid dehydrogenases 70 amino acid sequences with at least 30% identity to the gene encoding cortisol 20β-HSDH from B. desmolans ATCC 43058 are aligned in a phylogenetic tree constructed using FastTree v2.1.7. Steroid-17,20-desmolase expressing strains are denoted by a red “x” through a C-C bond. Predicted active-site residues are depicted adjacent to strain designations

Page 33: Copyright 2018 Seán Martin Mythen

23

displayed 20β-HSDH activity (86). Early studies have shown that Bifidobacterium spp.

are among the first bacteria to colonize the gastrointestinal tract of breast-fed infants

and are regularly used as probiotics in products such as milk, yogurt, and dietary

supplements (87). The recent phylogenetic analysis of the 20β-HSDH gene from B.

desmolans ATCC 43058 and B. scardovii DSM 13734 (ATCC BAA-773) resulted in the

identification of several Bifidobacterium strains that are predicted to express 20β-HSDH

activity (85). The fact that some bifidobacteria strains encode and express 20β-HSDH,

but not steroid-17,20-desmolase is interesting. It is hypothesized that 20β-HSDH

activity in bifidobacteria may reduce the conversion of cortisol to 11β-OHAD in guts

dominated by 20α-HSDH expressing bacteria such as C. scindens ATCC 35704.

We identified particular B. adolescentis strains that possess a gene predicted to

encode 20β-HSDH based on the phylogenetic tree with B. desmolans ATCC 43058.

Specifically, B. adolescentis strain L2-32 was selected for further study because this

strain is predicted to encode 20β-HSDH. B. adolescentis is a gram-positive, non-motile,

non-spore forming, obligate anaerobic bacteria. Because of the important pathways

20β-HSDHs mediate in cortisol metabolism and the properties conferred intrinsically by

Bifidobacteria, the 20β-HSDH expressed by B. adolescentis strain L2-32 has the

potential to be of great pharmaceutical interest.

OBJECTIVES

There are a plethora of steroids circulating throughout the human host which are

constantly being modified by microbes in the gut. The overall impact these steroids

have on the surrounding microbial community and the host is still understudied and

predicted to be of significant biomedical consequence. Currently, it is difficult to

Page 34: Copyright 2018 Seán Martin Mythen

24

bioinformatically determine the function of many predicted steroid metabolizing genes,

therefore, it is important that these genes are biochemically characterized so that their

functions can be tested. The desAB genes cluster with the desC (20α-HSDH) and desE

(20β-HSDH) genes in C. scindens and B. desmolans, respectively, and are strong

candidates for genes encoding steroid-17,20-desmolase. By characterizing the desAB

genes and other steroid metabolizing genes predicted to be HSDHs of various stereo

and regiospecificity, we can begin to understand and potentially control the steroid

profile by mediating receptor activation and downstream effects such as metabolism,

immune responses, and inflammation in the host. In an effort to elucidate the

importance of steroid metabolizing genes, we will:

1. Characterize a gene cluster predicted to encode multiple HSDHs in

Eggerthella:CAG 298, a gut metagenomic sequence.

2. Determine whether the desA and desB genes encode steroid-17,20-

desmolase. Develop an enzyme-coupled continuous spectrophotometric

assay to simplify enzyme characterization of DesAB.

3. Characterize 20β-HSDH from Bifidobacterium adolescentis strain L2-32 and

attempt to crystallize and determine the structure.

Page 35: Copyright 2018 Seán Martin Mythen

25

CHAPTER 2: Targeted synthesis and characterization of a gene-cluster encoding

NAD(P)H-dependent 3α-, 3β-, and 12α-hydroxysteroid dehydrogenases from

Eggerthella CAG:298, a gut metagenomic sequence

ABSTRACT

Gut metagenomic sequences provide a rich source of microbial genes, the majority of

which are annotated by homology or unknown. Genes and gene pathways that encode

enzymes catalyzing biotransformation of host bile acids are important to identify in gut

metagenomic sequences due to the importance of bile acids on gut microbiome

structure and host physiology. Hydroxysteroid dehydrogenases (HSDH) are pyridine

nucleotide-dependent enzymes with stereo-specificity and regio-specificity for bile acid

and steroid hydroxyl groups. HSDHs have been identified in several protein families

including medium chain and short chain dehydrogenase/reductase families as well as

aldo-keto reductase family. These protein families are large and contain diverse

functionalities making prediction of HSDH-encoding genes difficult, necessitating

biochemical characterization. We located a gene-cluster in Eggerthella CAG:298

predicted to encode three HSDHs (CDD59473, CDD59474, CDD59475) and

synthesized the genes for heterologous expression in E. coli. We then screened bile

acid substrates against the purified recombinant enzymes. CDD59475 is a novel 12α-

HSDH and we determined that CDD59474 (3α-HSDH) and CDD59473 (3β-HSDH)

constitute novel enzymes in an iso-bile acid pathway. Phylogenetic analysis of these

HSDHs with other gut bacterial HSDHs and closest homologues in the database

revealed predictable clustering of HSDHs by function and identify several likely HSDH

Page 36: Copyright 2018 Seán Martin Mythen

26

sequences from bacteria isolated or sequenced from diverse mammalian and avian gut

samples.

Bacterial HSDHs have the potential to significantly alter the physicochemical

properties of bile acids with implications for increased/decreased toxicity for gut bacteria

and the host. The generation of oxo-bile acids is known to inhibit host enzymes involved

in glucocorticoid metabolism and may alter signaling through nuclear receptors such as

farnesoid X receptor and G-protein coupled receptor TGR-5. Biochemistry or similar

approaches are required to fill in many gaps in our ability to link a particular enzymatic

function with a nucleic acid or amino acid sequence. In this regard, we have identified a

novel 12α-HSDH, and a novel set of genes encoding an iso-bile acid pathway (3α-

HSDH, 3β-HSDH) involved in epimerization and detoxification of harmful secondary bile

acids.

INTRODUCTION

Metagenomic sequencing of microbial DNA in environmental samples, including

human stool, has provided a massive gene catalogue to the microbiologist, only an

estimated 50% of which have been annotated; much of these homology-based, and a

miniscule number of these characterized biochemically (88). The Human Microbiome

Project (89) and Metagenomics of the Human Intestinal Tract (MetaHIT) have provided a

‘second human genome’ to decode (90). One set of functional pathways within this

‘second human genome’ is the gut sterolbiome (2) which participates in the endocrine

function of the host as well as, in the case of bile acids, regulation of microbiome structure

and functions. Gut bacterial metabolism of bile acids is known to affect signaling through

the farnesoid-X-receptor (91), G-coupled protein receptor TGR5 (57), muscarinic receptor

Page 37: Copyright 2018 Seán Martin Mythen

27

(92), and Vitamin-D-receptor (93). In addition, the formation of oxo-bile acids has been

shown to inhibit the enzyme 11β-hydroxysteroid dehydrogenase-1 (11β-HSD1), which

has been reported to alter gut microbiome community structure (94).

The oxidation of bile acid hydroxyl groups by gut bacteria also allows epimerization

by the concerted action of two regio-specific enzymes differing in stereo-specificity such

that the conversion of, for instance, 7α-hydroxylated chenodeoxycholic acid (CDCA) to

7β-hydroxylated ursodeoxycholic acid (UDCA) bile proceeds through a stable 7-

oxolithocholic acid intermediate (95). Epimerization of the 3α-hydroxy results in formation

of so called iso-bile acids (3β-hydroxyl) (96) which are more hydrophilic and less toxic

than the 3α-hydroxyl host-derived bile acid (26, 97), as is also the case with UDCA which

is used therapeutically in hepatobiliary diseases (98). The least studied pathway is the

epimerization of the 12α-hydroxyl 12-oxo 12β-hydroxyl of cholic acid derivatives (99,

100).

Eggerthella lenta (formerly Eubacterium lentum) is a gram-positive, non-spore

forming, non-motile, obligate anaerobic bacterium whose growth is stimulated by arginine

(70, 71). E. lenta is a particularly important participant in the gut sterolbiome, various

strains of which have the potential to encode enzymes that metabolize cardiac glycosides

(72, 73), 21-dehydroxylate corticosteroids (74), hydrolyze conjugated bile acids (75), and

oxidize bile acid hydroxyl groups (76, 77). E. lenta strains have previously been grouped

according to their pattern of hydroxysteroid dehydrogenase (HSDH) activities; some

strains having no HSDH activity (Group V), with others expressing 3α-HSDH, 3β-HSDH,

7α-HSDH, and 12α-HSDH (Group III) (77). E. lenta and other members of the

Coriobacteriaceae family have been correlated with changes in host liver metabolome,

Page 38: Copyright 2018 Seán Martin Mythen

28

specifically through reduction in serum triglycerides, cholesterol excretion, and

metabolism of xenobiotics (78).

The E. lenta sterolbiome offers a potential link between associations with strains

of this species and changes in host liver metabolome and physiology. Alteration of steroid

structure, and absorption of E. lenta steroid metabolites (and derivatives from additional

microbial biotransformations) from the colon and transport to the liver via the portal

circulation may alter nuclear receptor and G-protein coupled receptor activation, as well

as affect liver enzyme activities, leading to altered lipid and steroid metabolism. Currently,

the sterolbiome of E. lenta is poorly characterized. Genes encoding HSDHs, particularly

7α-HSDH and 12α-HSDH, have yet to be identified. It may be expected that diverse

isoforms of currently identified 3α-HSDH and 3β-HSDH may exist, requiring functional

characterization of genes whose protein products are predicted to encode these

enzymes, namely those within the family of short chain dehydrogenases/reductases

(SDR) (82). A major barrier to such mechanistic study is the need to identify and

characterize much of the E. lenta sterolbiome.

In the current study, we queried the non-redundant database with the amino acid

sequence from the only characterized 12α-HSDH to date, identified in Clostridium sp.

strain ATCC 29733/VPI C48-50 (99). We located an open reading frame (ORF) in a gut

metagenome of Eggerthella CAG:298, a sequence from Metagenomics of the Human

Intestinal Tract (MetaHIT), that is part of a gene cluster with additional NAD(P)H-

dependent dehydrogenases in the short-chain dehydrogenase/reductase (SDR) and

aldo-keto reductase protein families. Utilizing targeted gene synthesis, we identified and

cloned these three synthesized genes for heterologous expression in E. coli. We report

Page 39: Copyright 2018 Seán Martin Mythen

29

the characterization of a gene encoding a novel 12α-HSDH as well as two genes

encoding an iso-bile acid pathway (3α-HSDH, 3β-HSDH), distinct from those previously

reported (26).

MATERIALS AND METHODS

Bacterial strains and chemicals- Escherichia coli DH5α and E. coli BL21-

CodonPlus(DE3) RIPL competent cells. The pET-46 Ek/LIC vector kit was obtained

from Novagen (San Diego, CA). The QIAprep Spin Miniprep kit was obtained from

Qiagen (Valencia, CA). Isopropyl β-D-1-thiogalactopyranoside (IPTG) was purchased

from Gold Biotechnology (St. Louis, MO). Bile acid substrates were purchased from

Steraloids, Inc (Newport, RI, USA).

Targeted gene synthesis, cloning, expression and purification recombinant

proteins- Gene fragments (GenBlock®) encoding 3α-HSDH, 3β-HSDH and 12α-HSDH

from E. lenta CAG:298 (Bio Project PRJEB816) were synthesized by Integrated DNA

Technologies (IDT) (Coralville, IA, USA). Gene fragments were amplified with primers

synthesized by IDT (Table 2.1), using the Phusion high fidelity polymerase (Stratagene,

La Jolla, CA) and cloned into pET-46b vector (Novagen), as described by the

manufacturer’s protocol. Cultivation, plasmid isolation and sequence confirmation were

as previously described (85). HSDH enzymes were expressed as previously described

(85). The recombinant proteins were then purified using TALON® Metal Affinity Resin

(Clontech Laboratories, Mountain View, CA) as per manufacturer’s protocol. The

recombinant protein was eluted using an elution buffer composed of 20 mM Tris-HCl, 150

mM NaCl, 20% glycerol, 10mM 2-mercaptoethanol pH 7.9 and 250 mM imidazole. The

protein purity was assessed by sodium dodecyl sulfate-polyacrylamide gel

Page 40: Copyright 2018 Seán Martin Mythen

30

Table 2.1 Oligonucleotides used for amplification and directional cloning

Page 41: Copyright 2018 Seán Martin Mythen

31

electrophoresis (SDS-PAGE), and protein bands were visualized by staining with

Coomassie brilliant blue G-250. Subunit molecular mass was calculated using three

independent SDS-PAGE gels with Biorad Precision Plus ProteinTm Kaleidoscope

Prestained Protein Standards (Bio Rad Laboratories, Inc, Hercules, CA) using the image

processing software ImageJ (https://imagej.nih.gov/ij/index.html). Recombinant protein

concentrations were calculated based on their molecular mass and extinction coefficients

(Table 2.1). Briefly, deduced amino acid sequence was used as input for Expasy’s

ProtParam tool (https://web.expasy.org/protparam/) and subunit mass and extinction

coefficient (mM-1 cm-1) are utilized to determine enzyme concentration (mg ml-1) in a

Nanodrop 2000c with 10 mm pathlength cuvette at 280 nm.

Enzyme Assays- The buffers used to study the pH profiling of each enzyme

contained 150 mM NaCl, 20% glycerol and 10 mM 2-mercaptoethanol and varied as

follows: 50 mM sodium citrate, (pH 4.0–6.0), 50 mM sodium phosphate (pH 6.5–7.5), 50

mM Tris-Cl (pH 8–9), and 50 mM glycine-NaOH (pH 10–11). Linearity of enzyme activity

with respect to time and enzyme concentration was determined aerobically by monitoring

the oxidation/reduction of NAD(P)(H) at 340 nM (ε=6,220 M-1.cm-1) in the presence of bile

acid and steroid substrates. The standard reaction mixtures were 50 mM sodium

phosphate buffer, pH 6.5 (CDD59473 reductive direction) or 50 mM sodium phosphate

buffer, pH 7.0 (CDD59474 in reductive direction, CDD59475 in both oxidative and

reductive direction) or 50 mM Tris-HCl buffer pH 9.5 (CDD59473 oxidative direction) or

pH 10 (CDD59474 (oxidative direction) each containing 150 µM pyridine nucleotide

cofactor and initiated by addition of the enzyme (50 nM). Kinetic parameters were

estimated by fitting the data to the Michaelis-Menten equation by non-linear regression

Page 42: Copyright 2018 Seán Martin Mythen

32

method using the enzyme kinetics module in GraphPad Prism (GraphPad Software, La

Jolla, CA). Substrate-specificity studies were performed with 150 µM pyridine nucleotide,

and 50 µM bile acid substrate at the respective pH optimum for the enzyme and each

reaction direction.

Thin Layer Chromatography (TLC) and Mass Spectrometry (MS) of Bile Acid

Metabolites- The standard reaction mixtures with different pyridine nucleotide were setup

as described above and incubated overnight at room temperature. The reactions were

stopped by adding 100 μL of 1N HCl. The reaction mixtures were extracted by vortexing

with 2x volumes of ethyl acetate for 1-2 mins and then the organic phase was recovered.

The organic phase was evaporated under nitrogen gas. The residue was dissolved in 50

μL ethyl acetate and spotted along with standards on TLC plate (Silica Gel IB2-F Flexible

TLC Sheet 20 x 20 cm with 250 µm analytical layer, J.T. Baker, Avantor Performance

Materials, LLC. PA, USA). A mobile phase consisting of 70:20:2 v/v/v toluene:1,4-

dioxane:acetic acid was used. Bile acid metabolites separated on the TLC plate were

visualized by spraying 10% phosphomolybdic acid in ethanol and heated at 100⁰C for 15

min. The spots corresponding to substrate and product were isolated, extracted, and

underwent MS analysis. LC-MS analysis was run on a Shimadzu LCMS-IT-TOF System

(Shimadzu Corporation, Kyoto, Japan). The mass spectrometer was operated with an

electrospray ionization (ESI) source in negative ionization mode. The nebulizer gas

pressure was set at 150kPa with a source temperature of 200°C and the gas flow at 1.5

L/min. The detector voltage was 1.65kV. High-purity nitrogen gas was used as collision

cell gas. The mass spectrogram data was processed with the LC solution Workstation

software.

Page 43: Copyright 2018 Seán Martin Mythen

33

Phylogenetic reconstruction of bacterial SDR-family and aldo-keto

reductase family protein sequences- Protein sequences were aligned with MUSCLE

using the web service from the EMBL-EBI (http://www.ebi.ac.uk/Tools/msa/muscle/). The

alignment was manually explored to verify the absence of inaccuracies. Phylogeny

reconstruction was performed using both FastTree v2.1 (101, 102) and IQ-TREE, a time-

efficient tool to generate maximum likelihood phylogenies (103). The two methods were

used for comparative purposes. FastTree ran locally using default settings. IQ-TREE ran

locally using the following options: -st AA -m TEST -bb 1000 -alrt 1000 (-st, sequence

type; -m, automatic model selection; -bb, number of bootstrap replicates). The generated

trees were exported in newick format and uploaded to iTOL v3 for visualization (104). The

trees were compared using Phylo.io (105). Sequence information is found in Table 2.2.

RESULTS

Identification of a gene cluster predicted to encode multiple hydroxysteroid

dehydrogenases in a metagenomic sequence- 12α-HSDHs (EC 1.1.1.176) have been

reported in gut bacteria such as Clostridium perfringens (106), C. leptum (107), E. lenta

(76), Clostridium sp. strain ATCC 29733/VPI C48-50 (99). To date, the only reported 12α-

HSDH whose gene sequence is known is that of Clostridium sp. strain ATCC 29733/VPI

C48-50 (108). Previously, a phylogenetic analysis of the 12α-HSDH amino acid sequence

(ERJ00208.1) was performed, suggesting this function may be widespread in gut

microbiota (12); however, biochemical characterization has yet to confirm this. An

updated query of the non-redundant database identified numerous gut microbial phyla,

particularly Firmicutes and Actinobacteria. One particular sequence, CDD59475,

belonging to Eggerthella CAG:298, and not reported previously (12),

Page 44: Copyright 2018 Seán Martin Mythen

34

Table 2.2 Sequences utilized in phylogenetic analysis

Phylogenetic tree identifier

Protein Locus Tag Function Organism Ref

3A_CDD59473 CDD59473 3α-HSDH Eggerthella sp. CAG:298 TS

3A_CDD59474 CDD59474 3α-HSDH Eggerthella sp. CAG:298 TS

12A_CDD59475 CDD59475 12α-HSDH Eggerthella sp. CAG:298 TS

3A_Elen_0690 Elen_0690 3α-HSDH Eggerthella lenta DSM 2243 (26)

3B_Elen_1325 Elen_1325 3b-HSDH Eggerthella lenta DSM 2243 (26)

3B_Elen_0198 Elen_0198 3b-HSDH Eggerthella lenta DSM 2243 (26)

3A_Rumgna_02133 Rumgna_02133 3α-HSDH Ruminococcus gnavus ATCC 29149 (26)

3B_Rumgna_00694 3B_Rumgna_00694 3b-HSDH Ruminococcus gnavus ATCC 29149 (26)

12A_COLAER_RS06615 COLAER_RS06615

putative 12α-HSDH Collinsella aerofaciens ATCC 25986 (12)

12A_COPCOM_01375 COPCOM_01375 putative 12α-HSDH Coprococcus comes ATCC 27758 (12)

7A_Ecoli BAA01384 7α-HSDH Escherichia coli (109)

3A_BAIA2Csi BAIA2_CLOSV 3α-HSDH Clostridium scindens ATCC 35704 (110, 111)

3A_BAIAChyl ACF20977 3α-HSDH Clostridium hylemonae DSM 15053 (112)

3A_BAIAChir EEA86309 3α-HSDH Clostridium hiranonis DSM 13275 (113)

3A_RUMHYD_02185 RUMHYD_02185 putative 3α-HSDH Blautia hydrogenotrophica DSM 10507 (12)

12A_Csp29733 ERJ00208.1 12α-HSDH Clostridium sp. strain ATCC 29733 (99, 108)

3A_DORFOR_00173 DORFOR_00173 putative 3α-HSDH Dorea formicigenerans ATCC 27755 (12)

3A_RUMTOR_01195 RUMTOR_01195 putative 3α-HSDH Ruminococcus torques ATCC 27756 (12)

3A_ RUMOBE_03494 RUMOBE_03494 putative 3α-HSDH Ruminococcus obeum ATCC 29174 (12)

7A_NP_810824Bthet NP_810824 putative 7α-HSDH Bacteroides thetaiotaomicron VPI-5482 (12, 95, 114)

3A_EDT24590.1Cperf EDT24590.1 putative 3α-HSDH Clostridium perfringens B str. ATCC 3626 (106), (115)

7A_AAA53556.1Csord AAA53556.1 7α-HSDH Clostridium sordellii ATCC 9714 (116)

7A_CsciAAB61151.1 AAB61151.1 7α-HSDH Clostridium scindens VPI 12708 (117)

Page 45: Copyright 2018 Seán Martin Mythen

35

Table 2.2 Continued

Phylogenetic tree identifier Protein Locus Tag Function Organism Ref

7A_CLOHIR_RS06190 CLOHIR_RS06190 putative 7α-HSDH

Clostridium hiranonis DSM 13275 (12)

7A_ BACOVA_01473

BACOVA_01473 putative 7α-

HSDH Bacteroides ovatus ATCC 8483 (12)

7A_BF9343_RS16010 7A_BF9343_RS16010

putative 7α-HSDH Bacteroides fragilis NCTC 9343 (12)

7A_BACFIN_RS13120 BACFIN_RS13120 putative 7α-HSDH

Bacteroides finegoldii DSM 17565 (12)

7A_ANACAC_03661 7A_ANACAC_03661 putative 7α-HSDH

Anaerostipes caccae DSM 14662 (12)

7A_CD630_00650 CD630_00650 putative 7α-HSDH Clostridium difficile 630 (12)

7B_COLAER_RS09325 COLAER_RS09325 7b-HSDH Collinsella aerofaciens ATCC 25986 (118), (119)

7B_RUMGNA_RS11340 RUMGNA_RS11340 7b-HSDH Ruminococcus gnavus ATCC 29149 (120)

12A_CLOHYLEM_04236

12A_CLOHYLEM_04236 12α-HSDH Clostridium hylemonae DSM

15053 (12), unpublished

data

12A_CLOSCI_02455 CLOSCI_02455 putative 12α-HSDH

Clostridium scindens ATCC 35704

(12), unpublished data

12A_COLAER_01483 COLAER_01483 putative 12α-HSDH

Collinsella aerofaciens ATCC 25986 (12)

20B_Bdes WP_051643274.1 20b-HSDH Butyricicoccus demolans ATCC 43058 (85)

20B_Ccad1 WP_051196374.1 20b-HSDH Clostridium cadaveris AGR2141 (85)

20B_Ccad2” WP_027640050 20b-HSDH Clostridium cadaveris AGR2141 (85)

20B_BScar BAQ31198.1 20b-HSDH Bifidobacterium scardovii DSM 13734 (85)

20B_Plym WP_024111275.1 20b-HSDH Propionimicrobium sp. BV2F7 (85)

WP_025022266.1 WP_025022266.1 aldo/keto reductase Lactobacillus hayakitensis TS

KRM19149.1 KRM19149.1 oxidoreductase Lactobacillus hayakitensis DSM 18933 TS

KRM65245.1 KRM65245.1 oxidoreductase Lactobacillus agilis DSM 20509 TS

WP_050611824.1 WP_050611824.1 aldo/keto reductase Lactobacillus agilis TS

WP_019206370.1 WP_019206370.1 aldo/keto reductase Lactobacillus ingluviei TS

Page 46: Copyright 2018 Seán Martin Mythen

36

Table 2.2 Continued

TS=This Study

Phylogenetic tree identifier

Protein Locus Tag Function Organism Ref

KRL88436.1 KRL88436.1 organophosphate reductase Lactobacillus ingluviei DSM 15946 TS

ASN60792.1 ASN60792.1 aldo/keto reductase Lactobacillus curvatus TS

WP_085844664.1 WP_085844664.1 aldo/keto reductase Lactobacillus curvatus TS

WP_054778020.1 WP_054778020.1 aldo/keto reductase Lactobacillus saniviri TS

BAN77682.1 BAN77682.1 conserved hypothetical protein Adlercreutzia equolifaciens DSM 19450 TS

WP_041715350.1 WP_041715350.1 SDR family oxidoreductase Adlercreutzia equolifaciens TS

CDD77593.1 CDD77593.1 putative uncharacterized protein Cryptobacterium sp. CAG:338 TS

CDD68244.1 CDD68244.1 putative uncharacterized protein Eggerthella sp. CAG:368 TS

CDB33831.1 CDB33831.1 putative uncharacterized protein Eggerthella sp. CAG:209 TS

CCY06514.1 CCY06514.1 putative uncharacterized protein Eggerthella sp. CAG:1427 TS

CDD56334.1 CDD56334.1 putative uncharacterized protein

Bacteroides pectinophilus CAG:437 TS

WP_044925038.1 WP_044925038.1 short-chain dehydrogenase Anaerostipes hadrus TS

WP_008116638.1 WP_008116638.1 NAD(P)-dependent oxidoreductase Bacteroides pectinophilus TS

WP_077326212.1 WP_077326212.1 short-chain dehydrogenase Anaerostipes hadrus TS

OKZ9660.1 OKZ9660.1 short-chain dehydrogenase Clostridiales bacterium Nov_37_41 TS

WP_055161728.1 WP_055161728.1 short-chain dehydrogenase Anaerostipes hadrus TS

WP_055258939.1 WP_055258939.1 short-chain dehydrogenase Anaerostipes hadrus TS

CCX88977.1 CCX88977.1 dehydrogenases with different specificities Clostridium sp. CAG:590 TS

WP_009265642.1 WP_009265642.1 short-chain dehydrogenase Lachnospiraceae bacterium 5_1_63FAA TS

Page 47: Copyright 2018 Seán Martin Mythen

37

shared 71% identity (E value = 1e-134) with ERJ00208.1. This predicted 266 amino acid

protein in the 3-ketoacyl-(acyl-carrier protein) reductase/SDR family was unique among

gene products that populated the BLAST query in that it is part of a three-gene cluster

containing two additional putative NAD(P)(H)-dependent oxidoreductases. Directly

upstream is a gene (BN592_00769) encoding a 250 amino acid protein (CDD59474.1)

annotated as a “7α-HSDH” by homology in the SDR-family, and upstream

(BN592_00768) a 280 amino acid protein (CDD59473.1) predicted to encode aldo/keto

reductase. Downstream of BN592_00768 is a gene (BN592_00767) predicted to encode

a gamma-glutamylcysteine synthetase, and upstream on the opposite strand of the gene-

cluster is an ORF (BN592_00771) encoding a hypothetical protein in the Phasin

superfamily involved in intracellular storage of polyhydroxyalkanoates (Figure 2.1).

Targeted gene synthesis, cloning, and overexpression of CDD59475 in E.

coli- Since Eggerthella CAG:298 is a metagenomic sequence and a microbial isolate is

not available, the genes were synthesized (GenBlock® IDT, Inc) and PCR amplified for

ligation-independent cloning (LIC) into pET46(+) for overexpression of recombinant N-

terminal His-tagged protein in E. coli (Figure 2.1). Purified recombinant CDD59475

separated as a single band on SDS-PAGE with a subunit molecular mass of 28.0 ± 0.6

kDa (theoretical molecular weight of 28.30 kDa) (Figure 2.1). We initially screened

CDD59475 dehydrogenase activity spectrophotometrically at 340 nm in the presence of

oxo-bile acids at a concentration of 50 µM, including 3-oxoDCA, 7-oxoDCA, and 12-

oxoDCA with 150 µM NADH or NADPH cofactor. We did not detect enzyme activity when

3-oxoDCA or 7-oxoDCA were substrates; however, CDD59475 clearly displayed

NAD(P)H-dependent activity in the presence of 12-oxoDCA (Figure 2.2C). To initially

Page 48: Copyright 2018 Seán Martin Mythen

38

Figure 2.1 SDS-PAGE of purified recombinant 3β-, 3α-, and 12α-hydroxysteroid dehydrogenases from a gene cluster in Eggerthella CAG:298

3β-HSDH

3α-HSDH

12α-HSDH

Page 49: Copyright 2018 Seán Martin Mythen

39

Figure 2.2 Representative TLC of bile acid reaction products from recombinant CDD59473 (A), CDD59474 (B), and CDD59475 (C) Standard reactions contained 50 μM bile acid substrates, 150 μM pyridine nucleotide, and 50 μM enzyme. Experiments also include a no enzyme control. Reactions were incubated at 37°C for 12 hours. Reaction products that co-migrated with authentic standards were scraped from the TLC plate, extracted with ethyl acetate and dried, and subjected to MS analysis after re-suspension in mobile phase. Experiments were repeated three times.

A.

B.

Page 50: Copyright 2018 Seán Martin Mythen

40

Figure 2.2 Continued

C.

Page 51: Copyright 2018 Seán Martin Mythen

41

confirm activity, we ran several substrates in the oxidative direction and separated the

reaction products by thin layer chromatography, the products of which were subjected to

mass spectrometry (Figure 2.3). In the presence of NADP+, CDD59475 was capable of

oxidizing cholic acid (CA; 3α,7α,12α-trihydroxy-5β-cholan-24-oic acid) and deoxycholic

acid (DCA; 3α,12α-dihydroxy-5β-cholan-24-oic acid), but not CDCA (3α,7α-dihydroxy-

5β-cholan-24-oic acid). Thin layer chromatography (TLC) of overnight reactions

confirmed the formation of a product, co-migrating with authentic DCA standard, from 12-

oxoDCA in the presence of NAD(P)H (Figure 2.2). The reaction product was scraped from

the TLC plate and mass spectrometry (MS) identified a major mass ion in negative mode

at 391.2831 m/z (DCA molecular weight is 392.57 amu) (Figure 2.3B).

Next, we optimized pH in both the oxidative and reductive directions. In the

oxidative direction, with DCA as a substrate and NADP+ as a co-factor, the optimum

enzyme activity was observed at pH 7.0. Enzyme activity declined sharply from pH 7.0 to

pH 8.0 (< 10% relative activity) and ~50% residual activity remained at pH 6.0 (Figure

2.4C). In the reductive direction, with 12-oxoDCA as a substrate and NADPH as a co-

factor, the optimum was also determined to be pH 7.0, with a similar sharp decline to <5%

residual activity at pH 8.0, but ~80% maximal activity between pH 6.5-6.0. We therefore

chose pH 7.0 to run steady-state kinetics in both directions.

Kinetic parameters are listed in Table 2.3; substrate-saturation curves and

Lineweaver-Burke plots are displayed in Figure 2.5. Recombinant CDD59475 had an

order of magnitude higher KM for DCA (139.71 ± 14.64 μM) relative to 12-oxoDCA (11.41

± 0.68 μM). The native 12α-HSDH characterized from Clostridium sp. strain ATCC

29733/VPI C48-50 had a comparable KM value for DCA in the presence of

Page 52: Copyright 2018 Seán Martin Mythen

42

Figure 2.3 Mass spectrometry of enzymatic reaction products A. DCA standard, scraped from TLC plate after resolution in solvent system 75:20:2 toluene:dioxane:glacial acetic acid (v/v/v). B. MS analysis of reaction product co-migrating with iso-DCA, catalyzed by rCDD59473 in the presence of NADPH and 3-oxoDCA (see Figure 2.2A) C. MS analysis of reaction product co-migrating with 3-oxoDCA, catalyzed by rCDD59474 in the presence of NADP+ and DCA (see Figure 2.2B). D. MS analysis of reaction product co-migrating with DCA, catalyzed by rCDD59475 in the presence of NADPH and 12-oxoDCA (see Figure 2.2C).

A.

B.

Page 53: Copyright 2018 Seán Martin Mythen

43

Figure 2.3 Continued

C.

D.

Page 54: Copyright 2018 Seán Martin Mythen

44

Figure 2.4 pH optima for purified recombinant CDD59473 (A), CDD59474 (B), and CDD59475 (C) in the oxidative and reductive directions. CDD59473 was tested with 50 μM 3-oxoDCA (reductive) and isoDCA (oxidative). CDD59474 was tested with 50 μM 3-oxoDCA (reductive) and DCA (oxidative). CDD59475 was tested with 50 μM 12-oxoDCA (reductive) and DCA (oxidative). See Materials and methods for buffer compositions. Experiments were repeated three times and represented as ± standard error of the mean (S.E.M.).

A.

B.

C.

Page 55: Copyright 2018 Seán Martin Mythen

45

Tabl

e 2.

3 Ki

netic

par

amet

ers

of re

com

bina

nt C

DD

5947

3, C

DD

5947

4, a

nd C

DD

5947

5

Page 56: Copyright 2018 Seán Martin Mythen

46

Figure 2.5 Kinetic analysis of recombinant CDD59473, CDD59474, and CDD59475 from metagenomic sequence of Eggerthella CAG:298 Michaelis-Menten and Lineweaver-Burk plot of initial velocity with varying concentrations of substrate in the oxidative (left panels) and reductive (right panels) directions. The oxidation/reduction of NAD(P)(H) was measured by continuous spectrophotometry at 340 nm for 5 min and the initial velocities were used to calculate kinetic constants (Table 2.3). Each data-point represents ± S.E.M. from four independent experiments.

Page 57: Copyright 2018 Seán Martin Mythen

47

NADP+ (12 μM). The catalytic efficiency (kcat/KM) for CDD59475 in the reductive direction

was nearly an order of magnitude greater than in the oxidative, but turnover number (kcat)

was on the same order in both directions.

In the oxidative direction, CDD59475 recognized substrates regardless of side-

chain conjugation to glycine or taurine or presence/absence of 7α-hydroxyl group, which

was reported for the native 12α-HSDH from Clostridium sp. strain ATCC 29733/VPI C48-

50. Interestingly, in the reductive direction, the enzyme only displayed 4.06% activity for

12-oxoCA relative to 12-oxoDCA, suggesting steric hindrance from the 7α-hydroxyl group

(Table 2.4).

Characterization of recombinant CDD59473 and CDD59474 purified from E.

coli- We then overexpressed and purified recombinant CDD59473 (34.2 ± 0.4 kDa;

theoretical 31.59 kDa) and CDD59474 (28.0 ± 0.6 kDa; theoretical 26.40 kDa) in E. coli

(Figure 2.1) and tested these enzymes against a panel of bile acids and their oxo-

derivatives. Separation of CDD59473 and CDD59474 reaction products by TLC revealed

formation of iso-DCA and DCA, respectively, when 3-oxoDCA was the substrate (Figures

2.2A and 2.2B). The CDD59473 reaction product co-migrated with authentic iso-DCA

(392.57 amu) with major mass ion of 391.2849 m/z in negative ion mode (Figure 2.3C),

consistent with iso-DCA formation. The reaction product from CDD59474 co-migrated

with DCA and product in the oxidative direction co-migrated with 3-oxoDCA (390.56 amu)

with major mass ion of 389.2688 m/z in negative ion mode (Figure 2.3D), consistent with

formation of 3-oxo-DCA. A pH optimum for CDD59474 in the oxidative direction between

9 and 10 was observed, and optimum in the reductive direction was pH 7.0. With

NADP(H) as co-substrate,

Page 58: Copyright 2018 Seán Martin Mythen

48

Table 2.4 Substrate-specificity of Eggerthella CAG:298 HSDHs

Page 59: Copyright 2018 Seán Martin Mythen

49

CDD59474 favored the reductive direction, but mass action should favor the oxidative

direction in the anaerobic environment of the gut (Table 2.3; Figure 2.3).

E. lenta DSM 2243T was shown to express two 3β-HSDHs (Elen_0198,

Elen_1325) in the SDR family, sharing 38.1% amino acid sequence identity, involved in

iso-DCA formation (26). By contrast, the aldo/keto reductase family member, CDD59473,

shares only 20% amino acid sequence identity with the 3β-HSDHs Elen_0198 and

Elen_1325. The pH optimum of CDD59473 in the oxidative direction was between pH 9-

10, as observed in CDD59474, with optimum at pH 6.5 in the reductive direction (Figure

2.4). The 3β-HSDH activity of CDD59473 displayed a KM ~2.5 fold higher, a Vmax an order

of magnitude lower, and kcat four orders of magnitude lower than CDD59474 (3α-HSDH)

when 3-oxoDCA was substrate. CDD59473 shares with Elen_0198 and Elen_1325 the

ability to form small quantities of DCA (Figure 2.2), and thus low epimerase activity.

Previous work (77) and unpublished data in our laboratory with cultures of E. lenta DSM

2243T demonstrate that iso-CDCA or iso-DCA are minor metabolites, and that the mono-

keto and diketo intermediates predominate.

The high KM and low turnover number and catalytic efficiency of CDD59473 in the

reductive direction (Table 2.2), which partially agrees with values previously reported for

E. lenta DSM 2243T 3β-HSDH, may further explain why iso-CDCA does not predominate

in culture (26). Taken together, this three-gene cluster accounts for three of the four

HSDH activities reported in E. lenta strains (26, 76, 77) (Figure 2.6).

Phylogeny of Eggerthella HSDH enzymes- Next, the phylogenetic relationship

of these novel Eggerthella HSDH enzymes to other E. lenta SDR family proteins, and

bacterial HSDH proteins was determined (Figure 2.7; Table 2.2). The 3β-HSDH,

Page 60: Copyright 2018 Seán Martin Mythen

50

Figure 2.6 Schematic representation of bile acid metabolism by Eggerthella lenta. Accession numbers of proteins we identified in Eggerthella CAG:298 and the reactions they catalyze.

Page 61: Copyright 2018 Seán Martin Mythen

51

Figure 2.7 iTOl representation of FastTree maximum likelihood phylogeny from MUSCLE alignment of bacterial hydroxysteroid dehydrogenases and the placement of CDD59473, CDD59474, and CDD59475. Phylogeny is the result of 100 bootstraps compared via Phylo.io. Clusters are shaded according to function (see Key). See Table 2.1 for additional sequence information.

Page 62: Copyright 2018 Seán Martin Mythen

52

CDD59473, had by far the longest branch length of any cluster in the tree, distinct from

other 3β-HSDHs characterized in E. lenta DSM 2243T and Ruminococcus gnavus ATCC

29149T (26). Proteins clustering with CDD59473 include predicted aldo/keto reductases

and SDR family members. A BLAST search of CDD59473 against the non-redundant

database revealed amino acid sequences (61-59% identity) that cluster with CDD59473

encoded by Lactobacillus sp. isolated from the gut of pigeon, such as L. agilis

(KRM65245.1; WP_050611824.1) and L. ingluviei (WP_019206370.1; KRL88436.1)

(121, 122), L. hayakitensis (WP_025022266.1; KRM19149.1) from healthy equine (123),

L. saniviri isolated form human feces (124), and L. curvatus isolated from kimchi

(WP_089557116.1; WP_085844664.1) (125) (Figure 2.7).

CDD59474 is located within a cluster of confirmed and predicted 3α-HSDHs and

shares a node and 65.46% amino acid sequence identity with a reported 3α-HSDH

(Rumgna_02133) (26). An enzyme EDT24590.1 from C. perfringens clusters closely with

Rumgna_02133, suggesting this may be a 3α-HSDH whose activity was previously

reported (106). Clustering closely with CDD59474 are uncharacterized, potential 3α-

HSDH candidates from species including the equol-utilizing bacterium Adlercreutzia

equolifaciens (BAN77692.1) which shares much of its protein-coding genes with E. lenta

strains (126), the “Eubacterium-like” arginine-utilizer Cryptobacterium (CDD77593.1)

involved in periodontal disease (127). Also represented in this cluster are the BaiA

enzymes from Clostridium scindens, C. hylemonae, and C. hiranonis involved in the multi-

enzyme bile acid 7α-dehydroxylation pathway responsible for conversion of CA to DCA

and CDCA to lithocholic acid (LCA). BaiA enzymes are NAD(H)-dependent SDR-family

enzymes with strict specificity for bile acid coenzyme A thioesters, which is unique among

Page 63: Copyright 2018 Seán Martin Mythen

53

bacterial HSDHs (110, 111). Our phylogeny also revealed distinct clustering of recently

reported 20β-HSDHs from gut and urinary tract isolates (85).

Predicted 12α-HSDHs (12) also formed a separate branch which includes the 12α-

HSDH, CDD59475, and the characterized 12α-HSDH from Clostridium sp. ATCC 29733

(99), (108) (Figure 2.7). We have recently confirmed 12α-HSDH activity in whole cells

and pure CLOHYLEM_04236 and CLOSCI_02455 from C. hylemonae DSM 15053T and

C. scindens ATCC 35704T, respectively (Doden H, Sallam SA, Devendran S, Ly L, Doden

G, Mythen SM, Daniel SL, Alves JMP, Ridlon JM, unpublished data). Collinsella

aerofaciens is reported to encode 7β-HSDH (COLAER_RS09325) (120) but is also

predicted to encode a 12α-HSDH (COLAER_RS06615) based on homology with

CDD59475. We also identified a putative 12α-HSDH in a strain of the butyrate-producing

species Anaerostipes hadrus, an organism reported to exacerbate colitis in DSS-treated

mice (128). Human gut metagenomic sequences from Clostridiales bacterium

Nov_37_41 (OKZ96620.1), Lachnospiraceae bacterium 5_1_63FAA

(WP_009265642.1), and Clostridium sp. CAG:590 (CCX88977.1) are also predicted to

encode 12α-HSDH. Interestingly, while the majority of sequences were represented by

Firmicutes and Actinobacteria, a human gut metagenomic sequence from a pectin-

degrading bacterium, Bacteroides pectinophilus CAG:437 (CDD56334.1) was also

represented.

Previously characterized 7α-HSDHs formed two distinct clusters in our

phylogenetic analysis: one containing sequences from Firmicutes (gram-positive), the

other Bacteroidetes and E. coli (gram-negative); both clusters sharing a common node.

Page 64: Copyright 2018 Seán Martin Mythen

54

While 7α-HSDH activity has been reported in numerous E. lenta strains, a gene encoding

7α-HSDH has yet to be reported in E. lenta (76, 77).

DISCUSSION

In the current study, we utilized targeted gene synthesis to clone and

heterologously express genes from a gene-cluster (BN592_00768-BN592_00770) in

Eggerthella CAG:298, hypothesized to encode multiple HSDHs (CDD59473, CDD59474,

CDD59475) (Figure 2.1; Table 2.2). Characterization of the purified recombinant enzymes

revealed that each enzyme catalyzed a distinct oxidation/reduction of bile acid hydroxyl

groups. CDD59473 catalyzed NADPH-dependent reduction of 3-oxoDCA yielding iso-

DCA (3β-hydroxyl) as the main reaction product, co-migrating on TLC and with mass

spectrum consistent with iso-DCA (Figures 2.2 and 2.3). CDD59473 is a member of the

aldo-keto reductase family, and is distantly related to previously characterized 3β-HSDHs

from E. lenta DSM 2243T (Elen_0198, Elen_1325) and R. gnavus (Rumgna_00694), all

of which are SDR-family enzymes (Figure 2.7). Interestingly, despite the divergence in

sequence, the epimerase activity of CDD59473 (Figure 2.2; Table 2.2) is consistent with

gut microbial 3β-HSDHs characterized previously (26).

The host generates α-hydroxyl bile acids at C3, C7, and in half of the bile acid pool

also at C12 (cholic acid). Each α-hydroxyl group and the carboxyl group (C24) is on the

same face of the ABCD ring system resulting in a polar, hydrophilic side (Figure 2.6).

Below the ABCD ring system is the hydrophobic face which interact with lipids and

cholesterol packaged into mixed micelles. The isomerization of the 3α-hydroxyl to 3β-

hydroxyl decreases the amphipathic nature of bile acids making them less effective

detergents, and also less toxic to bacteria as well as host cells (26), (128, 129). The high

Page 65: Copyright 2018 Seán Martin Mythen

55

KM value (Table 2.3) suggests that this enzyme functions at high concentrations of toxic

secondary bile acids. The low turnover number (kcat) and catalytic efficiency (kcat/KM) of

CDD59473 vs. SDR-family enzymes (Elen_0198, Elen_1325, Rumgna_00694) provide a

contrast that structural biology might illuminate in an attempt to optimize the reductase

direction of 3β-HSDH and also to establish the basis for the stereo-specificity of the

reaction that yields low 3α-HSDH (epimerase) activity. Currently, there are no structures

reported for gut bacterial 3β-HSDH. Our phylogenetic analysis identified multiple

Lactobacillus spp. isolated from diverse avian and mammalian GI tracts encoding putative

3β-HSDHs with sequence identities at or below 71% (Figure 2.7). A similar targeted gene-

synthesis approach would be useful in determining the function and catalytic potential of

these putative 3β-HSDHs. Isomerization at C3 of primary bile acids may be important in

precluding bile acid 7α-dehydroxylation, a microbial biochemical pathway responsible for

formation of DCA and LCA, bile acids that are causally associated with cancers of the

colon and liver (2).

In contrast to CDD59473, CDD59474 converted 3-oxoDCA to DCA and thus

BN592_00769 encodes 3α-HSDH (Figure 2.6). DCA is generated by the multi-step bile

acid 7α-dehydroxylation of CA by a few species of intestinal clostridia (2) and is found in

fecal water at ~50 µM (130) but can reach several hundred micromolar (25). With KM

values for CDD59474 at ~150 µM and specificity for DCA, these results indicate that at

high DCA concentrations, CDD59473 and CDD59474 work in concert to epimerize DCA

to iso-DCA, potentially as a means of detoxification in E. lenta. Interestingly, CDD59474

was clustered with 3α-HSDH from R. gnavus (Rumgna_02133) but more distantly related

to a 3α-HSDH characterized in E. lenta DSM 2243T (Elen_0690) (Figure 2.7) (26). E.

Page 66: Copyright 2018 Seán Martin Mythen

56

lenta has been reported to express 7α-HSDH (76, 77); however, genes encoding 7α-

HSDH in E. lenta have yet to be identified. CDD59474 was annotated as “7α-HSDH” due

to sequence homology, yet expression and characterization of this protein revealed

specificity for 3α-hydroxyl groups providing a cautionary tale and further reinforcing the

need to functionally characterize genes predicted to encode particular functions by

annotation.

To our knowledge, BN592_00770 is the only reported sequence so far

demonstrated to encode 12α-HSDH (CDD59475) in E. lenta (Figures 2.2 and 2.6; Tables

2.3 and 2.4), and sequence comparison with CDD59475 is expected to elucidate further

gut bacterial 12α-HSDHs (Figure 2.7). CDD59475 shares with 12α-HSDH from

Clostridium sp. strain ATCC 29733/VPI C48-50 and the 3α-HSDH CDD59474 a

conserved N-terminal pyridine nucleotide binding domain (GX3GXG) and active-site

catalytic triad (SYK) (Figure 2.8). Particular bile acid structures trigger germination of C.

difficile spores. Recent work on a germinant receptor (CspC) in C. difficile determined that

bile acid specificity was dependent on 12α-hydroxylation (131). Binding to this receptor

led to the release of Ca2+ dipicolinic acid from the inside of the spore and subsequent

influx of water, ultimately leading to growth into a vegetative cell (131). Gut bacterial 12α-

HSDH activity may therefore affect C. difficile germination.

Thus, our analysis demonstrates that E. lenta strains harbor distinct forms of 3α-

HSDH, 3β-HSDH and 12α-HSDH enzymes (Figures 2.1 & 2.6). Critical at the intersection

between next-generation sequencing data (metagenomics, metatranscriptomics) and

metabolomics is biochemistry, providing the ability to experimentally demonstrate that a

Page 67: Copyright 2018 Seán Martin Mythen

57

Figure 2.8 CLUSTAL alignment of biochemically characterized gut bacterial hydroxysteroid dehydrogenases Red asterisks (*) represent active-site residues conserved in SDR-family. “7BCsard” AET80684.1 7-beta-hydroxysteroid dehydrogenase Clostridium sardiniense; “7BRgnav” AGN52919.1 7-beta-hydroxysteroid dehydrogenase Ruminococcus gnavus; “7BCaero” CZQ36308.1 unnamed protein product Collinsella aerofaciens; “3AComtest” WP_003078312.1 3-alpha-hydroxysteroid dehydrogenase/carbonyl reductase Comamonas testosteroni; “12A29733” ERJ00208.1 oxidoreductase, short chain dehydrogenase/reductase family protein Clostridium sp. ATCC 29733; “3ACSci” AAC45414.1 3-alpha hydroxysteroid dehydrogenase Clostridium scindens; “7AEcoi” AKF67908.1 7-alpha-hydroxysteroid dehydrogenase, NAD-dependent Escherichia coli; “7ABfrag” AAD49430.2 7-alpha-hydroxysteroid dehydrogenase Bacteroides fragilis; “7ACADSO” 5EPO_A Chain A, The Three-dimensional Structure of Clostridium absonum 7alpha-hydroxysteroid dehydrogenase.

Page 68: Copyright 2018 Seán Martin Mythen

58

Figure 2.8 Continued

Page 69: Copyright 2018 Seán Martin Mythen

59

particular amino acid sequence catalyzes a particular biochemical reaction. A major task

in gut microbiology research today is linking nucleic and amino acid sequences in

databases to function, and determining how functions affect microbiome structure and

host physiology. A major question in the field of bile acid metabolism (132) that we are

currently working to address is why bacteria, such as E. lenta, encode numerous HSDHs

and oxidize bile acids in an anaerobic environment.

Page 70: Copyright 2018 Seán Martin Mythen

60

CHAPTER 3: The desA and desB genes from Clostridium scindens ATCC 35704

encode steroid-17,20-desmolase

ABSTRACT

Clostridium scindens is a gut microbe capable of removing the side-chain of cortisol,

forming 11β-hydroxyandrostenedione. A cortisol-inducible operon (desABCD) was

previously identified in C. scindens ATCC 35704 by RNA-Seq. The desC gene was shown

to encode a cortisol 20α-hydroxysteroid dehydrogenase (20α-HSDH). The desD encodes

a protein annotated as a member of the major facilitator family, predicted to function as a

cortisol transporter. The desA and desB genes are annotated as N-terminal and C-

terminal transketolases, respectively. We hypothesized that the DesAB forms a complex

and has steroid-17,20-desmolase activity. We cloned the desA and desB genes from C.

scindens ATCC 35704 in pETDuet for overexpression in E. coli. The purified recombinant

DesAB was determined to be a 142 ± 5.4 kDa heterotetramer. We developed an enzyme-

linked continuous spectrophotometric assay to quantify steroid-17,20-desmolase. This

was achieved by coupling DesAB-dependent formation of 11β-hydroxyandrostenedione

with the NADPH-dependent reduction of the steroid 17-keto group by a recombinant 17β-

HSDH from the filamentous fungus, Cochliobolus lunatus. The pH optimum for the

coupled assay was 7.0 and kinetic constants using cortisol as substrate were KM of 4.96

± 0.57 µM and kcat of 0.87 ± 0.076 min-1. Substrate-specificity studies revealed that

rDesAB recognized substrates regardless of 11β-hydroxylation, but had an absolute

requirement for 17,21-dihydroxy 20-ketosteroids.

Page 71: Copyright 2018 Seán Martin Mythen

61

INTRODUCTION

Clinical studies in the 1950s reported urinary excretion of 17-ketosteroids following

rectal infusion of cortisol for the treatment of ulcerative colitis. Concomitant neomycin

treatment precluded urinary 17-ketosteroid excretion, implicating gut bacteria in the

conversion of a C21 glucocorticoid to a C19 androstane/androgenic metabolite (61), (62).

Later, it was shown that rat and human fecal bacteria were capable of this cortisol side-

chain cleavage reaction, which became known as steroid-17,20-desmolase (63), (64).

Serial dilutions of human and rat fecal samples revealed that ~ 106 bacterial gram-1 of

wet-weight express steroid-17,20-desmolase activity (65). Subsequently, an organism

was isolated in pure culture that expressed steroid-17,20-desmolase and 20α-

hydroxysteroid dehydrogenase (20α-HSDH) activities, and was named Clostridium

scindens, whose species epithet means “to cut”, a reference to separation of the side-

chain from the steroid D-ring (66).

Cortisol-inducible steroid-17,20-desmolase was partially purified, and 20α-HSDH

was purified to apparent electrophoretic homogeneity (SDS-PAGE) by traditional

chromatographic separation (67), (68). The application of genome-wide transcriptomics

(RNA-Seq) identified a polycistronic cortisol-inducible operon (desABCD) which included

a gene encoding 20α-HSDH (desC) (69). The desA and desB genes are annotated as N-

terminal and C-terminal transketolases, respectively. Transketolase catalyzes the TPP-

dependent transfer of glycol aldehyde to an aldose acceptor. Comparison of the

transketolation reaction to that of steroid-17,20-desmolase suggests an analogous

reaction, and that desAB encodes a novel steroid transketolase (69). C. cadavaris and

B. desmolans were shown previously to generate 20β-dihydrocortisol as well as 11β-

Page 72: Copyright 2018 Seán Martin Mythen

62

hydroxyandrostenedione (84). When the genomes of C. cadavaris and B. desmolans

were sequenced and reported, the desAB genes were located via BLAST search, and

flanked a gene encoding a short-chain dehydrogenase (SDR) family protein. We

overexpressed the SDR gene from B. desmolans in E. coli and showed that the enzyme

is a pyridine nucleotide-dependent 20β-HSDH (85). The clustering of genes encoding

cortisol 20α-HSDH or 20β-HSDH with desAB in different gut microbial species suggests

that desAB encodes steroid-17,20-desmolase; however, this function has not yet been

demonstrated. Here, we report that the desAB genes encode steroid-17,20-desmolase.

In addition, we developed a novel, enzyme-coupled continuous spectrophotometric assay

to quantify steroid-17,20-desmolase activity.

MATERIALS AND METHODS

Bacterial strains and chemicals- Escherichia coli DH5α (turbo) competent cells

from New England Biolab (NEB) (Ipswich, MA) was used for cloning, and E. coli BL21-

CodonPlus (DE3) RIPL was purchased from Stratagene (La Jolla, CA) and used for

protein overexpression. The pETDuet and pET-51b(+) vectors were obtained from

Novagen (San Diego, CA). Restriction enzymes were purchased from NEB, QIAprep Spin

Miniprep kit was obtained from Qiagen (Valencia, CA). Isopropyl β-D-1-

thiogalactopyranoside (IPTG) was purchased from Gold Biotechnology (St. Louis, MO).

Streptactin resin was purchased from IBA GmbH, (Goettingen, Germany). Steroids were

purchased from Steraloids (Newport, RI). Amicon Ultra-15 centrifugal filter units with 30-

kDa- and 50-kDa-molecular-mass cutoffs (MMCOs) were obtained from Millipore

(Billerica, MA). All other reagents were of the highest possible purity and were purchased

from Fisher Scientific (Pittsburgh, PA).

Page 73: Copyright 2018 Seán Martin Mythen

63

Cloning of desA and desB from C. scindens and 17β-hydroxysteroid

dehydrogenase (17β-HSDH) from C. lunatus- Genomic DNA (gDNA) was extracted

from C. scindens ATCC 35704 using Fast DNA isolation kit from Mo-Bio (Carlsbad, CA,

USA) according to the manufacturer's protocol. The nucleotide sequence of the coding

region of a cDNA encoding 17β-HSDH from C. lunatus was entered into the Codon

Optimization Tool from Integrated DNA Technologies (IDT, Coralville, IA, USA). The

resulting sequence, whose deduced amino acid sequence was identical to the wild type

17β-HSDH sequence was synthesized as double-strand DNA (dsDNA) by gBlocks(R)

gene fragments service from IDT. C. scindens ATCC 35704 gDNA and gBlocks 17β-

HSDH dsDNA were used as templates to amplify the desAB and 17β-HSDH genes using

oligonucleotide primers reported in Table 3.1 Amplification of desAB and 17β-HSDH

inserts was performed on an Applied Biosystems ProFlex PCR System (Foster City, CA,

USA) using Phusion High-Fidelity DNA Polymerase (Stratagene, La Jolla, CA). Inserts

were gel-purified from agarose and cloned into the pETDuet and pET-51b(+) vectors,

respectively, using appropriate restriction enzymes to digest insert and vector and ligated

with DNA ligase. Recombinant plasmids were transformed into chemically competent E.

coli DH5α cells via the heat shock method, plated, and grown for 16 hours at 37°C on

lysogeny broth (LB) agar plates supplemented with ampicillin (100 µg/mL). A single

colony from each transformation was inoculated into LB medium (5 ml) containing

ampicillin (100 µg/ml) and grown to saturation. The cells were subsequently centrifuged

(3,220 x g, 15 min, 4°C) and plasmids were extracted from the resulting cell pellet using

the QIAprep Spin Miniprep kit (Qiagen, Valencia, CA). The sequences of the inserts were

Page 74: Copyright 2018 Seán Martin Mythen

64

Table 3.1 Oligonucleotides and synthetic gene sequences used in this study

Restriction sites are in bold and italics

Page 75: Copyright 2018 Seán Martin Mythen

65

determined via Sanger sequencing (W. M. Keck Center for Comparative and Functional

Genomics at the University of Illinois at Urbana-Champaign).

Gene expression and purification of recombinant steroid-17,20-desmolase

and 17β-HSDH- For protein expression, plasmids extracted containing correct insert from

the E. coli DH5α cells were transformed into E. coli BL-21 CodonPlus (DE3) RIPL

chemically competent cells via the heat shock method (42°C for 30 s) and grown

overnight at 37°C on LB agar plates supplemented with ampicillin (100 µg/mL) and

chloramphenicol (50 µg/mL). After 16 hours, five isolated colonies were used to inoculate

10 mL of fresh LB medium supplemented with antibiotics and grown at 37°C for 6 hours

with vigorous aeration. The pre-cultures were then added to fresh LB medium (1 L)

supplemented with the same antibiotics at the same concentrations and grown with

vigorous aeration at 37°C. At OD600 of 0.3, IPTG was added to each culture at a final

concentration of 0.1 mM and the temperature was decreased to 16°C. Following 16 hours

of culturing, cells were pelleted by centrifugation (4,000 x g, 30 min, 4°C) and re-

suspended in 30 mL of binding buffer (20 mM Tris-HCl, 150 mM NaCl, 20% glycerol, and

10 mM 2-mercaptoethanol at pH 7.9). The cell suspension was subjected to four

passages through an EmulsiFlex C-3 cell homogenizer (Avestin, Ottawa, Canada), and

the cell lysate was clarified by centrifugation at 20,000 x g for 30 min at 4°C. The

recombinant steroid-17,20-desmolase was then purified using TALON® Metal Affinity

Resin (Clontech Laboratories, Mountain View, CA) as per the manufacturer’s protocol.

The recombinant protein was eluted using an elution buffer composed of 20 mM Tris-HCl,

150 mM NaCl, 10% glycerol, 10 mM 2-mercaptoethanol, and 250 mM imidazole at pH

7.9. The recombinant 17β-HSDH was then purified using Strep-Tactin® resin (IBA GmbH,

Page 76: Copyright 2018 Seán Martin Mythen

66

Goettingen, Germany) as per manufacturing protocol. The recombinant protein was

eluted using an elution buffer composed of 20 mM Tris-HCl, 150 mM NaCl, 10% glycerol,

10 mM 2-mercaptoethanol, and 2.5 mM desthiobiotin at pH 7.0. The purity of the proteins

were assessed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-

PAGE), and protein bands were visualized by staining with Coomassie Brilliant Blue G-

250 Dye. The protein concentrations were calculated based on the computed molecular

mass and extinction coefficient. The observed subunit mass for each protein was

calculated by migration distance of purified protein to standard proteins using ImageJ

(https://imagej.nih.gov/ij/index.html).

Gel filtration chromatography- Gel filtration chromatography was carried out on

a Superose-6 10/300 analytical column (GE Healthcare, Piscataway, NJ) attached to an

ÄKTAxpress chromatography system (GE Healthcare, Piscataway, NJ) at 4°C. The

eluted recombinant steroid-17,20-desmolase from the metal affinity resin was

concentrated and loaded on to the column with a buffer composed of 50 mM Tris-Cl, 150

mM NaCl, 10% glycerol, and 10 mM 2-mercaptoethanol at pH 7.0. The native molecular

mass of the protein was calculated by peak retention volume of protein to retention

volume of standard proteins.

Liquid chromatography-mass spectrometry analysis- Steroid-17,20-

desmolase activity was determined by incubating 1 μM recombinant enzyme with 50 μM

cortisol in the presence of 1 µM MnCl2 and 10 µM TPP in 50 mM glycine-glycine buffer

(pH 7.0) at room temperature overnight. After incubation, the products were extracted by

vortexing the reaction with 2 volumes of ethyl acetate for 1-2 mins and then recovering

the organic phase. The organic phase was then evaporated under nitrogen gas. The

Page 77: Copyright 2018 Seán Martin Mythen

67

residue was dissolved in 50 μL methanol and analyzed using a Shimadzu UHPLC system

with DAD detector coupled with a Shimadzu LCMS-IT-TOF System (Shimadzu

Corporation, Kyoto, Japan). The LC operating conditions were as follows: LC column, C-

18 analytical column (Capcell Pak C18, Shiseido, Japan), 250 mm × 2 mm i.d., particle

size - 3 μm; mobile phase, H2O containing 0.1% formic acid (A), and acetonitrile

containing 0.1% formic acid (B); total flow rate of mobile phase, 0.2 mL/min; total run time

including equilibration, 41 min. The initial mobile phase composition was 70% mobile

phase A and 30% mobile phase B. The percentage of mobile phase B was changed

linearly over the next 5 min until 35%. Over the next 25 min, the percentage was increased

to 98% linearly. After that the percentage was maintained for 5 min, the mobile phase

composition was allowed to return to the initial conditions and allowed to equilibrate for 5

min. The injection volume was 10 μL. The DAD detector was used at a wavelength of 254

nm. Peak retention times and peak areas of samples were compared with standard

steroid molecules. The mass spectrometer (LCMS-IT-TOF) was operated with an

electrospray ionization (ESI) source in positive mode. The nebulizer gas pressure was

set at 150 kPa with the source temperature of 200 °C and the gas flow at 1.5 L min-1. The

detector voltage was 1.65 kV. High-purity nitrogen gas was used as collision cell gas. The

raw chromatogram and mass spectrogram data were processed with the LC solution

Workstation software (Shimadzu).

Continuous enzyme-coupled spectrophotometric assay- Steroid-17,20-

desmolase activity was measured aerobically at 25°C by monitoring the conversion of

NADPH to NADP+ 340 nM (ε=6,220 M-1.cm-1) reflecting DesAB-catalyzed conversion of

cortisol to 11β-hydroxyandrostenedione followed by NADPH-dependent 17β-HSDH

Page 78: Copyright 2018 Seán Martin Mythen

68

conversion of 11β-hydroxyandrostenedione to a 11β-testosterone. Control reactions were

run leaving out one component at a time. Linearity with respect to time and enzyme

concentration were determined. The standard reaction mixture contained 50 mM MOPS

buffer, pH 7, 50 μM of cortisol, 1 µM MnCl2, 10 µM TPP, 200 μM NADPH, 1 μM each

enzyme, and buffer to a final volume of 0.5 ml. The reaction was initiated by the addition

of the steroid-17,20-desmolase. The initial velocity of enzyme catalyzed reactions was

plotted against the substrate concentrations, and the kinetic parameters were estimated

by fitting the data to the Michaelis-Menten equation by non-linear regression method

using the enzyme kinetics module in GraphPad Prism (GraphPad Software, La Jolla, CA).

An alternative methodology was utilized for substrates lacking a 17-hydroxyl, or those

substrates whose 17-keto products had low specificity for recombinant C. lunatus 17β-

HSDH (rCL17HSDH), was to run the basic assay without rCL17HSDH and NADPH.

Reactions were terminated with 1N HCl and reaction products extracted and separated

by LC/MS as described above. A standard curve of peak area vs. concentration of the

reaction products was generated to quantify reaction rates.

Determination of optimal pH- The buffers used to study the pH profiling of the

coupled enzymatic assay contained 50 mM of buffering agent, 1 µM MnCl2, 10 µM TPP,

and 150 mM NaCl. Buffering agents used were as follows: Sodium Citrate (pH 6.0),

MOPS (pH 6.5 – 7.0), Glycine-Glycine (pH 6.5 – 8.5), Tris-Cl (pH 9), and Glycine-NaOH

(pH 9.5 – 10). To determine the optimal pH, purified recombinant steroid-17,20-

desmolase and 17β-HSDH at final concentrations of 1 μM each, were added to 50 μM

cortisol and 200 μM NADPH in 500 μL of buffer. The oxidation of NADPH was measured

Page 79: Copyright 2018 Seán Martin Mythen

69

continuously at 340 nM for 5 mins by spectrophotometry and initial velocities were

calculated.

RESULTS

Cloning, overexpression of desAB, and purification of recombinant DesAB-

The desA gene encodes a 296 amino acid protein (CLOSCI_00899) with predicted

thiamine pyrophosphate (TPP) binding site (H74, G154, N183, K253), metal-binding site

(D153, E155) and active-site (H37, H268) based on multiple sequence alignment with

characterized transketolases and organisms expressing steroid-17,20-desmolase

activity, or encoding a predicted desABE operon (Figure 3.1). The desB encodes a 327

amino acid protein (CLOSCI_00900) comprising the C-terminal portion of a typical full-

length transketolase, and includes conserved residues with transketolases as diverse as

Lactobacillus salivarius UCC118 (YP_536833) (133) and Homo sapiens (P29401.3) (134)

including TPP-binding (F95), and active-site residues (E70). Also included in this

alignment are DesA and DesB sequences from known steroid-17,20-desmolase

expressing strains such as gut bacteria B. desmolans (WP_031476415.1,

WP_031476417.1), C. cadaveris (WP_027640053.1, WP_027640052.1), urinary isolate

Propionimicrobium lymphophilum ACS-093-V-SCH5 (135, 136) (WP_016455355.1,

WP_016455356.1), and inferred by homology and desABE operon structure from urinary

tract isolate Arcanobacterium urinimassiliense (137) (WP_084789932.1,

WP_073996552.1).

Since key conserved transketolase catalytic and co-factor binding residues are

distributed across both the desA and desB genes, we chose the pETDuet expression

vector, designed for overexpression of two target genes in E. coli BL21(DE3)RIL as His-

Page 80: Copyright 2018 Seán Martin Mythen

70

Figure 3.1 Steroid-17,20-desmolase (DesAB) shares conserved catalytic and co-factor binding residues with transketolases Sequences in partial CLUSTAL-Omega alignment include Homo sapien transketolase (TKT_Hsap), transketolase from Lactobacillus salivarius (TKT_Lsa1), DesAB from C. scindens ATCC 35704 (desAB_Csci), DesAB from Butyricicoccus desmolans (desAB_Bdes), DesAB from Clostridium cadavaris (desAB_Ccad), DesAB from Propionimicrobium lymphophilum (desAB_Plym), and DesAB from Arcanobacterium urinimassiliense (desAB_Aaur). Conserved residues are highlighted based on functional role depicted by the color key.

Page 81: Copyright 2018 Seán Martin Mythen

71

tagged recombinant proteins. SDS-PAGE analysis of the elution fraction following affinity

chromatography revealed two bands at 32.7 kDa and 38.4 kDa, corresponding to the

predicted molecular mass of rDesA and rDesB, respectively (Figure 3.2A). Gel filtration

chromatography determined the native molecular mass of rDesAB to be 142 ± 5.4 kDa,

suggesting that the rDesAB is a heterotetramer (Figure 3.2B).

Development of a coupled, continuous spectrophotometric assay to

measure steroid-17,20-desmolase activity- Next, we incubated recombinant DesAB

overnight in the presence of cortisol, TPP and Mn2+ in glycine-glycine buffer, pH 7.0.

Extraction of the reaction buffer with ethyl acetate and separation by reverse-phase HPLC

revealed net conversion of cortisol (RT = 17.5 min) to a single product, with retention time

identical to 11β-hydroxyandrostenedione standard (RT =23.5 min) (Figure 3.3). The

control cortisol peak was subjected to electrospray ionization-time of flight-mass

spectrometry (ESI-TOF-MS) and resulted in a major mass ion in positive mode of 363.17

m/z, consistent with the molecular weight of cortisol (362.46 amu). The major mass ion

of the rDesAB product was 303.15 m/z, consistent with side-chain cleavage and formation

of 11β-hydroxyandrostenedione (302.414 amu). These results demonstrate that the

DesAB is the steroid-17,20-desmolase.

The steroid-17,20-desmolase assay previously utilized is discontinuous, time-

consuming, and expensive. We therefore sought another approach to quantify steroid-

17,20-desmolase activity using a continuous enzyme-coupled spectrophotometric assay.

The steroid-17,20-desmolase reaction yields a 17-ketosteroid product which, if coupled

to an NAD(P)H-dependent 17α- or 17β-HSDH, could be used to measure reaction

velocity by monitoring NAD(P)H oxidation at 340 nm (Figure 3.4A). Consultation of the

Page 82: Copyright 2018 Seán Martin Mythen

72

A. B.

Figure 3.2 SDS-PAGE of affinity purified rDesA and rDesB and native molecular mass estimate of rDesAB. A. SDS-PAGE of pETDuet co-expressed rDesA and rDesB (Lane DesAB), protein molecular weight marker (M). B. Native molecular size analysis of rDesAB by size-exclusion chromatography.

Page 83: Copyright 2018 Seán Martin Mythen

73

Figure 3.3 Liquid chromatography-mass spectrometry of rDesAB reaction products Standards include authentic cortisol and 11β-hydroxyandrostenedione. Reaction assay conditions include 1 μM recombinant enzyme with 50 μM cortisol in the presence of 1 µM MnCl2 and 10 µM TPP in 50 mM glycine-glycine buffer (pH 7.0) at room temperature overnight. Control reactions omitted rDesAB. Mass spectra corresponding to HPLC peaks depicted in inset.

Page 84: Copyright 2018 Seán Martin Mythen

74

Figure 3.4 SDS-PAGE of purified recombinant 17β-HSDH from Cochliobolus lunatus (rCL17HSDH) and schematic representation of enzyme-coupled continuous spectrophotometric steroid-17,20-desmolase activity assay. A. Schematic of conversion of cortisol to 11β-hydroxyandrostenedione, catalyzed by rDesAB, followed by NADPH-dependent conversion of 11β-hydroxyandrostenedione to 11β-hydroxytestosterone by rCL17HSDH. B. SDS-PAGE of rCL17HSDH. Protein molecular weight marker (M).

Page 85: Copyright 2018 Seán Martin Mythen

75

relevant literature revealed a well characterized NADP(H)-dependent 17β-HSDH from the

filamentous fungus Cochliobolus lunatus, whose 1.044 kB cDNA sequence is available

at NCBI (Accession number AAD12052.1) (17). We trimmed the 5’- and 3’- untranslated

regions (UTR) sequences and optimized the resulting 813 bp coding sequence for codon-

usage in E. coli with the Codon Optimizer tool from IDT and the gene insert was

synthesized using the GenBlock synthetic biology service from IDT. The insert was PCR

amplified using primers designed for bi-directional cloning into pET51b(+) vector, allowing

for expression of rCL17HSDH as a 32.2 kDa N-terminal streptavidin affinity-tagged

protein (Figure 3.4B).

rCL17HSDH showed NADPH-dependent activity against 11β-

hydroxyandrostenedione, but not cortisol in 50 mM glycine-glycine buffer, pH 7.0.

Oxidation of NADPH was observed at 340 nm only in the presence of rDesAB, cortisol,

and rCL17HSDH. Previous characterization of rCL17HSDH showed pH optimum

between 7.0 and 8.0 (18). We determined that DesAB was most active between pH 7.0

and pH 7.5 by observing conversion of cortisol to 11β-hydroxyandrostenedione between

pH 6.0 to 8.5 (data not shown). Therefore, the pH optima of both rDesAB complex and

rCL17HSDH overlap. After optimization of the coupled assay with respect to enzyme

concentration, we determined pH optimum for the coupled reaction was pH 7.0 in MOPS

buffer. There was a precipitous decline in activity from pH 6.5 to pH 6.0, nearly equivalent

to the loss of activity from pH 7.0 to 8.0. At pH 10.0, the coupled assay retained ~5% of

its maximum activity (Figure 3.5).

Page 86: Copyright 2018 Seán Martin Mythen

76

Figure 3.5 Biochemical characterization of rDesAB by enzyme-coupled continuous spectrophotometric assay. (Top) Effect of pH on steroid-17,20-desmolase activity (Bottom) Michaelis-Menten of rDesAB activity with increasing cortisol.

Page 87: Copyright 2018 Seán Martin Mythen

77

Determination of kinetic constants and substrate-specificity- We varied the

concentration of cortisol in the presence of 200 µM NADPH, 1 µM MnCl2, 10 µM TPP in

50 mM MOPs buffer pH 7.0. Michaelis–Menten saturation curve is shown in Figure 3.5.

The KM for the rDesAB is 4.96 ± 0.57 µM and Vmax of 12.76 ± 1.12 nmol.min-1.mg-1. The

turnover number (kcat) was determined to be 0.87 ± 0.076 min-1 and catalytic efficiency

(kcat/KM) was 0.17 ± 0.029 min-1.μM-1. Suspecting that an aldose-acceptor may be utilized

in the reaction, glyceraldehyde-3-phosphate and ribose-5-phosphate were tested at 50

µM, 100 µM, and 500 µM; however, we did not observe increased enzymatic activity.

We tested substrate-specificity using a combination of methods, depending on whether

steroids were 17-hydroxylated. The enzyme-coupled continuous spectrophotometric

assay was relegated to 17-hydroxylated substrates. Those reaction products from

substrates lacking 17-hydroxy were characterized by LC-MS. Furthermore, we tested

NADPH-dependent rCL17HSDH activity against the expected C-19 steroid-17,20-

desmolase products to determine whether we could make a direct comparison between

substrates whose products differed structurally. We determined that Ring-A reduced

products (50 µM concentration), such as 5β-androsterone and 11β-hydroxyandrosterone,

are not substrates for rCL17HSDH at, resulting in 8.06 ± 2.65% and 4.73 ± 0.68% activity,

respectively, relative to 11β-hydroxyandrostenedione (three technical replicates). These

substrates were therefore tested for activity using rDesAB with separation of steroids by

LC/MS.

We determined that 11β-hydroxylation is not necessary for steroid-17,20-

desmolase (Table 3.2). 20α-Hydroxy or 20β-hydroxy derivatives of cortisol were not

Page 88: Copyright 2018 Seán Martin Mythen

78

Table 3.2 Substrate specificity for rDesAB from C. scindens ATCC 35704

Page 89: Copyright 2018 Seán Martin Mythen

79

substrates. Those substrates lacking 17-hydroxyl group, and/or 21-hydroxyl group, such

as corticosterone and progesterone metabolites were not side-chain cleaved (Figures 3.6

- 3.7). Mass spectra are presented in Figures 3.8. Interestingly, tetrahydrocortisol and

tetrahydrodeoxycortisol were also not recognized by rDesAB, indicating 3α-hydroxy, 5β-

reduced metabolites are not substrates.

DISCUSSION

In the current study, we report that the desA and desB genes from C. scindens

ATCC 35704 encode steroid-17,20-desmolase. Recombinant DesAB displayed

substrate-specificity consistent with native steroid-17,20-desmolase: the enzyme

recognizes 20-keto-17, 21-dihydroxy steroids (67, 68). C. scindens ATCC 35704

expresses NAD(H)-dependent 20α-HSDH (69), while B. desmolans, C. cadavaris, and P.

lymphophilum express 20β-HSDH (85) in addition to steroid-17,20-desmolase. Oxidation-

reduction of the C20 oxygen is predicted to function as a regulatory “switch”,

interconverting cortisol between DesAB substrate and non-substrate forms. The desD

gene is co-expressed with desABC (69) and encodes a predicted 51.2 kDa transport

protein in the major facilitator superfamily. DesD is predicted to transport cortisol from the

gut lumen into the bacterial cytoplasm. RNA-Seq analysis also identified a cortisol-

inducible ABC-type multidrug transport system predicted to export 11β-

hydroxyandrostenedione (69).

The conversion of cortisol to 11β-hydroxyandrostenedione by C. scindens is an

important step in a multi-species gut microbial endocrine pathway that results in the

formation of 11β-hydroxyandrogens (69). Previous studies reveal a number of cortisol

Page 90: Copyright 2018 Seán Martin Mythen

80

Figu

re 3

.6 D

eter

min

atio

n of

ste

roid

-17,

20-d

esm

olas

e ac

tivity

aga

inst

sub

stra

tes

lack

ing

a 17

-hyd

roxy

, and

te

trahy

dro(

deox

y)co

rtiso

l by

LC/M

S.

A. H

PLC

chr

omat

ogra

phs

with

con

trol (

left

pane

l) an

d rD

esAB

cat

alyz

ed (r

ight

pan

el).

Cor

tisol

and

11-

deox

ycor

tisol

are

incl

uded

as

posi

tive

cont

rols

. HPL

C p

eaks

are

labe

led

with

cor

resp

ondi

ng m

ajor

mas

s io

ns (m

/z).

B. O

verla

y of

con

trol (

red)

and

rDes

AB c

atal

yzed

(blu

e) H

PLC

chr

omat

ogra

phs

with

tetra

hydr

ocor

tisol

(top

pa

nel)

and

tetra

hydr

odeo

xyco

rtiso

l (lo

wer

pan

el).

MS

data

are

incl

uded

as

inse

ts a

nd c

orre

spon

d to

maj

or

peak

s. S

tero

id m

etab

olite

s w

ere

not d

etec

ted

in a

ny m

inor

pea

ks.

A.

Page 91: Copyright 2018 Seán Martin Mythen

81

B. Fi

gure

3.6

con

tinue

d

Page 92: Copyright 2018 Seán Martin Mythen

82

Figure 3.7 Functional groups important for rDesAB specificity

Page 93: Copyright 2018 Seán Martin Mythen

83

Mass spectra data for 11-epicorticosterone

Mass spectra data for 11β-hydroxyprogesterone

Figure 3.8 Mass spectra of rDesAB catalyzed reactions. MS spectra are from major peaks depicted in Figure 3.6A. Conditions for MS and LC-MS are described in Materials and Methods in the main text.

Page 94: Copyright 2018 Seán Martin Mythen

84

Mass spectra data for tetrahydrocortisol

Mass spectra data for tetrahydrodeoxycortisol

Figure 3.8 Continued

Page 95: Copyright 2018 Seán Martin Mythen

85

Mass spectra data for corticosterone

Mass spectra data for cortisol with rDesAB

Figure 3.8 Continued

Page 96: Copyright 2018 Seán Martin Mythen

86

Mass spectra data for cortisol

Mass spectra data for 11β-androstenedione

Figure 3.8 Continued

Page 97: Copyright 2018 Seán Martin Mythen

87

Mass spectra data for desoxycortisol with rDesAB

Mass spectra data for desoxycortisol

Figure 3.8 Continued

Page 98: Copyright 2018 Seán Martin Mythen

88

Mass spectra data for androstenedione

Figure 3.8 Continued

Page 99: Copyright 2018 Seán Martin Mythen

89

metabolites following incubation with mixed human or rat fecal bacteria included

3α,11β,17β-trihydroxy-5β-androstane, 3α,11β,17β-trihydroxy-5α-androstane,

demonstrating the ability of gut microbes to form 5α-, 17β-reduced metabolites (63),

(64). Some of the taxa responsible for these biotransformations have been identified. C.

paraputrificum is known to express 3α-HSDH and 5β-reductase (138), C. innocuum is

reported to encode 3βHSDH and 5β-reductase (139), Clostridium sp. Strain J-1 encodes

3α-HSDH and 5α-reductase (138), capable of reducing the 3-oxo-Δ4 bonds in ring-A. 17α-

HSDH has been reported in C. scindens VPI 12708 (140) and has been detected in E.

lenta DSM 2243 and C592 (141). The genes encoding these reductases have yet to be

identified, but are likely to affect steroid-17,20-desmolase activity in the gut. E. lenta also

expresses 21-dehydroxylase (74), and 21-dehydroxylated metabolites are detected

following incubation of cortisol in fecal suspensions (63, 64). Due to the absolute

requirement for the C21-hydroxy group, steroid-17,20-desmolase and cortisol 21-

dehydroxylation by Eggerthella lenta are competing reactions in the gut (74).

Extrapolations from radiometric studies in non-human primates (142) and data

from humans (143) suggest that 15% of the 15 mg of cortisol are detected in the gut.

Based on these studies, the concentration of cortisol metabolites in the ~0.4L human

colon (144) is ~15 µM. 11β-hydroxyandrostenedione is now viewed as a pro-androgen,

whose downstream 5α-, 17β-reduced metabolites, are now thought to play an important

role in activation of androgen receptor (33, 145). The field of microbial endocrinology has

emerged and recognizes the need to characterize androgen production by C. scindens

(38, 146–150). Many host immune cells express androgen-receptor (151, 152), and it is

possible that C. scindens affects immune function through generation of androgens.

Page 100: Copyright 2018 Seán Martin Mythen

90

Androgens directly affect bacterial growth and virulence (153, 154). Intriguingly, steroids

have been shown to influence C. difficile germination (3). 11β-hydroxy-androgens are

also inhibitors of the host enzyme 11β-HSD2 (80, 155), but in kidney and vascular smooth

muscle may be pro-hypertensive (155). Inhibition of 11β-HSD2 in colonocytes may be

protective against colorectal cancer (156). Prostate tissue contains 11β-

hydroxyandrogens despite inhibition of host enzymes involved in synthesizing androgens

(chemical castration) (33, 145). The role of 11β-OHAD formation by the gut microbiota in

androgen accumulation in the prostate is unknown; however, we discovered that bacteria

isolated from urine also express the bacterial steroid-17,20-desmolase pathway (85), and

the majority of cortisol excretion from the body is via the urine (142, 143). Of note, we

recently reported identification of the steroid-17,20-desmolase gene cluster in the

genomes of urinary bacterial isolates and confirmed steroid-17,20-desmolase activity in

an isolate of Propionimicrobium lymphophylum (85). Verification that desA and desB

genes encode steroid-17,20-desmolase is an important step towards the application of

nucleic acid-based approaches such as quantitative polymerase chain, and metagenomic

sequencing in determining whether correlations exist between levels/expression of

desAB and host phenotypes and disease states.

As gut metagenomic sequencing becomes less costly, and more widespread,

assigning function to particular nucleic acid and deduced amino acid sequences will be

important for hypothesis generation, and testing. The desA and desB genes are a clear

case in point. The desABCD operon is annotated as participating in carbohydrate

metabolism, due largely to the desA and desB gene product homology to N-terminal and

C-terminal TPP-dependent transketolases, respectively (Figure 3.9). Early culture-based

Page 101: Copyright 2018 Seán Martin Mythen

91

Figure 3.9 Gene organization and domain architecture of steroid-17,20-desmolase

Page 102: Copyright 2018 Seán Martin Mythen

92

studies established that steroid-17,20-desmolase activity is induced by cortisol (66–68).

Utilization of RNA-Seq to identify differentially expressed genes, coupled with biochemical

characterization of the recombinant gene products was necessary to determine the

nucleic acid sequences encoding this bacteria steroid biotransforming pathway (69).

A recent report describes side-chain cleavage of cholesterol by Mycobacterium

tuberculosis and revealed a thiolase-derived enzyme, a steroid-aldolase, involved in

microbial steroid-degradation (157) (Figure 3.10). It is therefore possible that additional

side-chain cleaving gut microbial members will be identified that are capable of aldolase

cleavage of host and dietary steroids in the gut. This is in contrast to eukaryotic cells

which typically utilize oxygen-dependent P450 monooxygenases (158).

The low turnover number is within the range of previously reported prokaryotic

enzymes involved in secondary metabolism (159), and may indicate that selection

pressure may be less stringent for optimization of enzyme rate for functions that may not

contribute greatly to organismal fitness. Studies are now ongoing to understand important

amino acid residues involved in substrate-binding to steroid-17,20-desmolase, and

catalysis, as well as the identity and fate of the side-chain. Development of a novel

enzyme-coupled continuous spectrophotometric assay to quantify steroid-17,20-

desmolase activity against cortisol is expected to hasten results from these efforts.

Current limitations of the assay with respect to androgen specificity of rCL17HSDH may

be addressed in the future either through protein engineering, or by identifying naturally

occurring 17α- or 17β-HSDHs. In addition, future studies will be required to determine the

identity of the side-chain and whether an aldose acceptor is utilized in vivo. A detailed

understanding of the steroid-17,20-desmolase will be important in uncovering how

Page 103: Copyright 2018 Seán Martin Mythen

93

Figure 3.10 Schematic representation of side-chain cleavage reactions for steroid-17,20-desmolase from C. scindens ATCC 35704, steroid-transaldolase from Mycobacterium tuberculosis, and Cytochrome P450 monooxygenase (CYP11A1) in Eukaryotes.

Page 104: Copyright 2018 Seán Martin Mythen

94

anaerobic bacteria co-opted pentose-phosphate pathway enzymes for steroid side-

chain cleavage.

Page 105: Copyright 2018 Seán Martin Mythen

95

CHAPTER 4: Structural and biochemical characterization of 20β-hydroxysteroid

dehydrogenase from Bifidobacterium adolescentis strain L2-32

ABSTRACT

A novel 20β-hydroxysteroid dehydrogenase (20β-HSDH) from Bifidobacterium

adolescentis strain L2-32 has been enzymatically characterized and crystallized. 20β-

HSDH catalyzes the reduction of the carboxyl group at position 20 of the steroid cortisol,

thereby producing 20β-dihydrocortisol. The enzyme was determined to be tetrameric in

structure, with a subunit molecular mass of 32 ± .12 kDa, and exhibits a pH optimum of

5.0 in the reductive direction and 5.5 in the oxidative direction. Cortisol was determined

to be the preferred substrate with a KM of 24.07 μM. Vmax, kcat, and kcat/KM values were

determined to be 21.08 µmol.min-1.mg-1, 668.3 min-1, and 27.765 μM-1.min-1, respectively.

Mutation of the putative catalytic serine confirmed its functional role in reaction catalysis.

The enzyme was shown to have coenzyme specificity for NADH rather than NADPH. The

apo and holo forms of the enzyme were crystallized and shown to diffract at 2.1 and 2.5

Å resolution, respectively.

INTRODUCTION

Hydroxysteroid dehydrogenases (HSDH) are pyridine nucleotide-dependent

enzymes classified into three superfamilies which include aldo-keto reductase (AKR),

long/medium chain dehydrogenase/reductase (MDR), and short-chain

dehydrogenase/reductase (SDR) ((11, 69, 160) ). The gut microbiome is exposed to a

mixture of corticosteroids and bile acids and much research has focused on bacteria

strains of diverse taxa which express stereo- and regio-specific HSDHs in the SDR and

AKR families capable of oxidation and epimerization of bile acid hydroxyl groups at C-3,

Page 106: Copyright 2018 Seán Martin Mythen

96

C-7, and C-12 (161, 162). Recent attention has been focused on the bacterial cortisol

steroid-17,20-desmolase pathway in Clostridium scindens, capable of converting cortisol

to 11β-hydroxyandrostenedione (11 β-OHAD), and the zinc-dependent MDR family 20α-

HSDH which is predicted to enzymatically regulate the pathway (67, 69). 11β-OHAD is a

pro-androgen whose formation in the gut is predicted to alter host physiology (33).

Bacterial steroid-17,20-desmolase has also been reported in Clostridium cadaveris and

Butyricicoccus desmolans (formerly Eubacterium desmolans) and is associated with

NADH-dependent cortisol 20β-HSDH activity (85). Previous studies also identified strains

of Bifidobacterium adolescentis isolated from human and rat fecal dilutions which

expressed cortisol 20β-HSDH activity alone (86). Recently, the gene encoding gut

bacterial NADH-dependent 20β-HSDH activity was reported in B. desmolans (85).

Phylogenetic analysis identified numerous Bifidobacterium spp., including strains of B.

adolescentis, encoding putative cortisol 20β-HSDH (85).

Gut bacterial 20β-HSDH is a member of the SDR family of proteins (11, 85). The

SDR family constitute one of the largest of protein superfamilies distributed across all

three domains of life, comprising of functionally diverse members (12, 162–164). SDR

family enzymes include several EC classifications including lyases, isomerases, and

pyridine nucleotide-dependent oxidoreductases constituting the majority of enzymes.

Currently in Uniprot there are 308,095 entries and members only share ~20-30% residue

identity in pairwise comparisons. Typical SDR family enzymes are 250-350 amino acids

in length, with secondary structure consisting of one-domain Rossmann folds containing

6-7 beta-strands flanked by three alpha helices. Conserved N-terminal Gly-X3-Gly-X-Gly

motif critical in NAD(P)(H) binding, as well as catalytic tetrad composed of Ser, Tyr, Lys,

Page 107: Copyright 2018 Seán Martin Mythen

97

and Asn residues (11, 165). SDR family enzymes vary in substrate recognition, including

sugars, dyes, prostaglandins, porphyrins, alcohols, and steroids (11). As a consequence,

substrate-binding residues are not highly conserved. Differentiating HSDHs from other

functional classes of SDR family enzymes is important for identifying steroid biochemical

pathways in gut metagenomes. At present, this task is impeded by a lack of primary

amino acid sequences for several HSDHs whose activities have been reported. If we are

to determine the residues important in differentiating HSDHs into functional categories

(i.e. 7α-HSDH from 7β-HSDH and both from 20β-HSDH), the genes encoding these

enzymes have to be located.

The first SDR family enzyme crystallized was 3α,20β-HSDH from the soil

bacterium Streptomyces hydrogenans (166). The 3α,20β-HSDH is able to reversibly

oxidize the 3α-hydroxyl and 20β-hydroxyl groups of androgens and progestogens. The

S. hydrogenans 3α,20β-HSDH shares 31% amino acid identity with the 20β-HSDH from

B. adolescentis, which does not appear to function as a dual 3α-HSDH. It is important to

obtain a mechanistic understanding of 20β-HSDH because this enzyme is capable of

regulating the major gut microbial side-chain cleavage of cortisol steroid-17,20-

desmolase resulting in pro-androgens, as well as cortisol 21-dehydroxylation resulting in

formation of progestogens (74). Herein, we report the characterization of 20β-HSDH from

Bifidobacterium adolescentis strain L2-32.

MATERIALS AND METHODS

Bacterial strains, enzymes, reagents, and materials- Bifidobacterium

adolescentis, strain L2-32 was purchased through BEI Resources (Catalog No. HM-633).

Turbo Competent Escherichia coli DH5α (High Efficiency) cells were purchased from New

Page 108: Copyright 2018 Seán Martin Mythen

98

England Biolabs (NEB). BL21-CodonPlus (DE3)-RIPL Competent E. coli cells were

purchased from Agilent Technologies (Catalog No. 230280). Phusion High-Fidelity DNA

Polymerase, Restriction Endonucleases, and T4 DNA Ligase were purchased from NEB.

Isopropyl β-D-1-thiogalactopyranoside (IPTG) was purchased from Gold Biotechnology.

TALON His Tag Purification Resin was purchased from Takara Bio USA, Inc. Corning

Spin-X UF concentrators with a 10 kDa molecular weight cut-off (MWCO) were purchased

from MilliporeSigma. Steroids were purchased from Steraloids. Further materials were

purchased from Fisher Scientific and MilliporeSigma.

Whole cell steroid conversion assay- B. adolescentis, strain L2-32, B.

desmolans ATCC 43058, and C. scindens ATCC 35704 were cultivated in anaerobic

brain heart infusion (BHI) broth. Steroids were dissolved in methanol and added to sterile

anaerobic BHI medium at a concentration of 50 μM. The samples were then seeded

separately with 0.1 ml of each bacterial culture and incubated at 37°C for 48 hours. After

incubation, the products were extracted by vortexing the samples with 2 volumes of ethyl

acetate for 3 min followed by recovery of the organic phase. The organic phase was then

evaporated under nitrogen gas. The remaining residues were dissolved in 100 μL of

methanol and analyzed using a high-performance liquid chromatography (HPLC) system

(Shimadzu, Japan) equipped with C-18 analytical column (Capcell Pak c18, Shiseido,

Japan). The mobile phase contained acetonitrile/water with 0.01% formic acid and the

flow rate was maintained at 0.2 ml.min−1. A DAD detector was used at a wavelength of

254 nm. Peak retention times and peak areas of samples were compared with standard

steroid molecules.

Page 109: Copyright 2018 Seán Martin Mythen

99

Cloning and formation of the 20β-HSDH expression vector- Genomic DNA

was extracted and purified from B. adolescentis, strain L2-32 using the FastDNA Spin Kit

from MP Biomedicals. The gene encoding 20β-HSDH (accession number

WP_003810233.1) was amplified using Phusion High Fidelity DNA Polymerase and the

following primers: 5'-ATATATcatatgATGGCAGTTGAATCATCGCAGATTCCA-3' and 5'-

ATATATctcgagTTAGAAGACGCTGTAGCCGCCATCTAC-3'. The restriction

endonuclease sites for NdeI and XhoI are shown in lower-case respectively. The PCR

product was purified and double digested with NdeI and XhoI restriction endonucleases.

The digested PCR product was purified and cloned into the pET-28(+) (Novagen,

Madison, WI, USA) vector using T4 DNA Ligase to create the expression construct. The

pET-28(+) vector used had previously been re-engineered with an ampicillin resistance

gene. The resulting construct was transformed into Turbo Competent E. coli DH5α (High

Efficiency) cells by heat shock at 43°C for 35 sec and outgrown form 1 hour in Super

Optimal Broth with Catabolite Repression (SOC) (NEB) before growing on LB agar plates

supplemented with ampicillin (100 µg/ml) for 16 hours at 37°C. Single colonies from the

LB agar plates where inoculated into 5 mL of LB medium containing ampicillin (100 µg/ml)

and grown overnight at 37°C at 220 rpm. The cells were then centrifuged at 4000 x g for

10 minutes at 4°C and plasmids were extracted and purified from the pellets using the

QIAprep Spin Miniprep Kit. Both strands of the plasmid DNA were sequenced for

confirmation of the insert and absence of mutations using T7 promoter and T7 terminator

primers (W. M. Keck Center for Comparative and Functional Genomics at the University

of Illinois at Urbana-Champaign). Once overexpressed, the protein will include an N-

terminus 6xHis-tag and a thrombin cleavage site.

Page 110: Copyright 2018 Seán Martin Mythen

100

Site-directed mutagenesis- Plasmids containing the 20β-HSDH gene construct

were mutated using the QuikChange Lightning Site-Directed Mutagenesis Kit. Primers

containing the desired mutations were designed as follows: 5'-

CTGATCTTCACAGCCgcgCTGTCCGGACACAA-3' and 5'-

TTGTGTCCGGACAGcgcGGCTGTGAAGATCAG-3' for Serine 181 to Alanine (S181A);

and 5'-TCTTCACAGCCTCCCTGgcgGGACACAACGCGAACTA-3' and 5'-

TAGTTCGCGTTGTGTCCcgcCAGGGAGGCTGTGAAGA-3' for Serine 183 to Alanine

(S183A). Mutated bases are shown in lowercase. The resulting mutated constructs were

transformed into Turbo Competent E. coli DH5α (High Efficiency) cells by heat shock at

43°C for 35 sec and outgrown form 1 hour in Super Optimal Broth with Catabolite

Repression (SOC) (NEB) before growing on LB agar plates supplemented with ampicillin

(100 µg/ml) for 16 hours at 37°C. Single colonies from the LB agar plates where

inoculated into 5 mL of LB medium containing ampicillin (100 µg/ml) and grown overnight

at 37°C at 220 rpm. The cells were then centrifuged at 4000 x g for 10 minutes at 4°C

and plasmids were extracted and purified from the pellets using the QIAprep Spin

Miniprep Kit. Mutations were confirmed by sequencing of the mutated plasmid using T7

promoter and T7 terminator primers (W. M. Keck Center for Comparative and Functional

Genomics at the University of Illinois at Urbana-Champaign).

Overexpression and purification of wild-type and mutant 20β-HSDH

proteins- For protein expression, recombinant plasmids extracted from the E. coli DH5α

cells were transformed into BL21-CodonPlus (DE3)-RIPL Competent E. coli cells by heat

shock at 43°C for 35 sec and outgrown for 1 hour in Super Optimal Broth with Catabolite

Repression (SOC) (NEB) before growing on LB agar plates supplemented with ampicillin

Page 111: Copyright 2018 Seán Martin Mythen

101

(100 µg/ml) and chloramphenicol (50 µg/ml). After 16 hours of growth at 37°C, five

colonies were picked to inoculate 10 mL of LB medium containing ampicillin (100 µg/ml)

and chloramphenicol (50 µg/ml) and were grown for 6 hours at 37°C at 220 rpm. The

turbid pre-culture was then added to 1 L of LB medium containing the same antibiotics

and shaken at 37°C at 220 rpm. At an OD600 of 0.6, the culture was induced with 0.25

mM IPTG, after which, the culture was incubated at 16°C at 220 rpm for 16 h. Cells were

pelleted by centrifugation at 4,000 x g for 30 minutes at 4°C. The pellets were

resuspended in 25 mL of buffer containing 50 mM Tris-Cl and 300 mM NaCl at a pH of

7.9. The resuspended cells were homogenized using an EmulsiFlex C-3 homogenizer

by passing the samples through the homogenizer four times and the lysate was refined

by centrifugation at 20,000 x g for 30 minutes at 4°C. The soluble lysate was then purified

using metal chelate chromatography followed by removal of the 6xHis-tag by use of

thrombin. The lysate was further purified using anion exchange chromatography to

remove the excess thrombin. The resulting purified protein was analyzed using sodium

dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE).

Gel filtration chromatography- Gel filtration chromatography was performed on

a Superdex 200 analytical column (GE Healthcare, Piscataway, NJ) attached to an

ÄKTAxpress chromatography system (GE Healthcare, Piscataway, NJ) at 4°C. The

purified protein was concentrated and loaded onto the analytical column equilibrated with

50 mM tris-Cl and 300 mM NaCl at a pH of 7.5 and was eluted at a flow rate of 1 ml/min.

The native molecular mass of the 20β-HSDH protein was determined to by comparing its

elution volume with that of standard proteins.

Page 112: Copyright 2018 Seán Martin Mythen

102

Optimal pH determination- The 20β-HSDH pH optimum was measured

spectrophotometrically by monitoring the reductive biotransformation of cortisol to 20β-

dihydrocortisol as well as the reverse oxidative biotransformation of 20β-dihydrocortisol

to cortisol. For the reductive direction, measurements were taken by analyzing the

decrease of NADH at 340 nm. For the oxidative direction, measurements were taken by

analyzing the increase of NADH at 340 nm. The buffers used in the pH profiling contained

50 mM of buffering agent and 150 mM NaCl. Buffering agents used were as follows:

Sodium acetate trihydrate (pH 3.5 – 5.5), Sodium phosphate monobasic monohydrate

(pH 6.5 – 7.5), Tris (pH 8.0 – 9.0), Glycine (pH 10 – 11). For the reductive reaction 50

nM of enzyme was added to 50 μM cortisol and 150 μM NADH in 200 μL of buffer. For

the oxidative reaction 50 nM of enzyme was added to 50 μM 20β-dihydrocortisol and 150

μM NAD+ in 200 μL of buffer. For each reaction the decrease or increase of NADH was

measured continuously for five minutes and initial velocities were calculated.

Enzyme activity analyses and determination of kinetic parameters- 20β-

HSDH wild-type and mutant activity was measured spectrophotometrically by monitoring

the reductive biotransformation of cortisol to 20β-dihydrocortisol as well as the reverse

oxidative biotransformation of 20β-dihydrocortisol to cortisol by measuring the levels of

NADH at 340 nm. The standard reaction mixture contained 50 mM sodium acetate

trihydrate buffer (pH 5.0), 150 μM of cofactor (NADH), 10 nM of enzyme, and a varied

substrate concentration between 5 μM and 90 μM. GraphPad Prism was used to plot

data, create Michaelis-Menten saturation curves, Lineweaver-Burk plots, and to calculate

kinetic parameters.

Page 113: Copyright 2018 Seán Martin Mythen

103

Substrate-specificity assay- A saturating concentration of 50 μM of substrate as

determined by the enzyme activity analyses, was used as the concentration for analyzing

various other steroid substrates and their activity. Like the enzyme activity analysis, 20β-

HSDH activity was measured spectrophotometrically by monitoring the biotransformation

of substrate by measuring the levels of NADH at 340 nm. The standard reaction mixtures

contained sodium acetate trihydrate buffer (pH 5.0 or pH 5.5), 150 μM of cofactor (NADH

or NAD+), 50 nM of enzyme, and 50 μM of substrate.

Circular dichroism spectroscopy- Circular dichroism (CD) spectroscopy was

performed using Jasco model J-815 circular dichroism spectropolarimeter (Jasco,

Japan). Purified recombinant 20β-HSDH enzyme samples were analyzed from 190 to

260 nm in 10 mM potassium phosphate buffer (pH 7.9) at a concentration of 0.2 mg/mL.

Crystallization- 20β-HSDH crystallization conditions were obtained using the

Protein Crystallography System from Art Robbins Instruments (PHOENIX). Purified

recombinant wild-type 20β-HSDH protein was concentrated between a range of 15-20

mg/ml. Utilizing the PHOENIX and dispensing in 96 - 3 well Intelli-Plates, sitting drops

were prepared comprising of 30 μL of well condition and 300 nL of protein mixed with 300

nL of well condition. Plates were placed in 4°C and crystals were grown in condition 69

of PEGRX HT containing 20% v/v 2-propanol, 0.1 M MES monohydrate pH 6.0, and 20%

w/v Polyethylene glycol monomethyl ether 2,000. The crystals continued to grow up to

day 7 and no further optimization was necessary. The crystals were overlaid with paraffin

oil to prevent evaporation and 20% ethylene glycol as a cryoprotectant. Crystals were

picked and immediately flash-cooled in liquid nitrogen. X-ray data was collected at the

Life Sciences Collaborative Access Team (LSCAT) of the Advanced Photon Source at

Page 114: Copyright 2018 Seán Martin Mythen

104

Argonne National Laboratory. For mutant S181A 20β-HSDH, purified recombinant

protein was incubated with 2.5 mM NADH for 2 hours. Crystals were grown in condition

86 of PEG/Ion containing 0.05 M Citric acid, 0.05 M BIS-TRIS propane / pH 5.0, and 16%

w/v Polyethylene glycol 3,350. The condition was then optimized in hanging-drop format

using 18-20% w/v Polyethylene glycol 3,350. S181A crystals were collected using the

same process outlined for the wild-type 20β-HSDH.

RESULTS

Bifidobacterium adolescentis expresses 20β-HSDH activity- Based on our

previous phylogenetic analysis (85), we predicted that B. adolescentis strain L2-32

expresses cortisol 20β-HSDH (Figure 4.1). In order to determine whether B. adolescentis

strain L2-32 expresses 20β-HSDH activity, we grew this strain in BHI in the presence of

50 μM cortisol, 20α-dihydrocortisol, or 20β-dihydrocortisol for 48 hours. After 48 hours of

growth, we observed total conversion of cortisol to 20β-dihydrocortisol by B. adolescentis

strain L2-32 (Figure 4.2). When 20β-dihydrocortisol was a substrate, we did not observe

formation of cortisol, suggesting that the enzyme reaction equilibrium greatly favors the

reductive direction under anaerobic conditions. 20α-dihydrocortisol is not a metabolite

for B. adolescentis strain L2-32 or B. desmolans ATCC 43058. As controls, B. desmolans

ATCC 43058 and C. scindens ATCC 35704 were shown to metabolize 20β-

dihydrocortisol and 20α-dihydrocortisol, respectively. 20β-dihydrocortisol is not a

metabolite for C. scindens ATCC 35704. Steroid-17,20-desmolase is present in C.

scindens ATCC 35704 and B. Desmolans 43058. Consistent with this, net conversion of

20α-dihydrocortisol to 11β-OHAD was observed in cultures of C. scindens ATCC 35704,

and net conversion of 20β-dihydrocortisol to 11β-OHAD was observed in cultures

Page 115: Copyright 2018 Seán Martin Mythen

105

Figure 4.1 Reaction mechanism of 20β-HSDH 20β-HSDH oxidizes the coenzyme NADH to transform cortisol into 20β-dihydrocortisol. Cortisol + NADH ⇌ 20β-dihydrocortisol + NAD+. This reaction reduces the double bonded oxygen at C-20 (highlighted in red) on cortisol and is fully reversible.

Page 116: Copyright 2018 Seán Martin Mythen

106

Figure 4.2 Whole cell assays of B. adolescentis strain L2-32, B. desmolans ATCC 43058, and C. scindens ATCC 35704 (A) B. adolescentis completely converts cortisol into 20β-dihydrocortisol. 20α-dihydrocortisol is not metabolized. In a highly reductive environment, B. adolescentis does not convert 20β-dihydrocortisol into cortisol. (B) B. desmolans completely converts cortisol into 11β-OHAD and 20β-dihydrocortisol into 11β-OHAD. 20α-dihydrocortisol is not metabolized. (C) C. scindens completely converts cortisol into 11β-OHAD and 20α-dihydrocortisol into 11β-OHAD. 20β-dihydrocortisol is not metabolized.

Page 117: Copyright 2018 Seán Martin Mythen

107

of B. desmolans ATCC 43058. We did not observe formation of 11β-OHAD by B.

adolescentis strain L2-32 which lacks desA and desB genes, recently reported to encode

steroid-17,20-desmolase activity (167).

Identification and biochemical characterization of recombinant 20β-HSDH- In

our previously reported phylogenetic analysis, we identified a putative 20β-HSDH gene

from B. adolescentis strain L2-32 annotated as an SDR family oxidoreductase

(WP_003810233.1) with amino acid identity of 59% with the characterized 20β-HSDH

from B. desmolans ATCC 43058 (WP_051643274.1) (85). To determine whether the

SDR family oxidoreductase from B. adolescentis encodes a 20β-HSDH, the gene

(WP_003810233.1) was cloned into the pET-28(+) expression vector for overexpression

in E. coli BL21-CodonPlus (DE3)-RIPL competent cells. The N-terminus 6xHis-tagged

recombinant protein was purified via TALON® metal affinity chromatography. The

purified protein was resolved as a single band using SDS-PAGE and three independent

gels show an average subunit molecular mass of 32 ± 0.12 kDa (Figure 4.3A), on par with

the theoretical calculated subunit molecular mass of 31.7 kDa.

The native molecular mass of recombinant 20β-HSDH protein was determined

using gel filtration chromatography and elute at 1.036 Ve/Vo which corresponds to a

molecular mass of 132 ± 3 kDa, eluting shortly after γ-globulin (158 kDa) (Figure

4.3B). The gel filtration data, incorporated with the subunit molecular mass determined

by SDS-PAGE, suggests that the 20β-HSDH forms a homotetrameric quaternary

structure. 20β-HSDH displayed a pH optimum in the reductive direction at a pH of 5.0

and in the oxidative direction at a pH of 5.5 (Figure 4.4A). The Michaelis-Menten

saturation curve for the primary reaction of cortisol reduction performed by 20β-HSDH is

Page 118: Copyright 2018 Seán Martin Mythen

108

Figure 4.3 Purified 20β-hydroxysteroid dehydrogenase (20β-HSDH) from Bifidobacterium adolescentis strain L2-32 (A) SDS-PAGE gel of crude and purified 20β-HSDH showing a subunit size of 32 ± 0.12 kDa. (B) Native molecular size analysis of purified 20β-HSDH via size-exclusion chromatography.

A. B.

Page 119: Copyright 2018 Seán Martin Mythen

109

Figure 4.4 pH optimum and Michaelis-Menten saturation curve of wild-type 20β-HSDH (A) pH optimization of 20β-HSDH in the reductive (blue) and oxidative (red) direction. (B) Saturation curve showing the reduction of cortisol by wild-type 20β-HSDH.

A. B.

Page 120: Copyright 2018 Seán Martin Mythen

110

shown in Figure 4.4B. The saturation curve was used to calculate the kinetic

constants for wild type and mutant 20β-HSDH proteins and are shown in Table 4.1. 20β-

HSDH shows a KM of 24.07 μM. Vmax, kcat, and kcat/KM values were deduced to be 21.08

µmol·min-1·mg-1, 668.3 min-1, and 27.765 μM-1·min-1, respectively. Mutating the predicted

catalytic serine 181 to alanine (S181A) completely abolished activity. Mutating a nearby

serine 183 to alanine (S183) decreased the catalytic efficiency by 44% when compared

to the wild type 20β-HSDH. The site-directed mutants did not affect the secondary

structures of the wild-type protein, as shown by circular dichroism (CD) spectroscopy

(Figure 4.5).

Next, we determined substrate specificity for the wild type 20β-HSDH. Cortisol

was determined to be the primary substrate utilized by 20β-HSDH (Table 4.2). The

enzyme requires NAD(H) as a cofactor, but did not recognize NADP(H). Substrates

identical to cortisol, but lacking an 11β-hydroxyl group showed ~34% loss of relative

activity whereas loss of 17-hydroxyl resulted in ~60% loss in relative activity. Reduction

of the 3-oxo-Δ4,5 to 3α-hydroxy-5β results in loss of ~46% activity suggesting the A/B ring

orientation (trans vs. planar) and C-3 oxygen are recognized by the enzyme. Consistent

with whole-cell experiments, the enzyme displayed negligible activity in the oxidative

direction with NAD+ as co-factor. Substrates with 20α-hydroxyl groups are not substrates.

These results are consistent with substrate-specificity analysis of the purified recombinant

20β-HSDH from B. desmolans ATCC 43058 (85).

Crystallization of 20β-HSDH from Bifidobacterium adolescentis strain L2-32-

Wild type 20β-HSDH and mutant S181A were crystallized using the PHOENIX, followed

Page 121: Copyright 2018 Seán Martin Mythen

111

Tabl

e 4.

1 Ki

netic

par

amet

ers

of c

ortis

ol re

duct

ion

by w

ild-ty

pe a

nd m

utan

t 20β

-HSD

H

Prot

ein

Vm

axK

Mk

cat

Rel

ative

kca

tk

cat/K

MR

elat

ive k

cat/K

M

Wild

-type

21.0

8 +

1.13

124

.07

+ 3.

797

668.

3 +

50.5

3510

027

.765

+ 3

.648

100

S181

A0

00

00

0

S183

A4.

691

+ 0.

097

9.48

5 +

2.59

814

8.7

+ 11

.244

2215

.677

+ 2

.06

56Al

l rea

ctio

ns w

ere

perfo

rmed

in 5

0 m

M s

odiu

m a

ceta

te tr

ihyd

rate

buf

fer (

pH 5

.0) a

nd w

ere

repl

icat

ed a

tleas

t thr

ee ti

mes

kca

t/Km

valu

es a

re in

µM

-1.m

in-1

Rel

ative

kca

t/Km

valu

es a

re s

peci

fied

as a

per

cent

age

of th

e w

iild-

type

Rel

ative

kca

t val

ues

are

spec

ified

as

perc

enta

ge o

f the

wild

-type

kca

t val

ues

are

in m

in-1

Km

valu

es a

re in

µM

Vm

ax v

alue

s ar

e in

µm

ol.m

in-1

.mg-1

Page 122: Copyright 2018 Seán Martin Mythen

112

Figure 4.5 Circular dichroism spectroscopy CD spectra of purified recombinant wild-type 20β-HSDH, and S181A and S183 mutants from Bifidobacterium adolescentis strain L2-32.

Page 123: Copyright 2018 Seán Martin Mythen

113

Tabl

e 4.

2 Su

bstra

te s

peci

ficity

of 2

0β-H

SDH

St

eroi

dTr

ivial

Nam

eC

oenz

yme

Rel

ative

Act

ivity

%

4-pr

egne

n-11

β, 1

7, 2

1-tri

ol-3

, 20-

dion

eC

OR

TISO

LNA

DH

100.

00 ±

0.7

7

4-pr

egne

n-17

, 21-

diol

-3, 2

0-di

one

11-D

ESO

XYC

OR

TISO

LNA

DH

66.0

3 ±

0.75

4-pr

egne

n-11

β, 2

1-di

ol-3

, 20-

dion

eC

OR

TIC

OST

ERO

NENA

DH

39.3

2 ±

0.25

5β-p

regn

an-3

α, 1

1β, 1

7, 2

1-te

trol-2

0-on

eTE

TRAH

YDR

OC

OR

TISO

LNA

DH

44.2

9 ±

0.29

4-pr

egne

n-11

β, 1

7, 2

0β, 2

1-te

trol-3

-one

20β-

DIH

YDR

OC

OR

TISO

LNA

D+

.23

± 0.

04

4-pr

egne

n-17

, 20β

-dio

l-3-o

ne17

α, 2

0β-D

IHYD

RO

XYPR

OG

ESTE

RO

NENA

D+

.09

± 0.

05

4-pr

egne

n-11

β, 2

0β, 2

1-tri

ol-3

-one

20β-

DIH

YDR

OC

OR

TIC

OST

ERO

NENA

D+

NA

4-pr

egne

n-17

, 20α

-dio

l-3-o

ne17

α, 2

0α-D

IHYD

RO

XYPR

OG

ESTE

RO

NENA

D+

NA

4-pr

egne

n-11

β, 1

7, 2

0α, 2

1-te

trol-3

-one

20α-

DIH

YDR

OC

OR

TISO

LNA

D+

NA

NA

= N

o Ac

tivity

Valu

es re

pres

ent m

ean

± s.

e.m

. fro

m th

ree

tech

nica

l rep

licat

es

Page 124: Copyright 2018 Seán Martin Mythen

114

by larger scale hanging-drops (Figure 4.6). For the wild-type, crystals grew in condition

69 of PEGRX containing 20% v/v 2-propanol, 0.1 M MES monohydrate pH 6.0, and 20%

w/v Polyethylene glycol monomethyl ether 2,000. Wild-type 20β-HSDH crystals

appeared within 24-48 hours and were allowed to grow up to 7 days. Crystals were then

overlaid with paraffin oil to prevent evaporation and 20% ethylene glycol as a

cryoprotectant, picked, and immediately stored in liquid nitrogen.

Prior to crystallizing S181A, the pure recombinant S181A was incubated with 2.5

mM NADH for 2 hours. S181A crystals were initially grown in condition 86 of PEG/Ion

containing 0.05 M Citric acid, 0.05 M BIS-TRIS propane / pH 5.0, and 16% w/v

Polyethylene glycol 3,350. S181A was then optimized to grow in condition containing

0.05 M Citric acid, 0.05 M BIS-TRIS propane / pH 5.0, and 18-20% w/v Polyethylene

glycol 3,350. Thin plate crystals appeared within 48 hours and were allowed to grow up

to 14 days. Crystals were then overlaid with paraffin oil to prevent evaporation and 20%

ethylene glycol as a cryoprotectant, picked, and immediately stored in liquid nitrogen.

Wild-type 20β-HSDH and mutant S181A were diffracted at 2.1 and 2.5 Å resolution,

respectively.

DISCUSSION

B. adolescentis is among the first bacteria to colonize the gastrointestinal tract of

breast-fed infants; the species was first isolated from the stool of a healthy two-year-old

infant in 1996 (168, 169). B. adolescentis, is a gram-positive, non-motile, anaerobic

bacterium found among the normal human microbiota. The presence of B. adolescentis

has been associated with numerous health benefits, however, the mechanisms by which

this species and other bifidobacteria confer these benefits on the host are still yet to be

Page 125: Copyright 2018 Seán Martin Mythen

115

Figure 4.6 20β-HSDH crystals (A) Wild-type 20β-HSDH crystals. (B) S181A mutant 20β-HSDH crystals bound to NADH

A.

B.

Page 126: Copyright 2018 Seán Martin Mythen

116

fully understood. Bacterial 20β-hydroxysteroid dehydrogenase (20β-HSDH) is an NADH-

dependent oxidoreductase belonging to the short-chain dehydrogenases/reductases

(SDR) family isolated from B. adolescentis. 20β-HSDH catalyzes the reduction of the

carboxyl group at position 20 of the steroid cortisol, thereby producing 20β-

dihydrocortisol. 20β-HSDH plays an important role in steroid hormone and bile acid

metabolism in the gut.

There are currently 31 known Bifidobacterium spp. (69). Of those species, only a

select few strains exhibit this potential probiotic 20β-HSDH activity. Based on a non-

redundant protein search from a recently characterized 20β-HSDH (WP_051643274.1)

from B. desmolans ATCC 43058, a gene annotated to encode an SDR family

oxidoreductase (WP_003810233.1) with 59% homology to the 20β-HSDH gene from B.

desmolans ATCC 43058 was identified in B. adolescentis strain L2-32 (87). The 20β-

HSDH gene from B. adolescentis strain L2-32 was cloned, overexpressed, and purified

from E. coli BL21-CodonPlus (DE3)-RIPL competent cells. Using whole cell and

enzymatic assays, the microbe and purified protein were shown to have 20β-HSDH

activity identical to that of 20β-HSDH found in B. desmolans ATCC 43058.

We mutated the predicted catalytic serine 181 to alanine (S181A) and activity was

completely ablated. This allowed for crystallization studies that would bind NADH and

cortisol, but would not convert them to NAD+ and 20β-dihydrocortisol, respectively. We

also mutated serine 183 to alanine (S183A) which is one residues away from the catalytic

serine 181. This mutation decreased the turnover rate by 5x, but decreased the KM about

3x. Meaning that without serine 183, substrate is able to bind easier. It is predicted that

Page 127: Copyright 2018 Seán Martin Mythen

117

serine 183 is important in hydrogen binding. Overall, the catalytic efficiency of S183A is

only 64% as efficient as wild-type 20β-HSDH.

The purified recombinant 20β-HSDH and co-factor-bound mutant S181A from B.

adolescentis strain L2-32 were crystallized after screening 1,250 unique conditions. The

crystal structure of 20β-HSDH in apo and holo form will allow us to visualize secondary

structure changes when co-factor binds to the enzyme. This information brings us one

step closer to predicting the reaction mechanism by which 20β-HSDH proceeds and the

residues involved in the reaction. It will also show us the exact location NADH binds to

the enzyme. The next step is to crystalize the ternary complex of 20β-HSDH which

contains co-factor (NADH) and substrate (cortisol), elucidating the location cortisol binds

on the enzyme and what conformational changes are involved. HSDHs follow an ordered

bi-bi mechanism where the co-factor (NAD(H) or NADP(H)) binds to the enzyme first,

followed by the substrate. Once the reaction is completed, the product leaves first, and

the co-factor leaves last. It is predicted that 20β-HSDH follows the same ordered reaction

mechanism. Isothermal titration calorimetry will be performed to test this hypothesis.

There are many B. adolescentis strains, however, only a select few express 20β-

HSDH activity. When B. adolescentis strain L2-32 metabolizes cortisol utilizing 20β-

HSDH, it is predicted to compete with cortisol metabolism by steroid-17,20-desmolase

from C. scindens and 21-dehydroxylation from E. lenta. The reason is simple: there are

three alternative pathways for side chain metabolism of cortisol by gut microbes. If the

rate of 20β-dihydrocortisol formation is greater than the rate of steroid-17,20-desmolase

or 21-dehydroxylase, then these pathways are prevented because 20β-dihydrocortisol is

not a substrate for either of these enzymes and there is no way for them to utilize cortisol

Page 128: Copyright 2018 Seán Martin Mythen

118

while it is in that form. By understanding 20β-HSDH and how it interacts with other steroid

metabolizing pathways, we may be able to alter the composition of toxic steroid hormones

and bile acids.

SUMMARY OF RESEARCH

In this work, we utilized targeted gene synthesis to clone and heterologously

express genes from a gene-cluster (BN592_00768-BN592_00770) in Eggerthella

CAG:298, hypothesized to encode multiple HSDHs (CDD59473, CDD59474,

CDD59475). Characterization of the purified recombinant enzymes revealed that each

enzyme catalyzed a distinct oxidation/reduction of bile acid hydroxyl groups. CDD59473

catalyzed NADPH-dependent reduction of 3-oxoDCA yielding iso-DCA (3β-hydroxyl) as

the main reaction product, co-migrating on TLC and with mass spectrum consistent with

iso-DCA. The formation of 3β-hydroxyl decreases the amphipathic nature of bile acids

making them less effective detergents, and also less toxic to bacteria as well as host cells

(26), (128, 129). Currently, there are no structures reported for gut bacterial 3β-HSDH.

Our phylogenetic analysis identified multiple Lactobacillus spp. isolated from diverse

avian and mammalian GI tracts encoding putative 3β-HSDHs with sequence identities at

or below 71%. A similar targeted gene-synthesis approach would be useful in

determining the function and catalytic potential of these putative 3β-HSDHs.

Isomerization at C3 of primary bile acids may be important in precluding bile acid 7α-

dehydroxylation, a microbial biochemical pathway responsible for formation of DCA and

LCA, bile acids that are causally associated with cancers of the colon and liver (2).

In contrast to CDD59473, CDD59474 converted 3-oxoDCA to DCA and thus

BN592_00769 encodes 3α-HSDH. DCA is generated by the multi-step bile acid 7α-

Page 129: Copyright 2018 Seán Martin Mythen

119

dehydroxylation of CA by a few species of intestinal clostridia (2) and is found in fecal

water at ~50 µM (130) but can reach several hundred micromolar (25). CDD59473 and

CDD59474 work in concert to epimerize DCA to iso-DCA, potentially as a means of

detoxification in E. lenta. Interestingly, CDD59474 was clustered with 3α-HSDH from R.

gnavus (Rumgna_02133) but more distantly related to a 3α-HSDH characterized in E.

lenta DSM 2243T (Elen_0690) (26). E. lenta has been reported to express 7α-HSDH (76,

77), however, genes encoding 7α-HSDH in E. lenta have yet to be identified. CDD59474

was annotated as “7α-HSDH” due to sequence homology, yet expression and

characterization of this protein revealed specificity for 3α-hydroxyl groups providing a

cautionary tale and further reinforcing the need to functionally characterize genes

predicted to encode particular functions by annotation. To our knowledge, BN592_00770

is the only reported sequence so far demonstrated to encode 12α-HSDH (CDD59475) in

E. lenta, and sequence comparison with CDD59475 is expected to elucidate further gut

bacterial 12α-HSDHs.

Thus, our analysis demonstrates that E. lenta strains harbor distinct forms of 3α-

HSDH, 3β-HSDH and 12α-HSDH enzymes. Critical at the intersection between next-

generation sequencing data (metagenomics, metatranscriptomics) and metabolomics is

biochemistry, providing the ability to experimentally demonstrate that a particular amino

acid sequence catalyzes a particular biochemical reaction. A major task in gut

microbiology research today is linking nucleic and amino acid sequences in databases to

function, and determining how functions affect microbiome structure and host physiology.

A major question in the field of bile acid metabolism (132) that we are currently working

to address is why bacteria, such as E. lenta, encode numerous HSDHs and oxidize bile

Page 130: Copyright 2018 Seán Martin Mythen

120

acids in an anaerobic environment.

We also report that the desA and desB genes from C. scindens ATCC 35704

encode steroid-17,20-desmolase. Recombinant DesAB displayed substrate-specificity

consistent with native steroid-17,20-desmolase: the enzyme recognizes 20-keto-17,21-

dihydroxy steroids (67, 68). C. scindens ATCC 35704 expresses NAD(H)-dependent 20α-

HSDH (69), while B. desmolans, C. cadavaris, and P. lymphophilum express 20β-HSDH

(85) in addition to steroid-17,20-desmolase. Oxidation-reduction of the C-20 oxygen is

predicted to function as a regulatory “switch”, interconverting cortisol between DesAB

substrate and non-substrate forms.

The conversion of cortisol to 11β-hydroxyandrostenedione by C. scindens is an

important step in a multi-species gut microbial endocrine pathway that results in the

formation of 11β-hydroxyandrogens (69). E. lenta expresses 21-dehydroxylase (74), and

21-dehydroxylated metabolites are detected following incubation of cortisol in fecal

suspensions (63, 64). Due to the absolute requirement for the C21-hydroxyl group,

steroid-17,20-desmolase and cortisol 21-dehydroxylation by Eggerthella lenta are

competing reactions in the gut (74).

Extrapolations from radiometric studies in non-human primates (142) and data

from humans (143) suggest that 15% of the 15 mg of cortisol are detected in the gut.

Based on these studies, the concentration of cortisol metabolites in the ~0.4L human

colon (144) is ~15 µM. 11β-hydroxyandrostenedione is now viewed as a pro-androgen,

whose downstream 5α-, 17β-reduced metabolites, are now thought to play an important

role in activation of androgen receptor (33, 145). The field of microbial endocrinology has

emerged and recognizes the need to characterize androgen production by C. scindens

Page 131: Copyright 2018 Seán Martin Mythen

121

(38, 146–150). Many host immune cells express androgen-receptor (151, 152), and it is

possible that C. scindens affects immune function through generation of androgens.

Androgens directly affect bacterial growth and virulence (153, 154). Intriguingly, steroids

have been shown to influence C. difficile germination (3). 11β-hydroxy-androgens are

also inhibitors of the host enzyme 11β-HSD2 (80, 155), but in kidney and vascular smooth

muscle may be pro-hypertensive (155). Inhibition of 11β-HSD2 in colonocytes may be

protective against colorectal cancer (156). Prostate tissue contains 11β-

hydroxyandrogens despite inhibition of host enzymes involved in synthesizing androgens

(chemical castration) (33, 145). The role of 11β-OHAD formation by the gut microbiota in

androgen accumulation in the prostate is unknown; however, we discovered that bacteria

isolated from urine also express the bacterial steroid-17,20-desmolase pathway (85), and

the majority of cortisol excretion from the body is via the urine (142, 143). Of note, we

recently reported identification of the steroid-17,20-desmolase gene cluster in the

genomes of urinary bacterial isolates and confirmed steroid-17,20-desmolase activity in

an isolate of Propionimicrobium lymphophylum (85). Verification that desA and desB

genes encode steroid-17,20-desmolase is an important step towards the application of

nucleic acid-based approaches such as quantitative polymerase chain, and metagenomic

sequencing in determining whether correlations exist between levels/expression of

desAB and host phenotypes and disease states.

Finally, we characterized and crystallized purified recombinant 20β-HSDH from B.

adolescentis strain L2-32 in the apo and holo forms. This will allow for predicting of the

reaction mechanism by which 20β-HSDH proceeds and the residues involved in the

reaction. It will also show us the exact location NADH binds to the enzyme. The next

Page 132: Copyright 2018 Seán Martin Mythen

122

step is to crystalize the ternary complex of 20β-HSDH which contains co-factor (NADH)

and substrate (cortisol), elucidating the location cortisol binds on the enzyme and what

conformational changes are involved. As discussed, there are many B. adolescentis

strains, however, only a select few express 20β-HSDH activity. When B. adolescentis

strain L2-32 metabolizes cortisol utilizing 20β-HSDH, it is predicted to compete with

cortisol metabolism by steroid-17,20-desmolase from C. scindens and 21-

dehydroxylation from E. lenta. By understanding 20β-HSDH and how it interacts with

other steroid metabolizing pathways, we may be able to alter the composition of toxic

steroid hormones and bile acids (Figure 5.1).

Page 133: Copyright 2018 Seán Martin Mythen

123

Figure 4.7 Competing cortisol metabolizing reactions by B. adolescentis, E. lenta, B. desmolans, and C. scindens

Page 134: Copyright 2018 Seán Martin Mythen

124

FUTURE RESEARCH

This research has laid the groundwork for future studies aiming to understand the

enzymes and pathways involved in steroid metabolism by gut microbes. The majority of

steroid biotransforming enzymes that gut microbes employ have yet to be explored. The

characterization of the steroid-17,20-desmolase and various other important HSDHs is

only one small piece of the puzzle. Microbes within the gut form complex interactions,

resulting in a myriad of steroid metabolites in different proportions and concentrations

which affect the health of the host. A major task in gut microbiology research today is

linking nucleic and amino acid sequences in databases to functions, and determining how

functions affect microbiome structure and host physiology. By experimentally

demonstrating that a particular amino acid sequence catalyzes a particular biochemical

reaction we get one step closer to predicting the biochemical potential of the gut

microbiome.

Future short-term studies will aim to further characterize novel steroid metabolizing

enzymes. We will crystallize steroid-17,20-desmolase from C. scindens ATCC 35704

and work to further elucidate the structure and catalytic mechanism of steroid

transketolation. Currently there are no crystal structures for 3β-, and 12α- Hydroxysteroid

dehydrogenases. By identifying the 3α-, 3β-, and 12α-hydroxysteroid dehydrogenases

from Eggerthella CAG:298 we can begin to crystallize these enzymes to gain structural

insight into the stereo- and regio-specificity of HSDHs. The structure of 20β-HSDH from

B. adolescentis strain L2-32 will be determined using the diffraction data collected from

the apo and holo enzyme crystals obtained in these studies. The 20β-HSDH from B.

adolescentis strain L2-32 will also be crystallized in the ternary form to determine binding

Page 135: Copyright 2018 Seán Martin Mythen

125

locations of substrate to analyze important conformational changes. We will perform

isothermal titration calorimetry to determine the binding order and affinity of ligands for

the serine active site mutant S181A. Our existing collaboration with the Theoretical and

Computational Biophysics group at the Beckman institute on campus will facilitate the

development of catalytic models based on the crystal structure and molecular dynamics.

A major long-term goal in the field of microbial endocrinology is to be able to

engineer microbiotas to generate steroid end products associated with health to reduce

the concentration of metabolites deemed to harmful to the hosts. The application of germ-

free animals and defined microbial communities will be utilized to determine how this will

be accomplished.

Page 136: Copyright 2018 Seán Martin Mythen

126

REFERENCES

1. O’Hara, A. M., and Shanahan, F. (2006) The gut flora as a forgotten organ.

EMBO Rep. 7, 688–693

2. Ridlon, J. M., and Bajaj, J. S. (2015) The human gut sterolbiome: bile acid-

microbiome endocrine aspects and therapeutics. Acta Pharm. Sin. B. 5, 99–105

3. Liggins, M., Ramirez, N., Magnuson, N., and Abel-Santos, E. (2011)

Progesterone analogs influence germination of Clostridium sordellii and

Clostridium difficile spores in vitro. J. Bacteriol. 193, 2776–83

4. Maser, E., and Lanišnik Rižner, T. (2012) Steroids and microorganisms. J. Steroid

Biochem. Mol. Biol. 129, 1–3

5. Holst, J. P., Soldin, O. P., Guo, T., and Soldin, S. J. (2004) Steroid hormones:

Relevance and measurement in the clinical laboratory. Clin. Lab. Med. 24, 105–

118

6. Norman, A. W., and Henry, H. L. (2015) Steroid Hormones. in Hormones, pp. 27–

53, Elsevier, 10.1016/B978-0-08-091906-5.00002-1

7. Gustagsson, B. E., Norman, A., and Sjovall, J. (1960) Influence of E. coli infection

on turnover and metabolism of cholic acid in germ-free rats. Arch. Biochem.

Biophys. 91, 93–100

8. Macdonald, I. A., Bokkenheuser, V. D., Winter, J., McLernon, A. M., and

Mosbach, E. H. (1983) Degradation of steroids in the human gut. J. Lipid Res. 24,

675–700

9. Chapman, K., Holmes, M., and Seckl, J. (2013) 11Β-Hydroxysteroid

Dehydrogenases: Intracellular Gate-Keepers of Tissue Glucocorticoid Action.

Page 137: Copyright 2018 Seán Martin Mythen

127

Physiol. Rev. 93, 1139–206

10. Persson, B., and Kallberg, Y. (2013) Classification and nomenclature of the

superfamily of short-chain dehydrogenases/reductases (SDRs). Chem. Biol.

Interact. 202, 111–115

11. Oppermann, U., Filling, C., Hult, M., Shafqat, N., Wu, X., Lindh, M., Shafqat, J.,

Nordling, E., Kallberg, Y., Persson, B., and Jörnvall, H. (2003) Short-chain

dehydrogenases/reductases (SDR): The 2002 update. Chem. Biol. Interact. 143–

144, 247–253

12. Kisiela, M., Skarka, A., Ebert, B., and Maser, E. (2012) Hydroxysteroid

dehydrogenases (HSDs) in bacteria: a bioinformatic perspective. J. Steroid

Biochem. Mol. Biol. 129, 31–46

13. Watanabe, M., Fukiya, S., and Yokota, A. (2017) Comprehensive evaluation of

the bactericidal activities of free bile acids in the large intestine of humans and

rodents. J. Lipid Res. 58, 1143–1152

14. Hammond, G. L. (1997) Determinants of steroid hormone bioavailability. Biochem.

Soc. Trans. 25, 577–82

15. Hammes, A., Andreassen, T. K., Spoelgen, R., Raila, J., Hubner, N., Schulz, H.,

Metzger, J., Schweigert, F. J., Luppa, P. B., Nykjaer, A., and Willnow, T. E. (2005)

Role of endocytosis in cellular uptake of sex steroids. Cell. 122, 751–762

16. Bury, N. R., and Sturm, A. (2007) Evolution of the corticosteroid receptor

signalling pathway in fish. Gen. Comp. Endocrinol. 153, 47–56

17. Whirledge, S., and Cidlowski, J. A. (2010) Glucocorticoids, stress, and fertility.

Minerva Endocrinol. 35, 109–125

Page 138: Copyright 2018 Seán Martin Mythen

128

18. Harris, A., and Seckl, J. (2011) Glucocorticoids, prenatal stress and the

programming of disease. Horm. Behav. 59, 279–289

19. Rabin, D., Gold, P. W., Margioris, A. N., and Chrousos, G. P. (1988) Stress and

reproduction: physiologic and pathophysiologic interactions between the stress

and reproductive axes. Adv. Exp. Med. Biol. 245, 377–87

20. Collu, R., Gibb, W., and Ducharme, J. R. (1984) Effects of stress on the gonadal

function. J. Endocrinol. Invest. 7, 529–537

21. Sandor, T., Fazekas, A., and Robinson, B. (1976) The biosynthesis of

corticosteroids throughout the vertebrates. Gen. Comp. Clin. Endocrinol. Adrenal

Cortex. 1, 25–142

22. Morris, D. J. (2015) Why do humans have two glucocorticoids: A question of

intestinal fortitude. Steroids. 102, 32–38

23. Björntorp, P., and Rosmond, R. (2000) Obesity and cortisol. Nutrition. 16, 924–

936

24. Hewagalamulage, S. D., Lee, T. K., Clarke, I. J., and Henry, B. A. (2016) Stress,

cortisol, and obesity: a role for cortisol responsiveness in identifying individuals

prone to obesity. Domest. Anim. Endocrinol. 56, S112–S120

25. Hamilton, J. P., Xie, G., Raufman, J.-P., Hogan, S., Griffin, T. L., Packard, C. A.,

Chatfield, D. A., Hagey, L. R., Steinbach, J. H., and Hofmann, A. F. (2007)

Human cecal bile acids: concentration and spectrum. Am. J. Physiol. Gastrointest.

Liver Physiol. 293, G256-63

26. Devlin, A. S., and Fischbach, M. A. (2015) A biosynthetic pathway for a prominent

class of microbiota-derived bile acids. Nat. Chem. Biol. 11, 685–90

Page 139: Copyright 2018 Seán Martin Mythen

129

27. Fuller, P. J., Yao, Y., Yang, J., and Young, M. J. (2012) Mechanisms of ligand

specificity of the mineralocorticoid receptor. J. Endocrinol. 213, 15–24

28. Nuzzi, R., Scalabrin, S., Becco, A., and Panzica, G. (2018) Gonadal Hormones

and Retinal Disorders: A Review. Front. Endocrinol. (Lausanne). 9, 66

29. Simpson, E. R. (2003) Sources of estrogen and their importance. J. Steroid

Biochem. Mol. Biol. 86, 225–230

30. Melcangi, R. C., Panzica, G., and Garcia-Segura, L. M. (2011) Neuroactive

steroids: Focus on human brain. Neuroscience. 191, 1–5

31. Roy, a K., and Chatterjee, B. (1995) Androgen action., 10.1007/978-1-61779-

243-4

32. Grino, P. B., Griffin, J. E., and Wilson, J. D. (1990) Testosterone at high

concentrations interacts with the human androgen receptor similarly to

dihydrotestosterone. Endocrinology. 126, 1165–1172

33. Bloem, L. M., Storbeck, K.-H., Schloms, L., and Swart, A. C. (2013) 11β-

hydroxyandrostenedione returns to the steroid arena: biosynthesis, metabolism

and function. Molecules. 18, 13228–44

34. Pretorius, E., Arlt, W., and Storbeck, K. H. (2017) A new dawn for androgens:

Novel lessons from 11-oxygenated C19 steroids. Mol. Cell. Endocrinol. 441, 76–

85

35. Azziz, R., Black, V., Hines, G. A., Fox, L. M., and Boots, L. R. (1998) Adrenal

androgen excess in the polycystic ovary syndrome: Sensitivity and responsivity of

the hypothalamic-pituitary-adrenal axis. J. Clin. Endocrinol. Metab. 83, 2317–2323

36. Baker, J. M., Al-Nakkash, L., and Herbst-Kralovetz, M. M. (2017) Estrogen–gut

Page 140: Copyright 2018 Seán Martin Mythen

130

microbiome axis: Physiological and clinical implications. Maturitas. 103, 45–53

37. Clark, M. A., Harvey, R. A., Finkel, R., Rey, J. A., and Whalen, K. (2011)

Pharmacology, Illustrated Reviews, Wolters Kluwer Health

38. Neuman, H., Debelius, J. W., Knight, R., and Koren, O. (2015) Microbial

endocrinology: the interplay between the microbiota and the endocrine system.

FEMS Microbiol. Rev. 39, 509–21

39. Lieberman, A., and Curtis, L. (2017) In Defense of Progesterone: A Review of the

Literature. Altern. Ther. Heal. Med. 21, 21

40. Houten, S. M., Watanabe, M., and Auwerx, J. (2006) Endocrine functions of bile

acids. EMBO J. 25, 1419–1425

41. Dawson, P. A. (2011) Role of the intestinal bile acid transporters in bile acid and

drug disposition. Handb. Exp. Pharmacol. 201, 169–203

42. Ajouz, H., Mukherji, D., and Shamseddine, A. (2014) Secondary bile acids: An

underrecognized cause of colon cancer. World J. Surg. Oncol. 12, 1–5

43. Hardison, W. G. (1978) Hepatic taurine concentration and dietary taurine as

regulators of bile acid conjugation with taurine. Gastroenterology. 75, 71–5

44. Hofmann, A. F., and Hagey, L. R. (2008) Bile acids: Chemistry, pathochemistry,

biology, pathobiology, and therapeutics. Cell. Mol. Life Sci. 65, 2461–2483

45. Gérard, P. (2013) Metabolism of Cholesterol and Bile Acids by the Gut Microbiota.

Pathogens. 3, 14–24

46. Chiang, J. (2013) Bile Acid Metabolism and Signalling. Compr Physiol. 3, 1191–

1212

47. Ridlon, J., Kang, D., Hylemon, P., and Bajaj, J. (2014) Bile acids and the Gut

Page 141: Copyright 2018 Seán Martin Mythen

131

Microbiome. Curr. Opin. Gastroenterol. 30, 332–338

48. Hofmann, A. F., Cravetto, C., Molino, G., Belforte, G., and Bona, B. (1987)

Simulation of the metabolism and enterohepatic circulation of endogenous

deoxycholic acid in humans using a physiologic pharmacokinetic model for bile

acid metabolism. Gastroenterology. 93, 693–709

49. Zwicker, B. L., and Agellon, L. B. (2013) Transport and biological activities of bile

acids. Int. J. Biochem. Cell Biol. 45, 1389–1398

50. Midtvedt, T. (1974) Microbial bile acid transformation. Am. J. Clin. Nutr. 27, 1341–

7

51. Hofmann, A. (1999) The Continuing Importance of Bile Acids in Liver and

Intestinal Disease. Arch Intern Med. 159, 2647–2658

52. Evans, J. M., Morris, L. S., and Marchesi, J. R. (2013) The gut microbiome: The

role of a virtual organ in the endocrinology of the host. J. Endocrinol.

10.1530/JOE-13-0131

53. Wang, H., Chen, J., Hollister, K., Sowers, L. C., and Forman, B. M. (1999)

Endogenous bile acids are ligands for the nuclear receptor FXR/BAR. Mol. Cell. 3,

543–553

54. Inagaki, T., Moschetta, A., Lee, Y.-K., Peng, L., Zhao, G., Downes, M., Yu, R. T.,

Shelton, J. M., Richardson, J. A., Repa, J. J., Mangelsdorf, D. J., and Kliewer, S.

A. (2006) Regulation of antibacterial defense in the small intestine by the nuclear

bile acid receptor. Proc. Natl. Acad. Sci. 103, 3920–3925

55. Claudel, T., Staels, B., and Kuipers, F. (2005) The Farnesoid X receptor: A

molecular link between bile acid and lipid and glucose metabolism. Arterioscler.

Page 142: Copyright 2018 Seán Martin Mythen

132

Thromb. Vasc. Biol. 25, 2020–2031

56. Pols, T. W. H., Noriega, L. G., Nomura, M., Auwerx, J., and Schoonjans, K.

(2011) The bile acid membrane receptor TGR5: A valuable metabolic target. Dig.

Dis. 29, 37–44

57. Kawamata, Y., Fujii, R., Hosoya, M., Harada, M., Yoshida, H., Miwa, M.,

Fukusumi, S., Habata, Y., Itoh, T., Shintani, Y., Hinuma, S., Fujisawa, Y., and

Fujino, M. (2003) A G protein-coupled receptor responsive to bile acids. J. Biol.

Chem. 278, 9435–40

58. Holdeman, L. V., Good, I. J., and Moore, W. E. (1976) Human fecal flora: variation

in bacterial composition within individuals and a possible effect of emotional

stress. Appl. Environ. Microbiol. 31, 359–75

59. Marchesi, J. R., Adams, D. H., Fava, F., Hermes, G. D. A., Hirschfield, G. M.,

Hold, G., Quraishi, M. N., Kinross, J., Smidt, H., Tuohy, K. M., Thomas, L. V.,

Zoetendal, E. G., and Hart, A. (2016) The gut microbiota and host health: A new

clinical frontier. Gut. 65, 330–339

60. Qin, J., Li, R., Raes, J., Arumugam, M., Burgdorf, S., Manichanh, C., Nielsen, T.,

Pons, N., Yamada, T., Mende, D. R., Li, J., Xu, J., Li, S. S. S. S. S., Li, D., Cao,

J., Wang, B., Liang, H., Zheng, H., Xie, Y., Tap, J., Lepage, P., Bertalan, M.,

Batto, J., Hansen, T., Paslier, D. Le, Linneberg, A., Nielsen, H. B., Pelletier, E.,

Renault, P., Zhou, Y., Li, Y., Zhang, X., Li, S. S. S. S. S., Qin, N., Yang, H.,

Burgdorf, K. S., Manichanh, C., Nielsen, T., Pons, N., Levenez, F., Yamada, T.,

Mende, D. R., Li, J., Xu, J., Li, S. S. S. S. S., Li, D., Cao, J., Wang, B., Liang, H.,

Zheng, H., Xie, Y., Tap, J., Lepage, P., Bertalan, M., Batto, J., Hansen, T., Le

Page 143: Copyright 2018 Seán Martin Mythen

133

Paslier, D., Linneberg, A., Nielsen, H. B., Pelletier, E., Renault, P., Sicheritz-

Ponten, T., Turner, K., Zhu, H., Yu, C., Li, S. S. S. S. S., Jian, M., Zhou, Y., Li, Y.,

Zhang, X., Li, S. S. S. S. S., Qin, N., Yang, H., Wang, J. J., Brunak, S., Doré, J.,

Guarner, F., Kristiansen, K., Pedersen, O., Parkhill, J., Weissenbach, J., Bork, P.,

Ehrlich, S. D., Wang, J. J., Burgdorf, S., Manichanh, C., Nielsen, T., Pons, N.,

Yamada, T., Mende, D. R., Li, J., Xu, J., Li, S. S. S. S. S., Li, D., Cao, J., Wang,

B., Liang, H., Zheng, H., Xie, Y., Tap, J., Lepage, P., Bertalan, M., Batto, J.,

Hansen, T., Paslier, D. Le, Linneberg, A., Nielsen, H. B., Pelletier, E., Renault, P.,

Zhou, Y., Li, Y., Zhang, X., Li, S. S. S. S. S., Qin, N., and Yang, H. (2010) A

human gut microbial gene catalog established by metagenomic sequencing.

Nature. 464, 59–65

61. Nabarro, J. D. N., Moxham, A., Walker, G., and Slater, J. D. H. (1957) Rectal

Hydrocortisone. BMJ. 2, 272–274

62. WADE, A. P., SLATER, J. D. H., KELLIE, A. E., and HOLLIDAY, M. E. (1959)

URINARY EXCRETION OF 17-KETOSTEROIDS FOLLOWING RECTAL

INFUSION OF CORTISOL. J. Clin. Endocrinol. Metab. 19, 444–453

63. Winter, J., and Bokkenheuser, V. D. (1978) 21-dehydroxylation of corticoids by

anaerobic bacteria isolated from human fecal flora. J. Steroid Biochem. 9, 379–

384

64. Cerone-McLernon, A. M., Winter, J., Mosbach, E. H., and Bokkenheuser, V. D.

(1981) Side-chain cleavage of cortisol by fecal flora. Biochim. Biophys. Acta -

Lipids Lipid Metab. 666, 341–347

65. Bokkenheuser, V. D., Morris, G. N., Ritchie, A. E., Holdeman, L. V, and Winter, J.

Page 144: Copyright 2018 Seán Martin Mythen

134

(1984) Biosynthesis of androgen from cortisol by a species of Clostridium

recovered from human fecal flora. J. Infect. Dis. 149, 489–94

66. Winter, J., Morris, G. N., O’Rourke-Locascio, S., Bokkenheuser, V. D., Mosbach,

E. H., Cohen, B. I., and Hylemon, P. B. (1984) Mode of action of steroid

desmolase and reductases synthesized by Clostridium “scindens” (formerly

Clostridium strain 19). J. Lipid Res. 25, 1124–31

67. Krafft, A. E., Winter, J., Bokkenheuser, V. D., and Hylemon, P. B. (1987) Cofactor

requirements of steroid-17-20-desmolase and 20 alpha-hydroxysteroid

dehydrogenase activities in cell extracts of Clostridium scindens. J. Steroid

Biochem. 28, 49–54

68. Krafft, a E., and Hylemon, P. B. (1989) Purification and characterization of a

novel form of 20 alpha-hydroxysteroid dehydrogenase from Clostridium scindens.

J. Bacteriol. 171, 2925–32

69. Ridlon, J. M., Ikegawa, S., Alves, J. M. P., Zhou, B., Kobayashi, A., Iida, T.,

Mitamura, K., Tanabe, G., Serrano, M., De Guzman, A., Cooper, P., Buck, G. a,

and Hylemon, P. B. (2013) Clostridium scindens : a human gut microbe with a

high potential to convert glucocorticoids into androgens. J. Lipid Res. 54, 2437–

2449

70. Kageyama, A., Benno, Y., and Nakase, T. (1999) Phylogenetic evidence for the

transfer of Eubacterium lentum to the genus Eggerthella as Eggerthella lenta gen.

nov., comb. nov. Int. J. Syst. Bacteriol. 49 Pt 4, 1725–32

71. Bokkenheuser, V. D., Winter, J., Finegold, S. M., Sutter, V. L., Ritchie, A. E.,

Moore, W. E., and Holdeman, L. V (1979) New markers for Eubacterium lentum.

Page 145: Copyright 2018 Seán Martin Mythen

135

Appl. Environ. Microbiol. 37, 1001–1006

72. Saha, J. R., Butler, V. P., Neu, H. C., and Lindenbaum, J. (1983) Digoxin-

inactivating bacteria: identification in human gut flora. Science. 220, 325–7

73. Haiser, H. J., Gootenberg, D. B., Chatman, K., Sirasani, G., Balskus, E. P., and

Turnbaugh, P. J. (2013) Predicting and manipulating cardiac drug inactivation by

the human gut bacterium Eggerthella lenta(). Science. 341, 295–298

74. Feighner, S. D., and Hylemon, P. B. (1980) Characterization of a corticosteroid

21-dehydroxylase from the intestinal anaerobic bacterium, Eubacterium lentum. J.

Lipid Res. 21, 585–93

75. Wegner, K., Just, S., Gau, L., Mueller, H., Gérard, P., Lepage, P., Clavel, T., and

Rohn, S. (2017) Rapid analysis of bile acids in different biological matrices using

LC-ESI-MS/MS for the investigation of bile acid transformation by mammalian gut

bacteria. Anal. Bioanal. Chem. 409, 1231–1245

76. MacDonald, I. A., Jellett, J. F., Mahony, D. E., and Holdeman, L. V (1979) Bile salt

3 alpha- and 12 alpha-hydroxysteroid dehydrogenases from Eubacterium lentum

and related organisms. Appl. Environ. Microbiol. 37, 992–1000

77. Edenharder, R., and Mielek, K. (1984) Epimerization, oxidation and reduction of

bile acids by Eubacterium lentum. Syst. Appl. Microbiol. 5, 287–298

78. Claus, S. P., Ellero, S. L., Berger, B., Krause, L., Bruttin, A., Molina, J., Paris, A.,

Want, E. J., de Waziers, I., Cloarec, O., Richards, S. E., Wang, Y., Dumas, M.-E.,

Ross, A., Rezzi, S., Kochhar, S., Van Bladeren, P., Lindon, J. C., Holmes, E., and

Nicholson, J. K. (2011) Colonization-induced host-gut microbial metabolic

interaction. MBio. 2, e00271-10

Page 146: Copyright 2018 Seán Martin Mythen

136

79. Morris, D. J., and Souness, G. W. (1996) Endogenous 11 beta-hydroxysteroid

dehydrogenase inhibitors and their role in glucocorticoid Na+ retention and

hypertension. Endocr. Res. 22, 793–801

80. Morris, D. J., Latif, S. A., Hardy, M. P., and Brem, A. S. (2007) Endogenous

inhibitors (GALFs) of 11β-hydroxysteroid dehydrogenase isoforms 1 and 2:

Derivatives of adrenally produced corticosterone and cortisol. J. Steroid Biochem.

Mol. Biol. 104, 161–168

81. Ferrari, P. (2010) The role of 11β-hydroxysteroid dehydrogenase type 2 in human

hypertension. Biochim. Biophys. Acta. 1802, 1178–87

82. Persson, B., Hedlund, J., and Jörnvall, H. (2008) Medium- and short-chain

dehydrogenase/reductase gene and protein families : the MDR superfamily. Cell.

Mol. Life Sci. 65, 3879–94

83. Takada, T., Watanabe, K., Makino, H., and Kushiro, A. (2016) Reclassification of

Eubacterium desmolans as Butyricicoccus desmolans comb. nov., and

description of Butyricicoccus faecihominis sp. nov., a butyrate-producing

bacterium from human faeces. Int. J. Syst. Evol. Microbiol. 66, 4125–4131

84. Bokkenheuser, V. D., Winter, J., Morris, G. N., and Locascio, S. (1986) Steroid

Desmolase Synthesis by Eubacterium Desmolans and Clostridium Cadavaris.

Appl. Environ. Microbiol. 52, 1153–1156

85. Devendran, S., Méndez-García, C., and Ridlon, J. M. (2017) Identification and

characterization of a 20β-HSDH from the anaerobic gut bacterium Butyricicoccus

desmolans ATCC 43058. J. Lipid Res. 58, 916–925

86. Winter, J., Cerone-McLernon, A., O’Rourke, S., Ponticorvo, L., and

Page 147: Copyright 2018 Seán Martin Mythen

137

Bokkenheuser, V. D. (1982) Formation of 20β-dihydrosteroids by anaerobic

bacteria. J. Steroid Biochem. 17, 661–667

87. YOSHIOKA, H., FUJITA, K., SAKATA, H., MURONO, K., and ISEKI, K. (1991)

Development of the Normal Intestinal Flora and Clinical Significance in Infants

and Children. Bifidobact. Microflora. 10, 11–17

88. Joice, R., Yasuda, K., Shafquat, A., Morgan, X. C., and Huttenhower, C. (2014)

Determining microbial products and identifying molecular targets in the human

microbiome. Cell Metab. 20, 731–41

89. Turnbaugh, P. J., Ley, R. E., Hamady, M., Fraser-Liggett, C. M., Knight, R., and

Gordon, J. I. (2007) The Human Microbiome Project. Nature. 449, 804

90. Qin, J., Li, R., Raes, J., Arumugam, M., Burgdorf, K. S., Manichanh, C., Nielsen,

T., Pons, N., Levenez, F., Yamada, T., Mende, D. R., Li, J., Xu, J., Li, S., Li, D.,

Cao, J., Wang, B., Liang, H., Zheng, H., Xie, Y., Tap, J., Lepage, P., Bertalan, M.,

Batto, J.-M., Hansen, T., Le Paslier, D., Linneberg, A., Nielsen, H. B., Pelletier, E.,

Renault, P., Sicheritz-Ponten, T., Turner, K., Zhu, H., Yu, C., Li, S., Jian, M.,

Zhou, Y., Li, Y., Zhang, X., Li, S., Qin, N., Yang, H., Wang, J., Brunak, S., Doré,

J., Guarner, F., Kristiansen, K., Pedersen, O., Parkhill, J., Weissenbach, J., Bork,

P., Ehrlich, S. D., and Wang, J. (2010) A human gut microbial gene catalogue

established by metagenomic sequencing. Nature. 464, 59–65

91. Sayin, S. I., Wahlström, A., Felin, J., Jäntti, S., Marschall, H.-U., Bamberg, K.,

Angelin, B., Hyötyläinen, T., Orešič, M., and Bäckhed, F. (2013) Gut microbiota

regulates bile acid metabolism by reducing the levels of tauro-beta-muricholic

acid, a naturally occurring FXR antagonist. Cell Metab. 17, 225–35

Page 148: Copyright 2018 Seán Martin Mythen

138

92. Raufman, J.-P., Cheng, K., Saxena, N., Chahdi, A., Belo, A., Khurana, S., and

Xie, G. (2011) Muscarinic receptor agonists stimulate matrix metalloproteinase 1-

dependent invasion of human colon cancer cells. Biochem. Biophys. Res.

Commun. 415, 319–324

93. Makishima, M., Lu, T. T., Xie, W., Whitfield, G. K., Domoto, H., Evans, R. M.,

Haussler, M. R., and Mangelsdorf, D. J. (2002) Vitamin D receptor as an intestinal

bile acid sensor. Science. 296, 1313–6

94. Odermatt, A., Da Cunha, T., Penno, C. A., Chandsawangbhuwana, C., Reichert,

C., Wolf, A., Dong, M., and Baker, M. E. (2011) Hepatic reduction of the

secondary bile acid 7-oxolithocholic acid is mediated by 11β-hydroxysteroid

dehydrogenase 1. Biochem. J. 436, 621–629

95. Ferrandi, E. E., Bertolesi, G. M., Polentini, F., Negri, A., Riva, S., and Monti, D.

(2012) In search of sustainable chemical processes: cloning, recombinant

expression, and functional characterization of the 7α- and 7β-hydroxysteroid

dehydrogenases from Clostridium absonum. Appl. Microbiol. Biotechnol. 95,

1221–33

96. Hofmann, A. F., Sjövall, J., Kurz, G., Radominska, A., Schteingart, C. D., Tint, G.

S., Vlahcevic, Z. R., and Setchell, K. D. (1992) A proposed nomenclature for bile

acids. J. Lipid Res. 33, 599–604

97. Heuman, D. M. (1989) Quantitative estimation of the hydrophilic-hydrophobic

balance of mixed bile salt solutions. J. Lipid Res. 30, 719–30

98. Paumgartner, G., and Beuers, U. (2002) Ursodeoxycholic acid in cholestatic liver

disease: mechanisms of action and therapeutic use revisited. Hepatology. 36,

Page 149: Copyright 2018 Seán Martin Mythen

139

525–31

99. Macdonald, I. A., Jellett, J. F., and Mahony, D. E. (1979) 12alpha-Hydroxysteroid

dehydrogenase from Clostridium group P strain C48-50 ATCC No. 29733: partial

purification and characterization. J. Lipid Res. . 20, 234–239

100. Edenharder, R., and Pfützner, A. (1988) Characterization of NADP-dependent

12β-hydroxysteroid dehydrogenase from Clostridium paraputrificum. Biochim.

Biophys. Acta - Lipids Lipid Metab. 962, 362–370

101. Price, M. N., Dehal, P. S., and Arkin, A. P. (2009) FastTree: computing large

minimum evolution trees with profiles instead of a distance matrix. Mol. Biol. Evol.

26, 1641–50

102. Price, M. N., Dehal, P. S., and Arkin, A. P. (2010) FastTree 2--approximately

maximum-likelihood trees for large alignments. PLoS One. 5, e9490

103. Nguyen, L.-T., Schmidt, H. A., von Haeseler, A., and Minh, B. Q. (2015) IQ-TREE:

a fast and effective stochastic algorithm for estimating maximum-likelihood

phylogenies. Mol. Biol. Evol. 32, 268–74

104. Letunic, I., and Bork, P. (2016) Interactive tree of life (iTOL) v3: an online tool for

the display and annotation of phylogenetic and other trees. Nucleic Acids Res. 44,

W242-5

105. Robinson, O., Dylus, D., and Dessimoz, C. (2016) Phylo.io: Interactive Viewing

and Comparison of Large Phylogenetic Trees on the Web. Mol. Biol. Evol. 33,

2163–6

106. Macdonald, I. A., Meier, E. C., Mahony, D. E., and Costain, G. A. (1976) 3alpha-,

7alpha- and 12alpha-hydroxysteroid dehydrogenase activities from Clostridium

Page 150: Copyright 2018 Seán Martin Mythen

140

perfringens. Biochim. Biophys. Acta. 450, 142–53

107. Harris, J. N., and Hylemon, P. B. (1978) Partial purification and characterization of

NADP-dependent 12alpha-hydroxysteroid dehydrogenase from Clostridium

leptum. Biochim. Biophys. Acta. 528, 148–57

108. Aigner, A., Gross, R., Schmid, R., Braun, M., and Mauer, S. Novel 12 alpha-

hydroxysteroid dehydrogenases, production and use thereof. Filed March 25,

2009. Issued April 21, 2011

109. Tanaka, N., Nonaka, T., Tanabe, T., Yoshimoto, T., Tsuru, D., and Mitsui, Y.

(1996) Crystal structures of the binary and ternary complexes of 7 alpha-

hydroxysteroid dehydrogenase from Escherichia coli. Biochemistry. 35, 7715–30

110. Mallonee, D. H., Lijewski, M. A., and Hylemon, P. B. (1995) Expression in

Escherichia coli and characterization of a bile acid-inducible 3 alpha-

hydroxysteroid dehydrogenase from Eubacterium sp. strain VPI 12708. Curr.

Microbiol. 30, 259–63

111. Bhowmik, S., Jones, D. H., Chiu, H.-P., Park, I.-H., Chiu, H.-J., Axelrod, H. L.,

Farr, C. L., Tien, H. J., Agarwalla, S., and Lesley, S. A. (2014) Structural and

functional characterization of BaiA, an enzyme involved in secondary bile acid

synthesis in human gut microbe. Proteins Struct. Funct. Bioinforma. 82, 216–229

112. Ridlon, J. M., Kang, D.-J., and Hylemon, P. B. (2010) Isolation and

characterization of a bile acid inducible 7alpha-dehydroxylating operon in

Clostridium hylemonae TN271. Anaerobe. 16, 137–46

113. Wells, J. E., and Hylemon, P. B. (2000) Identification and characterization of a bile

acid 7alpha-dehydroxylation operon in Clostridium sp. strain TO-931, a highly

Page 151: Copyright 2018 Seán Martin Mythen

141

active 7alpha-dehydroxylating strain isolated from human feces. Appl. Environ.

Microbiol. 66, 1107–13

114. Savino, S., Ferrandi, E. E., Forneris, F., Rovida, S., Riva, S., Monti, D., and

Mattevi, A. (2016) Structural and biochemical insights into 7β-hydroxysteroid

dehydrogenase stereoselectivity. Proteins. 84, 859–65

115. Sherrod, J. A., and Hylemon, P. B. (1977) Partial purification and characterization

of NAD-dependent 7alpha-hydroxysteroid dehydrogenase from Bacteroides

thetaiotaomicron. Biochim. Biophys. Acta. 486, 351–8

116. Hirano, S., Masuda, N., Oda, H., and Mukai, H. (1981) Transformation of bile

acids by Clostridium perfringens. Appl. Environ. Microbiol. 42, 394–399

117. Coleman, J. P., Hudson, L. L., and Adams, M. J. (1994) Characterization and

regulation of the NADP-linked 7 alpha-hydroxysteroid dehydrogenase gene from

Clostridium sordellii. J. Bacteriol. 176, 4865–74

118. Baron, S. F., Franklund, C. V, and Hylemon, P. B. (1991) Cloning, sequencing,

and expression of the gene coding for bile acid 7 alpha-hydroxysteroid

dehydrogenase from Eubacterium sp. strain VPI 12708. J. Bacteriol. 173, 4558–

69

119. Liu, L., Aigner, A., and Schmid, R. D. (2011) Identification, cloning, heterologous

expression, and characterization of a NADPH-dependent 7β-hydroxysteroid

dehydrogenase from Collinsella aerofaciens. Appl. Microbiol. Biotechnol. 90, 127–

35

120. Lee, J.-Y., Arai, H., Nakamura, Y., Fukiya, S., Wada, M., and Yokota, A. (2013)

Contribution of the 7β-hydroxysteroid dehydrogenase from Ruminococcus gnavus

Page 152: Copyright 2018 Seán Martin Mythen

142

N53 to ursodeoxycholic acid formation in the human colon. J. Lipid Res. 54,

3062–9

121. Baele, M., Devriese, L. A., and Haesebrouck, F. (2001) Lactobacillus agilis is an

important component of the pigeon crop flora. J. Appl. Microbiol. 91, 488–91

122. Baele, M., Vancanneyt, M., Devriese, L. A., Lefebvre, K., Swings, J., and

Haesebrouck, F. (2003) Lactobacillus ingluviei sp. nov., isolated from the

intestinal tract of pigeons. Int. J. Syst. Evol. Microbiol. 53, 133–6

123. Morita, H., Nakano, A., Shimazu, M., Toh, H., Nakajima, F., Nagayama, M.,

Hisamatsu, S., Kato, Y., Takagi, M., Takami, H., Akita, H., Matsumoto, M.,

Masaoka, T., and Murakami, M. (2009) Lactobacillus hayakitensis, L.

equigenerosi and L. equi, predominant lactobacilli in the intestinal flora of healthy

thoroughbreds. Anim. Sci. J. 80, 339–46

124. Oki, K., Kudo, Y., and Watanabe, K. (2012) Lactobacillus saniviri sp. nov. and

Lactobacillus senioris sp. nov., isolated from human faeces. Int. J. Syst. Evol.

Microbiol. 62, 601–7

125. Jo, S.-G., Noh, E.-J., Lee, J.-Y., Kim, G., Choi, J.-H., Lee, M.-E., Song, J.-H.,

Chang, J.-Y., and Park, J.-H. (2016) Lactobacillus curvatus WiKim38 isolated

from kimchi induces IL-10 production in dendritic cells and alleviates DSS-induced

colitis in mice. J. Microbiol. 54, 503–9

126. Toh, H., Oshima, K., Suzuki, T., Hattori, M., and Morita, H. (2013) Complete

Genome Sequence of the Equol-Producing Bacterium Adlercreutzia equolifaciens

DSM 19450(T). Genome Announc. 1, e00742-13

127. Uematsu, H., Sato, N., Djais, A., and Hoshino, E. (2006) Degradation of arginine

Page 153: Copyright 2018 Seán Martin Mythen

143

by Slackia exigua ATCC 700122 and Cryptobacterium curtum ATCC 700683.

Oral Microbiol. Immunol. 21, 381–4

128. Hofmann, A. F., and Roda, A. (1984) Physicochemical properties of bile acids and

their relationship to biological properties: an overview of the problem. J. Lipid Res.

25, 1477–89

129. Roda, A., Hofmann, A. F., and Mysels, K. J. (1983) The influence of bile salt

structure on self-association in aqueous solutions. J. Biol. Chem. . 258, 6362–

6370

130. Ditscheid, B., Keller, S., and Jahreis, G. (2009) Faecal steroid excretion in

humans is affected by calcium supplementation and shows gender-specific

differences. Eur. J. Nutr. 48, 22–30

131. Francis, M. B., Allen, C. A., Shrestha, R., and Sorg, J. A. (2013) Bile acid

recognition by the Clostridium difficile germinant receptor, CspC, is important for

establishing infection. PLoS Pathog. 9, e1003356

132. Bokkenheuser, V. D., and Winter, J. (1980) Biotransformation of steroid hormones

by gut bacteria. Am. J. Clin. Nutr. 33, 2502–6

133. McCool, B. A., Plonk, S. G., Martin, P. R., and Singleton, C. K. (1993) Cloning of

human transketolase cDNAs and comparison of the nucleotide sequence of the

coding region in Wernicke-Korsakoff and non-Wernicke-Korsakoff individuals. J.

Biol. Chem. 268, 1397–404

134. Claesson, M. J., Li, Y., Leahy, S., Canchaya, C., van Pijkeren, J. P., Cerdeño-

Tárraga, A. M., Parkhill, J., Flynn, S., O’Sullivan, G. C., Collins, J. K., Higgins, D.,

Shanahan, F., Fitzgerald, G. F., van Sinderen, D., and O’Toole, P. W. (2006)

Page 154: Copyright 2018 Seán Martin Mythen

144

Multireplicon genome architecture of Lactobacillus salivarius. Proc. Natl. Acad.

Sci. U. S. A. 103, 6718–23

135. Lewis, D. A., Brown, R., Williams, J., White, P., Jacobson, S. K., Marchesi, J. R.,

and Drake, M. J. (2013) The human urinary microbiome; bacterial DNA in voided

urine of asymptomatic adults. Front. Cell. Infect. Microbiol. 3, 41

136. Stackebrandt, E., Schumann, P., Schaal, K. P., and Weiss, N. (2002)

Propionimicrobium gen. nov., a new genus to accommodate Propionibacterium

lymphophilum (Torrey 1916) Johnson and Cummins 1972, 1057AL as

Propionimicrobium lymphophilum comb. nov. Int. J. Syst. Evol. Microbiol. 52,

1925–7

137. Diop, K., Morand, A., Dubus, J. C., Fournier, P.-E., Raoult, D., and Fenollar, F.

(2017) “ Arcanobacterium urinimassiliense ” sp. nov., a new bacterium isolated

from the urogenital tract. New Microbes New Infect. 18, 15–17

138. Bokkenheuser, V. D., Winter, J., Cohen, B. I., O’Rourke, S., and Mosbach, E. H.

(1983) Inactivation of contraceptive steroid hormones by human intestinal

clostridia. J. Clin. Microbiol. 18, 500–4

139. Stokes, N. A., and Hylemon, P. B. (1985) Characterization of delta 4-3-

ketosteroid-5 beta-reductase and 3 beta-hydroxysteroid dehydrogenase in cell

extracts of Clostridium innocuum. Biochim. Biophys. Acta. 836, 255–61

140. de Prada, P., Setchell, K. D., and Hylemon, P. B. (1994) Purification and

characterization of a novel 17 alpha-hydroxysteroid dehydrogenase from an

intestinal Eubacterium sp. VPI 12708. J. Lipid Res. 35, 922–9

141. Harris, S. (Virginia C. U. (2017) Discovery and characterization of bile acid and

Page 155: Copyright 2018 Seán Martin Mythen

145

steroid metabolism pathways in gut-associated microbes. Ph.D. Diss.

142. Bahr, N. I., Palme, R., Möhle, U., Hodges, J. K., and Heistermann, M. (2000)

Comparative aspects of the metabolism and excretion of cortisol in three

individual nonhuman primates. Gen. Comp. Endocrinol. 117, 427–38

143. HELLMAN, L., BRADLOW, H. L., ADESMAN, J., FUKUSHIMA, D. K., KULP, J.

L., and GALLAGHER, J. F. (1954) The fate of hydrocortisone-4-C14 in man. J.

Clin. Invest. 33, 1106–15

144. Sender, R., Fuchs, S., and Milo, R. (2016) Revised Estimates for the Number of

Human and Bacteria Cells in the Body. PLoS Biol. 14, e1002533

145. Mostaghel, E. A. (2014) Beyond T and DHT - novel steroid derivatives capable of

wild type androgen receptor activation. Int. J. Biol. Sci. 10, 602–13

146. Koppel, N., and Balskus, E. P. (2016) Exploring and Understanding the

Biochemical Diversity of the Human Microbiota. Cell Chem. Biol. 23, 18–30

147. Xie, G., Wang, X., Zhao, A., Yan, J., Chen, W., Jiang, R., Ji, J., Huang, F., Zhang,

Y., Lei, S., Ge, K., Zheng, X., Rajani, C., Alegado, R. A., Liu, J., Liu, P.,

Nicholson, J., and Jia, W. (2017) Sex-dependent effects on gut microbiota

regulate hepatic carcinogenic outcomes. Sci. Rep. 7, 45232

148. McKenney, P. T., and Pamer, E. G. (2015) From Hype to Hope: The Gut

Microbiota in Enteric Infectious Disease. Cell. 163, 1326–1332

149. Gomez, A., Luckey, D., and Taneja, V. (2015) The gut microbiome in

autoimmunity: Sex matters. Clin. Immunol. 159, 154–62

150. Forger, N. G., Strahan, J. A., and Castillo-Ruiz, A. (2016) Cellular and molecular

mechanisms of sexual differentiation in the mammalian nervous system. Front.

Page 156: Copyright 2018 Seán Martin Mythen

146

Neuroendocrinol. 40, 67–86

151. Lai, J.-J., Lai, K.-P., Zeng, W., Chuang, K.-H., Altuwaijri, S., and Chang, C. (2012)

Androgen receptor influences on body defense system via modulation of innate

and adaptive immune systems: lessons from conditional AR knockout mice. Am.

J. Pathol. 181, 1504–12

152. Kissick, H. T., Sanda, M. G., Dunn, L. K., Pellegrini, K. L., On, S. T., Noel, J. K.,

and Arredouani, M. S. (2014) Androgens alter T-cell immunity by inhibiting T-

helper 1 differentiation. Proc. Natl. Acad. Sci. U. S. A. 111, 9887–92

153. Dubey, R. (ed.) (2012) Sex Hormones, InTech, 10.5772/1313

154. Kornman, K. S., and Loesche, W. J. (1982) Effects of estradiol and progesterone

on Bacteroides melaninogenicus and Bacteroides gingivalis. Infect. Immun. 35,

256–63

155. Morris, D. J., and Ridlon, J. M. (2017) Glucocorticoids and gut bacteria: “The

GALF Hypothesis” in the metagenomic era. Steroids. 125, 1–13

156. Zhang, M.-Z., Xu, J., Yao, B., Yin, H., Cai, Q., Shrubsole, M. J., Chen, X., Kon, V.,

Zheng, W., Pozzi, A., and Harris, R. C. (2009) Inhibition of 11beta-hydroxysteroid

dehydrogenase type II selectively blocks the tumor COX-2 pathway and

suppresses colon carcinogenesis in mice and humans. J. Clin. Invest. 119, 876–

85

157. Gilbert, S., Hood, L., and Seah, S. Y. K. (2018) Characterization of an Aldolase

Involved in Cholesterol Side Chain Degradation in Mycobacterium tuberculosis. J.

Bacteriol. 10.1128/JB.00512-17

158. Storbeck, K.-H., Swart, P., and Swart, A. C. (2007) Cytochrome P450 side-chain

Page 157: Copyright 2018 Seán Martin Mythen

147

cleavage: insights gained from homology modeling. Mol. Cell. Endocrinol. 265–

266, 65–70

159. Bar-Even, A., Noor, E., Savir, Y., Liebermeister, W., Davidi, D., Tawfik, D. S., and

Milo, R. (2011) The moderately efficient enzyme: evolutionary and

physicochemical trends shaping enzyme parameters. Biochemistry. 50, 4402–10

160. Moeller, G., and Adamski, J. (2006) Multifunctionality of human 17β-

hydroxysteroid dehydrogenases. Mol. Cell. Endocrinol. 248, 47–55

161. Persson, B., Kallberg, Y., Bray, J. E., Bruford, E., Dellaporta, S. L., Favia, A. D.,

Duarte, R. G., Jörnvall, H., Kavanagh, K. L., Kedishvili, N., Kisiela, M., Maser, E.,

Mindnich, R., Orchard, S., Penning, T. M., Thornton, J. M., Adamski, J., and

Oppermann, U. (2009) The SDR (short-chain dehydrogenase/reductase and

related enzymes) nomenclature initiative. Chem. Biol. Interact. 178, 94–98

162. Kavanagh, K. L., Jörnvall, H., Persson, B., and Oppermann, U. (2008) Medium-

and short-chain dehydrogenase/reductase gene and protein families : the SDR

superfamily: functional and structural diversity within a family of metabolic and

regulatory enzymes. Cell. Mol. Life Sci. 65, 3895–906

163. Oppermann, U. C. T., Filling, C., and Jörnvall, H. (2001) Forms and functions of

human SDR enzymes. Chem. Biol. Interact. 130–132, 699–705

164. Persson, B., Kallberg, Y., Oppermann, U., and Jörnvall, H. (2003) Coenzyme-

based functional assignments of short-chain dehydrogenases/reductases (SDRs).

Chem. Biol. Interact. 143–144, 271–278

165. Filling, C., Berndt, K. D., Benach, J., Knapp, S., Prozorovski, T., Nordling, E.,

Ladenstein, R., Jörnvall, H., and Oppermann, U. (2002) Critical residues for

Page 158: Copyright 2018 Seán Martin Mythen

148

structure and catalysis in short-chain dehydrogenases/reductases. J. Biol. Chem.

277, 25677–25684

166. Ghosh, D., Wawrzak, Z., Weeks, C. M., Duax, W. L., and Erman, M. (1994) The

refined three-dimensional structure of 3α,20β-hydroxysteroid dehydrogenase and

possible roles of the residues conserved in short-chain dehydrogenases.

Structure. 2, 629–640

167. Devendran, S., Mythen, S. M., and Ridlon, J. M. (2018) The desA and desB

genes from Clostridium scindens ATCC 35704 encode steroid-17,20-desmolase.

J. Lipid Res. 10.1194/jlr.M083949

168. Flint, H. J. and S. H. D. Bifidobacterium adolescentis, Strain L2-32

169. Lee, J.-H., and O’Sullivan, D. J. (2010) Genomic Insights into Bifidobacteria.

Microbiol. Mol. Biol. Rev. 74, 378–416