copyright by silvia a. arredondo 2009

158
Copyright by Silvia A. Arredondo 2009

Upload: others

Post on 07-Jul-2022

13 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Copyright by Silvia A. Arredondo 2009

Copyright

by

Silvia A. Arredondo

2009

Page 2: Copyright by Silvia A. Arredondo 2009

The Dissertation Committee for Silvia A. Arredondo Certifies that this is the

approved version of the following dissertation:

Engineering and Characterization of

Disulfide Bond Isomerases in Escherichia coli

Committee:

George Georgiou, Supervisor

Brent L. Iverson

Jennifer Maynard

Charles B. Mullins

Jon D. Robertus

Page 3: Copyright by Silvia A. Arredondo 2009

Engineering and Characterization of

Disulfide Bond Isomerases in Escherichia coli

by

Silvia A. Arredondo, B.E.

Dissertation

Presented to the Faculty of the Graduate School of

The University of Texas at Austin

in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

The University of Texas at Austin

May, 2009

Page 4: Copyright by Silvia A. Arredondo 2009

To our Lady of Lourdes

And to all the great people who have supported me in the pursuit of my dreams beginning

with my parents Carmen Escobar de Arredondo and Marco Antonio Arredondo

Page 5: Copyright by Silvia A. Arredondo 2009

Acknowledgements

I have not done this alone. It has taken the efforts of a considerable amount of

people to earn this Ph.D. Along the path, I have been blessed with wonderful people who

have supported me in so many ways and I will always be grateful to all of them. Above

all, I am thankful to God and María for placing these people in my way and for this

wonderful experience.

I thank my parents Carmen and Marco Antonio for all they have done for my

brother and me. My dad is responsible for our early encounters with science, but both of

them lovingly taught us discipline, hard work, perseverance and to believe in ourselves.

Special thanks to my brother and his family. Thanks to my abuelita Juanita, who always

celebrated my school successes starting with that one book she gave me when I passed 1st

grade with honors, and who went to heaven before I could finish my Ph.D. I also want to

express my appreciation to my extended family for their love and support, they are all

great. Special thanks to my Uncle Roberto and Aunt Alicia, without whom I would not

have been able to come to this country and start this adventure. Gracias a mis papás, mi

hermano, mis abuelitos, mis tíos, mis primos y todos mis sobrinos por su apoyo

incondicional, por su amor y por creer en mí, éste no es sólo mi logro sino también el

suyo. ¡Que Dios me los bendiga, hoy y siempre! I thank my long-standing friends,

Claudia, Penélope, Miriam, Ileana, Valeria, Karin, Lesly and Lorena for their love and

support, and all those great moments shared in school and in life; like my family, they too

believed I could do this. My heart goes out to Eddie, my angel.

I could not have possibly started college in this country without the help of Dr.

Michelle Hirsh and Dr. Patricio Paez, my godparents, who kindly provided economical

and emotional support to start my education in New York, my heart will be always

v

Page 6: Copyright by Silvia A. Arredondo 2009

grateful to this great family for all they did for me. To Professor Sandra Poster at BMCC,

thank you so much for making an extra effort to ensure I could continue my education

and for your friendship. Taking the Speech class with you has made such a difference in

my life (and to think…I initially refused to take it!).

Thanks to Dr. Andy Grove (Grove Foundation) who as a former immigrant

continues to remember how City College helped him and does the same for students like

me, I couldn’t have completed my undergraduate degree without the scholarship he

provided. Equally important is the support and guidance of professor Alexander Couzis at

CCNY who believed I could go beyond what I thought my limitations were. Alex, thanks

for not letting us say no to you, for showing us the way and making sure we chased our

destiny, for not giving up on us and for making an extra effort to help us succeed. Look at

us…we are doctors now! Thanks to Dr. Juan Duque for walking with me in the hardest of

times, for the friendship, the support, the laughs and the pan de queso which also nurtured

my soul. My heart goes out to Gajanand, who started graduate school as we did but had

more pressing matters to attend in heaven, may he rest in peace.

Here at UT Austin, I would like to thank Professors Jennifer Maynard, Jon

Robertus, Brent Iverson and Buddie Mullins for serving as members of my committee.

Especially Dr. Mullins for valuable guidance as a graduate advisor in helping me find a

wonderful advisor to work with; and Dr. Iverson for his willingness to contribute to our

learning process even when he is not an official co-advisor. I am deeply grateful to

Professor George Georgiou for giving me the opportunity of working with him and being

part of this great lab, for his patience, support, trust, understanding, optimism, insight and

friendship. George, thanks for caring about your students beyond the lab and projects, for

always being there for us, for teaching by example, not only about science and becoming

integral scientists, but also about life; thanks for guiding us like a father. Thanks to Lizzie

vi

Page 7: Copyright by Silvia A. Arredondo 2009

and Donna for assistance in all administrative matters, and especially for all the

“management skills lessons” and support. I had the opportunity to learn how to be a

mentor with two great undergraduate assistants Tiffany Chen and Laura Strong and I

thank them for their hard work, persistence, patience and friendship.

Graduate school can be a difficult experience but having the right people around

makes a significant difference. It has been great to work with the past and present

members of the BIGG lab and I am thankful for their assistance, science tips and

techniques, helpful scientific discussions, friendship and all the fun times. Thanks to

Navin, Karl, Lluis, Vasso, Tom, Phil, Jason, Karthik, Giorgios, Li, Haixin, Everett,

Pogson, Jack, Ming-Jeong, Yariv and Costas. Special thanks to my girlfriends for their

unconditional friendship and support all these years: Laura for mentoring me with

patience in my early years when I knew nothing about cloning, proteins or E. coli,

Danielle for great conversations and standing by me in challenging times, Eva for her

sweetness, caring and “many hugs”, Vasiliki for her trust, Marsha for her kindness,

Mridula for her loyalty, strength, help and company through this crazy adventure. In

addition, I want to thank my friends outside the lab who helped me keep a sense of

balance and a connection to my roots, even if it was hard to understand the nature of my

cell-sitting weekend activities: Alejandra, my Mexican sister, Luciana, Claudia, Fabiana,

David, Domingo, Abraham, Ruben, Elecho, Rodrigo, Rashmi and Moriah.

Despite the long days, nights and weekends (what’s a weekend?), failed

experiments, low yields, disappearing proteins, dead or overgrown cells, contaminated

transformations, incorrect folding and a series of frustrating episodes, graduate school has

been a wonderful experience and I have really enjoyed it. I am sincerely thankful to all

(listed here or not) who have contributed in one way or another to this experience and to

my growth as a scientist and as a human being. May God bless you all! ¡Mil Gracias!

vii

Page 8: Copyright by Silvia A. Arredondo 2009

Engineering and Characterization of

Disulfide Bond Isomerases in Escherichia coli

Publication No._____________

Silvia A. Arredondo, Ph.D.

The University of Texas at Austin, 2009

Supervisor: George Georgiou

Disulfide bond formation is an essential process for the folding and biological

activity of most extracellular proteins; however, it may become the limiting step when the

production of these proteins is attempted in heterologous hosts such as Escherichia coli.

The rearrangement of incorrect disulfide bonds between cysteines that do not normally

interact in the native structure of a protein is carried out by disulfide isomerase enzymes.

The disulfide isomerase present in the bacterial secretory compartment (the periplasmic

space) is the homodimeric enzyme DsbC. The objective of this dissertation was to

understand the key features of how DsbC catalyzes disulfide bond isomerization.

Chimeric disulfide isomerases comprising of protein domains that share a similar

function, or are homologous to domains of DsbC were constructed in an effort to

viii

Page 9: Copyright by Silvia A. Arredondo 2009

understand the effect of the domain orientation in the dimeric protein, and the need for a

substrate binding region in disulfide isomerases. We successfully created a series of

fusion enzymes, FkpA-DsbAs, which catalyze in vivo disulfide isomerization with

comparable efficiency to DsbC. These enzymes comprise of the peptide binding region of

the periplasmic chaperone FkpA, which is functionally and structurally similar to the

binding domain of DsbC but share no amino acid homology with it, fused to the bacterial

oxidase DsbA. In addition, these chimeric enzymes were shown to assist in the initial

formation of disulfide bonds, a function that is normally exhibited only by DsbA.

Directed evolution of the FkpA-DsbA proteins conferred improved resistance to CuCl2, a

phenotype dependent on disulfide bond isomerization and highlighted the importance of

an optimal catalytic site.

The bacterial disulfide isomerase DsbC is a homodimeric V-shaped enzyme that

consists of a dimerization domain, two α-helical linkers and two opposing catalytic

domains. The functional significance of the existence of two catalytic domains of DsbC is

not well understood yet. The fact that identical subunits naturally dimerize to generate

DsbC has so far limited the study of the individual catalytic sites in the homodimer. In

chapter 3 we discuss the engineering, in vivo function, and biochemical characterization

of DsbC variants covalently linked via (Gly3Ser) flexible linkers. We have either

inactivated one of the catalytic sites (CGYC), or entirely removed one of the catalytic

domains while maintaining the putative binding area intact. Our results support the

hypotheses that dual catalytic domains in DsbC are not necessary for disulfide bond

isomerization, but are important in terms of increasing the effective concentration of

ix

Page 10: Copyright by Silvia A. Arredondo 2009

catalytic equivalents, and that the availability of a substrate binding region is a

determining feature in isomerization.

Finally, we have carried out initial studies to map the residues and sequence

motifs that are recognized in substrate proteins that interact with DsbC. Although the

main putative binding region of DsbC has been localized within the limits of the

hydrophobic cleft that emerges from the interaction of the N-terminal domains of this

enzyme, and, a few native substrates have already been identified, no information on the

features of substrate proteins that are recognized by the enzyme has been reported. To

address this problem, we have screened two different, 15 amino-acid random peptide

libraries for binding to DsbC. We have successfully isolated several peptides with high

affinity for the enzyme. Possible consensus binding motifs were identified and their

significance in substrate recognition will be examined in future studies.

x

Page 11: Copyright by Silvia A. Arredondo 2009

Table of Contents List of Tables ................................................................................................................... xiii

List of Figures .................................................................................................................. xiv

Chapter 1: Introduction, Background and Biotechnological Applications......................... 1

1.1 Introduction............................................................................................................... 1 1.1.1 Tissue Plasminogen Activator: a case study...................................................... 5

1.2 Background............................................................................................................... 8 1.2.1 Protein Folding................................................................................................... 8 1.2.2 The Peptydyl-prolyl Isomerase FkpA................................................................ 9 1.2.3 Protein Disulfide Bonds................................................................................... 10 1.2.4 Thioredoxin Superfamily Features .................................................................. 14 1.2.5 Oxidative folding in Eukaryotes ...................................................................... 15

1.2.5.1 PDI ............................................................................................................ 16 1.2.6 Oxidative Folding in E. coli............................................................................. 18

1.2.6.1 DsbA ......................................................................................................... 22 1.2.6.2 DsbC ......................................................................................................... 23

1.3 Protein Engineering ................................................................................................ 27 1.4 Previous Work on the Engineering of Bacterial Disulfide Isomerases................... 28 1.5 Assays for Monitoring In Vivo and In Vitro Activity ............................................. 30

1.5.1 In Vivo Studies ................................................................................................. 30 1.5.1.1 Isomerase Activity –tPA Assay ................................................................ 30 1.5.1.2 Isomerase Activity –AppA Assay............................................................. 32 1.5.1.3 Isomerase Activity –CuCl2 Viability Assay ............................................. 32 1.5.1.4 Redox State Assay .................................................................................... 33 1.5.1.5 Oxidase Activity –PhoA Assay ................................................................ 34 1.5.1.6 Oxidase Activity –Motility Assay ............................................................ 34

1.5.2 In vitro Studies................................................................................................. 35 1.5.2.1 Isomerase Activity- RNase Refolding Assay ........................................... 35 1.5.2.2 Reductase Activity- Insulin Assay............................................................ 35 1.5.2.3 Chaperone Activity- Citrate Synthase Assay............................................ 36

1.6 Research Objectives................................................................................................ 37 Chapter 2: De Novo Design and Evolution of Artificial Disulfide Isomerase Enzymes

Analogous to the Bacterial DsbC...................................................................................... 39

2.1 Introduction............................................................................................................. 39 2.2 Materials and Methods............................................................................................ 44

2.2.1 Strains and Plasmids ........................................................................................ 44 2.2.2 In vivo Disulfide Bond Formation and Isomerization...................................... 46 2.2.3 Directed Evolution for Increased CuCl2 Resistance ........................................ 47 2.2.4 Expression, Purification, and Biochemical Assays.......................................... 48

xi

Page 12: Copyright by Silvia A. Arredondo 2009

2.3 Results..................................................................................................................... 49 2.3.1 Engineering the Artificial Disulfide Bond Isomerases .................................... 49 2.3.2 In vivo Studies.................................................................................................. 55 2.3.3 Evolution of FkpA-DsbA Chimeras Under Selective Pressure ....................... 58 2.3.4 Biochemical Studies......................................................................................... 64

2.4 Discussion ............................................................................................................... 66 Chapter 3: The Role of Dimerization on the Catalytic Properties of the Escherichia coli

Disulfide Isomerase DsbC ................................................................................................ 73

3.1 Introduction............................................................................................................. 73 3.2 Materials and Methods............................................................................................ 77

3.2.1 Strains and Plasmids ........................................................................................ 77 3.2.2 In vivo Disulfide Bond Formation and Isomerization...................................... 78 3.2.3 In Vivo Redox State ......................................................................................... 79 3.2.4 Expression, Purification, and Biochemical Assays.......................................... 80

3.3 Results..................................................................................................................... 81 3.3.1 Construction of a Linked Monomeric DsbC.................................................... 81 3.3.2 Effect of Alanine Substitutions in one Catalytic Active Site........................... 87 3.3.3 A Linked DsbC with a Single Active Thioredoxin Domain............................ 89

3.4 Discussion ............................................................................................................... 97 Chapter 4: Analyis of the Substrate Specificity of the Bacterial Disulfide Isomerase DsbC

......................................................................................................................................... 101

4.1 Introduction........................................................................................................... 101 4.2 Materials and Methods.......................................................................................... 106

4.2.1 Strains and Plasmids ...................................................................................... 106 4.2.2 Protein Expression, Purification and Labeling .............................................. 107 4.2.3 Peptide Synthesis and Purification................................................................. 107 4.2.4 Whole Cell ELISA......................................................................................... 108 4.2.5 Magnetic-Activated Cell Sorting ................................................................... 109 4.2.6 Flow-cytometric Analysis and Sorting .......................................................... 110

4.3 Results................................................................................................................... 111 4.3.1 The NLCPX-X15 Library .............................................................................. 111 4.3.2 The eCPX Library.......................................................................................... 115

4.4 Future Work .......................................................................................................... 121 Chapter 5: Conclusions and Recommendations ............................................................. 122

5.1 Recommendations for Future Studies................................................................... 127 Author’s Publication List ................................................................................................ 129

References....................................................................................................................... 130

Vita.................................................................................................................................. 143

xii

Page 13: Copyright by Silvia A. Arredondo 2009

List of Tables

Table 2.1 (A) Primers and (B) Bacterial strains............................................................... 44

Table 2.2. Physiological function of artificial disulfide isomerases ................................ 59

Table 2.3. In vitro activities of purified enzymes ............................................................ 65

Table 3.1. (A) Primers and (B) bacterial strains and plasmids ........................................ 77

Table 3.2. In vitro activity of the linked proteins............................................................. 88

Table 4.1. Bacterial strains, plasmids and primers......................................................... 106

Table 4.2. DsbC-binding peptides isolated from the NLCPX-X15 library.................... 113

Table 4.3. DsbC-binding peptides isolated from the eCPX-GGS-n-15mer library ....... 119

xiii

Page 14: Copyright by Silvia A. Arredondo 2009

List of Figures

Figure 1.1. Total sale trends in US market for biologic drugs........................................... 2

Figure 1.2. Comparison of the production platforms for recombinant protein therapeutics

............................................................................................................................................. 3

Figure 1.3. Comparative economic analysis of the production of human tissue

plasminogen activator in CHO cells and E. coli ................................................................. 7

Figure 1.4. Molecular structure of FkpA ......................................................................... 10

Figure 1.5. The disulfide bond formation reaction .......................................................... 11

Figure 1.6. Thiol-disulfide exchange reaction ................................................................. 12

Figure 1.7. The reduction/isomerizaion reaction ............................................................. 13

Figure 1.8. The thioredoxin fold ...................................................................................... 15

Figure 1.9. The crystal structure of yeast PDI ................................................................. 17

Figure 1.10. Translocation of a preprotein through the Sec-pathway.............................. 19

Figure 1.11. Disulfide bond formation in the E. coli periplasm ...................................... 20

Figure 1.12. Diagram of the kinetic barriers separating the DsbA/B and DsbC/D redox

systems.............................................................................................................................. 22

Figure 1.13. Molecular structure of DsbA ....................................................................... 23

Figure 1.14. Molecular structure of DsbC ....................................................................... 24

Figure 1.15. Human tissue plasminogen activator........................................................... 31

Figure 2.1. The structures of DsbC, DsbG and PDI ........................................................ 41

Figure 2.2. In vivo isomerase activity of FkpA................................................................ 51

Figure 2.3. Protein structures of DsbA and FkpA, and a predictive molecular model of

FkpA-DsbA2..................................................................................................................... 52

Figure 2.4. Oligomeric state of the FkpA-DsbA chimeras .............................................. 54

Figure 2.5. Expression level of the FkpA-DsbA chimeras compared to DsbC ............... 55

Figure 2.6. Disulfide bond isomerization in vivo............................................................. 56

Figure 2.7. Disulfide bond formation in vivo................................................................... 58

Figure 2.8. Amino acid sequences of the FkpA-DsbA2 derived mutants........................ 60

xiv

Page 15: Copyright by Silvia A. Arredondo 2009

Figure 2.9. FkpA-DsbA2 mutants obtained following mutagenesis and selection for

CuCl2 resistance. ............................................................................................................... 62

Figure 2.10. The effect of the C146 mutation in isomerase activity in vivo. ................... 63

Figure 2.11. CD spectra of purified FkpA-DsbA2 and FkpA-DsbA2m18...................... 64

Figure 3.1. Protein structure of DsbC, and predictive molecular models of mDsbC-

mDsbC and the single-active-site covalently linked mutants. .......................................... 82

Figure 3.2. Gel filtration FPLC of DsbC and derived proteins........................................ 83

Figure 3.3. Expression level of the covalently linked DsbC variants .............................. 84

Figure 3.4. In vivo activity of single-active-site mutants. ................................................ 86

Figure 3.5. In vivo redox state of DsbC and covalently linked proteins. ......................... 87

Figure 3.6. Molecular model of mDsbC-dim................................................................... 90

Figure 3.7. Gel filtration FPLC of mDsbC-dim............................................................... 91

Figure 3.8. Gel filtration FPLC of mDsbC-dim derivative proteins................................ 92

Figure 3.9. mDsbC-dim and mDsbC[H45D]-dim[D53H] ............................................... 93

Figure 3.10. Comparison of mDsbC(H45D)-dim(D53H) and mDsbC(H45D)-dim(H45D)

........................................................................................................................................... 95

Figure 3.11. In vivo activity of single-catalytic-domain mutants. .................................. 96

Figure 4.1. Ribbon diagram and surface representation of the hydrophobic cleft of DsbC

......................................................................................................................................... 101

Figure 4.2. Molecular structure of OmpX and of the circularly permuted version, CPX

......................................................................................................................................... 104

Figure 4.3. Gel filtration FPLC of DsbC and DsbC[H45D] .......................................... 112

Figure 4.4. Whole cell ELISA data for binding of purified DsbC to peptide PHB879 . 114

Figure 4.5. Mass spectrometry trace for the synthetic peptide PWFLQNWG .............. 115

Figure 4.6. Fluorescence distribution of the magnetically enriched peptide library...... 117

Figure 4.7. Comparison of the fluorescence intensity of the eCPX-GGS-n-15mer library

in sequential rounds of FACS. ........................................................................................ 118

Figure 4.8. Representative data of the fluorescence intensity of individual clones

displaying 15 amino acid peptides.................................................................................. 119

xv

Page 16: Copyright by Silvia A. Arredondo 2009

Chapter 1

Introduction, Background and Biotechnological Applications

1.1 Introduction

About 25 years ago, advances in molecular biology and genetic engineering led to

the beginning of a medical revolution marked by the introduction of the first human

recombinant protein used for treatment -recombinant human insulin (Humulin). Since

then, the utilization of recombinant proteins in clinical practice has dramatically

increased and now represents the core of the medical biotechnology industry (Pavlou and

Reichert, 2004). As of 2008, more than 130 different proteins or peptides had been

approved by the FDA and many more candidates are in development or clinical trials

(Leader et al., 2008). These therapeutics are used in the treatment of a wide range of

diseases including cancer, immune disorders, diabetes and endocrinology disorders,

infectious diseases and many others. Accordingly, the market trends for biologics in the

United States reflect a steady increase in sales going from $18.9 billion in 2002 to $44.1

billion in 2007 (Fig 1.1) and projected to reach $52.1 billion in 2010. It is also interesting

to note that in the 2006-2007 period the growth rate for protein therapeutics was almost 3

times the overall growth of the pharmaceutical industry (Aggarwal, 2008; Pavlou and

1

Page 17: Copyright by Silvia A. Arredondo 2009

Reichert, 2004). These trends are a strong indication that recombinant protein research

and development will continue to be in high demand in the coming years.

$18.9$23.6

$27.7$33.0

$39.9$44.1

0

10

20

30

40

50

2002 2003 2004 2005 2006 2007Year

US

Bio

logi

c D

rug

Sale

s ($

bill

ion)

Figure 1.1. Total sale trends in US market for biologic drugs (2002-2007) (Aggarwal, 2008).

There are several reasons for which using proteins for therapeutic treatment is

more advantageous than the use of small molecule drugs. Proteins are highly specific and

can perform rather complex functions that are typically not displayed by small chemical

drugs. Given this specificity, the risk of interference with other biological processes that

may result in adverse side effects is minimized. Therapeutic proteins are usually from

human origin or can be humanized and are often a well tolerated. In addition, proteins

can be used as effective replacement treatment in genetic diseases where genes are

mutated or deleted and gene therapy is not yet available. Finally, the commercial

advantages of these proteins involve generally faster FDA approval times and patent

protection benefits derived from the uniqueness of the molecules. The era of protein

therapeutics was precipitated by the use of recombinant proteins basically due to further

benefits over non-recombinant counterparts. These benefits include faster, less expensive

2

Page 18: Copyright by Silvia A. Arredondo 2009

production, reduction of exposure to animal or human diseases coming from the source of

expression and, versatility and control for improvement of function or specificity (Leader

et al., 2008).

There are several systems available for the production of recombinant protein

therapeutics. Selecting the best system for a particular protein depends on factors such as

scale-up, post-translational modifications, total production, speed of production set up

and regulatory issues. The most commonly investigated production platforms are

bacteria, yeast, filamentous fungi, insect cells, mammalian cells and, transgenic plants,

animals and insects; Fig 1.2 presents a general comparison of the attributes of these

different systems (Dyck et al., 2003).

Figure 1.2. Comparison of the production platforms for recombinant protein therapeutics (Dyck et al., 2003).

3

Page 19: Copyright by Silvia A. Arredondo 2009

If the protein of interest does not require complex post-translational modifications

and is expressed in a soluble form, the system of choice is the gram-negative bacterium

Escherichia coli. The advantages of E. coli are many, including growth on inexpensive

carbon sources, fast growth rate, ability to thrive at high cell densities, well characterized

genetics, wide range of genetic tools already developed for its engineering, and relatively

simple process scale-up (Baneyx and Mujacic, 2004; Kolaj et al., 2009).

Despite significant progress in the characterization and engineering of the cellular

pathways involved in the expression of heterologous protein production in E. coli, there

are still important challenges to be addressed. When recombinant proteins are expressed

in E. coli, if unable to reach their native conformation spontaneously or with assistance in

a timely fashion, they either aggregate into inclusion bodies or are degraded by proteases.

Inclusion bodies can accumulate in the cytoplasm or periplasm and contain about 80-95%

of the heterologous protein mixed with other foreign material (Baneyx and Mujacic,

2004; Kolaj et al., 2009). Inclusion bodies can simplify purification but, although many

proteins are currently produced in this way, the refolding process can be expensive and

labor intensive and, there is no guarantee of a high yield of biologically active protein

(Baneyx, 1999). A more attractive alternative for increasing downstream processing

efficiency is the secretion of the already properly folded target protein into the periplasm

for later release. In general, there are many factors that can be further investigated for

improving the secretion of recombinant proteins in bacteria, for example: optimization of

signal sequences that target proteins for secretion, coexpression of chaperones and

foldases to improve folding, creation of protease deficient strains to prevent degradation,

4

Page 20: Copyright by Silvia A. Arredondo 2009

and growth condition optimization (Baneyx, 1999; Georgiou and Segatori, 2005;

Schmidt, 2004).

In the studies presented in this document we have focused on investigating the

engineering and characterization of disulfide bond isomerase, DsbC, which assists the

proper folding of disulfide-containing proteins in the bacterial periplasm. The disulfide

bond formation machinery in E. coli is only equipped to handle a few disulfide bonds on

a polypeptide chain; however, many recombinant proteins of therapeutic interest have

several disulfide bonds that must be properly formed for biological activity. In recent

years, the understanding of disulfide bond formation in E. coli has been enriched with the

discovery of the disulfide bond formation pathways, enzymes and their mechanisms of

action. Nonetheless, there are still many basic principles to unveil and a lot of room for

optimization which can eventually lead to higher yields and dramatically lower costs for

the production of recombinant protein therapeutics in bacteria.

1.1.1 Tissue Plasminogen Activator: a case study.

The biotechnological relevance of disulfide bond isomerases is highlighted in the

case study briefly discussed in this section. Tissue plasminogen activator (tPA) is a serine

protease that acts as a thrombolytic agent and whose folding into the native state requires

the formation of 17 disulfide bonds. In order to dissolve blood clots, tPA converts

plasminogen into plasmin which, in turn, degrades the fibrin forming the thrombus. tPA

has the particular ability to bind to fibrin while activating plasminogen, minimizing the

probability of internal bleeding thus, making it more desirable than other anti-

5

Page 21: Copyright by Silvia A. Arredondo 2009

thrombolytic drugs for the treatment of acute myocardial infarction, deep vein

thrombosis, pulmonary embolism and stroke (Datar et al., 1993; van Zonneveld et al.,

1986). About 15 years ago, Datar et al. published a comparison of the process economics

for the production of tPA both in Chinese hamster ovary (CHO) cells and in Escherichia

coli, and their conclusions are still applicable to a great extent. At the time, the authors

estimated that it was possible to produce 33.5 mg/L of tPA in CHO cells and 460 mg/ L

in E. coli (as insoluble inclusion bodies), before purification. However, an economic

analysis determined that for an annual production requirement of 11 kg, the E. coli based

process was not viable at all! The reasons for this disparity lie in the inefficiency of

downstream processing for E. coli production. While the material costs were about a third

of those of the CHO process (Fig 1.3), solubilization, refolding and purification of tPA,

originally expressed as inclusion bodies, dramatically increased the equipment and

manufacturing costs of the process in bacteria (Fig 1.3). It was apparent that an E. coli

process for the expression of the protein in soluble, active form would be required to

eliminate the expensive downstream processing costs and thus make bacterial expression

commercially practical (Datar et al., 1993).

Five years later, Qiu et al. from the Georgiou lab, reported the successful

expression of active human tissue plasminogen activator in the E. coli periplasm.

Although secreting tPA into the periplasm had been tried before, the protein was

misfolded and inactive due to the incorrect pairing of disulfide bonds. Qiu et al. (1998)

showed that co-expression of DsbC, the bacterial disulfide bond isomerase, improved

disulfide bond formation in tpA resulting in the production of 180 μg/L of purified active

6

Page 22: Copyright by Silvia A. Arredondo 2009

enzyme (Qiu et al., 1998). While this was a big step towards bacterial expression of

proteins containing multiple disulfide bonds, the protein yield was still low and further

optimization is definitely necessary. Engineering of bacterial disulfide isomerases in

order to enhance activity with heterologous substrates was proposed as a promising

solution; the fact that PDI, the eukaryotic disulfide isomerase, is known to catalyze

isomerization at higher rates than DsbC in vitro (Segatori et al., 2004) suggests potential

for the improvement of the bacterial enzyme.

E. coli $70,887,000

Refolding Tanks 75%Fermentors 3%Fermentors Aux 1%Disposal Equip 14%Recovery Aux 7%

Mammalian $11,123,000

Fermentors 54%Fermentors Aux 16%Recovery Equip 24%Recovery Aux 6%

Equipment Cost

Annual Manufacturing Expense

Annual Material Costs

Mammalian $30,538,000

Fermentation 75%Recovery 25%

Ferm Materials 33% Ferm Materials 1%Utilities 17% Utilities 20%

Patent/Royal 20%Patent/Royal 17%Recovery Materials 8%Recovery Materials 11%Waste 16%Waste 10%Labor 22%Labor 2%Other 13%Other 3%Mammalian $70,927,000 E. coli $112,967,000

Fermentation 12%Recovery 88%

E. coli $10,004,000

Figure 1.3. Comparative economic analysis of the production of human tissue plasminogen activator in CHO cells and E. coli. Most of the expenditure for production of tPA in E. coli is associated with the downstream processing of inclusion bodies (Datar et al., 1993).

7

Page 23: Copyright by Silvia A. Arredondo 2009

1.2 Background

1.2.1 Protein Folding

Proteins are involved in nearly every biological process in all living organisms.

The large variety of catalytic reactions and structural functions necessary to sustain

human life is accomplished by hundreds of thousands of highly specialized proteins

(Dobson, 2004). To become biologically active, a protein initially synthesized by the

ribosome as a linear chain of amino acids, has to acquire a specific three-dimensional

conformation.

Anfinsen and co-workers in their “Thermodynamic Hypothesis” stated that the

final structure of a native protein in its normal physiological environment is the most

energetically favorable and, that it results from the totality of interatomic interactions. In

other words, the final native conformation of a polypeptide in a particular environment is

determined by its amino acid sequence (Anfinsen, 1973). Subsequent studies have

focused on the understanding of the folding kinetics of single-domain proteins. Two

distinct mechanisms have been proposed: in the diffusion-collision model, folding takes

place as a hierarchical set of events starting with the rapid formation of secondary

structures like α-helices and β-sheets and is followed by their collision and consolidation

into the native conformation; every “layer” of structure that forms is stabilized by the

previous “layer” (Fersht, 2008; Travaglini-Allocatelli et al., 2009). The nucleation-

condensation model suggests the folding of an initial nucleus comprised of a few

important residues that then precipitates the formation of the tertiary structure (Dobson,

2004; Travaglini-Allocatelli et al., 2009).

8

Page 24: Copyright by Silvia A. Arredondo 2009

Folding of a polypeptide in the cell may begin immediately as the N-terminus of

the amino acid chain is exiting the ribosome. Some proteins might be able to reach their

three-dimensional final conformations in the necessary biological timeframe without

assistance. However, given the very crowded cellular environment, many proteins are

prone to non-productive intermolecular aggregation and/or premature folding that can

lead to the formation of insoluble protein deposits (inclusion bodies), degradation of the

polypeptide, or toxicity to the cell (Christis et al., 2008; Young et al., 2004). To avoid

these problems, cells have evolved a complex and efficient system of chaperones and

folding catalysts. Molecular chaperones facilitate folding by binding to the accessible

hydrophobic areas of polypeptides in unfolded, intermediate or misfolded conformations,

thus preventing aggregation. Folding catalysts, on the other hand, are necessary to

accelerate rate-limiting conformational transitions, for example peptidyl bond

isomerization and disulfide bond formation (Christis et al., 2008; Kolaj et al., 2009).

1.2.2 The Peptydyl-prolyl Isomerase FkpA

The members of the peptidyl-prolyl cis-trans isomerase (PPIase) family of

enzymes are found both in eukaryotes and prokaryotes. PPIases catalyze the rapid

isomerization of a peptide bond that precedes a proline residue from the trans to the cis

position, often a rate-limiting step in protein folding. FkpA is a bacterial member of the

PPIase family of relevance to the work presented in chapter 2. This enzyme is found as a

homodimer located in the E. coli periplasm and its structure reveals a V-shaped

conformation resulting from the dimerization of its identical subunits (Fig 1.4). Each

9

Page 25: Copyright by Silvia A. Arredondo 2009

monomer has an N-terminal dimerization domain formed by three helices and a C-

terminal catalytic domain. The two catalytic domains are held apart by the long helices of

the dimerization domain, positioning the catalytic sites facing inwards towards each other

and allowing for some flexibility in their relative orientation. FkpA not only has PPIase

activity, but also displays chaperone activity. The chaperone activity has been localized

to the N-terminal domain and the catalytic activity is believed to reside independently in

the C-terminal domain (Ramm and Pluckthun, 2000; Saul et al., 2004).

catalytic domain

linker

dimerization domain

Figure 1.4. Molecular structure of FkpA (Saul et al., 2004).

1.2.3 Protein Disulfide Bonds

Other than the peptide bond, the most common covalent linkage in proteins is the

disulfide bond that chemically links the sulfurs of two cysteines, resulting from the

reaction between thiols (-SH) that have each lost one electron. Thiol-disulfide exchange

reactions are employed for the transfer of electrons, a process which is essential for life.

For example, thiol-mediated electron transfer plays an important role in processes like

10

Page 26: Copyright by Silvia A. Arredondo 2009

prevention of oxidative damage, detoxification and protein folding (Vlamis-Gardikas,

2008). A representative equation for the thiol-disulfide reaction is shown in Fig 1.5.

Figure 1.5. The disulfide bond formation reaction.

In the forward reaction oxidation takes place as each sulfur atom loses one

electron; the reverse is a reduction reaction. Since this is an oxidation-reduction process,

appropriate electron donors and acceptors are required for the reaction to proceed.

However, the formation or disruption of disulfide bonds not only depends on accessible

electron sinks and sources, but also on kinetic reactivity and suitable redox potentials (the

thermodynamic tendency to exchange electrons) (Gilbert, 1990).

The mechanism of the thiol-disulfide exchange reaction is illustrated in Fig 1.6. A

fully protonated thiol (-SH) is inactive; therefore, for the reaction to occur, the thiol must

be deprotonated, forming a thiolate (-S-). Thiolates result from the protonation-

deprotonation equilibrium of a thiol and the extent of deprotonation determines the rate

11

Page 27: Copyright by Silvia A. Arredondo 2009

constant of the reaction at neutral pH. As shown (Fig 1.6), disulfide exchange starts with

the nucleophilic attack of the reactive thiolate (-Sii-) followed by the formation of a

transition state where charge is shared by the three sulfurs. At this point, if the original

disulfide is asymmetrical, the best leaving group (S0-) will be the one with the lowest

pKa, and as it leaves, a new disulfide bond (Sii-Si) is formed (Gilbert, 1990; Vlamis-

Gardikas, 2008). The pKa values of the involved thiols are modulated by the residues

adjacent to the corresponding cysteines (Grauschopf et al., 1995; Huber-Wunderlich and

Glockshuber, 1998; Ren et al., 2009).

Figure 1.6. Thiol-disulfide exchange reaction (Vlamis-Gardikas, 2008).

Alternatively, if a another thiolate (Siii- or Siv

-) is nearby, the newly formed

disulfide (Sii-Si) could go through a nucleophilic attack itself as illustrated in Fig 1.7. The

cysteine corresponding to Siii- or Siv

- is termed the resolving cysteine and can come from

the molecule that initiated the overall reaction (enzyme) or from the target molecule

(substrate). In the first case, the reactive molecule will end up with a disulfide and the

12

Page 28: Copyright by Silvia A. Arredondo 2009

substrate will lose the original disulfide (Si-S0) and will be left in a reduced state. In the

second case, an isomerization reaction will take place in which the thiolate Siv-

establishes a new bond with either Si- or S0

- and the reactive molecule is then left

unmodified (Gilbert, 1990; Vlamis-Gardikas, 2008).

SubstrateSubstrate

Enzyme

H+

H+

H+

SubstrateSubstrate

Enzyme

Siii-

Siv-

SubstrateSubstrate

Enzyme

SubstrateSubstrate

Enzyme

Siv

Si SiiHSiiiH

Figure 1.7. The reduction/isomerizaion reaction (Vlamis-Gardikas, 2008).

Protein disulfide bonds are classified according to their function into structural,

catalytic, and allosteric disulfides. Structural disulfide bonds take part in the stabilization

of the three-dimensional native conformation of the protein by lowering the entropy of

the unfolded state. Once its structure is stabilized, a protein is less prone to inactivation

by oxidants and proteases (Arolas et al., 2006). Structural disulfides are considered not

only to contribute to the tertiary structure, but also to be part of the primary structure

13

Page 29: Copyright by Silvia A. Arredondo 2009

since they are covalent in nature and remain intact even after denaturation (Wetzel,

1986). Catalytic disulfide bonds participate in oxidation-reduction cycles characteristic of

the protein’s activity as an enzyme (Kadokura et al., 2003). Allosteric disulfides are

involved in non-enzymatic activity in which their cleavage controls the function of the

protein (Hogg, 2003; Mamathambika and Bardwell, 2008).

In principle, disulfide bonds can form spontaneously in the presence of molecular

oxygen, however that reaction happens at a rate much slower than what is required for

biological purposes. In 1966 Anfinsen discovered the first enzyme that can catalyze the

formation of disulfide bonds (PDI) (De Lorenzo et al., 1966). For disulfide bonds to form

within the cell, it is essential for the surrounding environment to be oxidative in nature.

Accordingly, disulfide-containing proteins are mostly found in the non-cytoplasmic

compartments of organisms. In eukaryotic cells, disulfide bond formation takes place

mostly in the endoplasmic reticulum where Protein Disulfide Isomerase (PDI) is the main

folding catalyst and chaperone in charge of introducing disulfides into proteins

(oxidation) and rearranging incorrectly formed disulfides (isomerization) (Wilkinson and

Gilbert, 2004). In bacteria, oxidation and isomerization are accomplished by two

independent pathways carried out by the Dsb (disulfide bond) family of oxidoreductases

in the periplasmic compartment.

1.2.4 Thioredoxin Superfamily Features

Although prokaryotic and eukaryotic thiol-disulfide oxidoreductases share only

minimal sequence homology, they all belong to the thioredoxin superfamily. This family

14

Page 30: Copyright by Silvia A. Arredondo 2009

of enzymes is characterized by two common features; the first is the structural element

known as thioredoxin fold which consists of a four-stranded β-sheet and three flanking α-

helices, illustrated in Fig 1.8 (Martin, 1995). The second feature is the catalytic site

located in the thioredoxin fold near the N-terminus of one of the α-helices, and defined

by the sequence Cys-X-X-Cys, where X represents any amino acid (Aslund and

Beckwith, 1999; Ren et al., 2009).

Figure 1.8. The thioredoxin fold (Martin, 1995).

1.2.5 Oxidative folding in Eukaryotes

In eukaryotic cells, the processes of protein modification, quality control and

folding –including disulfide bond formation, take place primarily in the lumen of the

endoplasmic reticulum (ER). Most of the proteins folded in this compartment contain

disulfide bonds; protein disulfide isomerase (PDI) catalyzes both the oxidation and

15

Page 31: Copyright by Silvia A. Arredondo 2009

isomerization of these disulfides. To remain in an active state, PDI is recycled by other

ER proteins: Ero1-Lα and Ero1-Lβ in mammalian cells, AERO1 and AERO2 in plants,

and Ero1p or Erv2p in yeast (Gruber et al., 2006). PDI is not the only oxidoreductase

found in the ER, there are 19 PDI homologues in humans and 4 in yeast that have at least

one thioredoxin fold. These homologues differ in domain organization, tissue localization

and substrate specificity, and many have yet to be characterized in detail (Appenzeller-

Herzog and Ellgaard, 2008; Christis et al., 2008; Gruber et al., 2006).

1.2.5.1 PDI

Among the members of its family, PDI is believed to be the most abundant,

constituting about 0.8% of the total cellular protein in yeast and mammalian cells (Gruber

et al., 2006; Wilkinson and Gilbert, 2004). PDI is a monomer comprised by five domains

(a, a’, b, b’ and c) and has two catalytic sites, with the motif CXXC, located in the a and

a’ domains. The cysteines of the active sites are separated by a Gly-His dipeptide. Even

though the complete structure of mammalian PDI has not yet been reported, the crystal

structure of yeast PDI (yPDI) is available and is presented in Fig 1.9 (Tian et al., 2006).

The molecular structure (Fig 1.9) reveals that the shape of yPDI resembles a “U”.

The residues lining the inside of the U are mostly hydrophobic in nature, hence it has

been suggested that substrate binding interactions occur in this region. As expected, the

catalytic domains a and a’ adopt a thioredoxin fold structure; the same fold is also

observed in the non-catalytic b and b’ domains with minor variations. In general, PDI is

16

Page 32: Copyright by Silvia A. Arredondo 2009

thought to have evolved from thioredoxin domain duplication (Gruber et al., 2006; Tian

et al., 2006).

Figure 1.9. The crystal structure of yeast PDI (Tian et al., 2006).

Extensive in vitro characterization of PDI and of mutants comprising of different

combinations of its domains indicates that while all thioredoxin-like domains are

necessary for isomerization activity in complex protein substrates, oxidation requires

only the a or a’ domain and, isomerization of peptides is catalyzed by a combination of

b’ with either a or a’ (Ellgaard and Ruddock, 2005; Wilkinson and Gilbert, 2004).

Catalysis of isomerization can occur through either of two mechanisms: in the

intramolecular pathway, a mixed disulfide is formed between PDI and its substrate; it is

then resolved by a cysteine coming from the substrate resulting in a native disulfide. In

the reduction/oxidation pathway, cycles of oxidation and reduction leading to a properly

17

Page 33: Copyright by Silvia A. Arredondo 2009

folded substrate take place (Appenzeller-Herzog and Ellgaard, 2008; Wilkinson and

Gilbert, 2004).

In addition to the catalysis of disulfide bond formation and isomerization, PDI is

also able to work as a chaperone by inhibiting the aggregation of proteins lacking

disulfide bonds; the catalytic cysteines are not required for this activity (Wilkinson and

Gilbert, 2004). Attempts to utilize PDI as an isomerase in bacteria have failed. When the

enzyme is expressed in the periplasm, it behaves mainly as an oxidase (Ostermeier et al.,

1996).

1.2.6 Oxidative Folding in E. coli

Contrary to oxidative folding in eukaryotes, where a single enzyme (PDI)

catalyzes both oxidation and isomerization, in E coli these processes are accomplished

independently by two enzymes, DsbA and DsbC, located in the periplasm. As protein

synthesis takes place in the cytoplasm, which has a reducing redox potential, disulfide-

containing proteins need to be translocated into the much more oxidizing periplasm in

order to reach their final native conformation.

In bacteria, there are two major pathways for the export of proteins across the

cytoplasmic membrane: the twin-arginine or Tat pathway which translocates proteins that

are already in their folded native state and/or have a co-factor; and, the Sec-pathway (Fig

1.10) that transports proteins in their unfolded conformations, including polypeptides

requiring the formation of disulfide bonds. Exported preproteins contain hydrophobic

signal sequences that are critical for recognition. Upon translocation, signal peptidase

18

Page 34: Copyright by Silvia A. Arredondo 2009

cleaves the signal peptide yielding a mature protein (Natale et al., 2008; Rusch and

Kendall, 2007).

Figure 1.10. Translocation of a preprotein through the Sec-pathway (Natale et al., 2008).

Once in the periplasm, if the mature protein requires the formation of disulfide

bonds, it undergoes oxidative folding by the “Dsb” family of enzymes through two

distinct pathways: oxidation and isomerization (Gleiter and Bardwell, 2008; Ito and

Inaba, 2008) (Fig 1.11). In the oxidation pathway, the oxidase DsbA, catalyzes disulfide

bond formation on the protein by donating its very unstable disulfide. In order for DsbA

to remain active, the disulfide must be recycled; this is accomplished by the membrane

protein DsbB which releases the electrons onto the respiratory chain through ubiquinone

(Collet and Bardwell, 2002; Glockshuber, 1999). For proteins containing only two

disulfide-forming cysteines, oxidative folding is complete after catalysis by DsbA.

However, when the formation of two or more disulfide bonds is required for the protein

to reach its native conformation, as is usually the case in heterologous proteins, the strong

19

Page 35: Copyright by Silvia A. Arredondo 2009

oxidizing character of DsbA often results in the incorrect pairing of cysteines. Therefore,

isomerization of non-native disulfides is necessary to accomplish proper folding.

In the isomerization pathway, the main bacterial isomerase, DsbC, rearranges the

incorrect disulfides found in misfolded proteins until the native bonds are reached. DsbC

needs to be in a reduced state in order to perform the catalysis of isomerization; the

protein that maintains DsbC active is DsbD, an inner membrane protein that is itself

reduced by thioredoxin from the cytoplasm (Kadokura et al., 2003; Messens and Collet,

2006).

DsbA DsbC

DsbB

DsbD

Thioredoxin

Oxidation Isomerization

SHSH

SHSH

SHSH

SS

S S

S S

SS

SS

S

S

SHSHSH

SH

SS

Periplasm

Cytoplasm

Cytochrome

Oxidases

Figure 1.11. Disulfide bond formation in the E. coli periplasm.

20

Page 36: Copyright by Silvia A. Arredondo 2009

In addition to DsbC, another disulfide isomerase (DsbG) has been identified in E.

coli, however, the role of DsbG in oxidative folding is less clear. Although DsbG is

structurally similar to DsbC and in certain cases can complement dsbC- cells in a DsbD

dependent manner, its isomerization activity is not comparable to DsbC and it is believed

to have much narrower substrate specificity (Bessette et al., 1999b).

It is reasonable to ask what is preventing these two pathways from establishing a

futile electron cycle; in other words, why is DsbC not oxidized by DsbB? Or DsbD by

DsbA? It has been reported that large kinetic barriers keep the oxidation and

isomerization pathways apart. As seen on figure 1.12, non-productive (i.e. non-

physiologically relevant) reactions between the enzymes of the two pathways are 103 to

107 fold slower than the DsbA-DsbB and DsbC-DsbD interactions. Even the DsbB-DsbC

reaction, the fastest among the non-productive thiol-disulfide electron transfer processes,

is about 1000 times slower than the physiologically relevant oxidation of DsbA by DsbB.

These kinetic barriers allow the co-existence of both pathways in the periplasmic

envelope (Bader et al., 2000; Rozhkova et al., 2004). In addition, in vivo studies and

structural evidence suggest that the dimerization of DsbC prevents its active sites from

being available for oxidation by DsbB. Binding of DsbC to DsbB would be expected to

cause one of the DsbC domains to clash onto the lipid bilayer membrane, hence making it

sterically unfavorable (Bader et al., 2001; Inaba et al., 2006). However, this barrier can be

circumvented by changing the orientation of the catalytic sites of DsbC making them

available for interaction with DsbB (Segatori et al., 2006).

21

Page 37: Copyright by Silvia A. Arredondo 2009

Figure 1.12. Diagram of the kinetic barriers separating the DsbA/B and DsbC/D redox systems (Rozhkova et al., 2004).

1.2.6.1 DsbA

The first protein discovered to be involved in disulfide bond formation in the

bacterial periplasm was DsbA, a 21-kDa monomeric protein composed of 189 residues

(Bardwell et al., 1991; Kamitani et al., 1992). Structurally, DsbA consists of a

thioredoxin-fold domain linked by a flexible hinge to a compact α-helical domain (Fig

1.13). The catalytic site, characterized by the CXXC motif, has the sequence Cys30-

Pro31-His32-Cys33 and is located at the N-terminus of the first α-helix of the

thioredoxin domain (Martin et al., 1993). Among all known thiol-disulfide

oxidoreductases, DsbA is the second most oxidizing, with a redox potential of -120 mV;

22

Page 38: Copyright by Silvia A. Arredondo 2009

its disulfide bond is extremely unstable and readily transferred to the free cysteines of

substrate proteins, rather indiscriminately (Ito and Inaba, 2008; Zapun et al., 1993).

DsbA is found predominantly oxidized in vivo in spite of its strong oxidizing

potential, due to the recycling action of the inner membrane protein DsbB. The latter

protein has two pairs of essential cysteines: Cys104-Cys130, involved in the 1:1

stoichiometric redox interaction with DsbA, and Cys41-Cys44 which take part in

transferring electrons to ubiquinone or menaquinone (Grauschopf et al., 2003; Guilhot et

al., 1995; Inaba et al., 2006).

Figure 1.13. Molecular structure of DsbA, the catalytic site is shown in yellow (Martin et al., 1993).

1.2.6.2 DsbC

The identification of DsbC by two different groups in 1994 (Missiakas et al.,

1994; Shevchik et al., 1994) led to the discovery of the isomerization pathway in E. coli.

DsbC is a homodimeric protein that consists of two 23.4 kDa subunits. Each subunit

contains four cysteines, two of these are part of a buried structural disulfide bond; and the

other two are located in the Cys98-Gly99-Tyr100-Cys101 catalytic site forming an

23

Page 39: Copyright by Silvia A. Arredondo 2009

unstable and reactive disulfide, with a redox potential, E0= –130mV (Zapun et al., 1995).

In contrast to DsbA, the catalytic cysteines of DsbC are found in a reduced state in vivo

(Joly and Swartz, 1997).

thioredoxin domain

linker

dimerization domain

Catalytic site

C98 G99 Y100 C101

Figure 1.14. Molecular structure of DsbC. The CXXC catalytic site is shown in yellow (McCarthy et al., 2000).

The crystal structure of DsbC (Fig 1.14) reveals that both monomers come

together resulting in a V-shape molecule. Each monomer is formed by two distinct

domains: the N-terminal dimerization domain (residues 1-61), and the C-terminal

catalytic domain (residues 78-216) where the thioredoxin fold and the CXXC motif are

located. These domains are connected by a helical hinged linker (residues 62-77). The

surface of the cleft formed by the dimerization of the two units consists of uncharged or

24

Page 40: Copyright by Silvia A. Arredondo 2009

hydrophobic residues that are hypothesized to be involved in substrate recognition

(McCarthy et al., 2000).

According to the thiol/disulfide reaction mechanisms described in section 1.2.3,

isomerization by DsbC has been proposed to occur as follows: first, Cys98 launches a

nucleophilic attack onto a non-native disulfide in the substrate protein resulting in an

unstable mixed disulfide complex between DsbC and its target. Then, this mixed

disulfide can be resolved in two ways, either through the attack of another cysteine from

the substrate, forming a more stable disulfide and releasing DsbC in a reduced state; or by

reacting with Cys101 resulting in oxidized DsbC (Collet et al., 2002; Messens and Collet,

2006). In the latter case, DsbD, a three-domain inner membrane protein, will re-activate

oxidized DsbC by reducing the catalytic site cysteines through the transport of electrons

originating from the cytoplasmic Thioredoxin (Haebel et al., 2002; Rietsch et al., 1997).

More recently, Vertommen et al. have proposed a revised version of the above

mechanism in which DsbC can function independently of DsbD and might be able to

catalyze both oxidation and isomerization by cycling from the reduced to the oxidized

state upon interaction with the substrate (Vertommen et al., 2008).

So far only a few native DsbC substrates have been identified: the penicillin

insensitive endopeptidase MepA, ribonuclease RNase I, and acid phosphatase AppA, all

of which have non-consecutive disulfides (Berkmen et al., 2005; Hiniker and Bardwell,

2004). Additionally, several studies have established that the co-expression of DsbC

improves the production of multi-disulfide heterologous proteins in E. coli, including

25

Page 41: Copyright by Silvia A. Arredondo 2009

human nerve growth factor, scFv antibodies and human tissue plasminogen activator, tPA

(Kurokawa et al., 2001; Qiu et al., 1998; Zhang et al., 2002).

Besides being an isomerization catalyst, DsbC has also been shown to function as

a chaperone, assisting the folding of proteins without disulfide bonds, such as D-

glyceraldehyde-3-phosphate dehydrogenase (GAPDH); the Cys98-Cys101 catalytic

cysteines are not required for this activity (Chen et al., 1999; Liu and Wang, 2001).

26

Page 42: Copyright by Silvia A. Arredondo 2009

1.3 Protein Engineering

Protein engineering utilizes genetic and, to a lesser extent, chemical biology

techniques for the structural modification of a protein in order to enhance properties such

as stability, activity and specificity; it attempts to adapt proteins to function effectively

under new circumstances (Brannigan and Wilkinson, 2002). The two most significant and

relevant approaches to protein engineering are i) rational design and ii) directed

evolution.

For rational design it is necessary to have information on the three-dimensional

structure of the protein and possibly on its activity and interactions with substrates. Site-

directed mutagenesis is used to explore the effects of individual (or sets of) amino acid

substitutions at specific residues; the contribution of these substitutions to function is then

measured. Techniques such as saturation mutagenesis or alanine-scanning (substitution of

every amino acid or alanines in the selected positions, respectively) are also used to

investigate particular residues or sites in greater detail (Brannigan and Wilkinson, 2002;

Tucker and Hooper, 2006).

Directed evolution involves generating a diversity of mutation combinations in a

rather random or semi-random fashion; error-prone PCR, single gene shuffling, gene-

family shuffling, incremental truncation of hybrid enzymes (ITCHY) and other

techniques are used for the creation of very large “libraries” of mutants (Brannigan and

Wilkinson, 2002). These libraries are then screened through high-throughput mechanisms

that lead to the isolation of candidates fit for the stringent selection conditions. In some

cases several rounds of mutation and selection are necessary for the successful isolation

27

Page 43: Copyright by Silvia A. Arredondo 2009

of mutants with the desired quality (Brannigan and Wilkinson, 2002; Tucker and Hooper,

2006).

1.4 Previous Work on the Engineering of Bacterial Disulfide Isomerases

Since the discovery and characterization of DsbC, several studies of particular

relevance to the engineering of disulfide isomerase and to the present work have been

reported and a brief summary is described here.

Mutants of DsbC in which the nucleophilic Cys98 or both, Cys98 and Cys101,

have been substituted are completely devoid of in vivo or in vitro isomerase activity (Liu

and Wang, 2001; Missiakas et al., 1994); replacing the resolving Cys101, however,

allows for low levels of isomerization (Rietsch et al., 1997). In addition, an in vitro

generated heterodimer of DsbC in which one catalytic site is inactivated by

carboxymethylation was reported to be inactive as an isomerase (Sun and Wang, 2000).

These findings led to the assumption that both catalytic sites and domains are essential

for isomerization.

The in vitro study of a truncated version of DsbC containing only the thioredoxin

domain, determined that the N-terminal sequence of DsbC (residues 1-65) is essential for

dimerization and substrate binding (Sun and Wang, 2000). Additionally, Bader et al. was

able to isolate mutants of DsbC capable of catalyzing oxidation in vivo; the mutations

prevented the DsbC monomers to associate, suggesting that dimerization protects DsbC

from interacting with DsbB (Bader et al., 2001). The DsbC monomer was also shown to

28

Page 44: Copyright by Silvia A. Arredondo 2009

be unable to catalyze isomerization in vitro (Ke et al., 2006). Studies of the crystal

structure of only the dimerization domain of DsbC indicate that this domain folds

independently of the catalytic portion of the full-length protein (Yeh et al., 2007).

Thioredoxin (PDI-like) and DsbA exhibit very weak disulfide isomerase activity

in vitro (Lundstrom et al., 1992; Segatori et al., 2004). In vivo, both proteins are readily

oxidized by DsbB and therefore they cannot catalyze disulfide bond isomerization (Jonda

et al., 1999). However, fusing these proteins to the dimerization domain of DsbC (DsbC-

TrxA and DsbC-DsbA chimeras) resulted in high disulfide isomerase activity in vivo. The

oxidized form of the chimeras was also shown to serve as a catalyst for de novo disulfide

bond formation (Segatori et al., 2004). In addition, previous studies in our lab in which

we sequentially deleted amino acids from the α-helical linker of DsbC, thus rotating the

catalytic domain with respect to the dimerization domain, highlighted the role of this

linker in preventing oxidation by DsbB (Segatori et al., 2006). Collectively, these studies

suggest that the identity of the catalytic domain in DsbC is not a particular determinant

for the catalysis of isomerization, as long as the dimerization domain is kept constant;

however, the positioning of the catalytic sites seems to make a difference in terms of

crossing the barrier between the DsbA/DsbB and DsbC/DsbD pathways.

29

Page 45: Copyright by Silvia A. Arredondo 2009

1.5 Assays for Monitoring In Vivo and In Vitro Activity

The initial characterization of DsbC was reported about fifteen years ago by two

independent groups, Creighton et al. and Raina et al. (Missiakas et al., 1994; Zapun et al.,

1995). Since then, a number of assays have been devised for, or adapted to, the study of

bacterial disulfide isomerases. To better understand these enzymes, it is necessary to

investigate their in vivo and in vitro activity. In vivo studies shed light on how the protein

functions in its natural, very crowded, environment as a member of a rather complex

network of interactions. In vitro or biochemical assays, on the other hand, reveal the

intrinsic characteristics of the isolated enzyme in greater detail since the experiments are

performed in a controlled and precise environment. Both methods complement each

other. In the following section, a brief description of the different in vivo and in vitro

methods utilized in the characterization of the engineered disulfide isomerases is

presented.

1.5.1 In Vivo Studies

1.5.1.1 Isomerase Activity –tPA Assay

Human tPA is a thrombolytic protein of therapeutical importance used in the

treatment of acute myocardial infarction and stoke. It is a 527 amino acid secreted serine

protease containing 35 cysteines that require oxidative folding to form 17 disulfide

bonds. Structurally, tPA has five different domains: a finger region, an epidermal growth

factor-like subdomain, two kringle domains and a catalytic domain (Fig A.1). In order to

dissolve thrombi, tPA converts plasminogen into plasmin which in turn degrades the

30

Page 46: Copyright by Silvia A. Arredondo 2009

fibrin forming the thrombus. An important physiological characteristic of tPA is the

ability to bind to fibrin while activating plasminogen, minimizing the probability of

random activation of plasminogen that may lead to internal bleeding. This makes the

molecule particularly interesting therapeutically and commercially (van Zonneveld et al.,

1986).

Figure 1.15. Human tissue plasminogen activator (tPA)

Expression of tPA in the periplasm of E. coli results in misfolded, therefore

inactive, protein; however, Qiu et al. demonstrated that the overexpression of DsbC

allows the production of high yields of full-length tPA with specific activity comparable

to naturally produced tPA. These findings led to the design of an enzymatic assay based

on the proteolytic activity of tPA in which the DsbC-dependent production of active tPA

31

Page 47: Copyright by Silvia A. Arredondo 2009

is monitored in vivo. In this assay, vtPA, a truncated version of tPA that consists of the

kringle 2 and catalytic domains and contains 18 cysteines that form 9 disulfide bonds, is

co-expressed with DsbC or the protein being analyzed for isomerase activity. Following

cell lysis and incubation with a sprectrozyme/plasminogen buffer, the yield of properly

folded vtPA is reported chromogenically by a change in absorbance observed as

spectrozyme reacts with the plasmin resulting from plasminogen cleavage by the

correctly folded and active vtPA (Lafond et al., 2003; Qiu et al., 1998).

1.5.1.2 Isomerase Activity –AppA Assay

The acid phosphatase phytase (AppA) is a periplasmic enzyme native to E. coli

which catalyzes hydrolysis of phosphate moieties. The crystal structure indicates the

presence of three consecutive and one nonconsecutive disulfide bonds. The proper

folding of AppA is dependent upon disulfide bond isomerization by DsbC and its activity

is easily detected using a colorimetric substrate. Although AppA is an attractive reporter

of isomerase activity, the assay is limited by a high background level (Berkmen et al.,

2005).

1.5.1.3 Isomerase Activity –CuCl2 Viability Assay

Wild type E. coli cells grow in media containing CuCl2 concentrations as high as

11mM, higher concentrations become lethal. When growing DsbA-deficient cells in the

presence of CuCl2, cysteine thiol oxidation is catalyzed by Cu2+ ions leading to aberrant

disulfide bond formation patterns. Isomerization by DsbC then becomes an essential

32

Page 48: Copyright by Silvia A. Arredondo 2009

process for survival as evidenced by the observation that while dsbA- cells grow identical

to wild type cells, dsbA- dsbC- cells are not viable at CuCl2 concentrations higher than

3mM (Hiniker et al., 2005). The range of viability of double mutant (dsbA- dsbC-) cells

expressing test proteins while exposed to high concentrations of CuCl2 is an indicator of

the in vivo isomerase activity of the protein.

1.5.1.4 Redox State Assay

In order to determine the redox state of oxidoreductases in vivo, the cell culture

expressing the protein of interest is treated with trichloroacetic acid to precipitate proteins

preventing further thiol:disulfide rearrangements. The pellets are then resuspended in the

presence of 4-acetamido-4'-maleimidylstilbene-2,2'-disulfonic acid, disodium salt (AMS)

which reacts with free thiols, increasing the molecular mass by 490 Daltons per thiol.

Comparison of the electrophoretic mobility of in vitro generated standards with the in

vivo protein samples reveals the redox state of the protein in question. The oxidized

standard is generated by omitting AMS exposure. The reduced standard is prepared by

pre-treating the protein with DTT. For the DsbC reduced standard, the thiols of the four

disulfide-forming cysteines in each subunit react with AMS increasing the M.W. by ~2

kDa per monomer. In vivo, however, the thiolates involved in the structural disulfide of

DsbC are not accessible, and the protein is therefore found in a partially reduced state in

which only the catalytic cysteine thiols have been modified by AMS increasing the M.W.

by ~ 1 kDa per monomer and resulting in a different mobility pattern (Joly and Swartz,

1997; Rietsch et al., 1997).

33

Page 49: Copyright by Silvia A. Arredondo 2009

1.5.1.5 Oxidase Activity –PhoA Assay

Alkaline phosphatase (PhoA) is a zinc-containing periplasmic protein that can

generate inorganic phosphate from different phosphorylated derivatives in E. coli. This

enzyme consists of two identical subunits and it requires the formation of two disulfide

bonds per monomer in order to be active. Thiol oxidation only takes place after the

protein has been exported into the periplasm and is dependent upon the action of DsbA.

The enzymatic activity of PhoA can be analyzed using a colorimetric substrate. When

dsbA- strains are grown under phosphate starvation conditions, no PhoA activity is

observed, however, if an enzyme capable of disulfide bond formation is overexpressed in

this background, the level of PhoA activity restoration can be used as an indicator of in

vivo oxidase activity (Akiyama and Ito, 1993; Brickman and Beckwith, 1975).

1.5.1.6 Oxidase Activity –Motility Assay

Escherichia coli cells are naturally motile. Each cell is propelled by about six

flagella assembled by a number of different proteins. Proper assemblage of the flagellum

P-ring requires the catalytic activity of DsbA for the formation of a disulfide bond in the

protein FlgI; failure to form this bond results in lack of motility for the cells (Dailey and

Berg, 1993). To assess oxidase activity, cells are spotted on minimal-media soft-agar

plates and incubated overnight. While dsbA- cells will remain confined to the original

spot, wild type cells or cells expressing an effective oxidase will form a halo indicative of

34

Page 50: Copyright by Silvia A. Arredondo 2009

their level of motility. The size of the halo qualitatively reports the level of in vivo

oxidase activity.

1.5.2 In vitro Studies

1.5.2.1 Isomerase Activity- RNase Refolding Assay

Bovine pancreatic ribonuclease A (RNase A) requires the formation of four

disulfide bonds to be properly folded and active. When previously reduced and denatured

RNase A is incubated with an effective isomerase in a GSH/GSSG redox buffer, its

oxidative renaturation can be measured spectrophotometrically by recording the

hydrolysis of the substrate cCMP. The concentration of active RNase at any point is

calculated from the change of absorbance with respect to time. The in vitro isomerase

activity of a test protein is then reported as the rate of activated RNase per molar

concentration of test protein (Lyles and Gilbert, 1991).

1.5.2.2 Reductase Activity- Insulin Assay

Insulin contains two interchain disulfide bonds in its molecular structure. When

these disulfide bonds are broken by reduction, the β-chain aggregates and a white

precipitate forms. The in vitro reductase activity of an enzyme can be quantified by

monitoring the time-dependent change in turbidity of a solution containing the test

protein and insulin in the presence of DTT; activity is reported as the absorbance rate

corrected for the initial lag time of the reaction (Holmgren, 1979a; Holmgren, 1979b;

Martinez-Galisteo et al., 1993).

35

Page 51: Copyright by Silvia A. Arredondo 2009

1.5.2.3 Chaperone Activity- Citrate Synthase Assay

Citrate synthase (CS) is a homodimeric mitochondrial protein that catalyzes the

first step of the citric acid cycle, the condensation of oxaloacetic acid and acetyl-CoA to

citrate and coenzyme A. CS is readily inactivated upon incubation at 43º forming

aggregates that render the solution turbid. Aggregation can be monitored by

spectrometry. Molecular chaperones are able to suppress the aggregation process

effectively. The in vitro chaperone properties of the test enzymes are analyzed by

monitoring the restoration of the catalytic activity of thermally denatured CS (Buchner et

al., 1998).

36

Page 52: Copyright by Silvia A. Arredondo 2009

1.6 Research Objectives

The ultimate objective of these studies is optimizing the E. coli disulfide bond

formation machinery for the efficient folding of heterologous proteins containing

complex disulfide bond patterns that might be of therapeutic or commercial interest.

Additionally, it is of fundamental interest to enhance our understanding of the basic

scientific principles involved in disulfide bond formation and isomerization. Since the

rearrangement of non-native disulfide bonds represents the rate-limiting step during the

folding of multidisulfide containing proteins, we are particularly interested in developing

an improved disulfide bond isomerase. The questions we sought to address in this study

were (i) what are the basic functional components of an isomerase and can we construct

an artificial isomerase de novo based on knowledge of the function and interactions of

these basic functional components? (ii) Are two catalytic sites or domains strictly

necessary for successful isomerization in vivo? and (iii) What residues in the substrate are

recognized by the isomerase?

Previous studies performed in the Georgiou lab led to the hypothesis that the

structural features necessary for isomerization are: the presence of two catalytic domains,

the formation of a peptide-binding cleft, and the proper relative orientation of the

catalytic sites (Segatori et al., 2006; Segatori et al., 2004). Our goal in chapter 2 was

investigating whether these three features are necessary and sufficient for the catalysis of

isomerization through the engineering and characterization of a series of completely

artificial isomerases that structurally comprise of domains having the respective functions

but have no amino acid homology to DsbC. We were able to determine that, in fact, the

37

Page 53: Copyright by Silvia A. Arredondo 2009

resulting chimeras exhibit high levels of isomerization activity, comparable to that of

DsbC, but were also able to serve as oxidases replacing DsbA in the catalysis of disulfide

bond formation. Directed evolution was utilized with the objective of further enhancing

isomerase activity and the associated Cu+2 resistance. The mutations identified highlight

the importance of specific residues in the catalysis of isomerization.

In chapter 3, our objective was to determine whether the presence of two catalytic

domains in DsbC is strictly necessary for efficient isomerization. Therefore, we

constructed and characterized a series of heterodimers containing only one active site or

only one thioredoxin domain. The results of this study indicate that indeed one catalytic

site or domain is sufficient for isomerization activity both, in vivo and in vitro, as long as

the substrate binding site is kept intact. Furthermore, our studies suggest that the

determining feature of the isomerase is the availability of a three dimensional substrate

binding site resulting from the dimerization of the DsbC units.

As mentioned above, DsbC forms a V-shape structure that contains a hydrophobic

cleft which plays an essential role in substrate recognition and binding. Before we started

this investigation, no specific amino acid residues had been identified as important for

substrate binding. The purpose of the study presented in chapter 4 was the identification

and analysis of specific residues or amino acid motifs involved in the DsbC-substrate

binding interaction. Two peptide libraries have been screened in order to isolate peptides

that strongly bind to DsbC; further analysis of the isolated binders will give us important

clues in understanding how substrate interaction takes place.

38

Page 54: Copyright by Silvia A. Arredondo 2009

Chapter 2

De Novo Design and Evolution of Artificial Disulfide Isomerase

Enzymes Analogous to the Bacterial DsbC

2.1 Introduction

As described in the previous chapter, the formation of protein disulfide bonds in

the E. coli periplasm involves two complementary, but also competing processes:

cysteine-thiol oxidation, and isomerization of non-native disulfide bonds. The DsbA-

DsbB, and DsbC/DsbG-DsbD enzyme systems catalyze, respectively, the formation and

isomerization of protein disulfides (Collet and Bardwell, 2002; Kadokura et al., 2003).

DsbA is a soluble periplasmic enzyme that catalyzes the rapid formation of disulfide

bonds in substrate polypeptides. The membrane protein DsbB recycles DsbA via its

extremely oxidizing active site disulfide (Grauschopf et al., 2003). Disulfide bonds tend

to form preferentially between consecutive cysteines in the nascent polypeptide, without

regard for the correct disulfide pairing that is found in the native conformation (Berkmen

et al., 2005). The isomerization of incorrect disulfide bonds in bacteria is catalyzed by

DsbC and DsbG. These enzymes are maintained in the reduced, catalytically active state

by the membrane protein DsbD, which in turn, receives electrons from the cytoplasmic

39

Page 55: Copyright by Silvia A. Arredondo 2009

thioredoxin 1, the product of the trxA gene (Kadokura et al., 2003). DsbG has a more

restricted substrate specificity in comparison to DsbC (Messens and Collet, 2006).

The rearrangement of non-native disulfide bonds in eukaryotic cells is catalyzed

by protein disulfide isomerase (PDI) in the endoplasmic reticulum. Although there is

little sequence homology between DsbC and PDI, biochemical and structural studies have

revealed a remarkable degree of similarity in the overall structure of the two enzymes,

suggestive of convergent evolutionary processes [Fig 2.1, (Tian et al., 2006)]. DsbC is a

“V” shaped homodimer with the dimerization region contributed by the N-terminal

domain of each monomer. The topology of PDI, although resembling a “U” structure, is

noticeably similar to the molecular architecture of DsbC. In both proteins the inside,

solvent-exposed surface is enriched in hydrophobic residues and forms a putative peptide

binding cleft. Disulfide bond isomerization is catalyzed by two thioredoxin domains,

each located at the two ends of the molecule, with the CXXC catalytic active sites facing

each other. In PDI the two active site catalytic domains are asymmetric, allowing the

protein to exist as a mixture of oxidized and reduced molecules that catalyze both

disulfide bond formation and isomerization in the ER (Kulp et al., 2006). In contrast,

DsbC is maintained in a fully reduced state in the periplasm. DsbC’s resistance to

oxidation by the DsbB-DsbA system is due to its dimeric structure and, in particular, to

the orientation of the active sites, and the length of the α-helical linker (Segatori et al.,

2006). The structural comparison of DsbC and DsbG reveals that the latter contains a

longer α-helical linker and, has negatively charged patches in its active site cleft as well

40

Page 56: Copyright by Silvia A. Arredondo 2009

as a charged groove in the N-terminal domain. Nonetheless, the overall structure of DsbG

closely resembles DsbC [Fig 2.1 (Messens and Collet, 2006).]

Figure 2.1. The structures of DsbC, DsbG and PDI exhibit similar domain arrangement.

In DsbC and its homologues, the N-terminal dimerization domains fulfill at least

two functions: first, they are able to fold independently of the catalytic domains (Yeh et

al., 2007) forming a putative peptide binding cleft that might be important for substrate

recognition and for the chaperone function of the enzyme (Sun and Wang, 2000; Zhao et

al., 2003); second, dimerization of the DsbC monomer is necessary to prevent oxidation

of the protein by DsbB (Bader et al., 2001). The dimerization domain is connected to the

catalytic thioredoxin domain by means of a long α-helical linker. In an earlier study

(Segatori et al., 2004), we showed that the DsbC thioredoxin domain could be replaced

by other proteins belonging to the same superfamily, such as the E. coli thioredoxin 1

(TrxA), which normally serves as a cytoplasmic reductant, or the periplasmic oxidant

DsbA. The resulting DsbC-TrxA and DsbC-DsbA chimeras were maintained in a

partially reduced state by DsbD and conferred high disulfide isomerase activity in vivo. In

41

Page 57: Copyright by Silvia A. Arredondo 2009

addition, the oxidized form of the chimeras was shown to serve as a catalyst for de novo

disulfide bond formation (Segatori et al., 2004).

Thioredoxin (PDI-like) and DsbA exhibit very weak disulfide isomerase activity

in vitro (Lundstrom et al., 1992; Segatori et al., 2004). In vivo, both proteins are readily

oxidized by DsbB and therefore they cannot catalyze disulfide bond isomerization (Jonda

et al., 1999). Yet, as mentioned above, fusion of DsbA or TrxA to the N-terminal

dimerization domain of DsbC gives rise to chimeras that efficiently catalyze disulfide

isomerization in the periplasm. This result highlights the importance of the DsbC

dimerization domain in endowing thioredoxin proteins with enhanced ability to catalyze

disulfide bond isomerization. We envision this isomerase activity arises as a consequence

of: (i) the proximity and higher effective concentration of active site cysteines within a

thioredoxin dimer, (ii) the presence of the putative peptide cleft that may be playing an

important role in the recognition of substrate proteins in need of isomerization, (iii) a

dimeric structure probably being a better substrate for DsbD thus allowing the chimeras

to be maintained in the catalytically-active reduced state. Exploring each of the above

possibilities independently via a classical mutagenesis approach is challenging.

Dimerization, interactions with DsbD, and the formation of the peptide cleft appear to be

highly coupled and therefore mutations that affect one of these processes may also affect

the other. Moreover, the amino acids critical for peptide substrate binding or for

interactions with DsbD have not been defined. Therefore, instead we used a synthetic

approach whereby we constructed and characterized a series of fusions of thioredoxin

domains to sequences that mediate dimerization with or without peptide binding site. We

42

Page 58: Copyright by Silvia A. Arredondo 2009

show that chimeras comprising of DsbA fused to the dimerization region and α-helical

linker of the periplasmic chaperone FkpA, that has no disulfide isomerase activity on its

own, recapitulate the activities displayed by DsbC. Under selective pressure the FkpA

chimeras could be evolved to exhibit higher disulfide bond isomerization activity in vivo,

allowing growth in the presence of high CuCl2 concentrations nearly identical to that

afforded by DsbC. On the other hand, selections aimed at rendering FkpA-DsbA resistant

to periplasmic oxidation by DsbA-DsbB resulted in the mutation of the C-terminal

cysteine in the active site, a mutation that eliminated the oxidase activity of the chimera,

but also reduced disulfide bond isomerization.

43

Page 59: Copyright by Silvia A. Arredondo 2009

2.2 Materials and Methods

2.2.1 Strains and Plasmids

The bacterial strains and plasmids used in this study are listed in Table 2.1B. The

fkpA-dsbA chimeric genes were constructed by overlap extension PCR using the primers

listed in Table 2.1A, digested with XbaI and HindIII and cloned into pBAD33 (Guzman

et al., 1995). All the chimeric genes contain a sequence that encodes a C-terminal

hexahistidine tag. For protein purification purposes the gene fusions were digested with

XbaI and HindIII, cloned into pET28(a), and transformed into E. coli BL21(DE3) cells.

Table 2.1 (A) Primers and (B) Bacterial strains utilized in this study.

PrimersXbaIDsbCss.f GAGCTCGAATTCTCTAGATTAAAGAGGAGAAAGGTACCCATGATGAAGAAAGGTTTTAT1.f ATGAAGAAAGGTTTTATGTTGTTTACTT2.r AAGCCTGAAAACGCCGCTAACAAAGTAAACAACATAAAACCTT3.f GTTAGCGGCGTTTTCAGGCTTTGCTCAGGCTGCTGAAGCTGCA4.r TGTCAGCAGCTGTAGCAGGTTTTGCAGCTTCAGCAGCCTGAGC5.f AACCTGCTACAGCTGCTGACAGCAAAGCAGCGTTCAAAAATGA6.r TGCATAAGCTGATTTCTGATCGTCATTTTTGAACGCTGCTTTG7.f GATCAGAAATCAGCTTATGCACTGGGTGCCTCGCTGGGTCGTT8.r TCTTTTAGAGAGTTTTCCATGTAACGACCCAGCGAGGCACCCA9.f CATGGAAAACTCTCTAAAAGAACAAGAAAAACTGGGCATCAAA10.f CGATCAGCTGATCTTTATCCAGTTTGATGCCCAGTTTTTCTTG11.f TGGATAAAGATCAGCTGATCGCTGGTGTTCAGGATGCATTTGC12.r GTCGGAGAGTTTGCTCTTATCAGCAAATGCATCCTGAACACCA13.f GATAAGAGCAAACTCTCCGACCAAGAGATCGAACAGACTCTAC14.r TTCACGCGAGCTTCGAATGCTTGTAGAGTCTGTTCGATCTCTT15.f AGCATTCGAAGCTCGCGTGAAGTCTTCTGCTCAGGCGAAGATG16.r CGTTATCAGCCGCGTCTTTTTCCATCTTCGCCTGAGCAGAAGA17.f AAAAAGACGCGGCTGATAACGAAGCAAAAGGTAAAGAGTACCG18.r TTTCTCTTTGGCAAATTTCTCGCGGTACTCTTTACCTTTTGCTfkpA11-dsbA2.r GTAGTGTACTGTTTACCATCTTCATACTGTTTCTCTTTGGCAAATTTCfkpA11-dsbA3.r CAGGGTAGTGTACTGTTTACCATCTTCATATTTCTCTTTGGCAAATTTCfkpA11-dsbA4.r CAGGGTAGTGTACTGTTTACCATCTTCTTTCTCTTTGGCAAATTTCfkpA11-dsbA5.r CAGGGTAGTGTACTGTTTACCATCTTTCTCTTTGGCAAATTTCfkpAdsbA2.f GAAATTTGCCAAAGAGAAACAGTATGAAGATGGTAAACAGTACACTACfkpAdsbA3.f GAAATTTGCCAAAGAGAAATATGAAGATGGTAAACAGTACACTACCCTGfkpAdsbA4.f GAAATTTGCCAAAGAGAAAGAAGATGGTAAACAGTACACTACCCTGfkpAdsbA5.f GAAATTTGCCAAAGAGAAAGATGGTAAACAGTACACTACCCTGDsbAHisHindIII.r TTTTTAAGCTTTTAGTGGTGGTGGTGGTGGTGTTTTTTCTCGGACAGATATTTCAErrPXbaIDsbC.f TCGAACGCTCTAGATTAAAGAGGAGAAAGGTACCCATGATGErrPDsbAHisHindIII.r TATTGCCCAAGCTTTTAGTGGTGGTGGTGGTGGTG

A

44

Page 60: Copyright by Silvia A. Arredondo 2009

BStrains Relevant Genotype SourceXL1-blue recA1 endA1 gyrA96 thi-1 hsdR17 supE44 relA1 Stratagene

lac [F´ proAB lacIqZΔM15 Tn10 (Tetr)]DHB4 araD139 (araA-leu)7679 (codB-lac)X74 galE15 galK16 Laboratory Collection

rpsL150 relA1 thi phoA (PvuII) phoR malF3 F [lac+ (lacI) pro]MC1000 araD139 (araA-leu)7679 (codB-lac)X74 galE15 galK16 Ref (1)

rpsL150 relA1 thi F [lac+ (lacI) pro]LM106 MC1000 dsbA::kan5 Laboratory CollectionLM102 MC1000 dsbB::kan5 Ref (2)PB351 SF100 Δ degP::kan Δ dsbC Ref (3)PB401 SF100 Δ dsbA Ref (3)BL21(DE3) F ompT dcm (DE3) Laboratory CollectionMB706 DHB4 Δ

1. Calhoun, D. H., Wallen, J. W., Traub, L., Gray, J. E. & Kung, H. F. (1985) J Bacteriol 161, 128-32. 2. Masip, L., Pan, J. L., Haldar, S., Penner-Hahn, J. E., DeLisa, M. P., Georgiou, G., Bardwell, J. C. & Collet, J. F. (2004) Science 303, 1185-9. 3. Bessette, P. H., Qiu, J., Bardwell, J. C., Swartz, J. R. & Georgiou, G. (2001) J Bacteriol 183, 980-8. 4. Bessette, P. H., Cotto, J. J., Gilbert, H. F. & Georgiou, G. (1999) J Biol Chem 274, 7784-92.

dsbC, dsbA::kan Beckwith laboratory

Plasmids Relevant Genotype SourcepET-28(a) T7 expression vector, C-terminal 6x histidine tag NovagenpBADdsbC dsbC from Escherichia coli in pBAD33 Laboratory CollectionpBADdsbA dsbA from Escherichia coli in pBAD33 Laboratory CollectionpBAD-FkpA fkpA from Escherichia coli in pBAD33 This workpTrcStIIvtPA tPA(Δ6-175) with StII leader in pTrc99A Ref (4)pBAD-FkpA-DsbA2 FkpA(1-114) fused to DsbA(2-189) in pBAD33 This workpBAD-FkpA-DsbA3 FkpA(1-114) fused to DsbA(3-189) in pBAD33 This workpBAD-FkpA-DsbA4 FkpA(1-114) fused to DsbA(4-189) in pBAD33 This workpBAD-FkpA-DsbA5 FkpA(1-114) fused to DsbA(5-189) in pBAD33 This workpBAD-FkpA-DsbA2m7 FkpADsbA2 H145Y in pBAD33 This workpBAD-FkpA-DsbA2m18 FkpADsbA2 D20G D95N H145Y in pBAD33 This workpBAD-FkpA-DsbA2m33 FkpADsbA2 M35T A111V C146S H154Y in pBAD33 This workpBAD-FkpA-DsbA2m33S146C FkpADsbA2m33 S146C in pBAD33 This workpBAD-FkpA-DsbA2C146A FkpADsbA2 C146A in pBAD33 This workpET28-FkpA-DsbA2 FkpA(1-114) fused to DsbA(2-189) in pET28 This workpET28-FkpA-DsbA3 FkpA(1-114) fused to DsbA(3-189) in pET28 This workpET28-FkpA-DsbA4 FkpA(1-114) fused to DsbA(4-189) in pET28 This workpET28-FkpA-DsbA5 FkpA(1-114) fused to DsbA(5-189) in pET28 This workpET28-FkpA-DsbA2m7 FkpADsbA2 H145Y in pET28 This workpET28-FkpA-DsbA2m18 FkpADsbA2 D20G D95N H145Y in pET28 This workpAppA EcoRI/XbaI cut appA cloned into pBAD18 AmpR Beckwith laboratory

45

Page 61: Copyright by Silvia A. Arredondo 2009

2.2.2 In vivo Disulfide Bond Formation and Isomerization

E. coli PB351 (SF100 ΔdegP::kan, ΔdsbC) and E. coli PB401 (SF100 ΔdsbA)

were co-transformed with pBAD33 derivatives encoding the fkpA-dsbA chimera genes,

and with pTrcStIIvtPA, a pTrc99a derivative encoding the vtPA gene fused to the stII

leader peptide (Bessette et al., 1999a). Cell growth and vtPA assays were performed as

described earlier (Segatori et al., 2004).

To measure in vivo oxidase activity overnight cultures were grown in LB medium

with 50 μg/ml of kanamycin and 25 μg/ml of chloramphenicol, and subcultured 1:100 in

low phosphate minimal medium containing MOPS salts, 0.2% glycerol, 0.2% glucose,

0.2% casein amino acids, and 0.5 μg/ml thiamine, with 50 μg/ml of kanamycin and 25

μg/ml of chloramphenicol. When the cell density reached OD600=0.4, arabinose was

added to a final concentration of 0.2% w/v. After two hours of growth, 30 μl of cells

were mixed with 20 μl of a buffer containing 0.4 M iodoacetamide and lysis buffer (B-

PERTM, Pierce) in a 1:2 ratio. The activity of alkaline phosphatase was determined as

described previously (Brickman and Beckwith, 1975). Phytase activities were also

determined as described (Berkmen et al., 2005).

CuCl2 resistance was evaluated in E. coli MB706 (DHB4 ΔdsbC, dsbA::kan).

Cells were grown on Brain Heart Infusion (BHI) medium at 37°C with 25 μg/ml of

chloramphenicol, and 50 μg/ml of kanamycin. Cell counts were normalized by dilution,

and plated on BHI medium with 0.2% arabinose, antibiotics as above, and CuCl2

concentrations between 0 to 12 mM. MB706 cells do not grow in the presence of CuCl2

46

Page 62: Copyright by Silvia A. Arredondo 2009

at concentrations higher than 3mM; isomerization, resulting from the overexpression of

DsbC, restores growth to wt levels (10-11mM CuCl2.)

2.2.3 Directed Evolution for Increased CuCl2 Resistance

The fkpA-dsbA2 gene was subjected to random mutagenesis by error-prone PCR

(Fromant et al., 1995), and digested and ligated into pBAD33 as described in section

2.2.1. The ligation product was transformed into E. coli XL1-blue yielding a library size

of 106 clones with a 0.6% error rate. The plasmids were purified and re-transformed into

E. coli MB706. In order to screen for CuCl2 resistance, cells were collected, diluted and

plated on agar plates with BHI medium, 25 µg/ml of chloramphenicol, 50 µg/ml of

kanamycin, and containing 11 mM CuCl2. After approximately 20 hours, colonies were

observed; 32 of these colonies were picked randomly and the fkpA-dsbA2 DNA was

sequenced. In addition, the MB706 cells harboring the library were plated on BHI

containing 10mM CuCl2, resulting colonies were grown overnight in LB medium and

screened for motility by making 1:50 dilutions in minimal medium (M9 salts, 0.1%

casein amino acids, 2mM MgSO4, 5 μg/ml thiamine, 0.2% glycerol supplemented with

50 μg/ml of kanamycin and 25 μg/ml of chloramphenicol) and growing at 37°C for an

additional 6 h; plates containing the same medium with 0.3% agar and 0.2% arabinose

were spotted with 3 μl of the normalized cultures and incubated at 37°C for 20 h. The

presence of halos around the culture spots was indicative of motility as a result of in vivo

oxidase activity.

47

Page 63: Copyright by Silvia A. Arredondo 2009

2.2.4 Expression, Purification, and Biochemical Assays

For the purification of the FkpA-DsbA chimera proteins, the appropriate genes

were cloned behind the T7 promoter in pET28(a) at XbaI and HindIII sites, and

transformed into E. coli BL21(DE3). Protein expression and purification was performed

as previously described (Segatori et al., 2004). All proteins used in this study were more

than 95% pure as judged by Coomassie stained SDS-PAGE electrophoresis. The rate of

insulin reduction, the renaturation of reduced, denatured RNase A, and the protection of

citrate synthase from thermal inactivation were determined according to published

procedures (Buchner et al., 1998; Holmgren, 1979b; Lyles and Gilbert, 1991; Segatori et

al., 2004).

48

Page 64: Copyright by Silvia A. Arredondo 2009

2.3 Results

2.3.1 Engineering the Artificial Disulfide Bond Isomerases

Initially, we examined whether simply forcing the dimerization of thioredoxin can

give rise to a protein having disulfide isomerization activity. The E. coli thioredoxin with

a PDI-like active site (CGHC), designated here as TrxA”PDI-like”, has a redox potential

E0= -235mV (Krause et al., 1991; Lundstrom et al., 1992) which is higher than that of the

wt thioredoxin (E0=-270 mV) (Chivers et al., 1996). TrxA”PDI-like” becomes oxidized

by DsbB and cannot catalyze disulfide bond reduction or isomerization in vivo (Jonda et

al., 1999). TrxA”PDI-like” was forced to dimerize by fusing it to the C-terminal of the

GCN4 coiled-coil dimerization domain (Pack and Pluckthun, 1992). The in vivo disulfide

isomerase activity of the dimerized TrxA”PDI-like” was evaluated by determining the

yield of correctly folded vtPA, a truncated version of human tissue plasminogen activator

consisting of the kringle 2 and protease domains, and containing a total of 9 disulfide

bonds. The folding of vtPA, secreted into the periplasm using a bacterial signal peptide,

is limited by disulfide isomerization and requires co-overexpression of DsbC (Qiu et al.,

1998). In cells expressing DsbC under the control of the arabinose promoter, and grown

with 0.2% w/v arabinose, active vtPA accumulates at a 25-fold higher level relative to

control cells (without plasmid). Periplasmic expression of dimerized TrxA”PDI-like”

under identical conditions did not result in any increase in vtPA yield over the isogenic

background cells (i.e. cells containing the empty vector). Gel filtration chromatography

of the purified protein revealed that it was present exclusively as a dimer, as expected

(data not shown). Next, in order to provide better spatial separation between the two

49

Page 65: Copyright by Silvia A. Arredondo 2009

active sites in the chimera, the α-helical linker that separates the catalytic thioredoxin

domain from the dimerization domain in DsbC was inserted between the GCN4 leucine

zipper and TrxA”PDI-like”. Expression of the latter construct resulted in a modest, but

statistically significant, increase in the yield of vtPA to a level about 20% of that afforded

by DsbC (data not shown).

The above results indicate that the fusion of a dimerization domain followed by

an α-helical linker endows thioredoxin exported to the periplasm with a low but

detectable level of disulfide isomerase activity. Natural disulfide isomerase enzymes in

addition to having two catalytic domains also contain putative peptide binding sites that

confer chaperone activity and may be involved in interactions with substrate proteins.

Similar to the topology of DsbC, the cis/trans peptidyl-prolyl isomerase FkpA folds into

a homodimer, with each monomer consisting of an N-terminal dimerization domain and a

C-terminal catalytic domain, joined by a long α-helical linker. The two domains are

structurally and functionally independent. FkpA exhibits chaperone activity, which, as in

DsbC, is conferred by its dimerization domain, although its peptide-binding cleft is less

hydrophobic (Saul et al., 2004). However, unlike DsbC, FkpA does not display disulfide

isomerization activity, therefore its overexpression results in background levels of

correctly folded vtPA (Fig 2.2).

50

Page 66: Copyright by Silvia A. Arredondo 2009

0

5

10

15

20

25

DsbC FkpA --

Rel

ativ

e vt

PA A

ctiv

ity

SF100 dsbC

Figure 2.2. In vivo isomerase activity of FkpA.Yield of active vtPA in dsbC cells transformed with pTrcStIIvtPA and pBAD derivatives encoding DsbC or FkpA compared to cells harboring only pTrcStIIvtPA.

The dimerization domain and α-helical linker of FkpA were fused to TrxA“PDI-

like” or to DsbA. DsbA, the periplasmic catalyst of oxidation, is also a thioredoxin

superfamily enzyme, but shares little amino acid homology with the thioredoxin domain

of DsbC (6% identity and 13% aa similarity). DsbA was selected because the redox

potential of its catalytic cysteine pair (E0=-120 mV) approximates that of DsbC (E0=-130

mV). The N-terminal dimerization domain of FkpA (residues 1-114) including the long

α-helical linker (aa 70-114), which is the minimal protein domain exhibiting chaperone

activity (Saul et al., 2004), was fused to either TrxA”PDI-like” or to DsbA. Previously,

Segatori et al. (Segatori et al., 2006) showed that deletions in the α-helical linker of

DsbC, which alter the face to face alignment of the active site cysteines in the two

domains, render the protein susceptible to oxidation by DsbB. To explore the effect of the

51

Page 67: Copyright by Silvia A. Arredondo 2009

fusion junction sequence, we constructed a series of chimeras in which the residue 114 of

FkpA is fused to consecutive amino acids of the α-helical N-terminal region of the

thioredoxin domain protein (Fig 2.3). Specifically, fusions were constructed to the 1st,

2nd, 3rd or 4th amino acid of TrxA”PDI-like” or the 2nd, 3rd, 4th, or 5th residue of mature

DsbA, generating, respectively, FkpA-TrxA1 to FkpA-TrxA4 and FkpA-DsbA2 to FkpA-

DsbA5. Since the fusion junction is created by joining two α-helical segments, it is likely

that the resulting chimeric region assumes helical secondary structure, in which case

sequential deletion of amino acids should lead to a change in the orientation of the C-

terminal catalytic domain by an angle of 100º.

Figure 2.3. Protein structures of DsbA and FkpA, and a predictive molecular model of FkpA-DsbA2. The amino acid sequence at the fusion region of the two proteins is shown.

52

Page 68: Copyright by Silvia A. Arredondo 2009

Several of the chimeric proteins were purified to near homogeneity. Gel filtration

FPLC revealed that the proteins elute as a single peak with identical retention times,

which however were shorter relative to that of DsbC (Fig 2.4A). The fusions had an

apparent M.W. of around 100 kDa which is higher than the predicted M.W. for the

dimers (67 kDa). Chemical cross-linking (Bessette et al., 1999b) further confirmed that

the chimeras form dimers, consistent with the known ability of the FkpA (1-114) domain

to dimerize (Saul et al., 2004) (Fig 2.4B).

Superdex 200

0

100

200

300

400

500

600

0 5 10 15 20

ml

UV (mAU)

FkpA-DsbA2FkpA-DsbA2m7FkpA-DsbA2m18DsbC

A

Flow 0.5 ml/min

53

Page 69: Copyright by Silvia A. Arredondo 2009

B

0mM 0.1mM 0.2mM 0.3mM 0.4mM 0.5mM

FkpA-DsbA2

31.3 kDa

40.4 kDa

82.7 kDa

Monomer

Dimer

DsbC

0mM 0.1mM 0.2mM 0.3mM 0.4mM 0.5mM

31.3 kDa

40.4 kDa

82.7 kDa

Monomer

Dimer

Figure 2.4. Oligomeric state of the FkpA-DsbA chimeras. (A) Gel filtration FPLC of DsbC, FkpA-DsbA2, FkpA-DsbA2m7 and FkpA-DsbA2m18. Purified proteins were run on a SuperdexTM 200 column in PBS-10% glycerol buffer. (B) Chemical crosslinking: The oligomeric state of DsbC and FkpA-DsbA2 was examined by incubating the purified proteins in the presence of increasing concentrations (0-0.5 mM) of the crosslinking reagent ethylene glycol-bis(succinimidylsuccinate) (EGS) in DMSO and analyzing the products by SDS-PAGE.

54

Page 70: Copyright by Silvia A. Arredondo 2009

2.3.2 In vivo Studies

Following induction of protein expression with arabinose, all the FkpA chimeras

accumulated to nearly identical levels in vivo, albeit at a slightly lower level than DsbC,

as determined by Western blotting with a polyclonal antibody that recognizes the C-

terminal hexahistidine tag present in all the fusions (representative data for the FkpA-

DsbA fusions shown in Fig 2.5).

25 kDa

37 kDa

DsbC

FkpA-DsbA2

FkpA-DsbA3FkpA-DsbA4FkpA-DsbA5FkpA-D

sbA2m7

FkpA-DsbA2m18

Figure 2.5. Expression level of the FkpA-DsbA chimeras compared to DsbC. Overnight cultures were subcultured, grown to OD600 = 0.8 and induced with arabinose (0.2% final concentration). Three hours after induction, cells were collected and normalized according to OD600. Western blot analysis was carried out using a monoclonal antipolyhistidine HRP-conjugated antibody.

While the FkpA-TrxA1-4 chimeras exhibited higher isomerase activity in the

vtPA assay relative to TrxA”PDI-like” dimerized via a Leu zipper, they were nonetheless

less efficient in the folding of vtPA relative to the FkpA-DsbA enzymes. Specifically the

yield of active vtPA by the former proteins was between 35-50% of the level obtained in

55

Page 71: Copyright by Silvia A. Arredondo 2009

cells expressing DsbC (unpublished data). In contrast, FkpA-DsbA2, FkpA-DsbA3, and

FkpA-DsbA4 conferred yields of enzymatically active vtPA that were 80, 55, and 67% of

the yields obtained by expressing DsbC under identical conditions (Fig 2.6). The ability

of the chimeras to increase the yield of vtPA is also dependent on reduction by DsbD. In

dsbD cells expressing either DsbC or the FkpA-DsbA chimeras, only a low yield of vtPA

was obtained (data not shown).

0

5

10

15

20

25

FkpA-DsbA2

FkpA-DsbA3

FkpA-DsbA4

FkpA-DsbA5

DsbC --

Rel

ativ

e vt

PA A

ctiv

ity

SF100 dsbCSF100 dsbA

Figure 2.6. Disulfide bond isomerization in vivo. Yield of active vtPA in dsbC (grey bars) or dsbA (black bars) cells. E. coli PB351 and PB401 (respectively) were transformed with pTrcStIIvtPA and pBAD derivatives encoding the respective FkpA-DsbA fusion proteins and grown in LB media. Protein synthesis was induced as described in Materials and Methods, and the yield of active vtPA was determined 3 h after induction. Relative activities were obtained by dividing the ΔA405 of each strain (subtracted of the background consisting of a strain not expressing vtPA) by the ΔA405 of a strain expressing vtPA alone.

56

Page 72: Copyright by Silvia A. Arredondo 2009

Importantly, expression of FkpA-DsbA enzymes resulted in maximal yields of phytase

(AppA), a native E. coli enzyme whose folding in the periplasmic space is normally

dependent on isomerization by DsbC (see Figure 2.9 for representative results with

FkpA-DsbA2). Thus, the FkpA-DsbA chimera can catalyze the rearrangement of protein

disulfide bonds in both homologous and heterologous proteins in the E. coli periplasm.

The folding of vtPA, in addition to requiring the overexpression of DsbC, is also

dependent on the presence of DsbA. In E. coli strain SF100 dsbA only background levels

of active vtPA are observed. Multicopy expression of DsbC, failed to increase the yield

of active vtPA, since DsbC cannot normally serve as a protein oxidant (Fig 2.6).

However, the FkpA-DsbA chimeras allowed the folding of vtPA in a dsbA background,

albeit with a lower efficiency relative to wt E. coli.

These chimeras also catalyzed the formation of disulfide bonds in alkaline

phosphatase (PhoA). The catalytically active form of PhoA is a homodimer with two

intramolecular disulfide bonds per monomer. The formation of the disulfide bonds in

PhoA is dependent on the action of the DsbA-DsbB electron transfer relay (Kamitani et

al., 1992). Overexpression of the FkpA-DsbA proteins partially restored alkaline

phosphatase activity in dsbA mutants (Figure 2.7); FkpA-DsbA3 exhibited the highest

PhoA activity (70% of MC1000) while FkpA-DsbA5 gave the lowest activity

corresponding to 30% of MC1000. These results indicate that the FkpA-DsbA enzymes

simultaneously serve as catalysts for both disulfide bond formation and isomerization in

the periplasmic space.

57

Page 73: Copyright by Silvia A. Arredondo 2009

0

20

40

60

80

100

120

FkpA-DsbA2

FkpA-DsbA3

FkpA-DsbA4

FkpA-DsbA5

DsbC -- MC1000

Alk

alin

e Ph

osph

atas

e A

ctiv

ity (%

)

MC1000 dsbAMC1000 dsbBMC1000

Figure 2.7. Disulfide bond formation in vivo. PhoA activity in E. coli MC1000 dsbA (white bars) and MC1000 dsbB (black bars). The alkaline phosphatase activity in the parental isogenic strain MC1000 is shown by the gray bar.

2.3.3 Evolution of FkpA-DsbA Chimeras Under Selective Pressure

We wondered whether selective pressure can be applied to isolate FkpA-DsbA

mutants displaying enhanced disulfide isomerization activity. Because we do not have a

reliable method to quantitatively measure the expression yield of multi-disulfide proteins

such as vtPA in high throughput, we instead took advantage of the finding that disulfide

bond isomerization activity is required for growth of dsbA strains in media containing

elevated concentrations of CuCl2 (Hiniker et al., 2005). In the absence of DsbA, cysteine

thiol oxidation is catalyzed by Cu+2 ions, leading to aberrant disulfide bond formation

that increase the need for isomerization by DsbC. Wt strain DHB4 forms normal colonies

in media containing up to 10 mM CuCl2 and smaller colonies in the presence of 11 mM

58

Page 74: Copyright by Silvia A. Arredondo 2009

CuCl2, whereas the dsbA dsbC mutant MB706 cannot grow with more than 3 mM CuCl2

(Table 2.2). MB706 expressing the FkpA-DsbA enzymes formed normal colonies with 4

mM CuCl2 and smaller colonies at 6 mM, conditions under which cells containing empty

plasmid could not grow (Table 2.2). FkpA-DsbA2, which exhibited the highest isomerase

activity in the expression of vtPA, allowed growth with up to 8 mM CuCl2.

Table 2.2. Physiological function of artificial disulfide isomerases

CuCl2Strain Plasmid 3mM 4mM 6mM 7mM 8mM 10mM 11mM

DHB4 +++ +++ +++ +++ +++ ++-- ++--MB706 pBAD33 +++ ----- ----- ----- ----- ----- -----

B706 pBAD33-dsbC +++ +++ +++ +++ +++ ++-- +---B706 pBAD33-fkpA-dsbA2 +++ +++ +++ +---- +---- ----- -----B706 pBAD33-fkpA-dsbA3 +++ +++ +---- +---- ----- ----- -----B706 pBAD33-fkpA-dsbA4 +++ +++ +---- ----- ----- ----- -----

MB706 pBAD33-fkpA-dsbA5 +++ +++ +---- ----- ----- ----- -----

MB706 pBAD33-fkpA-dsbA2m7 +++ +++ +++ +++ +++ +--- +---B706 pBAD33-fkpA-dsbA2m18 +++ +++ +++ +++ +++ ++-- ++--

Viability ranging from wild-type size single colonies (+++) to no colonies (----)

MM

MM

M

Error prone PCR mutagenesis was used to introduce random mutations to the

fkpA-dsbA2 gene giving rise to a library of 106 independent transformants in MB706. The

library was plated in BHI medium containing 11 mM CuCl2. DNA was isolated from 32

colonies that grew and sequencing revealed 16 distinct clones (Fig 2.8). Remarkably,

12/16 clones shared a point mutation (H145Y) that resulted in the substitution of the

DsbA catalytic site histidine (CPHC), by a tyrosine (CPYC), the same amino acid found

in the DsbC catalytic site (CGYC).

59

Page 75: Copyright by Silvia A. Arredondo 2009

LiLiLi

FkpA-DsbA2 MKKGFMLFTLLAAFSGFAQAAE D DD E E E LDKD QD AD SDQEIE F. . . . . . .... .. .. ...... .. . .. . . . .... .. .. ...... .. . . . . . .... .. .. ...... .. . .. . . . .... .. .. ...... .. . .. . . . .... .. .. ...... .. . . . . . .... .. .. ...... .. . .. . . . .... .. .. ...... .. . . . . . .... .. .. ...... .. . .. . . . .... .. .. ...... .. . .. D . . . .... .. .. ...... .. . .. . . . .... .. .. ...... .. . .. . . . .... .. V. ...... .. . .. . . . .... .. .. ...... .. . .. . . . .... .D .. .. .... .. . .. . . . .... .. .. ...... .. . .. . . . .... .E .. ...... .

E D D E E E E ED E E FEE D E GD KD VE. . . . . . .. . . ... . . .. .. ... . . . . . .. . . ... . . .. .. ... . . . . . .. . . ... . . .. .. ... . . .

LibA2mLiLi

LibA2mLiLi

LibA2mLiLiLibA2m

FkLiLibA2m

LiLiLibA2m

AAKPATAA SKAAFKN QKSAYALGASLGRYM NSLK Q KLGIK QLIAGV AF KSKL QTLQA EARVKSSAQAbA2m1 ..................... ........ ....... G............... .... . ..... ...... .. .... ..... ..........bA2m2 ...................T. ........ ....... ............... .... . ..... ...... .. .... ..... ..........bA2m3 ..................... ........ ....... G............... .... . ..... ...... .. .... ..... ..........

LibA2m7 ..................... ........ ....... ............... .... . ..... ...... .. .... ..... ..........19 ...................T. ........ ....... ............... .... . ..... ...... .. .... ..... ..........

bA2m18 ..................... ........ ....... G............... .... . ..... ...... .. .... ..... ..........bA2m20 ..................... ........ ....... ............... .... . ..... ...... .. .... ..... ..........

LibA2m1a ..................... ........ ....... G............... .... . ..... ...... .. .... ..... ..........3a ..................... ........ ..T.... ............... .... . ..... ...... .. .... ..... ..........

bA2m9a ................Y..T. ........ ....... ....... ....... .... . ..... ...... .. .... ..... ..........bA2m7a ...................T. ........ ....... ............... .... . ..... ...... .. .... ..... ..........

LibA2m6a ..................... ........ ....... ............... .... . ..... ...... .. .... ..... ..........8a ..................... ......P. ....... ............... .... . ..... ...... .. .... ..... ..........

bA2m4a ....................T ........ ....... ............... .... . ..... ... . .. .....N ..... ..........bA2m5a ..................... ........ ....... ............... .... . ..... ...... .. .... ..... ...M......

12a .......L............. ........ ....... ............... .... . ..... ..T... .. .... ..... ..........

pA-DsbA2 KM K AA N AKGK YR KFAK KQY GKQYTTL KPVAGAPQVL FFSFF PH YQ VLHIS NVKKKLP GVKMTKYHVNFMG LG LTQAWAVAMALG DKbA2m1 .. .N.. . .... .. .... ... ....... .......... ..... .Y .. ..... ....... ............. .. ............ ..

2 .. .... . .... .. .... ... ....... .......... ..... .Y .. ..... ....... ............. .. ............ ..LibA2m3 .. .N.. . .... .. .... ... ....... .......... ..... .Y .. ..... ....... ............. .. ............ ..bA2m7 .. . .. . ....

C C. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .

W

.

LiLiLi

... .... ... ....... .......... ..... .Y .. ..... ....... ............. .. ............ ..bA2m19 .. .... . .... .. .... ... ....... .......... ..... .Y .. ..... ....... ............. .. ..H......... ..

18 .. . .. . .... .. .... ... ....... .......... ..... .Y .. ..... ....... ............. .. ............ ..LibA2m20 .. . .. . .... .. .... ... ....... .......... ..... .Y .. ..... ....... ............. .. ............ ..bA2m1a .. . .. . .... .. ....

. . .. . . ... . . .. .. ... . . . . . .. . . ... . . .. .. ... N . . . . . .. . . ... . . .. .. ... . . . . . . .. . . ... . . .. .. ... N . . . .

LiLiLi

Li

Fk

.... ....... .......... ..... .Y .. ..... ....... ............. .. ............ ..bA2m3a .. . .. . .... .. .... ... ....... .......... ..... .Y .. ..... ....... ............. .. ............ ..bA2m9a .. . .. . .... .. .... ... ....... .......... ..... .Y .. ..... ...I... ............. .. ............ ..

LibA2m7a .. . .. . .... .. .... ... ....... .......... ..... .Y .. ..... ....... ............. .. ............ ..bA2m6a .. . .. . .... .. .... ... ....... .....V.... ..... .Y .. ..... ....... ............. .. ............ ..bA2m8a .. . .. . .... .. .... ... ....... .......... ..... .. .. ..... ....... ............. .. ............ ..bA2m4a .. . .. . .... .. .... ... ....... .......... ..... .. .. ..... ....... ............. .. ............ ..

LibA2m5a .. . .. . .... .. .... ... .......

.. . . ... . . .. .. ... . . . . . . .. . . ... . . .. .. ... . . . . . . .. . . ... . . .. .. ... . . . . . . .. . . ... . . .. .. ... . . . . . . .. . . ... . . .. .. ... . . . . . . .. . . ... . . .. .. ... . . . . . . .N . . ... . . .. .. ... . . . . . . ..

LiLiLi

LiLiLi

LibA2mLiLi

LibA2mLiLi

....V...... ..... .. .. ..... ....... ............. .. ............ ..bA2m12a .. . .. . .... .. .... ... ....... .......... L.... .. .. ..... ....... .....N....... .. ............ ..

pA-DsbA2 VTVPLF GVQKTQTIRSAS IR VFINAGIKG Y AA NSFVVKSLVAQQ KAAA VQLRGVPAMFVNGKYQLNPQG TSN VFVQQY DTVKYL KK* LibA2m1 ...... ............ .. ......... . ..C............ .... ..................... ... ...... ...... ..HHHHHH*bA2m2 ...... ............ .. ......... . ............... .... ..................... ... ...... ...... ..HHHHHH*bA2m3 ...... ............ .. ...T..... . ............... .... ..................... ... ...... ...... ..HHHHHH*bA2m7 ...... ............ .. ......... . ............... .... ..................... ... ...... ...... ..HHHHHH*

LibA2m19 ...... ............ .. ......... . ............... .... ..................... ... ...... ...... ..HHHHHH*bA2m18 ...... ............ .. ......... . ............... .... ..................... ... ...... ...... ..HHHHHH*bA2m20 ...... ............ .. ..... ... . ............T.. .... ..................... ... ...... ...... ..HHHHHH*bA2m1a ...... ............ .. .....C... . ............... .... ........... ......... ... ...... ...... ..HHHHHH*

LibA2m3a ...... ............ .. ......... . ............... .... ...........D......... ... ..I... ...... ..HHHHHH*9a ...... ............ .. ......... . .................... ..................... ... ...... ...... ..HHHHHH*

bA2m7a ...... ............ .. ......... . ............... .... ..................... ... ...... ...... ..HHHHHH*bA2m6a ...... ............ .. ......... . .....I......... .... ..................... ... ...R.. ...... ..HHHHHH*

LibA2m8a ...... ............ .. ......... . ............... .... ..................... ... ...... ...... ..HHHHHH*4a .....L ............ ..

. ... . . .. .. ... . . . . . . .. . . ... . . .. .. ..

E D D EE D E D MD MD A SE. . . .. . . . .. .. . ... . . .. . . . .. .. . ... . . .. . . . .. .. . ... . . .. . . . .. .. . ... . . .. . . . .. .. . ... . . .. . . . .. .. . ... . . .. . . . .. .. . ... . . .. . . . . .. .. . ... . . .. . . . .. .. . ... . . .. . . .. .. . ... . . .. . K . .. .. . ... . . .. . . . .. T. . ... . . .. . . . .. .. . ... . .......... . ...S........... .... ..................... ... ...... ...... ..HHHHHH*

bA2m5a ...... ...........P .. ......... . T.............. .... ..................... A.. ...... ...... ..HHHHHH*bA2m12a ...... ............ .. ......... . ............... .... ..................... ... ...... ...... ..HHHHHH*

.. . . . .. .. . ... . . .. . . . .. T. . ... . . .. . . . .. .. . ..

Figure 2.8. Amino acid sequences of the FkpA-DsbA2 derived mutants isolated by applying directed evolution and screened for increased CuCl2 resistance.

60

Page 76: Copyright by Silvia A. Arredondo 2009

One mutant, enzyme FkpA-DsbA2m7, which contained only the H145Y

mutation, and FkpA-DsbA2m18, which had two additional mutations, D20G and D95N,

were examined further. The overexpression of both mutant enzymes in the periplasm of

MB706 gave rise to cells exhibiting copper resistance to a degree nearly identical to that

of DsbC. In fact, FkpA-DsbA2m18 overexpressing strain exhibited larger colonies at 11

mM CuCl2 (Table 2.2). However, the two mutant enzymes did not confer any further

increase in the yield of vtPA or phytase (AppA), relative to the parental FkpA-DsbA2

protein (Fig 2.9).

D20G

D95N

H145Y

FkpA-DsbA2m18

A

0

5

10

15

20

25

30

35

FkpA-DsbA2

FkpA-DsbA2m7

FkpA-DsbA2m18

DsbC --

Rel

ativ

e vt

PA A

ctiv

ity

SF100 dsbC

B

61

Page 77: Copyright by Silvia A. Arredondo 2009

0

20

40

60

80

100

120

140

FkpA-DsbA2

FkpA-DsbA2m7

FkpA-DsbA2m18

DsbCpBAD33

Rel

ativ

e A

ppA

Act

ivity

DHB4 dsbC

C Figure 2.9. FkpA-DsbA2 mutants obtained following mutagenesis and selection for CuCl resistance.2 (A) Molecular model of the FkpA-DsbA2m18 mutant. (B) Yield of active vtPA in E. coli PB351 (SF100 ΔdsbC). (C) AppA activity assayed as described in Materials and Methods. The AppA activity was determined by measuring A410. One unit was defined as 1,000 x A410 per min/ml (Berkmen et al., 2005).

One difference between FkpA-DsbA and the native disulfide isomerase DsbC is

that the latter is resistant to oxidation by the DsbA-DsbB system in vivo. We wondered

whether under selective pressure it would be possible to isolate FkpA-DsbA2 mutants

that are as resistant to oxidation as DsbC. For this purpose, 180 mutant colonies that grew

in 10 mM CuCl2 BHI plates were picked at random and screened for inability to catalyze

protein oxidation. The latter was evaluated using cell motility as readout. MB706 cells

lack dsbA and are not motile because the oxidation of the essential flagellar component

FlgI is impaired. As shown in Fig 2.7, FkpA-DsbA2 exhibits significant oxidase activity

and therefore, as expected, MB706 cells expressing this enzyme are motile. Out of the

180 CuCl2-resistant colonies tested, several clones were found to be non-motile and then

62

Page 78: Copyright by Silvia A. Arredondo 2009

screened further for their ability to form active alkaline phosphatase. One clone, FkpA-

DsbA2m33, which displayed no alkaline phosphatase activity, was found to contain four

different mutations: M35T, A111V, C146S and H154Y. C146 is the C-terminal cysteine

in the catalytic site (CPHC) of the DsbA domain of FkpA-DsbA2m33. DsbA lacking the

C-terminal cysteine cannot function as a periplasmic oxidant since it becomes trapped in

a mixed disulfide with DsbB and cannot be recycled to its oxidized active form (Guilhot

et al., 1995; Kishigami et al., 1995b). FkpA-DsbA2 containing only the mutation C146A

exhibited the same phenotype as FkpA-DsbA2m33 (in terms of alkaline phosphatase

activity, motility and active vtPA yield, Fig 2.10). This allowed us to conclude that

FkpA-DsbA2m33 is unable to restore motility or alkaline phosphatase activity because

the substitution of C146 impairs the oxidase activity.

0

5

10

15

20

25

FkpA-DsbA2

FkpA-DsbA2C146A

FkpA-DsbA2m33

FkpA-DsbA2m33S146C

--

Rel

ativ

e vt

PA A

ctiv

ity SF100 dsbC

Figure 2.10. The effect of the C146 mutation in isomerase activity in vivo. Yield of active vtPA in dsbC cells transformed with pTrcStIIvtPA and pBAD derivatives encoding the respective FkpA-DsbA2 variants.

63

Page 79: Copyright by Silvia A. Arredondo 2009

2.3.4 Biochemical Studies

FkpA-DsbA2-5, FkpA-DsbA2m7, FkpA-DsbA2m18 and DsbC as a control were

purified by IMAC. CD spectroscopy analyses (Fig 2.11) revealed that FkpA-DsbA2 had

an α-helical content of 46%, which compares favorably with the calculated value of 42%

deduced from the crystal structure of DsbA and of the FkpA dimerization domain. In

addition, the CD spectrum of FkpA-DsbA2m18 was identical to that of the parental

protein, FkpA-DsbA2 (Fig 2.11).

-1.2

-1

-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

200 210 220 230 240 250

Wavelength (nm)

Ellip

ticit

y

FkpA-DsbA2

FkpA-DsbA2m18

Figure 2.11. CD spectra of purified FkpA-DsbA2 and FkpA-DsbA2m18. CD spectroscopy of purified FkpA-DsbA2 and FkpA-DsbA2m18 were performed using an Aviv model 202SF circular dichroism spectrometer measuring the CD signal between 250 and 200 nm. Secondary structure elements content was calculated analyzing the CD spectra with the software CDPro (Data provided by Dr. Laura Segatori).

64

Page 80: Copyright by Silvia A. Arredondo 2009

The insulin reduction activity (Holmgren, 1979b) of the FkpA-DsbA chimeras

was only 7-10% of that of DsbC (Table 2.3). For comparison, DsbA exhibits about 10%

of the activity of DsbC in this assay (Segatori et al., 2004). All the chimeras, including

FkpA-DsbA2m7 and FkpA-DsbA2m18 (data not shown), exhibited low disulfide

isomerase activity in the refolding of reduced RNAse A. Finally, chaperone activity was

evaluated based on the protection of citrate synthase from thermal inactivation. The

artificial disulfide isomerases exhibited robust chaperone activity, with inactivation half

times (t1/2) ranging from 64 to 121% of that displayed by DsbC, and presumably

conferred by the FkpA domain, (Table 2.3).

Table 2.3. In vitro activities of purified enzymes (data provided by Dr. Laura Segatori and Dr. Hiram Gilbert).

RNAse Insulin Citrate SynthaseRefolding*† Reduction* Inactivation

Enzyme μM/min/μM Enz *10-3 ΔA650nm/min2 CS t1/2 (min) μ (min-1)

— — — 0.94 + 0.01 0.51 + 0.01DsbC 0.067+ 0.012 5.81 + 0.21 3.15 + 0.05 0.22 + 0.02

FkpA-DsbA2 0.012 + 0.001 0.45 + 0.11 2.46 + 0.04 0.17 + 0.06FkpA-DsbA3 0.013 + 0.001 0.43 + 0.13 2.01 + 0.05 0.36 + 0.06FkpA-DsbA4 0.014 + 0.003 0.55 + 0.12 3.82 + 0.06 0.45 + 0.22FkpA-DsbA5 0.01 + 0.001 0.38 + 0.16 2.71 + 0.07 0.26 + 0.09

*Data are expressed as mean + SD (n = 3-6)† The activities were determined from a plot of isomerization velocity against enzyme concentrations. t½ = inactivation half time; μ = inactivation rate.

65

Page 81: Copyright by Silvia A. Arredondo 2009

2.4 Discussion

A comparison of the bacterial disulfide isomerases DsbC/G and the yeast PDI

reveals a longer separation of the active sites in the bacterial enzymes (26, 38 and 60 Å

for yPDI, DsbC and DsbG respectively), a difference in the surface area of the

hydrophobic cleft, and, finally, the presence of one or more α-helices within the bacterial

thioredoxin motif catalytic domains (Gruber et al., 2006). Unlike the bacterial enzymes,

where the two catalytic sites are symmetric and catalytically identical, the a and a’

domains of PDI are asymmetric and functionally distinct (Tian et al., 2006).

In contrast to the disulfide isomerases, monomeric thioredoxin motif enzymes,

such as the E. coli thioredoxin, glutaredoxin, DsbA, the a and a’ domains of PDI, or the

catalytic domain of DsbC exhibit either low or no disulfide isomerase activity (Hawkins

et al., 1991; Lundstrom-Ljung and Holmgren, 1995; Ruoppolo et al., 2003). How is it

then that the thioredoxin catalytic domains are able to catalyze disulfide bond

isomerization in the context of PDI or DsbC? Earlier we had shown that fusion to the N-

terminal dimerization domain of DsbC converts DsbA into an efficient disulfide

isomerase. The acquisition of disulfide isomerization activity in DsbA fused to DsbC [aa

1-59] could be a general consequence of: (i) the dimerization of thioredoxin domains,

which is required for maintaining the protein in a reduced state, and could also be

important for catalysis, by virtue of the high effective concentration of active site thiols,

and/or (ii) the proximity of the catalytic domain to a putative peptide binding site.

In order to circumvent the many difficulties of applying classical mutagenesis

techniques, here we used a synthetic approach to delineate the key features that are

66

Page 82: Copyright by Silvia A. Arredondo 2009

required for catalysis of disulfide bond isomerization. We constructed and characterized

in some detail, fusion proteins in which either a thioredoxin mutant having the PDI active

site or DsbA were fused to two different dimerization domains, namely either the α-

helix--GCN4 leucine zipper which merely serves as a dimerization device or the N-

terminal of FkpA; the latter, also forms a peptide binding cleft that is important for

cis/trans proline isomerization in FkpA. The engineered disulfide isomerases displaying

highest in vivo activity, FkpA-DsbAs, were constructed utilizing the dimerization and α-

helical linker domains of FkpA and the thioredoxin catalytic domain of the cysteine

oxidase DsbA. In a manner analogous to DsbC, the ability of the chimeric enzymes to

catalyze disulfide bond isomerization was dependent on recycling by the membrane

electron transfer enzyme DsbD. The amino acid at the fusion junction affected the in vivo

isomerase activity as well as the sensitivity to oxidation by DsbB and the ability to serve

as periplasmic protein oxidants. The most successful design, FkpA-DsbA2, restored the

yield of active vtPA to 80% and of the native phytase AppA to 100% of the amount

observed with DsbC expressed at the same level. FkpA-DsbA2 also conferred increased

resistance to CuCl2 to E. coli dsbA dsbC cells and mutants that allowed growth in the

presence of high CuCl2 concentrations were isolated.

Collectively, these findings provide several important and unanticipated insights

on the mechanism and evolution of disulfide isomerases. First, forcing the dimerization

of thioredoxin domains gives rise to enzymes that catalyze the folding of multidisulfide

proteins in vivo, an activity not displayed by monomeric thioredoxin. The introduction of

an α-helical linker between the GCN4 and TrxA”PDI-like” domains is important for this

67

Page 83: Copyright by Silvia A. Arredondo 2009

activity. Specifically, expression of a fusion of TrxA“PDI-like” to the α-helical linker of

DsbC and dimerization coiled-coil domain of GCN4 resulted in a vtPA yield 5-times

higher than that of control cells, or 20% of the level observed with DsbC. In contrast,

exported thioredoxin without a dimerization domain is maintained in an oxidized form in

the periplasm, and therefore does not support disulfide bond isomerization (Debarbieux

and Beckwith, 1998; Jonda et al., 1999). Thus, the presence of two thioredoxin domains,

properly spaced via α-helical linkers, seems to be a key prerequisite for the ability to

catalyze disulfide bond isomerization.

Second, fusing the thioredoxin catalytic domain to an α-helical linker and, in

addition, to a suitable dimerization region, which folds into a DsbC-like “V-shaped”

topology and presents an internal substrate binding surface (unrelated to that of DsbC),

gave rise to fusion enzymes with improved in vivo activity. The catalytic activity of the

FkpA chimeras was dependent on the identity of the thioredoxin domain, with FkpA-

DsbA resulting in more efficient enzymes compared to FkpA-TrxA”PDI-like” chimeras.

A number of oxidoreductases have been observed to contain minor changes in this

structural domain – such as different active site dipeptide sequence, small helical

insertions, or fusion with different structural domains – which have been introduced

during the evolution to fine-tune the functionality of different enzymes. The thioredoxin

motif in the DsbA molecule contains a three helix insertion adjacent to the catalytic site.

The helical insertion creates an exposed hydrophobic site that may play a role in peptide

binding (Chivers et al., 1997; Gruber et al., 2006; Martin, 1995; McCarthy et al., 2000).

The higher in vivo activity displayed by the FkpA-DsbA chimeras may either be due to

68

Page 84: Copyright by Silvia A. Arredondo 2009

subtle structural differences in the orientation of the respective catalytic domains or to the

influence of the helical insertion in the thioredoxin fold of DsbA. In addition, the

electrostatic potential of the FkpA and DsbC clefts are different, with the former being

more negatively charged (Saul et al., 2004). It is thus possible that the hydrophobic

surface formed by the helical insertion in DsbA serves to extend the cleft in the FkpA

domain and assist in the binding of proteins requiring disulfide rearrangement.

Accordingly we showed that FkpA-DsbA2 exhibits chaperone activity and therefore is

able to bind protein substrates. Even though the FkpA-DsbA2 chimera could substitute

for all the known in vivo functions of DsbC (folding of heterologous and native

multidisulfide proteins, CuCl2 resistance), in vitro it exhibited much lower activity in the

refolding of reduced RNAse A and in the reduction of insulin (Table 2.3). The lower in

vitro activity of FkpA-DsbA2 in comparison to DsbC likely reflects differences in the

substrate specificity of the two enzymes.

Third, we show that under selective pressure FkpA-DsbA2 could be evolved to

confer increased resistance of dsbA dsbC cells to CuCl2 Earlier, Hiniker and Bardwell

had shown that resistance to CuCl2 requires disulfide bond isomerization in the periplasm

(Hiniker et al., 2005). A single amino acid His->Tyr substitution in the CPHC active site

of the DsbA domain of FkpA-DsbA2 conferred increased cellular fitness to CuCl2

challenge. As a result, the active site was changed to CPYC which is identical to that of

the second E. coli disulfide isomerase, DsbG. In addition the CPYC motif is found in

certain other DsbC homologues (e.g. P. aeruginosa,) and also in the E. coli glutaredoxins

1 and 3 (Foloppe and Nilsson, 2004). It is noteworthy that the active site of DsbC also

69

Page 85: Copyright by Silvia A. Arredondo 2009

contains an aromatic amino acid next to the C-terminal Cys residue (DsbC active site:

CGYC). The sequence of amino acids in the dipeptide between the two Cys in the active

site plays a key role in determining the redox potential of the enzyme (Grauschopf et al.,

1995; Huber-Wunderlich and Glockshuber, 1998). While the dipeptide sequences in

DsbA and DsbC are different, their redox potentials with glutathione are quite similar

(Eo,DsbA= -120 mV, Eo,DsbC= -130 mV). The consensus mutation in the FkpA-DsbA2 that

led to increased CuCl2 resistance changed His145 (corresponding to His32 in DsbA) with

Tyr. Huber-Wunderlich et al. showed that the His32Tyr substitution in DsbA (CPYC)

decreases its redox potential by 35 mV to E0=-157 mV and leads to a thermodynamic

stabilization of the oxidized state over the reduced state, which is opposite to wt DsbA

(Huber-Wunderlich and Glockshuber, 1998). While changes in the thermodynamic

properties of the DsbA domain may be responsible for its improved ability to confer

CuCl2 resistance, it is also possible that the His32Tyr mutation was selected for kinetic

reasons, for example improved interaction with – and reduction by – DsbD. Regardless of

the specific mechanism, it is noteworthy that under selective pressure the active site of

the artificial enzyme seems to converge to the sequence found in the native disulfide

isomerase, DsbC.

Fourth, the FkpA-DsbA chimeras catalyze both isomerization, and protein

oxidation, and could partially complement dsbA mutant strains in a DsbB-dependent

fashion. Analysis of the data in Figures 2.6 and 2.7 reveals that the oxidase activity and

isomerase activity in a dsbA background are correlated and depend on the amino acid of

the fusion junction between the α-helical linker from FkpA and DsbA. Similarly,

70

Page 86: Copyright by Silvia A. Arredondo 2009

deletions of one or more amino acids in the helical region of DsbC from various species

disrupts the barrier that normally prevents its oxidation by DsbB, and allows it to catalyze

the formation of disulfide bonds in addition to isomerization (Segatori et al., 2006). An

attempt to select for FkpA-DsbA variants that are protected from oxidation by

DsbB/DsbA, and therefore more closely resemble DsbC, led only to the isolation of

mutant proteins lacking the C-terminal catalytic cysteine. The absence of the C-terminal

Cys in FkpA-DsbA2 eliminates oxidase activity but also reduces disulfide isomerization.

Maintaining high isomerization activity and resistance to oxidation seems to be a

complex trait that hinges both on the orientation of the catalytic sites (Segatori et al.,

2006) and other subtle features that prevent the interaction of the enzyme with DsbB.

In summary, using a protein engineering approach, we have been able to assemble

functional disulfide isomerases via the fusion of domains derived from proteins of

unrelated function. We show that the dimerization of thioredoxin domains alone is not

sufficient to grant isomerase activity. However, the insertion of a linker between a coiled

coil dimerization domain and the thioredoxin catalytic domain gave rise to proteins with

basal disulfide isomerase activity. Substantially more active enzymes could be obtained

by using a dimerization domain from FkpA which contains a suitably located peptide

cleft and displays an overall topological organization analogous to that of DsbC.

Interestingly, even though the synthetic disulfide isomerase chimeras we report here bear

no sequence similarity with DsbC, when placed under selective pressure for CuCl2

resistance resulted in better fitness through mutations in the C-X-X-C motif within the

active site, indicating that an optimal active site is also critical for high catalytic activity.

71

Page 87: Copyright by Silvia A. Arredondo 2009

Our studies are consistent with the notion that vestigial disulfide isomerases could

have readily originated by gene duplication of a thioredoxin catalytic domain that later

acquired an appropriate substrate binding site, as well as structural features that prevented

wasteful shuttling of electrons between the oxidation and the reduction pathways.

72

Page 88: Copyright by Silvia A. Arredondo 2009

Chapter 3

The Role of Dimerization on the Catalytic Properties of the

Escherichia coli Disulfide Isomerase DsbC

3.1 Introduction

Disulfide bonds are critical for the proper folding and structural stability of many

exocytoplasmic proteins. The Dsb family of thiol:disulfide oxidoreductase enzymes

catalyzes oxidative protein folding in the periplasm of Escherichia coli by means of two

independent pathways (Gleiter and Bardwell, 2008; Ito and Inaba, 2008; Kadokura et al.,

2003). In the DsbA-DsbB oxidation pathway, DsbA, a very strong oxidant, catalyzes the

formation of disulfide bonds on newly translocated proteins (Zapun et al., 1993). The

DsbA disulfide is rapidly recycled by DsbB, a membrane protein that transfers electrons

from DsbA onto quinones (Grauschopf et al., 2003; Guilhot et al., 1995; Kishigami et al.,

1995b). In the DsbC-DsbD isomerization pathway, non-native disulfides are reduced or

rearranged by DsbC. DsbC is maintained in a reduced, catalytically-active state, via the

transfer of electrons from the inner membrane protein DsbD that in turn accepts electrons

from Thioredoxin 1 and, ultimately from NADPH (via thioredoxin reductase) within the

cytoplasm (Rietsch et al., 1996; Rietsch et al., 1997). Large kinetic barriers keep the

oxidation and isomerization pathways isolated, preventing the establishment of a futile

73

Page 89: Copyright by Silvia A. Arredondo 2009

cycle of electron transfer. Accordingly, reactions between enzymes of the two pathways,

for example the oxidation of DsbC by DsbB or the reduction of DsbA by DsbD, are 103

to 107 fold slower than the physiologically relevant DsbA-DsbB and DsbC-DsbD

reactions (Rozhkova et al., 2004). Nonetheless, the kinetic barrier between DsbB and

DsbC can be breached by introducing mutations that result in structural changes in DsbC

(Bader et al., 2001; Segatori et al., 2006).

DsbC is a homodimer with each monomer comprising an N-terminal dimerization

domain and a C-terminal thioredoxin-like catalytic domain fused by an α-helical linker.

The crystal structure of DsbC reveals that the two monomers come together to form a V

shaped protein. The inner surface of the resulting cleft is patched with uncharged and

hydrophobic residues suggesting an important role in the binding of substrate proteins.

The active sites comprising the sequence C98-G100-Y101-C101 in each of the

monomeric subunits are located in the arms of the “V” facing each other (McCarthy et

al., 2000). Isomerization involves an attack onto a substrate disulfide by C98 resulting in

the formation of a mixed disulfide, which then is resolved by either another cysteine from

the substrate or by C101 from DsbC (Collet and Bardwell, 2002; Ritz and Beckwith,

2001). Besides its isomerase activity, DsbC is also known to display chaperone activity

preventing protein aggregation during refolding (Chen et al., 1999). In E. coli, disulfide

bond isomerization is the limiting step in the oxidative folding of many heterologous

proteins that contain multiple cysteines. Overexpression of DsbC has been shown to

enhance the yield of proteins such as human nerve growth factor, human tissue

74

Page 90: Copyright by Silvia A. Arredondo 2009

plasminogen activator (tPA) and immunoglobulins (Kurokawa et al., 2001; Qiu et al.,

1998; Zhang et al., 2002).

DsbC is topologically analogous to the eukaryotic protein disulfide isomerase

(PDI). The structural similarities between the two enzymes may have resulted from

convergent evolution by thioredoxin-like domain replication in the case of PDI and

domain recruitment in DsbC (Gruber et al., 2006; Tian et al., 2006). PDI comprises two

thioredoxin-like catalytic domains (a and a’) separated by two non-catalytic domains (b

and b’), in addition to a c domain (Wilkinson and Gilbert, 2004). In PDI, the catalytic

domains are different and functionally nonequivalent (Lyles and Gilbert, 1994). Substrate

binding is mediated primarily by the b’ domain; the two catalytic domains a and a’ can

catalyze oxidation of small model peptides indicating that they must also have low

substrate binding affinity (Hatahet and Ruddock, 2007).

The DsbC monomer is essentially devoid of RNAse A isomerase activity (Ke et

al., 2006). Sun et al. (2000) reported that DsbC with one catalytic site impaired by

carboxymethylation is also inactive but, in separate studies Zapun et al. (1995) did not

detect cooperativity between the two catalytic sites indicating that they function

independently of each other. Moreover, unlike PDI, the significance of the putative

peptide binding cleft of DsbC on disulfide isomerization has not been ascertained. We

recently showed that while DsbA, or TrxA with a PDI active site dipeptide (CGHC)

display very little isomerase activity in vitro and in vivo (Jonda et al., 1999; Lundstrom et

al., 1992; Segatori et al., 2004), upon fusion to a dimerization region that provides a

putative substrate binding surface (the E. coli peptidyl proline isomerase FkpA) they

75

Page 91: Copyright by Silvia A. Arredondo 2009

acquire the ability to assist the folding of periplasmically expressed multi-disulfide

heterologous proteins (Arredondo et al., 2008).

In the present work we engineered heterodimer-like covalently linked DsbC

derivatives in which one of the catalytic sites has been inactivated, or one of the catalytic

domains has been entirely removed while maintaining the intact peptide binding cleft

(which is normally formed by association of the N-terminal domains of the two

monomers). We show that a DsbC forced-monomer with one functional active site, or

with one thioredoxin domain only, both display significant isomerization activity.

Interestingly, the latter variant is partially reduced in vivo indicating that the presence of

both thioredoxin domains is important for the avoidance of protein oxidation by DsbB.

76

Page 92: Copyright by Silvia A. Arredondo 2009

3.2 Materials and Methods

3.2.1 Strains and Plasmids

The bacterial strains and plasmids used in this study are listed in table 3.1B

published as supplemental material. Genes encoding DsbC fusions containing a

(GGGS)3GSA or (GGGS)7SA linker peptide were assembled by amplifying the

individual monomer genes via PCR using the primers shown in table 3.1A, digesting with

XbaI-PvuII or PvuII-HindIII, and ligating into pBAD30. The complete fusion genes were

then digested with XbaI and HindIII, and cloned either into pBAD33 (Guzman et al.,

1995), for in vivo testing, or into pET28(a) for protein purification purposes.

Table 3.1. (A) Primers and (B) bacterial strains and plasmids utilized in this study.

APrimers SequenceXbaIDsbCss.f GAGCTCGAATTCTCTAGATTAAAGAGGAGAAAGGTACCCATGATGAAGAAAGGTTTTATDsbCHisHindIII.r TTTTTAAGCTTTTAGTGGTGGTGGTGGTGGTGTTTACCGCTGGTCATTTTTTGdimstopHindIII.r TTTTTAAGCTTTTAGTGGTGGTGGTGGTGGTGATTGGTGACATTGACCGGAGCCG DsbC-AGYC.f GATATTACCGCGGGTTACTGCCACAAACTGDsbC-AGYC.r CAGTTTGTGGCAGTAACCCGCGGTAATATCDsbC-AGYA.f CCGTGTTTACTGATATTACCGCGGGTTACGCGCACAAACTGCATGAGCAAATGGCDsbC-AGYA.r GCCATTTGCTCATGCAGTTTGTGCGCGTAACCCGCGGTAATATCAGTAAACACGGDsbC-CGYA.f CCGTGTTTACTGATATTACCTGTGGTTACGCGCACAAACTGCATGAGCAAATGGCDsbC-CGYA.r GCCATTTGCTCATGCAGTTTGTGCGCGTAACCACAGGTAATATCAGTAAACACGGDsbC-H45D.r GGCCCCTGAATGATATCTTTACCATCATCDsbC-D53H.r GTGCCACTAACATGATACATTGGCCCLinkDsbCa.f GGCGGTGGCAGCGGTGGAGGCTCCGGCGGAGGTAGCGGTTCAGCTGATGACGCGGCALinkDsbCb.f GGTAGCGGTTCAGCTGATGACGCGGCAATTCAACAAACDsbCLink.r GCCGGAGCCTCCACCGCTGCCACCGCCTTTACCGCTGGTCATTTTTTGGTGTTCGTCLinker30a.r CCGCTCCCGCCACCGGATCCGCCTCCAGAACCTCCGCCGGAGCCTCCACCGCTGCCACCLinker30b.r CGCGTCATCAGCTGAACCGCTACCGCCAGAGCCACCTCCGCTCCCGCCACCGGATCCGC

77

Page 93: Copyright by Silvia A. Arredondo 2009

B

Strains Relevant Genotype SourceSF100 F- ΔlacX74 galE galK thi rpsL (strA ) ΔphoA (PvuII ) Δ(ompT- entF ) Ref (1MC1000 araD139

)Δ(araA-leu )7679 Δ(codB-lac )X74 galE15 galK16 rpsL150 relA1 thi Lab Collection

BL21(DE3)pLysS F- ompT hsdS B(rB- mB

-) gal dcm (DE3) pLysS (CamR) NovagenPB351 SF100 Δ degP Δ dsbC Ref (2)PB403 SF100 Δ degP dsbA::kan Lab CollectionLM106 MC1000 dsbA::kan5 Lab CollectionLM102 MC1000 dsbB::kan5 Ref (3)

Plasmids Relevant Genotype SourcepET-28(a) T7 expression vector, C-terminal 6x histidine tag NovagenpTrcStIIvtPA tPA (Δ6-175) with StII leader in pTrc99A Ref (4)pBADdsbC dsbC from Escherichia coli in pBAD33 Lab CollectionpBAD-mDsbC-mDsbC two dsbC genes fused via (GGGS)3GSA linker in pBAD33 This workpBAD-DsbC[C98A] dsbC C98A in pBAD33 This workpBAD-mDsbC-mDsbC[C98A] dsbC fused to dsbC C98A via (GGGS)3GSA linker in pBAD33 This workpBAD-DsbC[C98A/C101A] dsbC C98A C101A in pBAD33 This workpBAD-mDsbC-mDsbC[C98A/C101A] dsbC fused to dsbC C98A C101A via (GGGS)3GSA linker in pBAD33 This workpBAD-DsbC[C101A] dsbC C101A in pBAD33 This workpBAD-mDsbC-mDsbC[C101A] dsbC fused to dsbC C101A via (GGGS)3GSA linker in pBAD33 This workpBAD-mDsbC-dim dsbC fused to dsbC( Δ65-216) via (GGGS)3GSA linker in pBAD33 This workpBAD-mDsbC[H45D]-dim[D53H] dsbC H45D fused to dsbC D53H (Δ65-216) via (GGGS)7SA linker in pBAD33 This workpET28-DsbC dsbC from Escherichia coli in pET-28(a) Lab CollectionpET28-mDsbC-mDsbC two dsbC genes fused via (GGGS)3GSA linker in pET-28(a) This workpET28-DsbC[C98A] dsbC C98A in pET-28(a) This workpET28-mDsbC-mDsbC[C98A] dsbC fused to dsbC C98A via (GGGS)3GSA linker in pET-28(a) This workpET28-DsbC[C98A/C101A] dsbC C98A C101A in pET-28(a) This workpET28-mDsbC-mDsbC[C98A/C101A] dsbC fused to dsbC C98A C101A via (GGGS)3GSA linker in pET-28(a) This workpET28-DsbC[C101A] dsbC C101A in pET-28(a) This workpET28-mDsbC-mDsbC[C101A] dsbC fused to dsbC C101A via (GGGS)3GSA linker in pET-28(a) This workpET28-mDsbC-dim dsbC fused to dsbC( Δ65-216) via (GGGS)3GSA linker in pET-28(a) This workpET28-mDsbC[H45D]-dim[D53H] dsbC H45D fused to dsbC D53H (Δ65-216) via (GGGS)7SA linker in pET-28(a) This workpET28-mDsbC[H45D]-dim[H45D] dsbC H45D fused to dsbC H45D (Δ65-216) via (GGGS)7SA linker in pET-28(a) This work

1. Baneyx, F., and G. Georgiou 1990. J Bacteriol 172, 491-494 3. 2. Bessette, P. H., Qiu, J., Bardwell, J. C., Swartz, J. R. & Georgiou, G. (2001) J Bacteriol 183, 980-8. 3. Masip, L., Pan, J. L., Haldar, S., Penner-Hahn, J. E., DeLisa, M. P., Georgiou, G., Bardwell, J. C. & Collet, J. F. (2004) Science

303, 1185-9. 4. Bessette, P. H., Aslund. Beckwith & Georgiou, G. (1999) Proc. Natl. Acad. Sci. U.S.A. 96, 13703-13708

3.2.2 In vivo Disulfide Bond Formation and Isomerization

E. coli PB351 (SF100 ΔdegP ΔdsbC) and PB403 (SF100 ΔdegP dsbA::kan) were

co-transformed with pBAD33 plasmids encoding the DsbC or mDsbC-mDsbC variants,

and with pTrcStIIvtPA, a derivative of pTrc99A encoding the tPA (Δ6-175) gene fused to

the StII signal peptide (Bessette et al., 1999a). PB351 cells were grown in LB medium

and PB403 cells in minimal medium (M9 salts, 0.1% casein amino acids, 2mM MgSO4, 5

78

Page 94: Copyright by Silvia A. Arredondo 2009

μg/ml thiamine, 0.4% glycerol with 50 μg/ml of kanamycin) supplemented with 25

μg/ml of chloramphenicol and 100 μg/ml of ampicillin at 37 ºC. Induction of protein

synthesis, cell harvesting and assays for tPA activity were performed as described earlier

(Segatori et al., 2004). For the evaluation of in vivo oxidase activity of DsbC and its

variants, E. coli LM106 (MC1000 dsbA::kan5) and LM102 (MC1000 dsbB::kan5) were

transformed with the appropriate pBAD33 derivatives, grown in LB medium with 50

μg/ml of kanamycin and 25 μg/ml of chloramphenicol, and subcultured 1:50 in low

phosphate minimal medium (MOPS salts, 0.2% glycerol, 0.2% casein amino acids, and

0.5 μg/ml thiamine, with 50 μg/ml of kanamycin and 25 μg/ml of chloramphenicol).

When cultures reached an A600 of 0.4-0.5, arabinose was added to a final concentration of

0.2% w/v. After two hours of growth, 30 μl of cells were mixed with 20 μl of buffer (0.4

M iodoacetamide and lysis buffer (B-PERTM, Pierce) in a 1:2 ratio) in a 96 well plate and

incubated at room temperature for 30 minutes. To measure alkaline phosphatase activity,

200 μl of 0.2 M Tris pH 8 and 0.25 mg/ml of p-Nitrophenyl phosphate disodium

hexahydrate (pNPP) (Sigma, St.Louis, MO) were added and the change in absorbance at

A405 was recorded while incubating at room temperature for 2 hours.

3.2.3 In Vivo Redox State

Overnight cultures of E. coli PB351 (SF100 ΔdegP ΔdsbC) harboring the

appropriate pBAD33 derivatives were diluted 1:100 in LB medium containing 25 μg/ml

of chloramphenicol and grown to A600 ~0.6-0.7, then expression was induced by adding

arabinose to a final concentration of 0.0002%. 2 hours after induction, a culture aliquot

79

Page 95: Copyright by Silvia A. Arredondo 2009

with an A600 equivalent to 1.0 was collected and mixed with an equal volume of 20%

trichloroacetic acid (TCA) to simultaneously quench disulfide exchange and precipitate

the proteins in the sample (Kishigami et al., 1995a). 15mM 4-acetamido-4'-

maleimidylstilbene-2,2'-disulfonic acid, disodium salt (AMS) was added and after a 45

minute incubation at room temperature, proteins were pelleted and then solubilized in 75

μl of Laemmli 1X buffer without reducing agents (Joly and Swartz, 1997). A reduced

protein standard was prepared by first incubating 0.02 mg of purified protein with

100mM DTT for 1 hr followed by filtration through Sephadex G-25 for DTT removal

and by TCA precipitation, AMS treatment, and solubilization, as above. Oxidized

purified protein standards were generated by the same procedure except that incubation

with DTT and AMS were omitted. The reduced and oxidized protein bands were resolved

by SDS-PAGE electrophoresis on a 12% gel and protein bands detected by Western

blotting with an anti-histidine antibody (Sigma, Israel) at 1:20,000 dilution.

3.2.4 Expression, Purification, and Biochemical Assays

For the purification of the DsbC and mDsbC-mDsbC proteins, pET28(a) plasmids

encoding the appropriate genes were transformed into E. coli BL21(DE3). Protein

expression and purification was performed as previously described (Segatori et al., 2004).

All proteins used in this study were more than 95% pure as judged by Coomassie stained

SDS-PAGE electrophoresis. The kinetics of insulin reduction and the renaturation of

reduced and denatured RNase A, were determined according to published procedures

(Holmgren, 1979b; Lyles and Gilbert, 1991; Martinez-Galisteo et al., 1993).

80

Page 96: Copyright by Silvia A. Arredondo 2009

3.3 Results

3.3.1 Construction of a Linked Monomeric DsbC

A covalently linked DsbC was constructed by creating a gene fusion consisting of

a dsbC gene in frame with a sequence encoding a 15aa Gly-Ser linker and a second dsbC

gene. Gly-rich linkers have been used to join the two subunits of homodimeric proteins

into a single polypeptide without disrupting activity (Bizub et al., 1991; Liang et al.,

1993; Predki and Regan, 1995; Robinson and Sauer, 1996; Robinson and Sauer, 1998). In

the DsbC crystal structure, the distance between the C-terminus of one subunit and the N-

terminus of the second subunit is approximately 36 Å (McCarthy et al., 2000). To avoid

introducing strain in the DsbC-DsbC fusion we employed a (Gly3Ser)3GlySerAla linker

estimated to span 57 Å in its extended conformation based on the theoretical 3.8 Å per

residue planar length. The resulting fusion protein was designated mDsbC-mDsbC (Fig

3.1).

81

Page 97: Copyright by Silvia A. Arredondo 2009

DsbC

mDsbC-mDsbC

mDsbC-mDsbC[C98A]

mDsbC-mDsbC[C98A/C101A]

mDsbC-mDsbC[C101A]

-CN- CGYC GGGSGGGSGGGSGGGS CGYC

-CN- CGYC CGYC

-CN- CGYC GGGSGGGSGGGSGGGS CGYA

-CN- CGYC GGGSGGGSGGGSGGGS AGYA

-CN- CGYC GGGSGGGSGGGSGGGS AGYC

Figure 3.1. Protein structure of DsbC, and predictive molecular models of mDsbC-mDsbC and the single-active-site covalently linked mutants. Dimerization domains are shown in red, thioredoxin domains in blue and active catalytic sites in yellow.

The covalently linked DsbC was secreted into the periplasmic space via the DsbC

signal peptide and purified by metal affinity chromatography. Gel filtration FPLC

revealed that mDsbC-mDsbC elutes as a single peak with a retention time identical to that

of DsbC (Fig 3.2).

82

Page 98: Copyright by Silvia A. Arredondo 2009

0

50

100

150

200

250

300

0 5 10 15 20ml

mA

UDsbCmDsbC-mDsbCmDsbC-mDsbC[C98A]mDsbC-mDsbC[C98A/C101A]mDsbC-mDsbC[C101A]

Figure 3.2. Gel filtration FPLC of DsbC and derived proteins. Purified proteins were run on a SuperdexTM 200 column in PBS-10% glycerol buffer.

mDsbC-mDsbC exhibited slightly lower activity in the oxidative refolding of

reduced RNAse A and in the reduction of insulin (84% and 78%, respectively) relative to

wt DsbC; however these differences in activity were not statistically significant. The in

vivo disulfide isomerase activity of mDsbC-mDsbC in the E. coli periplasm was

evaluated by determining the yield of correctly folded and active vtPA secreted via the

StII signal peptide (Segatori et al., 2004). Western blot analysis revealed that mDsbC-

mDsbC and DsbC were expressed at the same level in E. coli (Fig 3.3).

83

Page 99: Copyright by Silvia A. Arredondo 2009

25 kDa

37 kDa

50 kDa

1. mDsbC-mDsbC 2. mDsbC-mDsbC[C98A] 3. mDsbC-mDsbC[C98A/C101A] 4. mDsbC-mDsbC[C101A] 5. mDsbC-dim 6. mDsbC[H45D]-dim[D53H] 7. DsbC

1 2 3 4 5 6 7

Figure 3.3. Expression level of the covalently linked DsbC variants as shown by western blot analysis. Overnight cultures of PB351 (SF100 ΔdegP ΔdsbC) cells grown in LB medium were sub-cultured, grown to OD600 ~ 0.8 and induced with arabinose (0.2% final concentration). After 2 hours of induction, samples were collected, and an equal number of A600 units were then loaded on each lane. Western blot analysis was carried out using a monoclonal antipolyhistidine HRP-conjugated antibody.

The vtPA activity obtained by co-expressing mDsbC-mDsbC was essentially

identical to that observed with wild type DsbC expressed from the same promoter

indicating that the two proteins have identical activity in vivo (Fig 3.4A). Earlier we had

shown that the orientation of the thioredoxin catalytic domains relative to each other in

the DsbC homodimer is critical for avoiding misoxidation by DsbB. Single amino acid

deletions in the α-helical linker that joins the catalytic and the dimerization domains

render the protein susceptible to oxidation by DsbB. In turn, oxidized DsbC accumulates

in the periplasm where it catalyzes the formation of disulfide bonds in substrate proteins

and thus can partially complement dsbA mutants (Segatori et al., 2006). We evaluated

84

Page 100: Copyright by Silvia A. Arredondo 2009

whether the covalent linkage of two DsbC monomers might have perturbed the tertiary

structure of the molecule to make it susceptible to oxidation. Alkaline phosphatase

(PhoA) contains two disulfide bonds that are essential for activity (Kamitani et al., 1992);

in MC1000 dsbA cells grown in low phosphate media, the PhoA level is about 3% of that

observed in the wt cells (Fig 3.4B). While overexpression of wt DsbC does not restore

PhoA activity, the expression of DsbC variants with a perturbed tertiary structure results

in PhoA levels >30% of those in wt cells (Segatori et al., 2006). In contrast, expression

of mDsbC-mDsbC conferred only background levels of PhoA activity. In addition, and

unlike the DsbC variants that are susceptible to oxidation, mDsbC-mDsbC was unable to

catalyze the in vivo folding of vtPA in a dsbA background.

We also determined the in vivo redox state of the covalently linked DsbC.

Briefly, cells were grown in LB media, proteins were precipitated with TCA to prevent

thiol:disulfide rearrangements, and the TCA pellets were treated with 4-acetamido-4'-

maleimidylstilbene-2,2'-disulfonic acid, disodium salt (AMS). AMS reacts with free

thiols increasing the molecular mass by 490 Daltons per thiol. When purified DsbC is

pre-treated with DTT, the thiols of the four disulfide-forming cysteines in each subunit

react with AMS increasing the M.W. by ~2 kDa per monomer. In vivo, however, the

thiolates involved in the structural disulfide of DsbC are not accessible, and the protein is

therefore found in a partially reduced state in which only the catalytic cysteine thiols have

been modified by AMS increasing the M.W. by ~ 1 kDa per monomer (Joly and Swartz,

1997; Rietsch et al., 1997). Comparison of the electrophoretic mobility of the in vitro

generated standards with the in vivo samples of DsbC and mDsbC-mDsbC revealed that

85

Page 101: Copyright by Silvia A. Arredondo 2009

both proteins are found exclusively in the reduced state (Fig 3.5). Collectively, these

results demonstrate that the covalent linkage of two DsbC subunits via the 15mer flexible

linker has no effect on the structure of the enzyme and its ability to resist oxidation by

DsbB (Fig 3.4B).

0

20

40

60

80

100

120

1 2 3 4 5 6 7 8 9

vtPA

Act

ivity

(%)

SF100 dsbC-SF100 dsbA-

1. DsbC 3. DsbC[C98A] 5. DsbC[C98A/C101A] 7. DsbC[C101A] 9. no plasmid

2. mDsbC-mDsbC 4. mDsbC-mDsbC[C98A] 6. mDsbC-mDsbC[C98A/C101A] 8. mDsbC-mDsbC[C101A] 10. MC1000

A

B

0

20

40

60

80

100

1 2 3 4 5 6 7 8 9 10Alk

alin

e Ph

osph

atas

e A

ctiv

ity (%

)

MC1000 dsbA-MC1000 dsbB-MC1000

Figure 3.4. In vivo activity of single-active-site mutants. (A) Yield of active vtPA in dsbC (grey bars) or dsbA (black bars) cells. E. coli PB351 and PB403 (respectively) were transformed with pTrcStIIvtPA and pBAD33 encoding the respective DsbC derivatives. Protein synthesis was induced as described above and the yield of active vtPA at 3 h after induction was determined. Relative activities were obtained by dividing the ΔA405 of each strain by the ΔA405 of the strain expressing wild type DsbC. (B) PhoA activity in E. coli MC1000 dsbA (white bars) and MC1000 dsbB (black bars). The alkaline phosphatase activity in the parental isogenic strain MC1000 is shown by the grey bar.

86

Page 102: Copyright by Silvia A. Arredondo 2009

OX RE

DsbC mDsbC-mDsbC mDsbC-mDsbC[C98A/C101A]

OX RE OX RE

mDsbC-dim mDsbC[H45D]-dim[D53H]

OX RE OX RE

Figure 3.5. In vivo redox state of DsbC and covalently linked proteins. PB351 cell samples were precipitated with trichloroacetic acid (TCA). Derivatization of free thiols was performed using 15mM AMS. The reduced and oxidized standards were generated by the same procedure using 0.02 mg of purified protein. For the reduced standard, the protein was previously incubated with 100mM DTT. For the oxidized standard, addition of AMS was omitted. Following SDS-PAGE in a 12% gel, the proteins were detected by western blot using antihistidine HRP-conjugated antibody.

3.3.2 Effect of Alanine Substitutions in one Catalytic Active Site

In earlier studies, Sun et al. (2000) reported that an in vitro prepared heterodimer

of DsbC consisting of a wt subunit and a carboxymethylated subunit, is devoid of

isomerase activity suggesting that both active sites in the DsbC homodimer are required

for the catalysis of disulfide bond isomerization (Gleiter and Bardwell, 2008; Ritz and

Beckwith, 2001; Sun and Wang, 2000). We employed covalently linked DsbC fusions to

examine this hypothesis. The Cys residues in one of the two active sites of mDsbC-

mDsbC were substituted by Ala to give rise to mDsbC-mDsbC[C98A], mDsbC-

mDsbC[C98A/C101A] and mDsbC-mDsbC[C101A] (Fig 3.1). All the mDsbC-mDsbC

mutants eluted as single peaks with a retention time corresponding to an apparent

87

Page 103: Copyright by Silvia A. Arredondo 2009

molecular weight of 39 kDa that was identical to those of wild type DsbC and catalytic-

site mutants (Fig 3.2 and unpublished data). Substitution of one of the two nucleophilic

active site Cys in the linked DsbC monomer (mDsbC-mDsbC[C98A]) resulted in only a

modest reduction (~19%) in the rate of refolding of reduced RNAse A from 64 + 8 x 10-3

to 52 + 8 x 10-3 μM/min/μM enzyme. However, a more significant decrease (61%) was

observed in the insulin reduction activity (Table 3.2). Mutation of both the C98 and C101

residues of one of the two active sites in mDsbC-mDsbC resulted in a slight further

reduction in the isomerase activity relative to mDsbC-mDsbC[C98A] but had no

additional effect on the disulfide reductase rate. For comparison, we also purified and

characterized DsbC[C98A] and DsbC[C98A/C101A]. As expected, the substitution of

C98 or both Cys in the DsbC homodimer completely abolished the isomerase and the

insulin reductase activities ((Liu and Wang, 2001); Table 3.2).

Table 3.2. In vitro activity of the linked proteins.

Enzyme RNAse Refolding Insulin ReductionμM/min/μM Enzyme % Activity *10-3 ΔA650/min2 % Activity

DsbC 0.076 + 0.007 100 11.1 + 4.8 100

mDsbC-mDsbC 0.064 + 0.008 84 8.63 + 3.0 78

DsbC[C98A] 0.004 + 0.001 5 0.04 + 0.01 0

mDsbC-mDsbC[C98A] 0.052 + 0.008 68 3.37 + 1.0 30

DsbC[C98A/C101A] 0.001 + 0.001 1 0.04 + 0.01 0

mDsbC-mDsbC[C98A/C101A] 0.038 + 0.005 50 3.19 + 0.12 29

DsbC[C101A] 0.081 + 0.009 107 12.88 + 2.6 116

mDsbC-mDsbC[C101A] 0.079 + 0.005 104 8.57 + 2.0 77

mDsbC-dim 0.038 + 0.005 50 1.95 + 0.61 18

mDsbC[H45D]-dim[D53H] 0.024 + 0.005 32 1.67 + 0.13 15

mDsbC[H45D]-dim[H45D] 0.009 + 0.002 12 nd nd

88

Page 104: Copyright by Silvia A. Arredondo 2009

Consistent with the observation that a linked DsbC fusion having only one

nucleophilic cysteine exhibits disulfide isomerase activity in vitro, the co-expression of

mDsbC-mDsbC[C98A] or mDsbC-mDsbC[C98A/C101A] resulted in a vtPA yield that

was 80% of the value obtained by expressing wt DsbC or mDsbC-mDsbC (Fig 3.4A). In

contrast, neither DsbC[C98A] nor DsbC[C98A/C101A] was able to catalyze the

formation of active vtPA. Substitution of the resolving cysteine C101A alone decreased

disulfide isomerization in vivo to 20% but had no effect on either isomerase or reductase

activity in vitro (Fig 3.4A and Table 3.2). None of the active-site mutations in mDsbC-

mDsbC compromised its ability to resist oxidation by DsbB, as evidenced by the finding

that the proteins are present solely in the reduced state in vivo and also by their inability

to catalyze the formation of active PhoA in a dsbA strain background (Fig 3.5 and 3.4B).

3.3.3 A Linked DsbC with a Single Active Thioredoxin Domain

The finding that the mutational inactivation of one of the two catalytic sites in

DsbC results only in a mild reduction in the disulfide isomerase activity, suggests that

enzymes in which one of the two thioredoxin subdomains has been deleted in its entirety

might be similarly active. To test this hypothesis, we fused a full-length DsbC monomer

to a DsbC N-terminal dimerization domain (aa 1-64) using the (Gly3Ser)3GlySerAla

linker described above. The resulting molecule was designated mDsbC-dim and is shown

schematically in Fig 3.6.

89

Page 105: Copyright by Silvia A. Arredondo 2009

mDsbC-dim-CN- CGYC GGGSGGGSGGGSGGGS

Figure 3.6. Molecular model of mDsbC-dim. Dimerization domains are shown in red, thioredoxin domain in blue and catalytic site in yellow.

mDsbC-dim expressed at a level comparable to DsbC (Fig 3.3). Size exclusion

FPLC on a Superdex 200™ column revealed that the purified protein elutes as several

peaks. One peak with an apparent molecular weight of 22 kDa corresponded to the M.W.

of a monomeric species and accounted for approximately 60% of the A280. Additionally,

a second peak that eluted as a 58 kDa protein was observed, as was a third peak

corresponding to an even higher M.W. species. The 22 kDa and 58 kDa species were

isolated. Upon further incubation the 58 kDa species gave rise to an elution profile

similar to that of the starting material, whereas the 22 kDa species continued to elute as a

single peak even after incubation at 4 ºC for >1 week (Fig 3.7). These results indicate that

the purified protein exists as a mixture of dimers and H.M.W. multimers that convert

slowly to the monomeric form under the conditions tested.

90

Page 106: Copyright by Silvia A. Arredondo 2009

0

50

100

150

200

0 5 10 15 20

ml

mA

U

Peak APeak B

0

100

200

300

400

500

0 5 10 15 20ml

mA

UmDsbC-dim

A

B

Figure 3.7. Gel filtration FPLC of mDsbC-dim. Purified proteins were run on a SuperdexTM 200 column in PBS-10% glycerol buffer. Fractions corresponding to peaks A and B were collected and further analyzed by re-running them independently under identical conditions.

It has been established that in scFv antibodies, which comprise VH and VL

immunoglobulin domains joined via a (Gly4Ser)n repeat linker, the linker length plays an

important role in dictating the monomer:multimer equilibrium. Linkers with more than 12

residues have been reported to strongly favor the monomeric state (Huston et al., 1988;

91

Page 107: Copyright by Silvia A. Arredondo 2009

Kortt et al., 1994; Todorovska et al., 2001). To examine the role of the linker length on

dimerization we constructed mDsbC-dim proteins having longer (Gly3Ser) repeat linkers

corresponding to 17, 19 or 30 aa. However, the respective proteins also eluted as a

mixture of monomers and H.M.W. species, although the fusion with the 30-amino-acid

linker displayed a higher monomer to multimer ratio (Fig 3.8).

0

200

400

600

800

1000

1200

0 5 10 15 20ml

mDsbC-dimmDsbC-17dimmDsbC-19dimmDsbC-30dimU

mA

Figure 3.8. Gel filtration FPLC of mDsbC-dim derivative proteins containing 17, 19, and 30 amino-acid glycine-serine linkers.

Earlier, Bader et al. (2001) reported that a His45Asp substitution disrupts

dimerization due to the interruption of a salt bridge that forms between His45 and Asp53.

The monomeric DsbC[H45D] is susceptible to oxidation by DsbB and capable of

complementing dsbA null mutants in vivo. We hypothesized that the His45Asp

substitution in mDsbC-dim might prevent intermolecular dimerization. However, this

could also disrupt the intramolecular association of the N-terminal regions and the

92

Page 108: Copyright by Silvia A. Arredondo 2009

formation of the peptide binding cleft. We reasoned that to prevent the oligomerization of

mDsbC-dim it was necessary to weaken, but not completely disrupt, the interaction of the

N-terminal regions; because the two N-terminal halves are covalently joined by the

linker, their high effective concentration should overcome the loss of binding energy due

to such mutations. To test this hypothesis we constructed and characterized a number of

variants containing amino acid substitutions at the DsbC dimerization interface. One of

these variants, mDsbC[H45D]-dim[D53H], was constructed by fusing DsbC[H45D] to

DsbC[Δ65-216, D53H] using (Gly3Ser)7SerAla (Fig 3.9).

0

100

200

300

400

500

600

0 5 10 15 20ml

mA

U

DsbCmDsbC-dimmDsbC[H45D]-dim[D53H]

DsbC mDsbC[H45D]-dim[D53H]

HIS

ASP

A

B

Figure 3.9. mDsbC-dim and mDsbC[H45D]-dim[D53H] (A) The crystal structure of DsbC at the dimerization interface is shown. Residues His45 (green) and Asp53 (purple) interact to establish salt bridges. In mDsbC[H45D]-dim[D53H], His45 is replaced by Asp in the first dimerization domain (purple) and Asp53 is replaced by His in the second dimerization domain (green) in order to disrupt inter-dimerization but promote intra-dimerization. (B) Gel filtration FPLC of mDsbC-dim and mDsbC[H45D]-dim[D53H] as compared to DsbC. Purified proteins were run on a SuperdexTM 200 column in PBS-10% glycerol buffer.

93

Page 109: Copyright by Silvia A. Arredondo 2009

For comparison, we also constructed the mDsbC[H45D]-dim[H45D] variant, in

which both histidines that normally form the His45-Asp53 salt bridges have been

replaced with Asp; electrostatic repulsion of Asp45 and Asp53 is expected to disrupt the

association of the two half dimerization domains increasing its hydrodynamic radius (Fig

3.10). Consistent with this hypothesis we found that, even though the predicted molecular

weights of mDsbC[H45D]-dim[D53H] and mDsbC[H45D]-dim[H45D] are very similar

(32.89 and 32.85 kDa respectively), the two proteins displayed distinctly different

retention by gel filtration FPLC, eluting as single peaks with apparent M.W. of about 27

kDa and 40 kDa respectively (Fig 3.10). These findings indicate that the mDsbC[H45D]-

dim[H45D] has a larger hydrodynamic radius implying that the dimerization domains

have failed to associate, unlike mDsbC[H45D]-dim[D53H]. mDsbC[H45D]-dim[H45D]

catalyzed RNase refolding only at a near-background level. In contrast, the DsbC variant

consisting of a thioredoxin domain and an intact dimerization domain (mDsbC[H45D]-

dim[D53H]) displayed 32% of the activity of wild type DsbC. In comparison with

mDsbC-mDsbC[C98A/C101A], which also has only one active catalytic site,

mDsbC[H45D]-dim[D53H] displayed 63% of the activity of the former, indicating that

the presence of the second thioredoxin domain makes a contribution to the catalytic

activity, perhaps by mediating weak interactions with substrate proteins (Table 3.2).

94

Page 110: Copyright by Silvia A. Arredondo 2009

0

50

100

150

200

0 5 10 15 20ml

mA

U

mDsbC[H45D]-dim[D53H]

mDsbC[H45D]-dim[H45D]

A

B

C

mDsbC[H45D]-dim[D53H]

HIS

ASP

mDsbC[H45D]-dim[H45D]

25 kDa

37 kDa

50 kDa

1. mDsbC[H45D]-dim[D53H] 2. mDsbC[H45D]-dim[H45D] 3. DsbC

1 2 3

Figure 3.10. Comparison of mDsbC(H45D)-dim(D53H) and mDsbC(H45D)-dim(H45D). (A) Molecular models of the dimerization interface. (B) Gel filtration FPLC. (C) SDS-PAGE.

Expression of the single thioredoxin domain variant mDsbC[H45D]-dim[D53H]

assisted the in vivo folding of active vtPA in SF100 degP dsbC cells to 75% of the level

observed with DsbC (Fig 3.11A). Given the absence of one thioredoxin domain, it was

95

Page 111: Copyright by Silvia A. Arredondo 2009

not surprising to find that unlike DsbC, mDsbC[H45D]-dim[D53H] was also able to

catalyze the correct folding of vtPA in a dsbA background at a 40% level (Fig 3.11A), an

indication of dual oxidase and isomerase activity. Disulfide bond formation in vivo was

further confirmed by the partial restoration of PhoA activity; this activity was dependent

upon the presence of DsbB (Fig 3.11B). Consistent with these dual oxidase and

isomerase activities, mDsbC[H45D]-dim[D53H] was found to exist as a mixture of

reduced and oxidized species at a ~75:25 ratio in the periplasm of E. coli (Fig 3.5).

0

20

40

60

80

100

1 2 3 4

Alk

alin

e Ph

osph

atas

e A

ctiv

ity (%

)

MC1000 dsbA-MC1000 dsbB-MC1000

0

20

40

60

80

100

1 2 3

vtPA

Act

ivity

(%) SF100 dsbC-

SF100 dsbA-

A

B

1. mDsbC-dim 3. no plasmid

2. mDsbC[H45D]-dim[D53H] 4. MC1000

Figure 3.11. In vivo activity of single-catalytic-domain mutants. (A) The yield of active vtPA in dsbC (grey bars) or dsbA (black bars) cells was determined as described in Materials and Methods. (B) PhoA activity in E. coli MC1000 dsbA (white bars) and MC1000 dsbB (black bars). The alkaline phosphatase activity in the parental isogenic strain MC1000 is shown by the gray bar.

96

Page 112: Copyright by Silvia A. Arredondo 2009

3.4 Discussion

Here we report the construction and characterization of covalently linked

derivatives of DsbC joined via (Gly3Ser) repeat flexible linkers. Linking the two

monomers of DsbC via a sufficiently long linker resulted in an enzyme that displayed

essentially identical catalytic and biochemical properties with the authentic dimeric

DsbC. We showed that linked enzymes in which one or both of the active site Cys had

been replaced with Ala exhibited significant isomerase activity. Similar to DsbC, the

mDsbC-mDsbC[C98A], mDsbC-mDsbC[C98A/C101A] and mDsbC-mDsbC[C101A]

variants were completely reduced in vivo and did not support the oxidation of PhoA in a

dsbA background. In contrast, and consistent with earlier findings, substitution of the two

N-terminal C98, or of both C98 and C101 in DsbC, completely eliminated isomerase

activity; whereas replacing the resolving cysteine, C101, on both active sites affected this

activity only partially ((Liu and Wang, 2001; Rietsch et al., 1997; Zapun et al., 1995);

Fig 3.4 and Table 3.2). Earlier, Sun et al. (2000) sought to characterize the enzymatic

properties of chimeric DsbC formed by mixing active and carboxymethylated monomers.

Analysis of the in vitro hybrid mixture led to speculation that both active sites are

required for isomerization (Gleiter and Bardwell, 2008; Sun and Wang, 2000). Our

findings now reveal that the two catalytic sites of DsbC catalyze disulfide bond

isomerization in RNAse A independently of each other. Thus, mDsbC-mDsbC[C98A]

and mDsbC-mDsbC[C98A/C101A] display half the activity of wt DsbC or of (mDsbC-

mDsbC) simply because at the same molar concentration the latter enzyme contains two

active sites. These results are consistent with the observation that substituting the N-

97

Page 113: Copyright by Silvia A. Arredondo 2009

terminal cysteine or both cysteines in either one of the two PDI catalytic sites leads to

partial isomerase activity at varying degrees (Lyles and Gilbert, 1994; Tian et al., 2006;

Vuori et al., 1992; Walker et al., 1996).

The high level of in vitro isomerase activity displayed by mutants in which the

resolving cysteine C101 had been replaced with Ala (DsbC[C101A] and mDsbC-

mDsbC[C101A]) indicates that mixed disulfides with RNAse A can be resolved

efficiently via the attack of another cysteine in the substrate. Alternatively, isomerization

may involve the formation of mixed disulfides with glutathione which is present in the

assay buffer. In contrast, the resolving cysteine, C101, has to participate in disulfide

reduction and therefore the finding that DsbC[C101A] and mDsbC-mDsbC[C101A]

display high insulin reduction activity was unexpected. However this data are consistent

with earlier findings that a DsbC variant with C101A and also C163A, the cysteine that

normally forms the intramolecular bond of DsbC, exhibits 30% activity in this assay (Liu

and Wang, 2001). It should be noted that the insulin reduction assay is performed in a

buffer that contains excess DTT which likely can substitute for C101 in the resolution of

mixed disulfides between the enzyme and the substrate.

By engineering the dimer interface of the linked DsbC we constructed a variant,

mDsbC[H45D]-dim[D53H] consisting of a single thioredoxin domain linked to an intact

dimerization domain. This protein displayed isomerase activity with RNase A, and in the

E. coli periplasm where it enhanced the folding of vtPA. The absence of a second

catalytic domain also allowed this protein to catalyze disulfide bond formation,

supporting the oxidation of PhoA and the formation of active vtPA in a dsbA strain

98

Page 114: Copyright by Silvia A. Arredondo 2009

background. Consistent with the in vivo activity of mDsbC[H45D]-dim[D53H], we

showed that the protein is maintained in a predominantly reduced state.

The cleft forming upon interaction of the N-terminal dimerization domains of

DsbC monomers, lined with uncharged and hydrophobic residues, is considered to be

involved in the non-covalent binding of substrate proteins (McCarthy et al., 2000; Sun

and Wang, 2000). Previous studies reporting that the DsbC dimerization domain folds

into its native conformation independent of the presence of the catalytic domain (Yeh et

al., 2007), together with our findings that disrupting the His-Asp interactions in

mDsbC[H45D]-dim[H45D] results in an extended polypeptide, support the premise that

mDsbC[H45D]-dim[D53H] in fact folds into a DsbC-like overall conformation, most

likely preserving the substrate binding region. The level of in vitro isomerase activity

observed with mDsbC[H45D]-dim[D53H] correlates well to that of the single-active-site

mutant and is substantially higher relative to the background observed with monomeric

DsbC, the active-site mutant DsbC[C98A], or mDsbC[H45D]-dim[H45D] in which the

N-terminal dimerization regions likely fail to interact (8, 5 and 12% respectively) ((Ke et

al., 2006); Table 3.2; unpublished data). This data is in agreement with the published

studies on PDI reporting that the minimal structure for simple disulfide isomerization

includes one catalytic domain (a or a’) in combination with the non-catalytic b’ domain,

the latter identified as the primary substrate binding site (Hatahet and Ruddock, 2007;

Wilkinson and Gilbert, 2004).

The emergence of DsbB-dependent oxidase activity in mDsbC[H45D]-

dim[D53H] which is structurally equivalent to DsbC lacking one thioredoxin domain,

99

Page 115: Copyright by Silvia A. Arredondo 2009

support the steric incompatibility model proposed by Inaba et al. 2006, wherein

interaction of DsbC with DsbB would result in its second thioredoxin domain clashing

onto the membrane (Inaba et al., 2006), and the experimental evidence showing that

monomeric DsbC and α-helix deletion mutants of DsbC with re-oriented catalytic

domains are able to catalyze DsbB-dependent oxidation (Bader et al., 2001; Segatori et

al., 2006).

In summary, through the engineering of heterodimer-like mutants of DsbC we

have demonstrated that only a single catalytic site or domain is essential for catalysis of

isomerization in bacteria. In addition, we propose that the determining factor that

converts the association of thioredoxin-fold-containing monomers from a mere pair of

oxidoreductases in close proximity into an efficient isomerase is the emergence of a

substrate binding region, only present when both N-terminal domains interact. We have

also provided further experimental support to the notion that the kinetic barrier separating

both pathways in bacteria derives from steric complications that dissipate with structural

modifications. These findings contribute to understanding the role that dimerization plays

in the catalysis of isomerization, highlight the indispensable features of DsbC, and

expand the basis for the engineering of DsbC into a more PDI-like protein capable of

catalyzing the oxidative folding of heterologous proteins in the bacterial periplasm in a

highly efficient manner.

100

Page 116: Copyright by Silvia A. Arredondo 2009

Chapter 4

Analysis of the Substrate Specificity of the Bacterial

Disulfide Isomerase DsbC

4.1 Introduction

As discussed in previous chapters, the crystal structure of DsbC reveals that its

two subunits come together through the interactions of the N-terminal domains. The inner

surface of the cleft that forms by the association of the monomers accounts for 45% of

the solvent-accessible surface area of the dimer and is lined with 46 uncharged or

hydrophobic residues. Having the dimensions 40Å x 40Å x 25Å, this cleft is sufficiently

large to accommodate small protein domains and it is believed to play and important role

in the recognition of unfolded polypeptides (Fig 4.1,(McCarthy et al., 2000)).

Figure 4.1. Ribbon diagram and surface representation of the hydrophobic cleft of DsbC that may account for substrate binding and chaperone activity, observed from the top of the V-shaped molecule (McCarthy et al., 2000).

101

Page 117: Copyright by Silvia A. Arredondo 2009

Substrate recognition is essential for DsbC to effectively catalyze disulfide bond

isomerization and to assist folding as a chaperone, however, little is known about the

binding interactions between this enzyme and its substrates. Sun et al. (2000), determined

that the C-terminal domain (residues 66-216) of DsbC, by itself, is a monomer not able to

act as a chaperone or catalyze isomerization. They concluded that the N-terminal

dimerization domain (residues 1-65) is indispensable for activity and substrate binding

(Sun and Wang, 2000). The relevance of the dimerization domain of DsbC in peptide

binding was further demonstrated by Segatori et al. (2004) and Zhao et al. (2003),

independently, through the construction and analysis of chimeras consisting of the DsbC

N-terminal domain fused to different thioredoxin domains (Trx, DsbA, PDIa) that do not

naturally display isomerase or chaperone activity; the fusion proteins, however, were able

to function as isomerases and chaperones (Segatori et al., 2004; Zhao et al., 2003). More

recently, as described in chapter 3, we found evidence that even though substrate binding

is predominantly mediated by the N-terminal domains, the α-helical linker and the

catalytic domains may also contribute to binding, albeit to a lower degree, since the

absence of these structures decreases in vitro isomerase activity of RNAse A by ~15%

(Arredondo, 2009).

Similar results have been reported for PDI, the eukaryotic disulfide isomerase, in

which the primary binding site has been localized in the b’ domain but the presence of

secondary substrate-binding sites in the a and a’ domains has also been proposed (Klappa

et al., 1998). The PDI-substrate binding interactions are not fully understood either,

however, the major binding site has been mapped to a small hydrophobic pocket within

102

Page 118: Copyright by Silvia A. Arredondo 2009

the b’ domain formed by residues Leu-242, Leu-244, Phe-258 and Ile-272, with the last

residue identified as the most critical (Pirneskoski et al., 2004). In addition, studies of the

b’ domain with a seven amino acid peptide phage-display library were reported to show

enrichment for hydrophobic residues in the displayed peptides, especially phenylalanine

(Hatahet and Ruddock, 2007). Although PDI’s substrate specificity is still undefined, for

PDIp, a homologue expressed exclusively in the pancreas, binding was reported to be

dependent upon the presence of a single tyrosine or tryptophan residue in the substrate

with no adjacent negative charge (Ellgaard and Ruddock, 2005; Ruddock et al., 2000).

Further understanding of the interactions between disulfide isomerases and their

substrates is crucial not only for elucidating the principles of isomerization and

chaperone-assisted folding in greater detail, but also as a powerful tool for the

engineering of substrate specificity. No information is available on the substrate

specificity of DsbC and no specific amino acid residues have been identified to be

involved in binding interactions.

In this chapter, we have sought to identify peptide sequences with affinity to

DsbC by screening two 15-amino acid random peptide libraries using bacterial display.

The libraries were created and provided by our collaborators in the Daugherty lab at the

University of California at Santa Barbara. Both libraries were constructed using a

bacterial cell surface display method in which the peptide library is inserted at the N-

terminus of the circularly permuted version of OmpX, a small, monomeric, β-barrel,

outer membrane protein native to E. coli that is highly expressed (Mecsas et al., 1995;

103

Page 119: Copyright by Silvia A. Arredondo 2009

Rice et al., 2006). The structural model for the circularly permuted OmpX (CPX) display

system is shown in figure 4.2.

L1

L3

L4Exterior

LPS

Outer Membrane

Periplasm

CPX

N

C

L1

L2

L3

L4Exterior

LPS

Outer Membrane

Periplasm

OmpX

Figure 4.2. Molecular structure of OmpX and of the circularly permuted version, CPX, in which the 15aa peptide library has been inserted at the N-terminus (shown in pink).

The particular design of this display scaffold was aimed at minimizing the effect

of the host protein (OmpX) in the binding interactions of the displayed peptide and its

target. It has been reported that when peptides that are originally isolated from a loop

insertion display system are analyzed in a soluble format, the binding affinity may be

significantly reduced. Presenting the peptide as a fusion to the N-terminus can overcome

this limitation by better resembling the behavior of a soluble peptide (Rice et al., 2006).

Several peptides with high affinity for proteins like streptavidin and vascular endothelial

growth factor (VEGF) have been isolated or tested using this novel bacterial display

method (Rice and Daugherty, 2008; Rice et al., 2006)

104

Page 120: Copyright by Silvia A. Arredondo 2009

Screening of the CPX peptide libraries, comprising ~109 individual clones,

through magnetic-activated cell sorting (MACS) and fluorescence-activated cell sorting

(FACS) led to the isolation of several peptides that are able to bind to DsbC, displaying

the sequence motif PWX4W as described below.

105

Page 121: Copyright by Silvia A. Arredondo 2009

4.2 Materials and Methods

4.2.1 Strains and Plasmids

The bacterial strains and plasmids used in this study are listed in Table 4.1. The

dsbC[H45D] gene was amplified via PCR using the listed plasmids, digested with XbaI

and HindIII and cloned into pET28(a). Both dsbC and dsbC[H45D] genes contain a

sequence that encodes a C-terminal hexahistidine tag for IMAC purification. The detailed

construction of the circularly permuted OmpX peptide libraries has been previously

described by Rice et al. (2006). Two different libraries were utilized in this study, CPX

and eCPX, the latter being an enhanced version of the former in terms of peptide display

efficiency. Both libraries were constructed using pBAD33 (Guzman et al., 1995) as the

vector and circularly permuted OmpX (CPX) as the scaffold for the insertion of the N-

terminal 15 amino acid peptides (Rice and Daugherty, 2008; Rice et al., 2006).

Table 4.1. Bacterial strains, plasmids and primers utilized in this study.

S

trains Relevant Genotype SourceMC1061 F- araD139 Δ(ara-leu )7696 galE15 galK16 Δ(lac )X74 rpsL (StrR) hsdR2 Lab Collection

(rK- mK+) mcrA mcrB1)PB1077 MC1061 dsbC::Kan Daugherty Lab

21(DE3)pLysS F- ompT hsdS B(rB- mB

-) gal dcm (DE3) pLysS (CamR

BL

) Novagen

Plasmids Relevant Genotype SourceET-28(a) T7 expression vector, C-terminal 6x histidine tag Novagen

pET28-DsbC dsbC from Escherichia colp

i

in pET-28(a) Lab Collection28-DsbC[H45D] dsbC H45D in pET-28(a) This work

cpXopt-temp circularly permuted OmpX lacking an N-terminal peptide insert in pBAD33 Daugherty LabpB33cpX-PHB879 circularly permuted OmpX displaying peptide PHB879 in pBAD33 Daugherty Lab

eCPX-GGS-n-15mer circularly permuted OmpX with random N-terminal 15 aa peptide insert Daugherty Labin pBAD33

rimers SequenceXbaIDsbCss.f GAGCTCGAATTCTCTAGATTAAAGAGGAGAAAGGTACCCATGATGAAGAAAGGTTTTAT

bCHisHindIII.r TTTTTAAGCTTTTAGTGGTGGTGGTGGTGGTGTTTACCGCTGGTCATTTTTTG

bC-H45D.r GGCCCCTGAATGATATCTTTACCATCATC

pETpB33

pB33

P

DsDs

106

Page 122: Copyright by Silvia A. Arredondo 2009

4.2.2 Protein Expression, Purification and Labeling

For the purification of DsbC and DsbC[H45D], the pET28(a) derivates expressing

the respective genes were transformed into E. coli BL21(DE3). Protein expression and

purification was performed as previously described (Segatori et al., 2004). The proteins

were more than 95% pure as judged by Coomassie stained SDS-PAGE electrophoresis.

Biotinylation of DsbC was performed by incubating 1 ml of the protein (2mg/ml

concentration) with EZ-Link® Sulfo-NHS-Biotin (Pierce, IL) according the

manufacturer’s instructions, resulting in an average of 5 biotin groups conjugated to each

DsbC molecule. For flow cytometry, DsbC was labeled with Alexaflour®488 (Molecular

Probes, OR), at ~ 2 moles of fluorophore per mole of protein, following the provider’s

protocol.

4.2.3 Peptide Synthesis and Purification

Solid phase peptide synthesis was performed in a CEM Liberty Automated

Synthesizer (CEM, NC) using the protected amino acids: Fmoc-Pro-OH, Fmoc-

Trp(Boc)-OH, Fmoc-Phe-OH, Fmoc-Leu-OH, Fmoc-Gln(Trt)-OH, Fmoc-Asn(Trt)-OH

and the Fmoc-Gly-Wang resin (EMD Biosciences, CA). Upon synthesis completion, the

resin-peptide product was washed multiple times with dichloromethane (DCM) to

remove any residual dimethylformamide (DMF), and dried in vacuum for 30 min. The

peptide was then cleaved and deprotected by mixing the dry resin with 10 ml of cleavage

cocktail (8.8 ml TFA, 0.5 g phenol, 0.5 ml ddH 0 and 0.2 ml TIPS) and shaking at room

temperature for 1hr. The peptide mix was precipitated from the filtered solution by

2

107

Page 123: Copyright by Silvia A. Arredondo 2009

adding 80ml of ether and incubating at 4ºC overnight. The final product was vacuum-

filtered.

To determine the purity of the crude peptide, a sample was analyzed by HPLC on

a C-18 reverse-phase column using the gradient: 5% MeOH-95% H O for 1 min,

increasing to 30% MeOH-70% H O over a period of 5 min, increasing to 80% MeOH-

20% H O over 60 min and finally increasing to 100% MeOH over 5 min. Once the

correct peak was identified by liquid chromatography-mass spectrometry (LC-MS)

traces, the peptide was purified using a C-18 reverse-phase semi-prep column with the

gradient: 5% MeOH-95% H O for 1 min, increasing to 30% MeOH-70% H O over a

period of 5 min, increasing to 55% MeOH-45% H O over 40 min and finally increasing

to 100% MeOH over 5 min. After lyophilization, the mass of the purified peptide was

confirmed by LC-MS.

2

2

2

2 2

2

4.2.4 Whole Cell ELISA

Overnight cultures of E. coli PB1077 (MC1061 dsbC::Kan) harboring pB33cpX-

PHB879 or the empty scaffold pB33cpXopt-temp were subcultured 1:50 in LB medium

with 50 μg/ml of kanamycin and 25 μg/ml of chloramphenicol and grown for 2 hrs with

shaking (250 rpm) at 37 ºC. The cultures were then transferred to 25 ºC and induced with

arabinose to a 0.04% (w/v) final concentration. After two hours, cell volumes

corresponding to OD600= 1 were collected and washed once with incubation buffer (PBS,

1% BSA). Cells were resuspended in 1ml of the same buffer and 80 μl aliquots were

transferred into a non-binding 96 well plate. Purified DsbC was added to the wells at

108

Page 124: Copyright by Silvia A. Arredondo 2009

different concentrations ranging from 0-2 μg/μl and the plate was incubated for 1 ½

hours at 4ºC. Cells were then washed twice with incubation buffer through

centrifugation-resuspension cycles. 200 μl of monoclonal antipolyhistidine HRP-

conjugated antibody (Sigma, Israel) at a 1:2000 dilution in 1% milk buffer was added to

each well and incubated for 1 hr at 4ºC. After washing twice, 0.2 ml of TMB+ substrate

(Dako, Denmark) was added to each well and the reaction was quenched with 0.1 ml of

1M H2SO4 . Cells were precipitated by centrifugation and A450 of the supernatant was

then measured.

4.2.5 Magnetic-Activated Cell Sorting

Overnight cultures of E. coli MC1061 cells harboring the eCPX library

(pB33eCPX-GGS-n-15mer) corresponding to 5 times the size of the library (OD600= 50)

were inoculated to a final concentration of OD600= 0.1 in LB medium with 25 μg/ml of

chloramphenicol and grown at 37ºC with shaking (250 rpm). When the culture reached

OD600= 0.5, it was transferred to 25ºC and induced with arabinose to a 0.04% (w/v) final

concentration. Following an additional incubation at 25 ºC for 50 min, a volume of cells

corresponding to OD600= 50 (5X the size of the library) was concentrated by

centrifugation (3000 g, 4ºC, 10 min) and resuspended in 2 ml of cold PBS. To deplete the

library from cells displaying streptavidin-binding peptides, the cells were mixed with ~ 2

x 108 Dynal MyOne SA beads (Invitrogen, Oslo, Norway) and incubated at 4ºC with light

rotation. Unbound cells were magnetically separated from the beads and incubated with

250 μl of biotinylated DsbC (2mg/ml) for 1 hr at 4ºC. Cells were then concentrated by

109

Page 125: Copyright by Silvia A. Arredondo 2009

centrifugation (3000 g, 4ºC, 8 min), resuspended in 3 ml of fresh PBS and mixed with ~ 2

x 108 fresh beads. After incubation for 30 min at 4ºC, magnetic separation was used to

wash the cells four times with LB. The cell-DsbC-beads were finally inoculated in 500

ml of LB with 0.2% glucose (w/v) and 25 μg/ml of chloramphenicol for overnight

growth.

4.2.6 Flow-cytometric Analysis and Sorting

Inoculation, growth and induction of E. coli MC1061 cells harboring the

magnetically enriched eCPX library (pB33eCPX-GGS-n-15mer) was carried out as

described in section 4.2.5 with the exception that the starting inoculum was equivalent to

OD600= 2. Following induction, a volume of cells corresponding to OD600= 0.2 was

concentrated by centrifugation (5,000 rpm, 5 min), resuspended in 100 μl of 2μM DsbC-

AF488 (DsbC conjugated to Alexaflour®488 (Molecular Probes, OR) in PBS + 1% BSA)

and incubated on ice for 30 minutes. The cells were then pelleted by centrifugation,

resuspended in 1 ml of fresh PBS and immediately analyzed and/or sorted via a

FACSAria cytometer with 488 nm excitation. Approximately 2-5% of the most

fluorescent events were collected and grown overnight in LB with 2% glucose (w/v) and

chloramphenicol for subsequent rounds of sorting.

110

Page 126: Copyright by Silvia A. Arredondo 2009

4.3 Results

4.3.1 The NLCPX-X15 Library

The CPX bacterial display system was created at the Daugherty Lab at UCSB by

fusing the native N and C termini of the outer membrane protein OmpX using a GGSG

linker and opening the backbone where the loop 2 was originally found. Then, a peptide

passenger was inserted in between two GGQSGQ linkers followed by the new N-

terminus sequence (Fig 4.1, (Rice et al., 2006)). The NLCPX-X15 library carries a

random 15 amino acid peptide insertion. This library was originally screened against

purified DsbC by Dr. Paul Bessette at the Daugherty Lab. After two rounds of magnetic

sorting (MACS) and two rounds of florescence sorting (FACS), 30 highly florescent

clones were isolated and sent to our lab for further characterization.

The putative binding site of DsbC is localized at the cleft that forms upon the

interaction of the N-terminal domains of the monomers, making dimerization essential

for substrate binding (McCarthy et al., 2000; Segatori et al., 2004; Sun and Wang, 2000;

Zhao et al., 2003) (Arredondo, 2009). Therefore, we reasoned that in order to closely

mimic real substrate interactions with DsbC, it was important to isolate peptides that bind

to DsbC in its natural dimeric conformation but that do not interact with the monomeric

form of the enzyme. Based on the findings of Bader et al, indicating that the mutation

H45D renders DsbC unable to dimerize due to the disruption of a salt bridge that forms

between H45 and D53 (Bader et al., 2001), DsbC[H45D] was constructed and purified.

Analysis by FPLC gel chromatography (Fig 4.3) revealed that DsbC[H45D] has an

elution pattern different from wt DsbC, with a shorter retention time and an apparent

111

Page 127: Copyright by Silvia A. Arredondo 2009

M.W. of 22 kDa which correlates well with the expected M.W. of 24 kDa corresponding

to a single monomeric unit of DsbC.

0

100

200

300

400

500

0 5 10 15 20ml

mA

U

DsbCDsbC[H45D]

Figure 4.3. Gel filtration FPLC of DsbC and DsbC[H45D]. Purified proteins were run on a SuperdexTM 200 column in PBS-10% glycerol buffer.

Fluorescently labeled DsbC and DsbC[H45D] were then used as probes by Dr.

Vasso Skouridou for the flow-cytometric analysis of the individual peptides. Thirteen of

the thirty peptides originally isolated were identified as binders to DsbC but not to

DsbC[H45], and a common motif consisting of the sequence PWX4W was observed. The

three peptides that showed the highest fluorescence ratio when compared to the empty

scaffold are shown in table 4.2.

112

Page 128: Copyright by Silvia A. Arredondo 2009

Table 4.2. DsbC-binding peptides isolated from the NLCPX-X15 library. The consensus motif is shown in yellow.

Peptide Sequence Fluorescence ratio [Peptide/Empty Scaffold]

PHB879 WPWFLQNWSDGLSFQ 3.1 + 0.1885 CLGMPWWCDNWMTNS 2.4 +PHB 0.1

HB903 GTPWYLELWDYNADR 2.6 +

The binding affinity of peptides PHB879, PHB885 and PHB903 was further

investigated through whole cell enzyme-linked immunosorbent assays (ELISA). In this

assay, PB1077 (MC1061 dsbC::Kan) cells expressing the scaffold displaying the

peptides, or only the scaffold, from the arabinose promoter in pBAD33 vectors were

collected and incubated with purified DsbC containing a C-terminus hexa-histidine

sequence. Detection of the bound DsbC was done through incubation with a HRP-

conjugated monoclonal antibody against the His tag, followed by HRP substrate reaction

and acid quenching. Upon confirmation of peptide PHB879 as the best binder, ELISA

was used to approximate the affinity of the peptide to DsbC at different concentrations of

the protein, resulting in a rough dissociation constant Kd of 100 nM (Fig 4.4).

P 0.3

113

Page 129: Copyright by Silvia A. Arredondo 2009

0

0.5

1

1.5

2

2.5

3

0.1 1 10 100 1000

nM DsbC per ~ 10^7 cells

Abs

450

PHB879 peptide

Empty scaffold

Figure 4.4. Whole cell ELISA data for binding of purified DsbC to peptide PHB879 containing the sequence WPWFLQNWSDGLSFQ, and to the empty scaffold as a negative control.

Peptide PHB879, as originally displayed in the cell, is defined by the amino acid

sequence: WPWFLQNWSDGLSFQ; however, based on the common motif PWX4W

derived from the comparison of the stronger binders, we hypothesized that the actual

residues that interact with DsbC are within the very hydrophobic sequence PWFLQNW.

To explore this hypothesis and to precisely determine the affinity of this sequence of

amino acids to DsbC, it was necessary to have the peptide in a soluble form. Therefore,

the seven amino acid peptide was synthesized by the solid phase synthesis method

(SPPS) (Merrifield, 1964) in an automated synthesizer using Fmoc-protected amino acids

and a Gly-resin. Following reverse-phase HPLC purification, the expected mass of the

peptide corresponding to 1047.4 Da was confirmed by mass spectrometry LC-MS (Fig

4.5).

114

Page 130: Copyright by Silvia A. Arredondo 2009

0

10

20

30

40

50

60

70

80

90

100

1 2 3 4 5 6 7 8 9 10 11

m/z

Rel

ativ

e A

bund

ance

200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400 1500 1600m/z

1047.4

1048.4

1069.51153.4

1070.5708.3 836.5 933.3498.3209.2 817.5 1209.7 1370.9 1474.71086.41033.4 1330.6524.3 664.9 1571.2648.4455.3239.2 397.3330.0

PHB879 PWFLQNWG = 1047.4

1047.4

Figure 4.5. Mass spectrometry trace for the synthetic peptide PWFLQNWG.

The next step will be the precise determination of the binding affinity of

PWFLQNW to DsbC using fluorescence anisotropy followed by alanine-scanning of the

sequence, where each amino acid residue is systematically substituted with alanine in

order to identify the key residues in the binding interaction.

4.3.2 The eCPX Library

Although the original CPX scaffold is effective in displaying passenger peptides,

the efficiency observed is not the same as when peptides are inserted in the loop of native

OmpX, due to reduced membrane localization of the circularly permuted version. The

lower display efficiency translates into longer induction times in order to obtain sufficient

115

Page 131: Copyright by Silvia A. Arredondo 2009

peptide display for screening by fluorescence-activated cell sorting (FACS); and in some

cases, may result in isolation of clones that are better expressed rather than clones with

higher binding affinity. To address these issues, Rice et al, created an enhanced version

of this system, known as eCPX, using a combination of random and targeted

mutagenesis; the new display system is as efficient as the non-permuted OmpX in

displaying peptides. Based on the eCPX scaffold, a second random peptide library was

created and provided by the Daugherty lab (Rice and Daugherty, 2008).

The eCPX-GGS-n-15mer library consists of 8.3 x 109 independent clones

displaying random peptides 15 aa in length. Magnetic-activated cell sorting (MACS) with

streptavidin-coated beads was initially utilized for screening the library. Approximately

4x 1010 MC1061 cells harboring the peptide library were grown and induced as described

above. In order to avoid the isolation of peptides with affinity for streptavidin, the cells

were incubated with the beads and magnetically separated. After incubating the unbound

cells with biotinylated DsbC, fresh beads were added to capture the DsbC-E. coli

complexes. Significant enrichment of E. coli cells displaying peptides with affinity

towards DsbC was obtained after two rounds of magnetic sorting as confirmed by flow

cytometric analysis (Fig 4.6). A two-fold increase in the mean fluorescence of the

MACS-enriched library was observed relative to cells expressing the empty scaffold as a

negative control.

116

Page 132: Copyright by Silvia A. Arredondo 2009

M2 library

Negative control

Figure 4.6. Fluorescence distribution of the magnetically enriched peptide library for affinity towards purified DsbC conjugated to the fluorescent probe Alexaflour®488.

Following magnetic sorting, the library was subjected to three rounds of

fluorescent-activated cell sorting (FACS). For each round, approximately ~108 cells were

labeled with DsbC conjugated to the fluorophore Alexaflour®488 (Molecular Probes,

OR) and immediately analyzed and sorted using a FACSaria instrument. For the first two

rounds of sorting, 2μM DsbC-AF488 was used for labeling the cells and 2-5% of the most

fluorescent population was collected. For the third round, 1μM DsbC-AF488 was used and

the sorting gate was further restricted to 1.5%. The mean fluorescence of the peptide

library consistently increased in each round of sorting as shown in Fig 4.7.

117

Page 133: Copyright by Silvia A. Arredondo 2009

FL intensity

Even

ts

100 101 102103

104

MACS library

Round 1

Round 2

Round 3

M 12.8

M 18.7

M 19.5

M 21.2

Figure 4.7. Comparison of the fluorescence intensity of the eCPX-GGS-n-15mer library in sequential rounds of FACS. Data represent the mean fluorescence (M) of 10,000 E. coli events.

The events collected from the third round of FACS were plated on LB-agar and

grown overnight. 20 clones were randomly selected, grown and induced for peptide

display. After labeling with 1μM DsbC-AF488, the clones were individually scanned for

increased fluorescence in comparison to the empty scaffold. In figure 4.8, the

fluorescence intensity of three of the best binders is shown. Six of the clones with highest

signal with fluorescent DsbC were sequenced and are shown in table 4.3.

118

Page 134: Copyright by Silvia A. Arredondo 2009

Neg control SAA001

SAA019SAA005

Figure 4.8. Representative data of the fluorescence intensity of individual clones displaying 15 amino acid peptides with affinity towards DsbC isolated after two rounds of magnetic sorting and three rounds of FACS.

Table 4.3. DsbC-binding peptides isolated from the eCPX-GGS-n-15mer library.

Peptide Sequence Fluorescence ratio

[Peptide/Empty scaffold]SAA001 QCAVDQWKHHWLNGW 38.2SAA005 RWVCVHEIESWLAGW 7.2SAA007 CDMDGMKDWRSGWPV 5.0SAA013 SGWCSVASHSWMVDW 3.1SAA019 SCQVRLPITALQWRW 10.7SAA004 LGGPYCPVGVWSPMR 2.9

119

Page 135: Copyright by Silvia A. Arredondo 2009

The data shown in table 4.3 reveals there is a recurring cysteine residue in all the

peptides isolated. The presence of a common cysteine in these peptides suggests the

possibility that the high affinity for DsbC may derive from the ability of the peptide to

establish a disulfide bond with the catalytic cysteines in the isomerase and get trapped.

Interestingly, a common set of amino acids is observed in three of the six peptides,

WX2GW; and a similar arrangement, WX3W, is found in a fourth peptide. Whether the

affinity towards DsbC can be attributed mostly to the interaction with the common

sequence motif, to a combination of the motif and the formation of a disulfide bond, or to

additional interactions with other residues, will be determined in the near future.

Further characterization of the isolated peptides will require assessment of their

affinity towards an inactive version of DsbC (lacking the catalytic cysteines) and to the

monomeric DsbC[H45D] in order to eliminate the possibility of disulfide bond formation

and binding outside the putative substrate-recognition cleft formed by dimerization.

Additionally, a bigger pool of clones will be analyzed and sequenced; and, if necessary,

the library may be sorted again including negative selection against inactive DsbC

variants.

120

Page 136: Copyright by Silvia A. Arredondo 2009

4.4 Future Work

Although the investigation of the amino acid specificity of the DsbC binding

pocket is well underway as presented in this chapter, further studies will be necessary for

successful completion of the project. Once additional consensus motifs have been

identified and/or confirmed to have affinity for DsbC, shorter peptide versions containing

the defined motifs, and alanine-scanning variants, may be inserted in the empty scaffold

and further analyzed in the cell-display format via flow cytometry and/or ELISA. For the

precise determination of the binding affinity of the candidate sequences, fluorescence

anisotropy, using synthetic peptides will be necessary. To that end we have already begun

the synthesis of putative peptide substrates.

To validate and expand our findings, we will perform comparison studies of the

consensus motifs with the sequences and three-dimensional structures (if available) of the

known DsbC substrates, MepA, RNaseI and AppA, and of heterologous proteins which

folding is known to be assisted by the bacterial isomerase, for example tPA and RNase A.

Finally, the isolation of robust binding motifs will be useful as probes for

screening random mutagenesis libraries of DsbC to find mutants with increased affinity

that can shed light on the residues of DsbC that interact with the substrate.

121

Page 137: Copyright by Silvia A. Arredondo 2009

Chapter 5

Conclusions and Recommendations

Compiled in this dissertation are three interrelated projects that share as a central

objective the progressive understanding of the mechanism of action of the disulfide bond

isomerase in E. coli. The overall motivation of this work is the engineering of an

optimized version of the disulfide bond isomerase for the efficient production of

heterologous proteins containing complex disulfide patterns in bacteria, which requires

an in depth knowledge of these enzymes. Initially, we examined the significance of

thioredoxin catalytic domains, their optimal positioning and the availability of a substrate

binding region as key features for isomerization (Segatori et al., 2004). Subsequently, we

addressed the questions of whether two identical catalytic domains are strictly necessary

for catalysis and what are the minimal components of an enzyme that is able to catalyze

disulfide isomerization. Finally, we directed our efforts towards elucidating the substrate

binding interactions of the bacterial disulfide isomerase DsbC.

As discussed in chapter 2, we engineered artificial disulfide isomerases utilizing

domains derived from proteins that do not exhibit such a function. Fusion of DsbA,

serving as the catalytic domain, to the dimerization domain of FkpA, providing a

substrate binding region, resulted in chimeric enzymes able to catalyze in vivo disulfide

bond formation to varying degrees. It is important to note that simple fusion of

122

Page 138: Copyright by Silvia A. Arredondo 2009

thioredoxin domains without a suitable substrate binding area did not result in proteins

displaying efficient isomerase activity. Each version of the FkpA-DsbA fusion proteins

had the thioredoxin domains likely to be positioned at slightly different orientation as a

result of sequential amino acid deletions in the α-helical linker. In addition, and unlike

DsbC, these proteins were able to catalyze disulfide bond formation; activity, we believe,

also derived from the unaligned orientation of the catalytic domains. The best chimera,

FkpA-DsbA2 catalyzed in vivo isomerization at ~ 80% of the level observed with DsbC,

and was subjected to directed evolution with the purpose of simultaneously enhancing

isomerization while reducing oxidation to more closely resemble the behavior of DsbC.

Although better physiological fitness in the presence of CuCl2 was conferred by the

evolved enzyme through a mutation in the dipeptide of the active site increasing its

similarity to DsbC, no isomerase activity improvement was observed in the folding of

vtPA. Additionally, elimination of oxidase activity resulted only from the mutation of the

resolving cysteine in the CXXC motif. Collectively, these studies indicate that the correct

orientation of the catalytic domains is in fact significant for isomerization, that the

dimerization of thioredoxin domains alone is not sufficient to grant isomerase activity

and that providing a substrate binding region with similar topology to DsbC is crucial for

the catalysis of isomerization. Moreover, the evolution of FkpA-DsbA2 pointed towards

the importance of having an optimal active site for high catalytic activity.

Through the engineering of heterodimer-like versions of DsbC, as described in

chapter 3, we were able to determine that dual catalytic domains are not required for

isomerization. A monomeric version of DsbC was initially constructed by fusing both

123

Page 139: Copyright by Silvia A. Arredondo 2009

subunits of the enzyme via a Gly3Ser flexible linker. Upon inactivation of one of the C98-

GY-C101 catalytic sites in the linked DsbC, in vivo isomerization was observed to remain

at high levels in comparison to the wild type protein. The intrinsic isomerase activity of

the active site mutants, however, decreased to about half of the activity of DsbC, a

reasonable effect given that, with the same molar protein concentration, only half of the

catalytic equivalents were present in that molecule. Furthermore, we were able to

determine that removal of an entire catalytic domain still results in substantial in vitro

activity. However, the DsbC monomeric subunit, which lacks the dimerization domain,

cannot catalyze the isomerization of RNase A. Interestingly, deletion of one thioredoxin

domain also resulted in appreciable levels or oxidase activity, not naturally observed in

DsbC. In short, the results derived from the construction and analysis of the heterodimer-

like DsbC mutants demonstrated that only a single catalytic site or domain is essential for

catalysis of isomerization in bacteria, and led us to propose that the determining factor

that converts the association of thioredoxin-fold-containing monomers from a mere pair

of oxidoreductases in close proximity into an efficient isomerase is the emergence of a

substrate binding region, only present when both N-terminal domains interact

It is worth emphasizing two common findings derived from the separate studies

described in chapters 2 and 3. The first was the emergence of oxidase activity as a result

of structural modifications in these enzymes, breaching the oxidation/isomerization

barrier, either by changing the orientation of the catalytic domains, or by entirely

removing one domain. Our work provides further experimental evidence supporting the

notion that the kinetic barrier separating the two pathways in bacteria originates from

124

Page 140: Copyright by Silvia A. Arredondo 2009

steric effects. These results agree well with the steric incompatibility model proposed by

Inaba et al. 2006, wherein interaction of DsbC with DsbB were suggested to result in its

second thioredoxin domain clashing onto the membrane (Inaba et al., 2006), and with

previous experimental evidence showing that monomeric DsbC and α-helix deletion

mutants of DsbC with re-oriented catalytic domains are able to catalyze DsbB-dependent

oxidation (Bader et al., 2001; Segatori et al., 2006).

The second major conclusion of our investigations is the relevance of a suitable

substrate binding region for the efficient catalysis of isomerization. Initially, we

demonstrated that simply fusing two thioredoxin domains does not confer isomerase

activity unless an appropriate substrate recognition region is provided. Then, we showed

that the minimal structure for efficient isomerization must include the N-terminal

domains from the two subunits, and that these domains need to interact to establish the

substrate binding cleft.

Given the importance of the substrate-enzyme interaction for the catalysis of

isomerization, it is surprising that there is little information regarding its biochemical

properties. Hence, the studies reported in chapter 4 are focused on the analysis of the

substrate specificity of DsbC and the identification of particular amino acid residues or

motifs involved in binding. We have isolated several peptides with affinity for DsbC and

potential consensus motifs have been determined. Future work will be directed at

elucidating the interactions of DsbC and its substrates in greater detail.

In summary, through the engineering and characterization of disulfide bond

isomerases in bacteria, we have not only delineated the key features of catalysis of

125

Page 141: Copyright by Silvia A. Arredondo 2009

isomerization, but also contributed to the understanding of the role that dimerization and

substrate interactions play in this process. We believe that our results provide a firm basis

for the engineering and optimization of a bacterial disulfide isomerase capable of

catalyzing efficient oxidative folding of heterologous proteins for biotechnological

applications.

126

Page 142: Copyright by Silvia A. Arredondo 2009

5.1 Recommendations for Future Studies

Since the general objective is the construction of a better bacterial platform for the

efficient production of eukaryotic, multidisulfide proteins, it follows that the co-

expression of a eukaryotic disulfide isomerase might be preferable to the bacterial

enzyme. Unfortunately, when the eukaryotic PDI is expressed in the bacterial periplasm,

it acts as an oxidase and not as an isomerase (Ostermeier et al., 1996). As discussed in

chapter 2, the use of a directed evolution approach for the enhancement of the FkpA-

DsbA2 fusion protein combined with the CuCl2 viability assay led to the isolation of a

point mutation that rendered the protein more similar to DsbC. This same approach could

prove beneficial in isolating mutants of PDI that may or may not be closer to DsbC but

that are able to act as isomerases in bacteria. And even if a better isomerase is not

immediately isolated, a mutant with moderate activity could point towards residues or

motifs worthy of further exploration.

Alternatively, since isomerization is so heavily dependent on the availability of an

appropriate substrate binding region, and recycling mechanisms to maintain PDI active as

an isomerase are not present in the periplasmic space of E. coli, creating a chimera

utilizing the b and b’ domains from PDI as the main substrate binding site and the

thioredoxin domains of DsbC (and possibly the α-helical linkers) as the catalytic

domains, may improve heterologous substrate recognition while maintaining

isomerization activity.

The new linked monomeric versions of DsbC open the door to many other

possibilities by allowing unique changes within the context of the full protein. Perhaps

127

Page 143: Copyright by Silvia A. Arredondo 2009

the linked DsbC proteins could be used to create PDI-DsbC chimeras through the

incremental truncation of hybrid enzymes system (ITCHY) or by constructing a library

through random gene shuffling (Brannigan and Wilkinson, 2002). However, the absence

of robust high-throughput screening methods that can be used for the isolation of better

isomerases from large libraries may hinder the progress in this area. Therefore, it is

recommended that time and effort is invested in developing improved methods to screen

for the enhancement of isomerization at high-throughput scales.

Finally, once residues and motifs involved in binding are characterized in detail,

as discussed in chapter 4, a random mutagenesis library of DsbC can be created and

screened in order to isolate DsbC variants with better affinity for the motif. These

mutants will shed light onto the location of the substrate recognition regions within

DsbC. The identification of these main binding sites will open the possibility of using

saturation mutagenesis to engineer the substrate specificity of the enzyme to catalyze the

folding of a wide range of substrate proteins with or without multiple disulfide bonds.

128

Page 144: Copyright by Silvia A. Arredondo 2009

Author’s Publication List

Arredondo S, Chen T, Gilbert H, Georgiou G. (2009) The Role of Dimerization in the

Catalytic Properties of the Escherichia coli Disulfide Isomerase DsbC.

(submitted)

Arredondo S, Segatori L, Gilbert H, Georgiou G. (2008) De novo Design and Evolution

of Artificial Disulfide Isomerase Enzymes Analogous to the Bacterial DsbC. J.

Biol. Chem., 283, 31469-31476.

Segatori L, Murphy L, Arredondo S, Kadokura H, Gilbert H, Beckwith J, Georgiou G.

(2006) Conserved Role of the Linker α-Helix of the Bacterial Disulfide Isomerase

DsbC in the Avoidance of Misoxidation by DsbB. J. Biol. Chem., 281, 4911-

4919.

129

Page 145: Copyright by Silvia A. Arredondo 2009

References

Aggarwal, S. (2008) What's fueling the biotech engine-2007. Nat Biotechnol, 26, 1227-

1233.

Akiyama, Y. and Ito, K. (1993) Folding and assembly of bacterial alkaline phosphatase in

vitro and in vivo. J Biol Chem, 268, 8146-8150.

Anfinsen, C.B. (1973) Principles that govern the folding of protein chains. Science, 181,

223-230.

Appenzeller-Herzog, C. and Ellgaard, L. (2008) The human PDI family: versatility

packed into a single fold. Biochim Biophys Acta, 1783, 535-548.

Arolas, J.L., Aviles, F.X., Chang, J.Y. and Ventura, S. (2006) Folding of small disulfide-

rich proteins: clarifying the puzzle. Trends Biochem Sci, 31, 292-301.

Arredondo, S., Segatori, L., Gilbert, H.F. and Georgiou, G. (2008) De Novo Design and

Evolution of Artificial Disulfide Isomerase Enzymes Analogous to the Bacterial

DsbC. J Biol Chem, 283, 31469-31476.

Aslund, F. and Beckwith, J. (1999) The thioredoxin superfamily: redundancy, specificity,

and gray-area genomics. J Bacteriol, 181, 1375-1379.

Bader, M.W., Hiniker, A., Regeimbal, J., Goldstone, D., Haebel, P.W., Riemer, J.,

Metcalf, P. and Bardwell, J.C. (2001) Turning a disulfide isomerase into an

oxidase: DsbC mutants that imitate DsbA. Embo J, 20, 1555-1562.

Bader, M.W., Xie, T., Yu, C.A. and Bardwell, J.C. (2000) Disulfide bonds are generated

by quinone reduction. J Biol Chem, 275, 26082-26088.

Baneyx, F. (1999) Recombinant protein expression in Escherichia coli. Curr Opin

Biotechnol, 10, 411-421.

Baneyx, F. and Mujacic, M. (2004) Recombinant protein folding and misfolding in

Escherichia coli. Nat Biotechnol, 22, 1399-1408.

Bardwell, J.C., McGovern, K. and Beckwith, J. (1991) Identification of a protein required

for disulfide bond formation in vivo. Cell, 67, 581-589.

130

Page 146: Copyright by Silvia A. Arredondo 2009

Berkmen, M., Boyd, D. and Beckwith, J. (2005) The nonconsecutive disulfide bond of

Escherichia coli phytase (AppA) renders it dependent on the protein-disulfide

isomerase, DsbC. J Biol Chem, 280, 11387-11394.

Bessette, P.H., Aslund, F., Beckwith, J. and Georgiou, G. (1999a) Efficient folding of

proteins with multiple disulfide bonds in the Escherichia coli cytoplasm. Proc

Natl Acad Sci U S A, 96, 13703-13708.

Bessette, P.H., Cotto, J.J., Gilbert, H.F. and Georgiou, G. (1999b) In vivo and in vitro

function of the Escherichia coli periplasmic cysteine oxidoreductase DsbG. J Biol

Chem, 274, 7784-7792.

Bizub, D., Weber, I.T., Cameron, C.E., Leis, J.P. and Skalka, A.M. (1991) A range of

catalytic efficiencies with avian retroviral protease subunits genetically linked to

form single polypeptide chains. J Biol Chem, 266, 4951-4958.

Brannigan, J.A. and Wilkinson, A.J. (2002) Protein engineering 20 years on. Nat Rev Mol

Cell Biol, 3, 964-970.

Brickman, E. and Beckwith, J. (1975) Analysis of the regulation of Escherichia coli

alkaline phosphatase synthesis using deletions and phi80 transducing phages. J

Mol Biol, 96, 307-316.

Buchner, J., Grallert, H. and Jakob, U. (1998) Analysis of chaperone function using

citrate synthase as nonnative substrate protein. Methods Enzymol, 290, 323-338.

Chen, J., Song, J.L., Zhang, S., Wang, Y., Cui, D.F. and Wang, C.C. (1999) Chaperone

activity of DsbC. J Biol Chem, 274, 19601-19605.

Chivers, P.T., Laboissiere, M.C. and Raines, R.T. (1996) The CXXC motif: imperatives

for the formation of native disulfide bonds in the cell. Embo J, 15, 2659-2667.

Chivers, P.T., Prehoda, K.E. and Raines, R.T. (1997) The CXXC motif: a rheostat in the

active site. Biochemistry, 36, 4061-4066.

Christis, C., Lubsen, N.H. and Braakman, I. (2008) Protein folding includes

oligomerization - examples from the endoplasmic reticulum and cytosol. Febs J,

275, 4700-4727.

131

Page 147: Copyright by Silvia A. Arredondo 2009

Collet, J.F. and Bardwell, J.C. (2002) Oxidative protein folding in bacteria. Mol

Microbiol, 44, 1-8.

Collet, J.F., Riemer, J., Bader, M.W. and Bardwell, J.C. (2002) Reconstitution of a

disulfide isomerization system. J Biol Chem, 277, 26886-26892.

Dailey, F.E. and Berg, H.C. (1993) Mutants in disulfide bond formation that disrupt

flagellar assembly in Escherichia coli. Proc Natl Acad Sci U S A, 90, 1043-1047.

Datar, R.V., Cartwright, T. and Rosen, C.G. (1993) Process economics of animal cell and

bacterial fermentations: a case study analysis of tissue plasminogen activator.

Biotechnology (N Y), 11, 349-357.

De Lorenzo, F., Goldberger, R.F., Steers, E., Jr., Givol, D. and Anfinsen, B. (1966)

Purification and properties of an enzyme from beef liver which catalyzes

sulfhydryl-disulfide interchange in proteins. J Biol Chem, 241, 1562-1567.

Debarbieux, L. and Beckwith, J. (1998) The reductive enzyme thioredoxin 1 acts as an

oxidant when it is exported to the Escherichia coli periplasm. Proc Natl Acad Sci

U S A, 95, 10751-10756.

Dobson, C.M. (2004) Principles of protein folding, misfolding and aggregation. Semin

Cell Dev Biol, 15, 3-16.

Dyck, M.K., Lacroix, D., Pothier, F. and Sirard, M.A. (2003) Making recombinant

proteins in animals--different systems, different applications. Trends Biotechnol,

21, 394-399.

Ellgaard, L. and Ruddock, L.W. (2005) The human protein disulphide isomerase family:

substrate interactions and functional properties. EMBO Rep, 6, 28-32.

Fersht, A.R. (2008) From the first protein structures to our current knowledge of protein

folding: delights and scepticisms. Nat Rev Mol Cell Biol, 9, 650-654.

Foloppe, N. and Nilsson, L. (2004) The glutaredoxin -C-P-Y-C- motif: influence of

peripheral residues. Structure, 12, 289-300.

Fromant, M., Blanquet, S. and Plateau, P. (1995) Direct random mutagenesis of gene-

sized DNA fragments using polymerase chain reaction. Anal Biochem, 224, 347-

353.

132

Page 148: Copyright by Silvia A. Arredondo 2009

Georgiou, G. and Segatori, L. (2005) Preparative expression of secreted proteins in

bacteria: status report and future prospects. Curr Opin Biotechnol, 16, 538-545.

Gilbert, H.F. (1990) Molecular and cellular aspects of thiol-disulfide exchange. Adv

Enzymol Relat Areas Mol Biol, 63, 69-172.

Gleiter, S. and Bardwell, J.C. (2008) Disulfide bond isomerization in prokaryotes.

Biochim Biophys Acta, 1783, 530-534.

Glockshuber, R. (1999) Protein folding. Where do the electrons go? Nature, 401, 30-31.

Grauschopf, U., Fritz, A. and Glockshuber, R. (2003) Mechanism of the electron transfer

catalyst DsbB from Escherichia coli. Embo J, 22, 3503-3513.

Grauschopf, U., Winther, J.R., Korber, P., Zander, T., Dallinger, P. and Bardwell, J.C.

(1995) Why is DsbA such an oxidizing disulfide catalyst? Cell, 83, 947-955.

Gruber, C.W., Cemazar, M., Heras, B., Martin, J.L. and Craik, D.J. (2006) Protein

disulfide isomerase: the structure of oxidative folding. Trends Biochem Sci.

Guilhot, C., Jander, G., Martin, N.L. and Beckwith, J. (1995) Evidence that the pathway

of disulfide bond formation in Escherichia coli involves interactions between the

cysteines of DsbB and DsbA. Proc Natl Acad Sci U S A, 92, 9895-9899.

Guzman, L.M., Belin, D., Carson, M.J. and Beckwith, J. (1995) Tight regulation,

modulation, and high-level expression by vectors containing the arabinose PBAD

promoter. J Bacteriol, 177, 4121-4130.

Haebel, P.W., Goldstone, D., Katzen, F., Beckwith, J. and Metcalf, P. (2002) The

disulfide bond isomerase DsbC is activated by an immunoglobulin-fold thiol

oxidoreductase: crystal structure of the DsbC-DsbDalpha complex. Embo J, 21,

4774-4784.

Hatahet, F. and Ruddock, L.W. (2007) Substrate recognition by the protein disulfide

isomerases. Febs J, 274, 5223-5234.

Hawkins, H.C., Blackburn, E.C. and Freedman, R.B. (1991) Comparison of the activities

of protein disulphide-isomerase and thioredoxin in catalysing disulphide

isomerization in a protein substrate. Biochem J, 275 (Pt 2), 349-353.

133

Page 149: Copyright by Silvia A. Arredondo 2009

Hiniker, A. and Bardwell, J.C. (2004) In vivo substrate specificity of periplasmic

disulfide oxidoreductases. J Biol Chem, 279, 12967-12973.

Hiniker, A., Collet, J.F. and Bardwell, J.C. (2005) Copper stress causes an in vivo

requirement for the Escherichia coli disulfide isomerase DsbC. J Biol Chem, 280,

33785-33791.

Hogg, P.J. (2003) Disulfide bonds as switches for protein function. Trends Biochem Sci,

28, 210-214.

Holmgren, A. (1979a) Reduction of disulfides by thioredoxin. Exceptional reactivity of

insulin and suggested functions of thioredoxin in mechanism of hormone action. J

Biol Chem, 254, 9113-9119.

Holmgren, A. (1979b) Thioredoxin catalyzes the reduction of insulin disulfides by

dithiothreitol and dihydrolipoamide. J Biol Chem, 254, 9627-9632.

Huber-Wunderlich, M. and Glockshuber, R. (1998) A single dipeptide sequence

modulates the redox properties of a whole enzyme family. Fold Des, 3, 161-171.

Huston, J.S., Levinson, D., Mudgett-Hunter, M., Tai, M.S., Novotny, J., Margolies, M.N.,

Ridge, R.J., Bruccoleri, R.E., Haber, E., Crea, R. and et al. (1988) Protein

engineering of antibody binding sites: recovery of specific activity in an anti-

digoxin single-chain Fv analogue produced in Escherichia coli. Proc Natl Acad

Sci U S A, 85, 5879-5883.

Inaba, K., Murakami, S., Suzuki, M., Nakagawa, A., Yamashita, E., Okada, K. and Ito,

K. (2006) Crystal structure of the DsbB-DsbA complex reveals a mechanism of

disulfide bond generation. Cell, 127, 789-801.

Ito, K. and Inaba, K. (2008) The disulfide bond formation (Dsb) system. Curr Opin

Struct Biol, 18, 450-458.

Joly, J.C. and Swartz, J.R. (1997) In vitro and in vivo redox states of the Escherichia coli

periplasmic oxidoreductases DsbA and DsbC. Biochemistry, 36, 10067-10072.

Jonda, S., Huber-Wunderlich, M., Glockshuber, R. and Mossner, E. (1999)

Complementation of DsbA deficiency with secreted thioredoxin variants reveals

134

Page 150: Copyright by Silvia A. Arredondo 2009

the crucial role of an efficient dithiol oxidant for catalyzed protein folding in the

bacterial periplasm. Embo J, 18, 3271-3281.

Kadokura, H., Katzen, F. and Beckwith, J. (2003) Protein disulfide bond formation in

prokaryotes. Annu Rev Biochem, 72, 111-135.

Kamitani, S., Akiyama, Y. and Ito, K. (1992) Identification and characterization of an

Escherichia coli gene required for the formation of correctly folded alkaline

phosphatase, a periplasmic enzyme. Embo J, 11, 57-62.

Ke, H., Zhang, S., Li, J., Howlett, G.J. and Wang, C.C. (2006) Folding of Escherichia

coli DsbC: characterization of a monomeric folding intermediate. Biochemistry,

45, 15100-15110.

Kishigami, S., Akiyama, Y. and Ito, K. (1995a) Redox states of DsbA in the periplasm of

Escherichia coli. FEBS Lett, 364, 55-58.

Kishigami, S., Kanaya, E., Kikuchi, M. and Ito, K. (1995b) DsbA-DsbB interaction

through their active site cysteines. Evidence from an odd cysteine mutant of

DsbA. J Biol Chem, 270, 17072-17074.

Klappa, P., Ruddock, L.W., Darby, N.J. and Freedman, R.B. (1998) The b' domain

provides the principal peptide-binding site of protein disulfide isomerase but all

domains contribute to binding of misfolded proteins. Embo J, 17, 927-935.

Kolaj, O., Spada, S., Robin, S. and Wall, J.G. (2009) Use of folding modulators to

improve heterologous protein production in Escherichia coli. Microb Cell Fact, 8,

9.

Kortt, A.A., Malby, R.L., Caldwell, J.B., Gruen, L.C., Ivancic, N., Lawrence, M.C.,

Howlett, G.J., Webster, R.G., Hudson, P.J. and Colman, P.M. (1994)

Recombinant anti-sialidase single-chain variable fragment antibody.

Characterization, formation of dimer and higher-molecular-mass multimers and

the solution of the crystal structure of the single-chain variable fragment/sialidase

complex. Eur J Biochem, 221, 151-157.

135

Page 151: Copyright by Silvia A. Arredondo 2009

Krause, G., Lundstrom, J., Barea, J.L., Pueyo de la Cuesta, C. and Holmgren, A. (1991)

Mimicking the active site of protein disulfide-isomerase by substitution of proline

34 in Escherichia coli thioredoxin. J Biol Chem, 266, 9494-9500.

Kulp, M.S., Frickel, E.M., Ellgaard, L. and Weissman, J.S. (2006) Domain architecture

of protein-disulfide isomerase facilitates its dual role as an oxidase and an

isomerase in Ero1p-mediated disulfide formation. J Biol Chem, 281, 876-884.

Kurokawa, Y., Yanagi, H. and Yura, T. (2001) Overproduction of bacterial protein

disulfide isomerase (DsbC) and its modulator (DsbD) markedly enhances

periplasmic production of human nerve growth factor in Escherichia coli. J Biol

Chem, 276, 14393-14399.

Lafond, R., Zhan, X. and Georgiou, G. (2003) Screening and selection strategies for

disulfide isomerase activity. Methods Mol Biol, 230, 239-257.

Leader, B., Baca, Q.J. and Golan, D.E. (2008) Protein therapeutics: a summary and

pharmacological classification. Nat Rev Drug Discov, 7, 21-39.

Liang, H., Sandberg, W.S. and Terwilliger, T.C. (1993) Genetic fusion of subunits of a

dimeric protein substantially enhances its stability and rate of folding. Proc Natl

Acad Sci U S A, 90, 7010-7014.

Liu, X. and Wang, C.C. (2001) Disulfide-dependent folding and export of Escherichia

coli DsbC. J Biol Chem, 276, 1146-1151.

Lundstrom-Ljung, J. and Holmgren, A. (1995) Glutaredoxin accelerates glutathione-

dependent folding of reduced ribonuclease A together with protein disulfide-

isomerase. J Biol Chem, 270, 7822-7828.

Lundstrom, J., Krause, G. and Holmgren, A. (1992) A Pro to His mutation in active site

of thioredoxin increases its disulfide-isomerase activity 10-fold. New refolding

systems for reduced or randomly oxidized ribonuclease. J Biol Chem, 267, 9047-

9052.

Lyles, M.M. and Gilbert, H.F. (1991) Catalysis of the oxidative folding of ribonuclease A

by protein disulfide isomerase: dependence of the rate on the composition of the

redox buffer. Biochemistry, 30, 613-619.

136

Page 152: Copyright by Silvia A. Arredondo 2009

Lyles, M.M. and Gilbert, H.F. (1994) Mutations in the thioredoxin sites of protein

disulfide isomerase reveal functional nonequivalence of the N- and C-terminal

domains. J Biol Chem, 269, 30946-30952.

Mamathambika, B.S. and Bardwell, J.C. (2008) Disulfide-linked protein folding

pathways. Annu Rev Cell Dev Biol, 24, 211-235.

Martin, J.L. (1995) Thioredoxin - A Fold For All Reasons. Structure, 3, 245-250.

Martin, J.L., Bardwell, J.C. and Kuriyan, J. (1993) Crystal structure of the DsbA protein

required for disulphide bond formation in vivo. Nature, 365, 464-468.

Martinez-Galisteo, E., Padilla, C.A., Garcia-Alfonso, C., Lopez-Barea, J. and Barcena,

J.A. (1993) Purification and properties of bovine thioredoxin system. Biochimie,

75, 803-809.

McCarthy, A.A., Haebel, P.W., Torronen, A., Rybin, V., Baker, E.N. and Metcalf, P.

(2000) Crystal structure of the protein disulfide bond isomerase, DsbC, from

Escherichia coli. Nat Struct Biol, 7, 196-199.

Mecsas, J., Welch, R., Erickson, J.W. and Gross, C.A. (1995) Identification and

characterization of an outer membrane protein, OmpX, in Escherichia coli that is

homologous to a family of outer membrane proteins including Ail of Yersinia

enterocolitica. J Bacteriol, 177, 799-804.

Merrifield, R.B. (1964) Solid-Phase Peptide Synthesis. 3. An Improved Synthesis Of

Bradykinin. Biochemistry, 3, 1385-1390.

Messens, J. and Collet, J.F. (2006) Pathways of disulfide bond formation in Escherichia

coli. Int J Biochem Cell Biol, 38, 1050-1062.

Missiakas, D., Georgopoulos, C. and Raina, S. (1994) The Escherichia coli dsbC (xprA)

gene encodes a periplasmic protein involved in disulfide bond formation. Embo J,

13, 2013-2020.

Natale, P., Bruser, T. and Driessen, A.J. (2008) Sec- and Tat-mediated protein secretion

across the bacterial cytoplasmic membrane--distinct translocases and

mechanisms. Biochim Biophys Acta, 1778, 1735-1756.

137

Page 153: Copyright by Silvia A. Arredondo 2009

Ostermeier, M., De Sutter, K. and Georgiou, G. (1996) Eukaryotic protein disulfide

isomerase complements Escherichia coli dsbA mutants and increases the yield of

a heterologous secreted protein with disulfide bonds. J Biol Chem, 271, 10616-

10622.

Pack, P. and Pluckthun, A. (1992) Miniantibodies: use of amphipathic helices to produce

functional, flexibly linked dimeric FV fragments with high avidity in Escherichia

coli. Biochemistry, 31, 1579-1584.

Pavlou, A.K. and Reichert, J.M. (2004) Recombinant protein therapeutics--success rates,

market trends and values to 2010. Nat Biotechnol, 22, 1513-1519.

Pirneskoski, A., Klappa, P., Lobell, M., Williamson, R.A., Byrne, L., Alanen, H.I., Salo,

K.E., Kivirikko, K.I., Freedman, R.B. and Ruddock, L.W. (2004) Molecular

characterization of the principal substrate binding site of the ubiquitous folding

catalyst protein disulfide isomerase. J Biol Chem, 279, 10374-10381.

Predki, P.F. and Regan, L. (1995) Redesigning the topology of a four-helix-bundle

protein: monomeric Rop. Biochemistry, 34, 9834-9839.

Qiu, J., Swartz, J.R. and Georgiou, G. (1998) Expression of active human tissue-type

plasminogen activator in Escherichia coli. Appl Environ Microbiol, 64, 4891-

4896.

Ramm, K. and Pluckthun, A. (2000) The periplasmic Escherichia coli peptidylprolyl

cis,trans-isomerase FkpA. II. Isomerase-independent chaperone activity in vitro. J

Biol Chem, 275, 17106-17113.

Ren, G., Stephan, D., Xu, Z., Zheng, Y., Tang, D., Harrison, R.S., Kurz, M., Jarrott, R.,

Shouldice, S.R., Hiniker, A., Martin, J.L., Heras, B. and Bardwell, J.C. (2009)

Properties of the thioredoxin-fold superfamily are modulated by a single amino

acid residue. J Biol Chem.

Rice, J.J. and Daugherty, P.S. (2008) Directed evolution of a biterminal bacterial display

scaffold enhances the display of diverse peptides. Protein Eng Des Sel, 21, 435-

442.

138

Page 154: Copyright by Silvia A. Arredondo 2009

Rice, J.J., Schohn, A., Bessette, P.H., Boulware, K.T. and Daugherty, P.S. (2006)

Bacterial display using circularly permuted outer membrane protein OmpX yields

high affinity peptide ligands. Protein Sci, 15, 825-836.

Rietsch, A., Belin, D., Martin, N. and Beckwith, J. (1996) An in vivo pathway for

disulfide bond isomerization in Escherichia coli. Proc Natl Acad Sci U S A, 93,

13048-13053.

Rietsch, A., Bessette, P., Georgiou, G. and Beckwith, J. (1997) Reduction of the

periplasmic disulfide bond isomerase, DsbC, occurs by passage of electrons from

cytoplasmic thioredoxin. J Bacteriol, 179, 6602-6608.

Ritz, D. and Beckwith, J. (2001) Roles of thiol-redox pathways in bacteria. Annu Rev

Microbiol, 55, 21-48.

Robinson, C.R. and Sauer, R.T. (1996) Covalent attachment of Arc repressor subunits by

a peptide linker enhances affinity for operator DNA. Biochemistry, 35, 109-116.

Robinson, C.R. and Sauer, R.T. (1998) Optimizing the stability of single-chain proteins

by linker length and composition mutagenesis. Proc Natl Acad Sci U S A, 95,

5929-5934.

Rozhkova, A., Stirnimann, C.U., Frei, P., Grauschopf, U., Brunisholz, R., Grutter, M.G.,

Capitani, G. and Glockshuber, R. (2004) Structural basis and kinetics of inter- and

intramolecular disulfide exchange in the redox catalyst DsbD. Embo J, 23, 1709-

1719.

Ruddock, L.W., Freedman, R.B. and Klappa, P. (2000) Specificity in substrate binding

by protein folding catalysts: tyrosine and tryptophan residues are the recognition

motifs for the binding of peptides to the pancreas-specific protein disulfide

isomerase PDIp. Protein Sci, 9, 758-764.

Ruoppolo, M., Orru, S., Talamo, F., Ljung, J., Pirneskoski, A., Kivirikko, K.I., Marino,

G. and Koivunen, P. (2003) Mutations in domain a' of protein disulfide isomerase

affect the folding pathway of bovine pancreatic ribonuclease A. Protein Sci, 12,

939-952.

139

Page 155: Copyright by Silvia A. Arredondo 2009

Rusch, S.L. and Kendall, D.A. (2007) Interactions that drive Sec-dependent bacterial

protein transport. Biochemistry, 46, 9665-9673.

Saul, F.A., Arie, J.P., Vulliez-le Normand, B., Kahn, R., Betton, J.M. and Bentley, G.A.

(2004) Structural and functional studies of FkpA from Escherichia coli, a cis/trans

peptidyl-prolyl isomerase with chaperone activity. J Mol Biol, 335, 595-608.

Schmidt, F.R. (2004) Recombinant expression systems in the pharmaceutical industry.

Appl Microbiol Biotechnol, 65, 363-372.

Segatori, L., Murphy, L., Arredondo, S., Kadokura, H., Gilbert, H., Beckwith, J. and

Georgiou, G. (2006) Conserved role of the linker alpha-helix of the bacterial

disulfide isomerase DsbC in the avoidance of misoxidation by DsbB. J Biol

Chem, 281, 4911-4919.

Segatori, L., Paukstelis, P.J., Gilbert, H.F. and Georgiou, G. (2004) Engineered DsbC

chimeras catalyze both protein oxidation and disulfide-bond isomerization in

Escherichia coli: Reconciling two competing pathways. Proc Natl Acad Sci U S

A, 101, 10018-10023.

Shevchik, V.E., Condemine, G. and Robert-Baudouy, J. (1994) Characterization of

DsbC, a periplasmic protein of Erwinia chrysanthemi and Escherichia coli with

disulfide isomerase activity. Embo J, 13, 2007-2012.

Sun, X.X. and Wang, C.C. (2000) The N-terminal sequence (residues 1-65) is essential

for dimerization, activities, and peptide binding of Escherichia coli DsbC. J Biol

Chem, 275, 22743-22749.

Tian, G., Xiang, S., Noiva, R., Lennarz, W.J. and Schindelin, H. (2006) The crystal

structure of yeast protein disulfide isomerase suggests cooperativity between its

active sites. Cell, 124, 61-73.

Todorovska, A., Roovers, R.C., Dolezal, O., Kortt, A.A., Hoogenboom, H.R. and

Hudson, P.J. (2001) Design and application of diabodies, triabodies and

tetrabodies for cancer targeting. J Immunol Methods, 248, 47-66.

Travaglini-Allocatelli, C., Ivarsson, Y., Jemth, P. and Gianni, S. (2009) Folding and

stability of globular proteins and implications for function. Curr Opin Struct Biol.

140

Page 156: Copyright by Silvia A. Arredondo 2009

Tucker, J.B. and Hooper, C. (2006) Protein engineering: security implications. The

increasing ability to manipulate protein toxins for hostile purposes has prompted

calls for regulation. EMBO Rep, 7 Spec No, S14-17.

van Zonneveld, A.J., Veerman, H., MacDonald, M.E., van Mourik, J.A. and Pannekoek,

H. (1986) Structure and function of human tissue-type plasminogen activator (t-

PA). J Cell Biochem, 32, 169-178.

Vertommen, D., Depuydt, M., Pan, J., Leverrier, P., Knoops, L., Szikora, J.P., Messens,

J., Bardwell, J.C. and Collet, J.F. (2008) The disulphide isomerase DsbC

cooperates with the oxidase DsbA in a DsbD-independent manner. Mol

Microbiol, 67, 336-349.

Vlamis-Gardikas, A. (2008) The multiple functions of the thiol-based electron flow

pathways of Escherichia coli: Eternal concepts revisited. Biochim Biophys Acta,

1780, 1170-1200.

Vuori, K., Myllyla, R., Pihlajaniemi, T. and Kivirikko, K.I. (1992) Expression and site-

directed mutagenesis of human protein disulfide isomerase in Escherichia coli.

This multifunctional polypeptide has two independently acting catalytic sites for

the isomerase activity. J Biol Chem, 267, 7211-7214.

Walker, K.W., Lyles, M.M. and Gilbert, H.F. (1996) Catalysis of oxidative protein

folding by mutants of protein disulfide isomerase with a single active-site

cysteine. Biochemistry, 35, 1972-1980.

Wetzel, R. (1986) Investigation of the Structural Roles of Disulfides by Protein

Engineering: a study of T4 Lysozyme. In Sarma, M.I.a.R. (ed.), Protein

Engieering. Applications in Science, Medicine and Industry. Academic Press,

Inc., Orlando, Vol. 1, pp. 257-274.

Wilkinson, B. and Gilbert, H.F. (2004) Protein disulfide isomerase. Biochim Biophys

Acta, 1699, 35-44.

Yeh, S.M., Koon, N., Squire, C. and Metcalf, P. (2007) Structures of the dimerization

domains of the Escherichia coli disulfide-bond isomerase enzymes DsbC and

DsbG. Acta Crystallogr D Biol Crystallogr, 63, 465-471.

141

Page 157: Copyright by Silvia A. Arredondo 2009

Young, J.C., Agashe, V.R., Siegers, K. and Hartl, F.U. (2004) Pathways of chaperone-

mediated protein folding in the cytosol. Nat Rev Mol Cell Biol, 5, 781-791.

Zapun, A., Bardwell, J.C. and Creighton, T.E. (1993) The reactive and destabilizing

disulfide bond of DsbA, a protein required for protein disulfide bond formation in

vivo. Biochemistry, 32, 5083-5092.

Zapun, A., Missiakas, D., Raina, S. and Creighton, T.E. (1995) Structural and functional

characterization of DsbC, a protein involved in disulfide bond formation in

Escherichia coli. Biochemistry, 34, 5075-5089.

Zhang, Z., Li, Z.H., Wang, F., Fang, M., Yin, C.C., Zhou, Z.Y., Lin, Q. and Huang, H.L.

(2002) Overexpression of DsbC and DsbG markedly improves soluble and

functional expression of single-chain Fv antibodies in Escherichia coli. Protein

Expr Purif, 26, 218-228.

Zhao, Z., Peng, Y., Hao, S.F., Zeng, Z.H. and Wang, C.C. (2003) Dimerization by

domain hybridization bestows chaperone and isomerase activities. J Biol Chem,

278, 43292-43298.

142

Page 158: Copyright by Silvia A. Arredondo 2009

Vita

Silvia A. Arredondo was born in Guatemala City, Guatemala, the daughter of

Marco Antonio Arredondo Montúfar and Carmen Escobar de Arredondo. She received

her high school education at the Instituto Belga Guatemalteco, Guatemala City,

Guatemala, where she obtained a teaching degree. Soon after, while working as an

elementary school teacher, she attended the school of engineering of the Universidad de

San Carlos de Guatemala as a part-time student. However, in 1999, she moved to New

York, New York, and attended the Borough of Manhattan Community College for a year.

Then, she transferred to the City College of New York, CUNY, where she graduated as

the 2003 valedictorian, receiving the degree of Bachelor of Engineering in Chemical

Engineering. In the Fall 2003, she entered the Graduate School of The University of

Texas at Austin to pursue a Doctorate degree in Chemical Engineering.

Permanent Address: 25 ave. 25-11 zona 5, Ciudad de Guatemala

This manuscript was typed by the author.

143