Download Mechanisms and applications of disulfide bond formation

Document related concepts

Ancestral sequence reconstruction wikipedia , lookup

Evolution of metal ions in biological systems wikipedia , lookup

Biochemistry wikipedia , lookup

Gene expression wikipedia , lookup

SR protein wikipedia , lookup

Biochemical cascade wikipedia , lookup

G protein–coupled receptor wikipedia , lookup

Magnesium transporter wikipedia , lookup

Expression vector wikipedia , lookup

Paracrine signalling wikipedia , lookup

Signal transduction wikipedia , lookup

Bimolecular fluorescence complementation wikipedia , lookup

Protein wikipedia , lookup

Interactome wikipedia , lookup

QPNC-PAGE wikipedia , lookup

Protein structure prediction wikipedia , lookup

Protein purification wikipedia , lookup

Metalloprotein wikipedia , lookup

Nuclear magnetic resonance spectroscopy of proteins wikipedia , lookup

Western blot wikipedia , lookup

Protein–protein interaction wikipedia , lookup

Two-hybrid screening wikipedia , lookup

Proteolysis wikipedia , lookup

Transcript
D 1282
OULU 2015
UNIVERSITY OF OUL U P.O. Box 8000 FI-90014 UNIVERSITY OF OULU FINLA ND
U N I V E R S I TAT I S
O U L U E N S I S
ACTA
A C TA
D 1282
ACTA
U N I V E R S I T AT I S O U L U E N S I S
Van Dat Nguyen
University Lecturer Santeri Palviainen
Postdoctoral research fellow Sanna Taskila
Van Dat Nguyen
Professor Esa Hohtola
MECHANISMS AND
APPLICATIONS OF DISULFIDE
BOND FORMATION
Professor Olli Vuolteenaho
University Lecturer Veli-Matti Ulvinen
Director Sinikka Eskelinen
Professor Jari Juga
University Lecturer Anu Soikkeli
Professor Olli Vuolteenaho
Publications Editor Kirsti Nurkkala
ISBN 978-952-62-0724-7 (Paperback)
ISBN 978-952-62-0725-4 (PDF)
ISSN 0355-3221 (Print)
ISSN 1796-2234 (Online)
UNIVERSITY OF OULU GRADUATE SCHOOL;
UNIVERSITY OF OULU,
FACULTY OF BIOCHEMISTRY AND MOLECULAR MEDICINE;
BIOCENTER OULU
D
MEDICA
ACTA UNIVERSITATIS OULUENSIS
D Medica 1282
VAN DAT NGUYEN
MECHANISMS AND APPLICATIONS
OF DISULFIDE BOND FORMATION
Academic dissertation to be presented with the assent
of the Doctoral Training Committee of Health and
Biosciences of the University of Oulu for public defence
in Auditorium F101 of the Department of Physiology
(Aapistie 7), on 6 February 2015, at 12 noon.
U N I VE R S I T Y O F O U L U , O U L U 2 0 1 5
Copyright © 2015
Acta Univ. Oul. D 1282, 2015
Supervised by
Professor Lloyd Ruddock
Reviewed by
Professor Neil Bulleid
Doctor Renaud Vincentelli
Opponent
Professor Jan Riemer
ISBN 978-952-62-0724-7 (Paperback)
ISBN 978-952-62-0725-4 (PDF)
ISSN 0355-3221 (Printed)
ISSN 1796-2234 (Online)
Cover Design
Raimo Ahonen
JUVENES PRINT
TAMPERE 2015
Nguyen, Van Dat, Mechanisms and applications of disulfide bond formation.
University of Oulu Graduate School; University of Oulu, Faculty of Biochemistry and
Molecular Medicine; Biocenter Oulu
Acta Univ. Oul. D 1282, 2015
University of Oulu, P.O. Box 8000, FI-90014 University of Oulu, Finland
Abstract
About one-third of mammalian proteins are secreted proteins and membrane proteins. Most of
these proteins contain disulfide bonds in their native state, covalent links formed between the thiol
groups of cysteine residues. In many proteins, disulfide bonds play an essential role in folding,
stabilizing structure and the function of the protein. Therefore, understanding the pathways of
disulfide bond formation is crucial for a wide range of medical processes and therapies. Disulfide
bond formation is catalyzed by the Protein Disulfide Isomerase (PDI) family. To date the
mechanisms of the PDIs in disulfide bond formation and pathways for disulfide bond formation
have not been fully characterized.
Here the structure of the substrate binding b’x domain of human PDI was determined. The
structure shows that the b' domain has a typical thioredoxin fold and that the x region can interact
with the substrate binding site of the b' domain. Specifically, the x region of PDI can adopt
alternative conformations during the functional cycle of PDI action and that these are linked to the
ability of PDI to interact with folding substrates.
In addition, this study showed that two human proteins, GPx7 and GPx8 are involved in
disulfide bond formation. The addition of GPx7 or GPx8 to a folding protein along with PDI and
peroxide allows the efficient oxidative refolding of a reduced denatured substrate protein.
Finally, this thesis includes the development of a system for the efficient production of
disulfide bond containing proteins in the cytoplasm of E. coli. It showed that the introduction of
Erv1p, a sulfhydryl oxidase and FAD-dependent catalyst of disulfide bond, allows the formation
of native disulfide bonds in the cytoplasm of E. coli even without the disruption of genes involved
in disulfide bond reduction. Introduction of Erv1p and a disulfide isomerase, e.g. PDI, allows the
efficient formation of natively folded eukaryotic proteins with multiple disulfide bonds in the
cytoplasm of E. coli. This system is able to express high levels of complex disulfide bonded
eukaryotic proteins.
Keywords: disulfide bond formation, Erv1p, GPx7, GPx8, hydrogen peroxide and
protein disulfide isomerase
Nguyen, Van Dat, Disulfidisidosten muodostumisen mekanismit ja sovellukset.
Oulun yliopiston tutkijakoulu; Oulun yliopisto, Biokemian ja molekyylilääketieteen tiedekunta;
Biocenter Oulu
Acta Univ. Oul. D 1282, 2015
Oulun yliopisto, PL 8000, 90014 Oulun yliopisto
Tiivistelmä
Noin kolmasosa kaikista nisäkkäiden proteiineista on solun ulkopuolelle eritettäviä proteiineja ja
kalvoproteiineja. Monet näistä proteiineista sisältävät natiivissa konformaatiossaan disulfidisidoksia, jotka ovat kovalenttisia sidoksia kysteiinitähteiden tioliryhmien välillä. Useissa proteiineissa näillä disulfidisidoksilla on keskeinen rooli proteiinin laskostumisessa, kolmiulotteisen
rakenteen stabiloinnissa sekä proteiinin toiminnassa. Disulfidisidosten muodostumisen taustalla
olevien mekanismien tunteminen onkin tärkeää monien lääketieteellisten prosessien ja hoitomenetelmien kannalta. Disulfidisidosten muodostumista katalysoivat proteiinidisulfidi-isomeraasi
(PDI) -perheeseen kuuluvat entsyymit. PDI entsyymien toimintamekanismeja ja disulfidisidosten muodostumisen reaktioreittejä ei kuitenkaan vielä tunneta tarkasti.
Tässä väitöskirjassa selvitettiin ihmisen PDI entsyymin substraattia sitovan b’x alayksikön
rakenne. Rakenteesta voidaan todeta b’ alayksikön laskostuminen tyypilliseen tioredoksiini
muotoon sekä x alueen interaktio b’ alayksikön substraattia sitovan kohdan kanssa. PDI entsyymin katalysoiman reaktioketjun aikana x alayksikkö voi muuttaa konformaatiotaan mahdollistaen PDI entsyymin interaktion laskostuvien substraattiproteiinien kanssa.
Tässä tutkimuksessa osoitettiin myös kahden ihmisen proteiinin, GPx7 ja GPx8 osallistuminen disulfidisidosten muodostumista katalysoiviin reaktioihin. GPx7 ja GPx8 entsyymien lisäys
laskostumisreaktioon yhdessä PDI:n ja vetyperoksidin kanssa mahdollistaa pelkistetyn, denaturoidun substraattiproteiinin tehokkaan, hapettaviin reaktioihin perustuvan uudelleenlaskostumisen natiiviin muotoonsa.
Osana tätä väitöstutkimusta kehitettiin menetelmä, joka mahdollistaa disulfideja sisältävien
proteiinien tehokkaan tuoton E.colin solulimassa. Menetelmässä sulfhydryylioksidaasina ja
FAD:sta riippuvana disulfidisidosten muodostumisen katalysaattorina toimiva Erv1p mahdollistaa disulfidisidosten muodostumisen E.colin solulimassa myös ilman solun pelkistävien reaktioreittien geneettistä poistamista. Erv1p yhdessä disulfidi-isomeraasin, kuten PDI, kanssa mahdollistaa oikein laskostuneiden, useita disulfidisidoksia sisältävien eukaryoottisten proteiinien
tehokkaan tuotannon E.colin solulimassa. Menetelmällä pystytään tuottamaan suuria määriä
monimutkaisia disulfidisidoksellisia proteiineja.
Asiasanat: disulfidisidos,
vetyperoksidi
Erv1p,
GPx7,
Gpx8,
proteiinidisulfidi-isomeraasi,
To my family
8
Acknowledgements
The present thesis was carried in the research group of Professor Lloyd Ruddock,
Biocenter Oulu, Faculty of Biochemistry and Molecular Medicine, University of
Oulu and was supported by Academy of Finland, Department of Biochemistry,
Biocenter Oulu, Oulu University Scholarship Foundation, University of Oulu.
First of all, I would like to express my sincere gratitude to my supervisor
Professor Lloyd Ruddock for giving me the great opportunity to learn and explore
science in his research group. His outstanding supervision and his endless
supports and encouragements were crucial to the progress of this thesis. I would
like to thank Professor Kalervo Hiltunen, the head of the department for
providing well facilitated research environment.
All members of Department of Biochemistry are warmly acknowledged for
creating such a good working atmosphere at the department. Special thanks to
Professor Prof. Rikkert K. Wierenga and Dr. Antti Haapalainen for teaching me
protein crystallography. I also would like to thank Professor Lars Ellgaard who
was my external supervisor. I would like to thank the co-authors of the
publications that form part of this thesis. Many thanks to my thesis committee
group members: Professor Olli Vuolteenaho, Dr. Ilya Skovorodkin, Dr. Jingdong
Shan, Dr. Jussi Vuoristo, Dr. Pekka Kilpeläinen.
I would like thank to the current and former members of the LR-group: Kirsi
Salo, Dr. Anna-Kaisa Lappi, Anna-Riikka Karala, Dr. Mirva Saaranen, Heli
Alanen, Dr. Feras Hatahet, Dr. Annamari Ruddock, Dr. Irina Raykhel, Kati
Korhonen, Marja Luukas, Herkko Sikkila for being very good friends and
colleagues. Special thanks to Anna-Kaisa Lappi and Kirsi Salo for their
encouragements and for pushing me forward with thesis writing. I would like
thank to Dr. Emmy Verschuren and Dr. Liisa Kauppi for their encouragements
during thesis writing.
I am deeply grateful to my parents, my parents in law, my brother and my
sister to their love and encouragements.
Finally, My deepest gratitude I want to express to my wife Quynh and our
beloved sons Truong, Nhat, Quoc.
Helsinki, December 2014
Van Dat Nguyen
9
10
Abbreviations
aa
ACN
Arg
Asp
BPTI
CCD
Cys
D
DHA
DNA
DNTB
DTT
EDTA
ER
FAD
Glu
Grx
GSH
GSSG
IAA
kDa
LB
Lys
M
MIMS
MHC
NNADPH
PCR
PDB
PDI
Q
R
ROS
Amino acid
Acetonitrile
Arginine
Aspartic acid
Bovine pancreatic trypsin inhibitor
Carboxy (terminal)Circular dichroism
Cysteine
Aspartic acid
Dehydroascorbate
Deoxyribonucleic acid
5,5-dithio-bis(2-nitrobenzoic acid)
Dithiothreitol
Ethylenediamidetetraacetic acid
Endoplasmic reticulum
Flavin adenine nucleotide
Glutamic acid
Glutaredoxin
Reduced glutathione
Oxidized glutathione
Iodoacetamide
Kilo Dalton
Luria-Bertani
Lysine
Methionine
Mitochondrial intermembrane space
Major histocompatibility complex
Amino (terminal)Nicotinamide adenine dinucleotide phosphate
Polymerase chain reaction
Protein data bank
Protein disulfide isomerase
Glutamine
Arginine
Reactive oxygen species
11
S
SDS-PAGE
Trx
UV
V
W
X
12
Serine
Sodium dodecyl sulphate polyacrylamide gel electrophoresis
Thioredoxin
Ultraviolet
Voltage
Tryptophan
Any amino acid
List of original publications
This thesis is based on the following publications, which are referred throughout
the text by their Roman numerals:
I
Nguyen VD*, Wallis K*, Howard MJ, Haapalainen AM, Salo KE, Saaranen MJ,
Sidhu A, Wierenga RK, Freedman RB, Ruddock LW, Williamson RA (2008)
Alternative conformations of the x region of human protein disulphide-isomerase
modulate exposure of the substrate binding b' domain. J Mol Biol 383(5):1144-55.
II Nguyen VD, Saaranen MJ, Karala AR, Lappi AK, Wang L, Raykhel IB, Alanen HI,
Salo KE, Wang CC, Ruddock LW (2011) Two endoplasmic reticulum PDI
peroxidases increase the efficiency of the use of peroxide during disulfide bond
formation. J Mol Biol 406(3):503-15.
III Hatahet F, Nguyen VD, Salo KE, Ruddock LW (2010) Disruption of reducing
pathways is not essential for efficient disulfide bond formation in the cytoplasm of E.
coli. Microb Cell Fact 9(67):1-9.
IV Nguyen VD, Hatahet F, Salo KE, Enlund E, Zhang C, Ruddock LW (2011) Preexpression of a sulfhydryl oxidase significantly increases the yields of eukaryotic
disulfide bond containing proteins expressed in the cytoplasm of E. coli. Microb Cell
Fact 10(1):1-13.
* equal contribution
13
14
Contents
Abstract
Tiivistelmä
Acknowledgements
9 Abbreviations
11 List of original publications
13 Contents
15 1 Introduction
17 2 Review of the literature
19 2.1 Protein folding......................................................................................... 19 2.2 Oxidative folding of proteins .................................................................. 21 2.3 Disulfide bond formation in the endoplasmic reticulum ......................... 22 2.3.1 Structure and function of human protein disulfide
isomerase family members ........................................................... 23 2.3.2 Reactions of protein disulfide isomerase ...................................... 29 2.3.3 Disulfide bond generation in the endoplasmic reticulum ............. 31 2.4 Disulfide bond formation in the mitochondrial inter-membrane
space ........................................................................................................ 37 2.5 Disulfide bond formation in the prokaryotic periplasm .......................... 39 2.6 Disulfide bond formation in the engineered cytoplasm of E. coli ........... 41 3 Aims of the present work
43 4 Materials and methods
45 5 Results
47 6 Discussion
49 6.1 Structure of the substrate binding domain of human protein
disulfide isomerase and the functional significance of a novel
structural interaction between the b’ domain and the x linker ................ 49 6.2 GPx7 and GPx8 are involved in disulfide bond formation in the
endoplasmic reticulum ............................................................................ 51 6.3 Applications of disulfide bond formation ............................................... 54 7 Conclusions
57 References
59 Original articles
75 15
16
1
Introduction
Disulfide bonds are covalent linkages formed between two cysteine residues in
proteins as a result of oxidation reactions. Disulfide bonds are a crucial post
translational modification for most secreted proteins and outer membrane
proteins. Most of them require disulfide bonds for their stability, folding, and
function. Catalysis of disulfide bond formation naturally occurs in the
endoplasmic reticulum (ER) of eurokaryotes, the inter-membrane space of
mitochrondria, and the periplasm of prokaryotes. These compartments have
optimal environments, catalysts, and oxidation pathways for folding of native
disulfide bond-containing proteins. Broadly speaking two steps are involved in
native disulfide bond formations: i) oxidation reactions to form disulfide bonds
and ii) subsequent rearrangement or isomerization of incorrect or non-native
disulfide bonds to obtain native disulfide bonds.
Native disulfide bond formation in human cells is catalyzed by the PDIfamily of proteins which contains 20 members. The archetypal member, PDI
(PDIA1), is the most studied member in the family. It is a multifunctional enzyme
whose most well-characterized function is as a catalyst of native disulfide bond
formation. PDI contain four domains plus two short regions, whose architecture is
defined: a-b-b’-x-a’-c. a and a’ are catalytic domains containing active sites, the
b and b’ domains lack active sites and are non-catalytic, x is a linker region and c
is a C-terminal acidic tail. The b’ domain is essential for substrate binding of PDI.
It is needed for catalysis of protein disulfide isomerization and for non-covalent
ligand binding.
Ero1 proteins are also key factors in disulfide bond formation pathways in the
ER. Ero1 works in conjunction with PDI and generates hydrogen peroxide as a
by-product when producing a disulfide bond. The hydrogen peroxide produced
could become a toxic product to the cell if it accumulates. However, potentially
hydrogen peroxide can be used to generate disulfide bonds. Hence possibly, Ero1
can generate two disulfide bonds in each catalytic cycle. The question is what is
the pathway in the ER that can convert hydrogen peroxide to disulfide bonds most
efficiently?
Native disulfide bond formation is often the rate-limiting step in protein
folding in vitro and in vivo. Hence the production of recombinant disulfide
bonded proteins is expensive on an industrial scale.
In the present study a number of parallel approaches were adopted to
understand the mechanisms of disulfide bond formation in the ER and to apply
17
knowledge of natural mechanisms for the efficient production of disulfide bond
containing proteins in the cytoplasm of prokaryotes. Specifically, to examine in
more detail the mechanism of action of PDI in catalyzing folding of disulfide
bond-containing proteins a fluorescence screening of 75 mutants of the b’x
domain was carried out and mutants that stabilize the b’x domain in different
conformational forms were identified. The crystal structure of the I272A mutant
of b’x domain was reported in this study, a structure that showed the x linker caps
the substrate binding site of the b’ domain. This report also presents evidence that
the two conformers are also significant in full-length PDI. In addition, this thesis
showed that two ER resident proteins, GPx7 and GPx8 together with PDI form an
efficient pathway to generate disulfide bonds from hydrogen peroxide in the ER.
Finally, the present study innovated new efficient methodologies for making
disulfide bond containing proteins by introduction of Erv1p, along with a catalyst
of disulfide bond isomerization, into the cytoplasm of E. coli.
18
2
Review of the literature
2.1
Protein folding
Proteins are made by a concerted process, with DNA being transcribed to mRNAs
and mRNAs translated into the amino acid sequence of proteins. The sequence of
a protein determines its native structure and function (Anfinsen CB 1973). In
nature, proteins are synthesized as linear polypeptides, the so called primary
structure, and then fold into their final secondary, tertiary, and quaternary native
structure depending on the primary structure of protein and on its biophysical
environment. Protein folding is driven by a number of non-covalent interactions
such as hydrogen bonding involving the polypeptide backbone or polar side
groups, electrostatic or ionic interactions between charged side groups, Van der
Waals forces between all atoms, and hydrophobic packing of nonpolar side
groups to exclude water. After folding the number of hydrophobic residues
exposed on the surface of a protein is normally minimized. Hence a correctly
folded protein usually has a closely packed hydrophobic core and hydrophilic
residues, such as charged and polar side chains, displayed on the surface of
protein (Pace et al. 1996). In the cell, the process of folding often initiates cotranslationally, so that the N-terminus of the protein begins to fold while the Cterminus of the protein is still being synthesized by the ribosome (Alberts et al.
2002). Across different species proteins fold in multiple cellular compartments
including the cytoplasm, the ER, the intermembrane space and matrix of
mitochondria and chloroplasts, and the extracytoplasmic space (e.g. the
periplasm). In eukaryotic cells, more than a half of the total proteins are folded in
the cytoplasm. This compartment, in both eukaryotes and prokaryotes, does not
have mechanisms for oxidative folding i.e. the formation of disulfide bonds
between the side chains of cysteine residues. In contrast, lysosomal proteins,
membrane proteins and secreted proteins are mainly folded in the secretory
pathway, specifically in the ER, in which oxidative folding does occur, in addition
to posttranslational modifications such as N-glycosylation. Both disulfide bond
formation and N-glycosylation make proteins more stable (Wickner & Schekman
2005). Proteins form into oligomers either after monomers of proteins have been
folded completely into native domains, or when the subunit is not properly folded
(Hurtley & Helenius 1989, Hartl & Hayer-Hartl 2009). Some large complex
proteins continue folding events after they have exported from the ER, for
19
example the Von Willebrand Factor (Sadler et al. 2009), mucins etc (Adler et al.
2013).
A cell may contain hundreds of thousands of proteins. For example the
human cell has more than twenty thousand protein-coding genes (Pennisi et al.
2012) and alternate splicing, alternate initiation of translation and post
translational modifications further increase the diversity. The total number of
proteins in a human cell is estimated to be between 250,000 to one million
(http://proteomics.cancer.gov/whatisproteomics). These proteins need to fold in a
complex environment, for example the protein concentration in the cytoplasm has
been estimated at being around 100 mg/ml (Zeskind et al. 2007). Inside the cell,
many proteins are synthesized simultaneously by ribosomes at very high speed. A
single eukaryotic protein is produced at a speed of 2-8 residues per second
(Schroder & Kaufman 2005, Kleizen et al. 2005), while in prokaryotes the rate of
transcription (around 60 nucleotides per second) is linked to the rate of translation
(around 20 amino acids per second) (Alberts et al. 2002). During folding, the
exposure of hydrophobic side-chains of newly synthesized polypeptides has to be
minimized otherwise inappropriate aggregative events will occur (Pace et al.
1996). Aggregation is an autocatalytic event, which will have severe deleterious
effects on the cell – including cell death. Hence protein synthesis and folding in
the cell needs to be well organized and well controlled, so that proteins are folded
accurately and efficiently.
In cells the folding of proteins is assisted by pathways which control the
speed of protein folding, and chaperones that prevent unwanted interactions
between nascent proteins and hence avoid aggregation of newly synthesized
proteins. Cells also have protein folding quality control pathways to make sure
that all proteins are folded correctly before they reach their destinations where
they act. Some proteins require posttranslational modifications such as disulfide
bond formation, phosphorylation, glycosylation, etc, before attaining their native,
functional state. Proteins that do not reach their native structure have to be
refolded or degraded. The cell needs to maintain the balance between protein
synthesis, protein folding and protein degradation/ turnover to avoid accumulation
of protein aggregates caused by excess unfolded protein that leads to cell stress
and disease. For example, accumulation of aggregated proteins leads to amyloidrelated illnesses such as Alzheimer’s (Harper et al. 1997, Knowles et al. 2014)
and familial amyloid cardiomyopathy or polyneuropathy (Hund et al. 2002), prion
infections (Burny et al. 1997), Huntington's and Parkinson's disease (Jucker &
Walker 2011). Similarly a high rate of degradation causes a number of
20
proteopathy diseases such as antitrypsin-associated emphysema (Silverman &
Sandhaus 2009), cystic fibrosis (Ratjen & Döring 2003) and the lysosomal
storage diseases (Hollak & Wijburg 2014).
2.2
Oxidative folding of proteins
Oxidative folding is the formation of disulfide bonds between cysteine residues in
proteins, i.e. the formation of a covalent linkage between two thiol groups. This is
a reversible reaction (Fig. 1). The disulfide bond is strong, but still it is a
relatively weak covalent bond. Disulfide bonds play an important role in the
folding and stability of some proteins, usually secreted proteins, outer-membrane
proteins, and lysosomal proteins (Gross et al. 2002, Winther & Thorpe 2014). It is
predicted that the majority of proteins which enter the eukaryotic secretory
pathway are disulfide bonded proteins (Chen et al. 2005). Oxidative folding
occurs in the periplasmic space of bacteria and in the ER, chloroplasts, and
mitochondria of eukaryotic cells. It is also reported to occur in the cytoplasm of
some extremophiles (Madonna et al. 2006, Limauno et al. 2014).
Fig. 1. Schematic presentation of reduction, oxidation and isomerization reactions.
Oxidative protein folding can be achieved by multiple pathways (Fig. 2)
(Ruddock 2012, Oka & Bulleid 2013). For example, dithiol groups can be
21
oxidized directly by oxygen, or by low molecular weight biochemical compounds
such as GSSG (Freedman 1995), by reactive oxygen species (ROS) (Cumming et
al. 2004), by dehydroascorbate (DHA) (Saaraanen et al. 2010), by sulfhydryl
oxidases such as the Ero1 family (via Protein Disulfide Isomerase (PDI); Frand et
al. 1999) or QSox family (Alon et al. 2012), by Vitamin K linked oxidoreductases
(Rutkevich & Williams 2012).
Fig. 2. Disulfide bond formation pathways in the ER. Dotted arrows represent potential
redox routes that had not been characterized experimentally in vivo before this study
started.
2.3
Disulfide bond formation in the endoplasmic reticulum
The ER is the second major organelle for protein folding in the eukaryote cell
after the cytosol. One third of all human proteins are translocated to the ER,
where they are folded before getting secreted through the Golgi apparatus (Chen
et al. 2005). Proteins that fold in the ER are often modified by disulfide bond
formation and glycosylation (Helenius & Aebi 2004). The ER has catalysts and
chaperones that assist protein folding, including oxidative protein folding. It is
estimated that there are about 1700 human proteins that are retained in the ER
22
(http://locate.imb.uq.edu.au/). The lumen of the ER has an optimal environment
for oxidative folding of proteins and has efficient oxidative pathways to generate
disulfide bonds including PDI-family members to assist native oxidative folding.
2.3.1 Structure and function of human protein disulfide isomerase
family members
PDI-family members are present in all eukaryotic species. To date, 20 members of
the human PDI-family have been identified (Fig. 3). They are characterized by
containing at least one thioredoxin fold containing domain, and having ER
localization (Hatahet & Ruddock 2009). Most PDIs catalyze the formation,
breakage and rearrangement of disulfide bonds, allowing proteins to quickly find
the correct arrangement of disulfide bonds in their fully folded states; they might
also have chaperone-like activities (Ellgaard & Ruddock 2005). Since there are
only 20 PDI proteins, but about 7000 proteins are folded in the human ER, each
member in the family probably catalyzes disulfide formation in multiple
substrates and different family members may either play different roles in the
process, or be involved in folding of different set of substrates i.e. have different
substrate specificities. PDI-family members probably contribute to different
stages of oxidative folding and work cooperatively to ensure the efficient
production of multi-disulfide proteins in the ER. So far the functions of individual
PDI-family members have not been fully characterized.
The most well characterized protein of the PDI-family is PDI or P4HB. It
comprises two homologous catalytic domains, a at the N-terminal and a’ at the Cterminal, which are connected by non-catalytic domains b and b’, The b’ domain
is connected with the a’ domain by a 19 amino acid long linker named x. The
protein also has a highly acidic C-terminal extension named c, and an ER
retention motif at the C-terminal –KDEL (Ellgaard & Ruddock 2005). It is a
highly abundant multifunctional enzyme. It was first shown to have oxidative
protein folding activity 50 years ago (Goldberger et al. 1964). It is also the beta
subunit of prolyl 4-hydroxylase (P4H), an enzyme that is involved in
hydroxylation of prolyl residues in preprocollagen (Pihlajaniemi et al. 1987) and
a subunit of microsomal triglyceride transfer protein complex (MTTP) (Pons et
al. 2011). Other known functions include i) being a chaperone; ii) ER associated
degradation (Molinari et al. 2002); iii) regulation of NAD(P)H oxidase
(Janiszewski et al. 2005, Liberman et al. 2008). ; iv) PDI is involved in loading
antigenic peptides into MHC class I molecules (Lee et al. 2009); v) PDI has been
23
shown to assist HIV entry by reducing a disulfide bond on the HIV gp120 protein
during HIV infection of CD4 positive cells (Ryser et al. 2005); vi) PDI was
identified to play an important role in glioma cell invasion, specifically deficiency
of PDI inhibited tumor cell migration and invasion (Goplen et al. 2006).
Fig. 3. The human PDI family (Hatahet & Ruddock 2009). Catalytic domains and noncatalytic domains with a thioredoxin fold are shown as elliptical circles, The active site
motif is indicated in each catalytic domain.
24
Structures of the a and b, domains have been solved by NMR (Kemmink et
al. 1996, Kemmink et al. 1999), an NMR structure of a’ domain has been
deposited to the Protein Data Bank (PDB entry code: 1X5C) and, as part of this
project, the structure of the b’x domain has been studied (Nguyen et al. 2008). All
four domains have a thioredoxin fold. The catalytic a and a’ domains include the
active site motif -CGHC-, that catalyzes disulfide reduction, oxidation, and
isomerization reactions (section 2.3.2). The b’ domain plays a role in substrate
recognition and binding (Klappa et al. 1998). The b domain plays a structural
role, connecting and orientating the a and b’ domains. The a and a’ domains
share nearly 37% identity with each other and are highly similar to thioredoxin.
The structure of full length yeast PDI (Pdi1p) (Tian et al. 2006) and the structure
of ERp57 in complex with tapasin (Dong et al. 2009) have been released. The
structure of Pdi1p shows that the four domains of PDI (a, b, b’, a’) form an Ushape in which active site motifs of the a and a’ domains face each other. The
structure of ERp57 shows high similarity to that of Pdi1p. Structures of other full
length mature PDI family members have been solved including ERp44 (Wang et
al. 2008) and ERp29 (Barak et al. 2009). The structure of AGR-3 has been
deposited by the Ruddock group to Protein Data Bank (PDB entry code: 3PH9).
AGR-3 displays a thioredoxin fold with an active site –CQYS- exposed on the
surface of protein. The structure of the bb’ domains of ERp72 was published and
shows significant similarity with the structure of the bb’ domains of ERp57
(Kozlov et al. 2009). Recently, structures of human PDI in an oxidized and
reduced form have been characterized (Wang et al. 2013; Fig. 4). These show
high similarity to the structure of Pdi1p and ERp57. Wang’s study showed that in
the reduced PDI, domains a, b, and b' stay in the same plane, but the domain a'
rotates ∼45° out, and the distance between the active sites is 27.6 Å. Whereas the
four domains in the oxidized PDI stay in the same plane, and the distance
between the active sites is 40.3 Å. Reduced PDI shows a closed conformation,
whereas oxidized PDI exists in an opened conformation with more exposed
hydrophobic areas and a larger cleft with potential for substrate binding.
25
Fig. 4. The overall structures of human PDI in oxidized form (4EL1), and in reduced
form (4EKZ) (Wang et al. 2013). The structures show the a-b-b’-x-a’ domain
architectures. The active site cysteines are shown in red, x linker is shown in blue. In
both the reduced and oxidized forms, the a-b-b’-x-a’ domains are arranged as a
horseshoe shape with two CGHC active sites in the a and a' facing each other.
In the human PDI-family, Erp57, PDIp, and PDI share the same domain
architecture of a-b-b’-a’-c, but appear to have significant other differences
(Ellgaard & Ruddock 2005). For example, while PDI recognizes non-native
proteins directly, ERp57 assists disulfide bond formation in glycoproteins by
interacting with the ER-resident lectins calnexin or calreticulin (Leach et al. 2002,
Urade et al. 2004). ERp57 also associates with Tapasin and is part of the major
histocompatibility complex (MHC) class I peptide-loading complex, which is
essential for formation and export of the antigen from the endoplasmic reticulum
to the cell surface (Garbi et al. 2006, Peaper et al. 2005). ERp57 also interacts
with another PDI-family member, ERp27 (Alanen et al. 2006), but the cellular
function of this interaction and that of ERp27 is still unknown.
Human PDIp has the same active site motif (CXXC) and domain architecture
as human PDI and Erp57 (a-b-b’-a’) (Ellgaard & Ruddock 2005). It shares 46%
amino acid sequence identity to human PDI (Desilva et al. 1997). PDIp is a
glycoprotein specifically expressed in pancreatic acinar cells (Desilva et al.
1997), suggesting that PDIp may function as a protein-folding catalyst for
secretory digestive enzymes (Fu et al. 2009). PDIp can bind 17β-oestradiol, and is
reported to be the predominant intracellular oestrogen storage protein in the
pancreas (Fu et al. 2012).
26
Erp72 is the second largest protein in the PDI family after ERdj5, and the
only protein in this family that contains five thioredoxin-like domains with three
catalytic domains, a structural domain and a substrate binding domain (Hatahet &
Ruddock 2009). In addition, ERp72 contains an N-terminal stretch of 17 aspartate
and glutamate residues, which is believed to bind Ca2+ (Van et al. 1993). ERp72 is
localized in the ER, but also has been detected on cell membranes (Chen et al.
2008). ERp72 can catalyze thiol-disulfide exchange reactions (Alanen et al.
2003a). ERp72 is associated with several PDI family members such as PDI, P5,
ERdj3 (Hatahet & Ruddock 2009).
P5 or ERp5 is one of the less well studied member of the PDI family (Hatahet
& Ruddock 2009). It contains two catalytic domains at N-terminal and a noncatalytic domain at C-terminal. It has been shown to be associated with Binding
immunoglobulin protein (BiP) (Jessop et al. 2009), PDI and ERp72 (Stockon et
al. 2003). ERp5 presents mainly on platelet intracellular membranes, but it is
rapidly recruited to the cell surface in response to a range of platelet agonists
(Jordan et al. 2005).
ERp44 contains three thioredoxin like domain (a-b-b’), two non-catalytic
domains and a catalytic domain with an unusual active site motif (CXXS) that
lacks the C-terminal cysteine usually found in other family members (Anelli et
al. 2002). The three domains (a-b-b′) of ERp44 form a V-shaped molecule (Wang
et al. 2008), with a flexible C-terminal tail that could have a regulatory role in
binding and releasing its substrate during the catalytic cycle (Wang et al. 2008). It
is highly expressed in secretory tissues, and is localized mainly to the ER-Golgi
intermediate compartment (Wang et al. 2007). This protein is associated with
Ero1 (Anelli et al. 2003), and inositol 1,4,5-triphosphate receptor type I (Higo et
al. 2005).
ERdj5 is the largest protein in the PDI-family. It contains six thioredoxin like
domains, five catalytic domains and a non-catalytic domain (Hatahet & Ruddock
2009). ERdj5 has been shown to have a reductase activity that cleaves the
disulfide bonds of misfolded proteins. Hence it is involved in the degradation of
misfolded ER proteins (Ushioda et al. 2008). ERdj5 catalyzes the removal of nonnative disulfides and correct folding of the LDL receptor (Oka et al. 2013).
ERp29 or ERp28 has two domains, but lacks a catalytic domain (Ferrari et al.
1998, Barak et al. 2009). It is widely expressed (Demmer et al. 1997). The Nterminal domain of ERp29 is a b-domain like with thioredoxin fold, whereas the
C-terminal domain shows an all α-helical fold (Barak et al. 2009). ERp29 has
been shown to play a role in thyroglobulin processing and a role in the entry of
27
polyomavirus into cells (Rainey-Barger et al. 2007). This protein has been shown
to promote ATF6 transport from ER to Golgi apparatus under ER stress
conditions (Hirsch et al. 2014).
ERp27 has only two non-catalytic domains, a b-like domain at the N-terminal
of the protein and a b’-like domain at the C-terminal. This protein is up-regulated
during ER stress. The b’-like domain of ERp27 can bind to peptides unfolded
substrates (Alanen et al. 2006), but does not interact with folded substrates
(Kober et al. 2013). The substrate-binding cleft of ERp27 on the b’-like domain
is homologous to PDI, but is able to adapt in size and hydrophobicity (Kober et
al. 2013). ERp27 interacts with ERp57 (Alanen et al. 2006) and presumably
presents unfolded substrates to ERp57.
ERp46 has three catalytic domains separated by long linkers, and forms an
opened V-shape (Kojima et al. 2014). It plays a role in protecting cells against
hypoxia (Sullivan et al. 2003). ERp46 can efficiently catalyze the formation of
disulfide bonds but lacks the ability to introduce native disulfide bonds selectively
(Sato et al. 2013). This protein probably acts during early oxidative protein
folding events in the ER (Kojima et al. 2014).
PDIr contains four domains, three C-terminal catalytic domains, and an Nterminal non-catalytic domain (Hayano et al. 1995). It has been shown to be
involved in the folding of α1-antitrypsin and N-linked glycoproteins (Horibe et al.
2004). The non-catalytic domain of PDIr plays a key role in the recognition of
protein partners and misfolded proteins (Vinaik et al. 2013).
PDILT contains four thioredoxin like domains two of which have unusual
active-site motifs SSKQS in the N-terminal domain a domain and WSKKC in the
C-terminal a’ domain. This protein is expressed highly in postmeiotic male germ
cells. PDILT associates with Ero1α (Van Lith et al. 2005), calmegin (Van Lith et
al. 2007), and the testis-specific homologue of calnexin that is required for sperm
fertility (Ikawa et al. 1997). Recently, it has been showed to interact with CALR3
in testicular germ cells and plays a crucial role in the disulfide formation in
ADAM3 (Tokuhiro et al. 2012).
ERp18 is one of the smallest of the human PDI-family members, containing a
single catalytic domain with a CGAC active site motif (Alanen et al. 2003b). It
shows low activity in catalyzing disulfide exchange reactions (Alanen et al.
2003b, Jeong et al. 2008). Erp18 has been shown to associate with hVKORC1
indicating its involvement in oxidative folding via vitamin K metabolic pathway
(Schulman et al. 2010).
28
AGR-2 (anterior gradient homolog 2), and AGR-3 (anterior gradient homolog
3), both are small single-domain proteins that show sequence homology with
ERp18, but they have very different active-site motifs, WCGAC (ERp18),
ECPHS (AGR-2), and DCQYS (AGR-3). The crystal structure of AGR-3 has
been solved (PDB entry code: 3PH9) and is highly similar to solution structure of
ERp18 (PDB entry code: 2K8V) (Rowe et al. 2009). The functions of both
proteins are not clear. However, AGR-2 has been shown to be localized within the
ER of intestinal secretory epithelial cells and is essential for in vivo production of
the intestinal mucin MUC2, a large cysteine rich glycoprotein that forms the
protective mucus gel lining the intestine (Park et al. 2009).
Transmembrane proteins (TMX), including TMX, TMX2, TMX3, TMX4,
TMX5. TMX, TMX2, TMX4, TMX5 contain only one thioredoxin-like domain
with unusual active site motifs while TMX3 contains three domains with a Nterminal catalytic domain (Hatahet & Ruddock 2009). TMX has been shown to
associate with MHC class I heavy chain and protect it from degradation (Matsuo
et al. 2009). Most of PDI family members except TMX and TMX4 were shown to
form a mixed disulfide bond with Ero1α and Ero1β, indicating their involvements
in oxidative folding (Benham et al. 2000, Appenzeller-Herzog et al. 2010).
Association of TMX, TMX4 with hVKORC1 indicates their involvement in
oxidative folding via vitamin K metabolism pathway (Schulman et al. 2010).
2.3.2 Reactions of protein disulfide isomerase
The active sites (CXXC) of PDI naturally exist in multiple different states; a
reduced dithiol state, an oxidized intramolecular disulfide state and an
intermolecular mixed disulfide state. The reduced state can catalyze the
reduction/isomerization of disulfide bonds in substrate proteins, while the
oxidized state is able to catalyze oxidation of dithiols in substrates (Gane et al.
1995, Chivers et al. 1997). PDI may catalyze isomerization reactions through
cycles of reduction and oxidation reaction, or directly though intramolecular
disulfide rearrangement (Schwaller et al. 2003).
In oxidation reactions (Fig. 5), two free thiols in substrate are oxidized by an
oxidized active site of PDI to form a disulfide bond in the substrate and a reduced
active site in PDI. In turn, PDI is re-oxidized to complete the catalytic cycle of
PDI. In the ER the PDI is oxidized usually by Ero1 family, or oxidized
glutathione, or DHA, or reactive oxygen species (see sections 2.2 and 2.3.3).
29
Fig. 5. Schematic presentation of the reactions of PDI. For clarity only one active site
of PDI is shown. An oxidized PDI catalyses disulfide bond formation in a reduced
substrate (oxidation); a reduced PDI reduces disulfide bonds in an oxidized substrate
(reduction) or can catalyse native disulfide bond formation in an incorrectly folded
substrate (isomerization).
In reduction reactions (Fig. 5), a disulfide bond in substrate is reduced by the
reduced active site of PDI which results in oxidation of the active site of PDI and
the formation of two free cysteines in the substrate. In turn, PDI is reduced by
reduced glutathione (GSH).
Another function of PDI is to catalyze disulfide bond isomerization reactions
(Fig. 5) in which disulfide bonds and free thiols in substrates are rearranged to
gain native disulfide bonds. This reaction happens either via cycles of reduction
30
and oxidation reactions, or via a direct intramolecular disulfide rearrangement
(Fig. 5). In the later, an incorrect disulfide bond in a misfolded protein is attacked
by the N-terminal active site cysteine of PDI to form a mixed disulfide between a
cysteine of the substrate and a cysteine of PDI’s active site. Next, the attack of a
second cysteine residue from the substrate protein results in the formation of a
more stable disulfide in the substrate protein. This is repeated until PDI is
released. If at any time the C-terminal active site cysteine of PDI is the
nucleophile, then the substrate undergoes a net reduction and PDI becomes
oxidized and has to be reduced in order to play a catalytic role in reduction or
direct isomerization.
Disulfide bond formation depends heavily on the redox state and pH of the
environment since the protonation state of thiol groups is heavily dependent on
the pH of surroundings and only the deprotonated thiolate state is a good
nucleophile. The pKa values of the active-site cysteine residues are very
important for PDI activities. The redox potential of PDI depends on the pKa
values of the two active site cysteine residues (Lappi et al. 2004). The pKa of the
N-terminal cysteine in the active site is often very low and may not vary
significantly during the catalytic cycle (Karala et al. 2010). In contrast, the pKa of
the C-terminal cysteine is often high and may vary considerably during PDI’s
reaction cycles. It has been shown that the pKa value of the C-terminal cysteine
of a domain is modulated by the movement of the side chain of R120 (Lappi et al.
2004). This allow PDI to efficiently catalyze both oxidation and isomerization
reactions (Karala et al. 2010).
2.3.3 Disulfide bond generation in the endoplasmic reticulum
To date, several pathways for disulfide bond formation in the ER have been
characterized (Fig. 2).
The sulfhydryl oxidase Ero1 and PDI are thought to form the major pathway
for protein disulfide bond formation in the ER (Fig. 6). Ero1 is a FAD-dependent
sulfhydryl oxidase tightly associated with the luminal face of the ER membrane
that uses molecular oxygen as an electron acceptor to oxidize itself. Oxidized
Ero1 then transfers disulfide to folding secretory proteins via reduced PDIs
(Frand et al. 1999, Benham et al. 2000). Ero1 family proteins are present in all
eukaryotes. Many simple eukaryotes like invertebrates, fungi, and yeast encode a
single Ero1, except S. pombe has two Ero1 isoforms (Frand et al. 1998, Kettner et
al. 2004), whereas mammalian cells have two Ero1 family members, Ero1α, and
31
Ero1β (Cabibbo et al. 2000, Tu et al. 2000, Gross et al. 2006). The role of Ero1 in
disulfide bond formation was first identified in yeast (Frand & Kaiser 1998,
Pollard et al.1998). It has been shown that yeast Ero1p directly oxidizes Pdi1p,
since it forms a disulfide bond with Pdi1p in vivo (Frand & Kaiser 1999). Yeasts
that are mutated in Ero1p are extremely sensitive to the reducing agent DTT,
whereas overexpression of Ero1p confers DTT resistance (Frand & Kaiser 1998,
Pollard et al.1998). Later on, the roles of mammalian Ero1 family members, both
Ero1α and Ero1β, in PDI oxidation in vivo were shown (Cabibbo et al. 2000,
Pagani et al. 2000). Two mouse Ero1 isoforms have similar in vitro biochemical
activities (Zito et al. 2010a). To date, no major differences in redox activities or
substrate interactions have been identified between the Ero1 isoforms (Sevier
2010). However, there is different in tissue distribution between Ero1α and Ero1β.
Ero1β presents selectively in the pancreas. Whereas, Ero1α appears in all tissues
(Zito et al. 2010a). Mice lacking both Ero1α and Ero1β exhibit slightly decreased
rates of disulfide bond formation (Zito et al. 2010a). A wide range of PDI family
members were shown to associate with Ero1α and Ero1β (Benham et al. 2000,
Appenzeller-Herzog et al. 2010).
Fig. 6. Schematic representations for Ero1 activity. Activity of Ero1 controls redox
homeostasis of the ER. Ero1 activity is regulated by the prescence of GSSG, oxidized
PDIs and oxidized substrates in the ER.
32
Ero1 contains two pairs of catalytic cysteines, an active-site pair located at
the C-terminus of Ero1 and a shuttle cysteine pair (Fig. 7) (Frand & Kaiser 2000).
Ero1 uses molecular oxygen to oxidize its active site, which in turn transfers
disulfide to the shuttle cysteine pair via internal dithiol-disulfide exchange, then
the oxidized shuttle cysteine pair oxidizes reduced PDI which in turn oxidizes
folding protein substrates (Frand & Kaiser 2000, Sevier & Kaiser 2006). Ero1
does not contain an ER retention signal, it is retained in the ER by associating
with the PDI-family member ERp44 (Anelli et al. 2003). Ero1 activity is tightly
regulated by non-catalytic regulatory cysteine pairs to control redox homeostasis
of the endoplasmic reticulum (Sevier et al 2007, Appenzeller-Herzog et al. 2008,
Tavender & Bulleid 2010a, Hansen et al. 2014). The regulatory disulfides are
formed under oxidizing ER conditions and their formation is thought to limit the
movement of the shuttle cysteines and hence limits the ability of Ero1 to shuttle
disulfides to PDI. In contrast, the regulatory disulfides are reduced under reducing
ER conditions leading to an increase of Ero1 activity (Sevier et al 2007,
Appenzeller-Herzog et al. 2008, Hansen et al. 2014).
Fig. 7. Schematic representation of the disulfide pattern in the active and inactive
forms of human Ero1α with catalytic disulfides (solid lines) and regulatory disulfides
(dashed lines) indicated (Appenzeller-Herzog et al. 2008, Inaba et al. 2010).
Altogether the data suggests that Ero1 is an essential route for disulfide bond
formation the ER of yeast, but in mammals Ero1 is non-essential (Zito et al.
2010a), but may still provide the major route for disulfide bond formation in the
ER. What are the alternative routes for oxidative folding in the ER of mammalian
cells?
Another possible source for disulfide bond formation in the ER is via
Quiescin Sulfhydryl Oxidases (QSOX) (Heckler et al. 2008). Like the Ero1
family they are FAD-dependent oxidases which can produce disulfides with the
33
corresponding reduction of molecular oxygen to hydrogen peroxide (Kodali &
Thorpe 2010). QSOX family members are present in most eukaryotes, but absent
in yeast and fungi (Thorpe et al. 2002, Kodali & Thorpe 2010). Human has two
QSOX orthologs, QSOX1 and QSOX2. QSOX1 and QSOX2 share 41% identity.
These proteins have been detected in the ER, the Golgi, and extracellularly, but
are primarily in the Golgi apparatus and secreted fluids (Kodali & Thorpe 2010).
QSOXs contain an FAD binding Erv/ALR sulfhydryl domain at the C-terminal,
and a thioredoxin domain with a CXXC active site (Alon et al. 2012). The
domain combination in QSOX family members makes them very efficient at
catalyzing disulfide bond formation in vitro (Kodali & Thorpe 2010). QSOX can
introduce disulfide bonds directly to reduced unfolded proteins much more
efficiently than it does to reduced active site of PDI (Hoober et al. 1999, Thorpe
et al. 2002). In vitro, QSOX cooperatively assist PDI in oxidative folding of
disulfide bond-containing protein (Rancy & Thorpe 2008).
Sulfhydryl oxidases including Ero1 and QSOX produce hydrogen peroxide as
a by-product during their catalytic cycles, with one molecule of hydrogen
peroxide being produced for each disulfide bond made. Hydrogen peroxide is
considered to be a toxic product to the cells when it accumulates. A secretory
plasma cell can generate 100000 disulfide bonds per second (Cenci et al. 2011),
which would result in the rapid generation of high concentrations of peroxide.
The ER does not have a catalase, nor a similar enzyme that could remove the
peroxide, raising the question which mechanism is used to remove hydrogen
peroxide produced in the ER during disulfide bond formation? In principal
hydrogen peroxide can react directly with the free thiol group of cysteine in
folding proteins or PDI to form cysteine sulfenic acid. In turn the cysteine
sulfenic acid can react with another free thiol of cysteine to form a disulfide or
react with another cysteine sulfenic acid to generate cysteine sulfinic acid (Fig. 8).
Fig. 8. Formation of disulfide bonds by using peroxide.
34
In vitro it has been shown that hydrogen peroxide can be used to efficiently
generate disulfide bonds in reduced folding substrates either directly or via PDI
(Karala et al. 2009). Hence, in vivo, the ER may use hydrogen peroxide generated
by Ero1 or QSOX to create disulfide bonds efficiently. If it is the case, for each
catalytic cycle Ero1 will produce two disulfide bonds from a single oxygen
molecule. That not only removes a potentially toxic byproduct, but is also more
economical for the cell. Recently studies have shown that ER-localized
peroxiredoxin IV can cooperate with PDI to convert hydrogen peroxide generated
by Ero1 to disulfide bonds (Tavender et al. 2010, Zito et al. 2010b).
Before the role of Ero1 in oxidative folding was discovered, oxidized
glutathione (GSSG) had been considered the primary route for disulfide bond
formation in the ER (Hwang et al 1992). Glutathione is the most abundant peptide
in animal cells and exists primarily in a reduced state (GSH) or an oxidized form
(GSSG). GSH is synthesized in cytoplasm, and imported into the ER by diffusion
or active transport. In the ER, GSSG is generated from GSH by the oxidative
folding machinery (Banhegyi et al. 1999, Appenzeller-Herzog 2011). GSH can
act as an electron donor in the reduction of peroxides, which is catalyzed by
glutathione peroxidases, or that of disulfides, a reaction catalyzed by
glutaredoxins (Winterbourn & Metodiewa 1999). A deletion of glutathione leads
to faults in native disulfide bond formation in the ER (Chakravarthi & Bullied
2004). GSH is required for the correct reduction or isomerization of non-native
disulfide pairings during oxidative protein folding (Chakravarthi et al. 2006). The
ratio and concentrations of GSH and GSSG determine the redox potential of
intracellular compartments. GSH/GSSG forms the primary redox buffer in the ER
(Gutscher et al. 2008). The ER is more oxidizing than the cytoplasm because of
having higher levels of GSSG (Hwang et al. 1992, Fewell et al. 2001). GSSG is
able to directly oxidize reduced PDI, or reduced folding proteins efficiently in
vitro (Karala et al. 2009). All together the data suggests that glutathione plays an
important role in oxidative folding in the ER, including in maintaining redox state
of the ER.
The metabolism of vitamin C and vitamin K are also considered as sources of
oxidizing equivalents for disulfide bond formation in the ER. Dehydroascorbic
acid (DHA) is an oxidized form of ascorbic acid (vitamin C). Cells contain
ascorbic acid at millimolar concentration, including in the ER. It is a primary
antioxidant in plasma and within cells. However, while DHA is actively imported
into the endoplasmic reticulum of cells via glucose transporters, mainly by
Glucose transporter 1 (Glut1), ascorbic acid cannot be imported (Welch et al.
35
1995), requiring the reduction of DHA to ascorbic acid in the ER lumen. This
reaction is regarded to be catalyzed by several enzymes, including PDI (Ruddock
2012). However, recently studies in vitro has shown that DHA can react more
rapidly with thiols in unfolded or partially folded proteins than with the thiols of
reduced active site in PDI (Saaranen et al. 2010). DHA is able to react with thiols
to form disulfide bonds and ascorbate. DHA can also be generated from the noncatalyzed antioxidant action of ascorbate in situ. Additionally, ascorbate in the ER
reaches millimolar concentrations in most tissues (May 2012). Therefore, the
ascorbate/DHA couple may be a significant route for disulfide bond formation in
the ER (Zito et al. 2012).
Another potential route for disulfide formation in the ER is via the vitamin K
epoxide reductase (VKOR) family of proteins which are involved in vitamin K
metabolism. Vitamin K is a class of closely related compounds. Vitamin K2
(menaquinone) is a cofactor of DsbB, a key disulfide bond producer in the
bacterial periplasm (Kadokura & Beckwith 2010; section 2.5). VKOR family
members are present in vertebrates, insects, plants, bacteria and archaea
(Goodstadt & Ponting 2004). Bacterial VKOR has been shown to catalyze
disulfide bond formation in the periplasm (Dutton et al. 2008). Mammals have
two VKOR family members, VKORC1 and VKORCL1. Both VKORs are ERresident transmembrane proteins with cysteine active site residues pointing
toward the ER lumen (Schulman et al. 2010). VKORC1 has been shown to be
involved in γ-carboxylation of glutamate residues of some secreted proteins
including blood clotting factors (Jin et al. 2007, Wallin et al. 2008). A CXXC
motif of VKOR is oxidized during this process (Wajih et al. 2005). Recently
VKOR has been shown to contribute to oxidation and disulfide bond formation
within the ER (Rutkevich & Williams 2012). In contrast to VKORC1,
VKORC1L1 has been poorly characterized to date. Altogether the data clearly
show that the mammalian VKOR could function as per the bacterial homologue
in disulfide bond formation. The question is, which protein in the ER can be a
redox partner of VKOR? PDI was initially thought to be a partner since it has
been shown to associate with VKOR (Wajih et al. 2007). However, a later study
has shown that other PDI-family members, TMX, TMX4, and ERp18, but not
PDI, crosslink to VKOR (Schulman et al. 2010).
36
2.4
Disulfide bond formation in the mitochondrial inter-membrane
space
While disulfide bond formation was originally thought to be limited to the ER in
eukaryotes, more recently it has been identified that disulfide bond formation
occurs in other eukaryotic cellular compartments as well, the best well
characterized of which is the inter-membrane space of mitochondria (Riemer et
al. 2009).
Mitochondria primary function is in the production of cellular energy, and
hence they are essential for cell growth and viability. More than 1400 proteins are
present in human mitochondria (Taylor et al. 2003, http://locate.imb.uq.edu.au/).
Of these only 13 proteins are encoded by mitochondrial DNA (Taylor & Turnbull
2005). Therefore most of proteins in mitochondria are encoded by nuclear DNA,
are synthesized in the cytoplasm and then are imported into mitochondria.
Proteins that are intended to be inside mitochondria normally have mitochondrial
signal sequences at their N-terminal (Herrmann et al. 2000). The mitochondrial
targeted proteins are recognized by TOM (Translocase of the Outer Membrane).
TOM cooperates with TIM (Translocase of the Inner Membrane) to translocate
proteins into the mitochondria (Koehler et al. 2004). Some proteins in
mitochondria were identified to contain disulfide bonds (Mesecke et al. 2005,
Webb et al. 2006, Herrmann & Kohl 2007). Hence mitochondria require
machineries to assist oxidative folding (Fig. 9).
37
Fig. 9. Oxidative protein folding in the IMS of mammalian mitochondria. Reduced
unfolded proteins from the cytosol are imported to the IMS via the TOM complex in an
reduced unfolded state. Mia40 is oxidized by ALR and traps reduced unfolded
imported proteins by disulfide links. The imported proteins then are released from
Mia40 in an oxidized folded state. Electrons are delivered from the flavin cofactor of
ALR to Cytochrome C and further into the respiratory chain.
The intermembrane space (IMS) of mitochondria has a disulfide relay system
that includes Mia40 (mitochondrial intermembrane space import and assembly
protein 40) and Erv1p in yeast, or Mia40 and ALR (FAD-linked sulfhydryl
oxidase) in mammals. This system promotes the import and folding of small
cysteine-containing proteins (e.g. small TIM, Cox-17, Cox-19) into the IMS
(Tokatlidis 2005, Daithankar et al. 2010). In the IMS, oxidized Mia40 forms
mixed disulfide bridges with reduced newly imported proteins. Subsequent
isomerization of these disulfide bridges permits the imported protein to be folded
in the IMS (Allen et al. 2005, Mesecke et al. 2005). There are a number of
similarities between oxidative protein folding in the ER and IMS. For example,
Mia40 acts like PDI in that it catalyzes the formation and rearrangement of
38
disulfides, however it does not have thioredoxin folded domain, and it has CXC
active site motif instead of CXXC (Banci et al. 2009). Erv1 and ALR are FADdependent sulfhydryl oxidases like Ero1. Also like Ero1 both contain two pairs of
catalytic cysteines, an active site cysteine pair, and a shuttle cysteine pair. Erv1
transfers a disulfide to reduced Mia40 which in turn oxidizes reduced substrates,
analogous to the Ero1-PDI-substrate protein relay in the ER (section 2.3.3). Like
other sulfhydryl oxidases, Erv1 can use molecular oxygen to oxidize its active site
and produce a hydrogen peroxide molecule. However, in vivo Erv1 activity is
strongly associated with the respiratory chain of mitochondria, therefore
preventing production of H2O2. In this pathway, electrons are passed from Erv1 to
oxygen molecule via cytochrome c and cytochrome c oxidase (Bihlmaieret et al.
2007).
2.5
Disulfide bond formation in the prokaryotic periplasm
In prokaryotes, the mechanism of oxidative folding is best characterized in the
periplasm of gram-negative bacteria, especially E. coli. Most proteins in E. coli
are folded in the cytoplasm, however some periplasmic and membrane proteins of
E. coli are folded in the periplasm space. It is estimated that more than 250
proteins enter the periplasm of E. coli and gain disulfide bonds there (Dutton et
al. 2008) e.g. PhoA, AppA. These proteins are synthesized in the cytoplasm, then
are imported to the periplasm as reduced unfolded proteins – in an analogous
manner to protein import and oxidative folding in the ER or IMS. Similarly, the
periplasm has oxidative folding pathways to support oxidative folding of proteins.
In the periplasmic space of E. coli native disulfide bonds are generated by two
pathways; i) an oxidation pathway comprised of DsbB and DsbA and ii) an
isomerization pathway linked to reduction pathways formed by DsbD, DsbC and
DsbG (Ito et al. 2008) (Fig. 10).
39
Fig. 10. Pathways for disulfide bond formation in the periplasm of E. coli. DsbA is
oxidized by DsbB. Oxidized DsbA oxidizes secretory proteins. DsbC and DsbG
catalyst disulfide bond isomerization to form native disulfide bonds. DsbD shuttles
electrons via thioredoxin from the cytoplasm to reduce DsbC and DsbG.
DsbB is a transmembrane protein with four transmembrane helices. It has two
critical cysteine pairs located in two periplasmic loops, the N-terminal active site
cysteine pair C41/C44 and the C-terminal shuffle cysteine pair C104/C130
(Kobayashi et al. 2001). During the catalytic cycle of DsbB, the C-terminal
shuttle cysteine pair is oxidized by the N-terminal active site cysteine pair,
resulting in a reduced active site cysteine pair. The oxidized shuttle cysteine pair
then transfers a disulfide to the soluble protein DsbA which in turn oxidizes
reduced folding substrates (Kadokura & Beckwith 2002, Zhou et al, 2008). This
is analogous to the mechanisms of disulfide bond formation in the ER and IMS
(sections 2.3.3 and 2.4) i.e. the primary catalyst of disulfide bond formation does
not act directly on folding substrates. Under aerobic conditions DsbB is oxidized
by oxidized ubiquinone, which then is oxidized by respiratory chain cytochrome
oxidases, finally cytochrome oxidases reduce oxygen. In anaerobic conditions
DsbB is oxidized by menaquinone, which in turn is oxidized by fumarate
reductase or nitrate reductase (Takahashi et al. 2004).
The DsbB/DsbA system introduces disulfide bonds into folding proteins,
however in many substrate proteins, especially those which contain nonsequential disulfides, the disulfide bonds introduced by DsbA may not be native
40
disulfides. Therefore, in order to reach the native state those incorrect disulfide
bonds need to be rearranged by a disulfide isomerization pathway which consists
of DsbC, DsbD, and DsbG.
DsbC and DsbG act as disulfide isomerases. Both form a V-shaped
homodimer with two catalytic domains facing each other (McCarthy et al. 2000,
Heras et al. 2004). Each monomer consists of an N-terminal dimerization domain
and a C-terminal catalytic domain with a thioredoxin fold and CXXC active site
(McCarthy et al. 2000, Heras et al. 2004). Both DsbC and DsbG have a substrate
binding cleft located between two arms of the V-shaped homodimer (Heras et al.
2004). However, the substrate binding cleft of DsbG is bigger than one of DsbC,
suggesting that DsbG works on different substrates than DsbC (McCarthy et al.
2000, Heras et al. 2004). In the isomerization pathway, the reduced active site of
DsbC or DsbG attacks an incorrect disulfide in a misfolded protein and forms a
mixed disulfide between DsbC and folding protein. This is analogous to the
isomerization reaction catalyzed by PDI, (see Fig. 5 and Fig. 10). In order to act
as isomerases, DsbC and DsbG have to be kept in their reduced state. Oxidized
DsbC and oxidized DsbG are reduced by the transmembrane protein DsbD
(Katzen et al. 2000), which in turn is reduced by cytoplasmic thioredoxins.
Thioredoxins are in turn reduced by cytoplasmic thioredoxin reductase
(Goldstone et al. 2001, Cho & Beckwith 2009).
Some bacterial species such as Mycobacterium tuberculosis and Streptomyces
coelicolor, do not have DsbB in their periplasmic oxidative pathway although
disulfide bond formation is detected in their periplasmic space. Instead their
periplasmic spaces have bacterial homolog of the eukaryotic enzyme vitamin K
epoxide reductase (VKOR) that can generate disulfide bonds (Dutton et al. 2008,
Dutton et al. 2010).
2.6
Disulfide bond formation in the engineered cytoplasm of E. coli
Naturally the cytoplasm of most bacteria has to be maintained as a reducing
environment to support specific reactions, e.g. reaction of ribonucleotide
reductase
which
catalyzes
the
reduction
of
ribonucleotides
to
deoxyribonucleotides (Elledge et al. 1992). The cytoplasm of E. coli contains two
main pathways (Fig. 11) that reduce disulfides, the thioredoxin (Trx) / thioredoxin
reductase (TrxR) pathway and the glutathione / glutaredoxin (Grx) / glutathione
reductase (GSR) pathway (Prinz et al. 1997).
41
Fig. 11. Disulfide reducing pathways of the cytoplasm in E. coli. NADPH reduces TrxR
which in turn reduces the thioredoxins TrxA and TrxC. Reduced thioredoxins reduce
disulfide bonds in substrate proteins e.g. RNR. NADPH is also used to reduce
glutathione reductase which in turn reduces oxidized glutathione (GSSG) to reduced
glutathione (GSH). GSH can either directly reduce disulfide bonds or can reduce
thioredoxins or glutaredoxins which in turn reduce disulfide bonds in substrates (Ritz
& Beckwith 2001).
E .coli has been widely used to produce recombinant proteins both on a
laboratory and an industrial scale. However, many proteins of interest to both
academics and for biopharmaceutical companies contain disulfide bonds.
Production of disulfide containing proteins is not possible in the normal
cytoplasm of E. coli. However, it has been shown that the disruption of both main
reducing pathways in the cytoplasm through a knockout of glutathione reductase
(gor) and thioredoxin reductase (trxB) combined with a lethality suppressor
mutation in peroxiredoxin alkyl hydroperoxidase (ahpC) leads to a significant
increase in the production of active disulfide bonded proteins such as vtPA, a
protein that contains 9 predominatly non-sequential disulfide bonds (Bessette et
al. 1999). Although those genetic modified strains can produce significant yields
of active disulfide bonded proteins in their cytoplasm, the level and the quality of
disulfide bonded protein production is far from that required by industry, probably
due to the lack of an active catalyst of disulfide bond formation in this system.
Therefore, for the cytoplasm of E. coli to be used as an efficient cell factory for
the production of disulfide bond containing proteins a better system is needed.
42
3
Aims of the present work
The aims of the thesis were to study mechanism of action of human PDI in
disulfide bond formation, disulfide bond formation pathways in the ER and to
generate an efficient system to produce disulfide bond containing proteins.
My specific goals were:
1.
2.
3.
to identify protein structure of the substrate binding domain (b’x) of PDI.
to characterize the function of two ER resident glutathione peroxidases
(GPx7 and GPx8), and their involvement in disulfide bond formation
pathways.
to develop systems for efficient folding of disulfide bonded proteins in the
cytoplasm of E. coli using current knowledge of disulfide bond formation
pathways.
43
44
4
Materials and methods
Detailed description of the materials and methods used in this study are found in
the original articles I, II, III, IV.
Method
Original article
Vector construction
II, III, IV
Protein expression
I, II, III, IV
Protein purification
I, II, III, IV
Mutagenesis
I
Spectroscopy
I
NMR spectroscopy
I
Crystallization, Data collection, structure determination and refinement
I, II
Subcellular localization
II
GPx activity measurements
II
BPTI refolding
II
Kinetics of GPx8 oxidation followed by stopped-flow fluorescence
II
Bimolecular fluorescence complementation
II
Oxygen consumption assay
II
Glutathione peroxidase activity measurements
II
Protein analysis and activity measurements from lysates
III, IV
Analysis of purified proteins
IV
45
46
5
Results
Detailed description of the results in this study are found in the original articles I,
II, III, IV.
Result
Original article
Preparations of recombinant human PDI b’x contain at least two conformers
I
Screening mutations that trap a single conformation of PDI b’x
I
NMR studies of PDI b’x constructs
I
Crystallography of mutant PDI b’x
I
Evidence for alternative conformations of the x region in full-length PDI
I
Human GPx7 and GPx8 are ER-resident proteins
II
Stopped-flow kinetics for the peroxide oxidation of GPx8
II
GPx7 and GPx8 are PDI-peroxidases and not glutathione peroxidases
II
GPx7 and GPx8 combined with PDI are efficient for peroxide-mediated oxidative protein II
folding
Crystal structure of human GPx8
II
GPx7 and GPx8 interact with Ero1α in vivo and modulate Ero1α activity in vitro
II
Alkaline phosphatase production
III
Phytase production
III
Production of a tissue plasminogen activator fragment
IV
Production of a tissue plasminogen activator fragment fusion protein
IV
Development of a pre-expression vector for Erv1p and a disulfide isomerase
IV
Media effects on the yield of active vtPA
IV
Production of other eukaryotic proteins with multiple disulfide bonds
IV
47
48
6
Discussion
6.1
Structure of the substrate binding domain of human protein
disulfide isomerase and the functional significance of a novel
structural interaction between the b’ domain and the x linker
Human PDI is a key enzyme in the ER in the catalysis of disulfide bond
formation in secretory proteins. PDI contains four domains a, b, b’, and a’, an x
linker between domains b’ and a’. When this study was carried out, structures of
full length human PDI and of the b’ domain had not been determined, only the
structure of the individual thioredoxin domains a, b, and a’ had been solved by
NMR (Kemmink et al. 1996, Kemmink et al. 1999, unpublished PDB entry code:
1X5C). The structure of full length yeast Pdi1p was published while this work
was ongoing (Tian et al. 2006).
The structure of full-length human PDI and that of the b’ domain by
crystallization and NMR had not been successfully determined probably due to
the high conformational flexibility of these proteins. The goal of my research was
to study the structure of the b’ domain of PDI. This study should lead to increased
understanding of the mechanisms of action of PDI as b’ is the major substrate
binding domain of PDI, and the only domain whose structure had not previously
been solved.
In this study we showed that the b’x exists in at least two conformational
states, which clearly differ in the structural relationship between the x region and
the b’ domain (paper I). We hypothesized it might be possible to generate mutants
that favorably stabilize one state and abolish the conformational heterogeneity
and by doing so to overcome the complexities introduced by the existence of
these states. Mutations that stabilize one state of conformation may also be
suitable for structural studies. To screen for mutants that have single
conformational state we used Trp fluorescence of purified recombinant mutants.
Our study demonstrated that recombinant b’x exists in two distinct conformations
that differ in tertiary structure, specifically in the environment of the unique
W347 residue in the x region. Mutants that well adopted capped and uncapped
conformation were identified. The I272A mutant of b’x has been crystallized and
the structure determined. In the crystal structure of b’x I272A, the x linker caps
the b’ domain and the side chain of W347 is buried in a hydrophobic pocket in the
b’ domain. The structure of b’ is highly similar to the structure of b’ domains of
49
other PDI-family members, including the bb’ fragment of human ERp57 (PDB
entry code: 2H81), the b’ domain of yeast Pdi1p (PDB entry code: 2BE5), the bb’
fragment from human PDI (PDB entry code: 2K18) and the b’ domain from the
thermophilic fungus Humicola insolens (PDB entry code: 2DJK). While the
crystal structure of b’x was being refined, a structure of full-length yeast PDI
(Pdi1p) was released, in which the four thioredoxin domains a, b, b’, a’ form an
U-shape structure (Tian et al. 2006). The structure of Pdi1p displays a highly
hydrophobic pocket on the surface of the b’ domain and there are also
hydrophobic patches on the surface of a, b, and a’ domains. These hydrophobic
patches on a, b, and a’ domains and the hydrophobic pocket on the b’ domain
form a continuous hydrophobic surface on PDI, which is thought to be important
for substrate recognition of PDI. In the crystal structure one molecule of Pdi1p is
occupying the central cavity of the U-shape structure of another Pdi1p molecule
i.e. it is acting like a substrate. Hence the yeast structure represents a
conformation in which a protein ligand is bound, displacing the x linker. The b’
domains of human PDI and yeast PDI show a similar structure fold and
homologous hydrophobic pocket. However, one significant difference between
the human and yeast PDI structures is in the orientation of the x region, that in the
x region caps the b’ domain of the human PDI, but the x region of Pdi1p does not
cap the hydrophobic pocket of the b’ domain (I Fig. 6). The crystal structure of
human PDI-family member (ERp44) also was published while this work was
ongoing (PDB entry code: 2R2J) (Wang et al. 2008). The ERp44 structure
displayed three thioredoxin domains with the architecture a-b-b’ and with a Cterminal tail which caps a hydrophobic pocket in the b’ domain. This is highly
similar to the x region capping the b’ domain of human PDI (I Fig. 7), in
particular, L361 in ERp44 interacts with the hydrophobic pocket of b’, as W347
does in PDI.
Previous studies have identified mutations that inhibit substrate binding of
b'x or full-length PDI using a combination of modelling and mutagenesis studies
(Klappa et al. 1998, Pirneskoski et al. 2004). However, the binding site identified
by the previous studies does not match that identified in our b’x crystal structure
or by b’ of Pdi1p (Tian et al. 2006). All of the mutations that were previously
reported to inhibit substrate binding instead shift the equilibrium towards the
capped conformer. They are not located in the hydrophobic pocket of the b’
crystal structure. Therefore, these mutations rather act indirectly to inhibit
substrate binding by decreasing the accessibility of the substrate binding site.
Analogous to this, studies on ERp44 showed the role of C-terminal tail in
50
regulating the activity of ERp44, in particular deletion of the C-terminal tail in
ERp44 increased the activity of ERp44 and its ability to act as a molecular
chaperone (Wang et al. 2008).
The conformational exchange, which caps the substrate binding site, does not
only happen in the isolated b’ domain. Studies of mutants in full length PDI
demonstrated that a similar conformational exchange happens in the full-length
protein with the x region capable of interacting reversibly with the principal
substrate binding site on b’ domain (paper I). The x region can adopt different
conformations and can potentially exchange between the capped conformation
and the uncapped conformation. In other words, the substrate binding site is
unavailable to bind substrates when it is capped by x, so that inhibits the activities
of PDI towards protein substrates. The conformational exchange in the b' domain
can be modulated by the addition of peptide ligands that can compete with x
binding at the substrate binding site (Byrne et al. 2009).
Recently, the crystal structures of hPDI containing all four domains (abb'xa')
in both the reduced and oxidized states of the active sites (Wang et al. 2013) have
been published. Those studies reveal a closed conformation of reduced hPDI, and
an open conformation of oxidized hPDI with more exposed substrate binding
sites. The redox-regulated conformational changes in hPDI are mainly located in
the b’xa’ region. The flexibility of this region is mainly dependent on the x linker.
The x linker in the crystal structure of reduced hPDI is close to the b’ domain; it
is highly similar to our b’x I272A structure. The C-terminal half of the x linker
and the a’ domain in oxidized hPDI twists about 45o in comparison to that of the
reduced hPDI. Hence the x linker is required for conformational changes of PDI,
which in turn depends on the redox-state of PDI.
Altogether the published data suggests that the x region of PDI can adopt
alternative conformations during the functional cycle of PDI and that these are
critical for access to the substrate binding site.
6.2
GPx7 and GPx8 are involved in disulfide bond formation in the
endoplasmic reticulum
Currently the major route for disulfide bond formation in the ER is thought to be
via Ero1 family members. This route, in which the flow of oxidizing equivalents
is from molecular oxygen to Ero1 family members and from these to PDI and
from PDI to substrates, is highly regulated (Sevier et al. 2007, AppenzellerHerzog et al. 2008, Baker et al. 2008). Regulation of Ero1 activity is important to
51
keep the optimal redox environment for oxidative folding of proteins.
Specifically, oxidative folding requires not only oxidation, but also isomerization
and reduction of incorrect disulfide bonds to reach the native state. If the redox
environment is too oxidizing isomerization and reduction are inhibited as both
need PDI in the reduced state.
Each catalytic cycle, Ero1 family members generate one disulfide bond and
one molecule of hydrogen peroxide (Gross et al. 2006). Hydrogen peroxide
production as a by-product would result in cellular oxidative stress unless it is
utilized efficiently or degraded. The ER does not have a recognizable catalase or a
similar enzyme for peroxide degradation. In addition, It has been shown that
disulfide bond formation does not induce oxidative stress, but newly synthesized
misfolded proteins in the ER lumen, activates the unfolded protein response
(UPR), causes oxidative stress, and induces apoptosis in vitro and in vivo
(Malhotra et al. 2008). These combined suggest peroxide may be used
functionally.
A previous study from our group using an in vitro BPTI refolding assay
showed that peroxide can efficiently oxidize dithiols in folding proteins to the
native state under physiologically relevant conditions. These studies demonstrated
that no significant oxidative side reactions occur so long as the peroxide
concentration is kept at submillimolar concentrations (Karala et al. 2009). This
productive oxidative folding can occur either via the direct reaction of peroxide
with a free thiol group on the folding protein to form a cysteine sulfenic acid with
subsequent formation of a disulfide bond or via a reaction with a thiol group in
the active site of PDI and subsequent transfer of the disulfide made from PDI to
the substrate protein. Therefore, potentially the ER should have pathways that can
use hydrogen peroxide to generate disulfide bonds. If it is the case, two disulfide
bonds can be generated per molecule of molecular oxygen with no net production
of harmful species.
While the previous in vitro studies showed that peroxide could be used to
make native disulfide bonds, the kinetics of the process were relatively slow and
required relatively high concentrations of peroxide (up to 0.5mM) to be efficient.
To be utilized in vivo a catalyst of this process might be expected. The results
from the study reported here (paper II) demonstrate a novel pathway for utilizing
the peroxide generated by Ero1 for native disulfide bond formation. Two human
glutathione peroxidase (GPx) family members, GPx7 and GPx8 were identified as
containing sequences that directed them to the ER and ER retention signals. Both
52
GPx7 and GPx8 were shown to be localized in the ER, with GPx7 being a soluble
protein and GPx8 being an ER transmembrane protein.
Mature purified recombinant GPx7 or the luminal domain of GPx8 both gave
only a marginally faster reaction than the non-catalyzed reaction in the coupled
GPx assay, indicating that neither GPx7 nor GPx8 are efficient glutathione
peroxidases. An amino acid sequence alignment of the human GPx family and
comparison of the GPx7 protein structure (unpublished PDB entry code: 2P31)
and our GPx8 crystal structure (PDB entry code: 3KIJ) with the GPx1 structure
(unpublished PDB entry code: 2F8A) shows that GPx7 and GPx8 lack the loop
that determines glutathione specificity in the glutathione peroxidase family (II
Fig. 1). This suggests that both may be members of the thioredoxin peroxidase or
TGPx family (Toppo et al. 2008). To date, the PDI-family contains the only
known thioredoxin superfamily proteins that are present in the ER. Therefore,
GPx7 and GPx8 may be PDI peroxidases that catalyze oxidation of reduced PDI
by peroxide. Indeed, this study (paper II) demonstrates that both proteins have
PDI peroxidase activity in vitro and hence presumably in vivo. The addition of
either GPx7 or GPx8 to a mixture of reduced denatured BPTI, PDI, and peroxide
increases the rate of oxidative folding of BPTI. Additionally, by monitoring
oxygen consumption in vitro this study indicates that the presence of GPx7 and
GPx8 increases Ero1α activity, suggesting that these PDI peroxidases interact
with Ero1 and could directly utilize peroxide generated by Ero1. Our data (paper
II) also shows interaction between GPx7 and GPx8 with Ero1α in the ER.
Furthermore, a recent study shows that GPx7 promotes oxidative protein folding
in the ER by directly utilizing Ero1α-generated H2O2. H2O2 oxidizes Cys57 of
GPx7 to sulfenic acid, which can be resolved by Cys86 of GPx7 to form an
intramolecular disulfide bond. In turn both intramolecular disulfide form and
sulfenic acid form of GPx7 can oxidize PDI (Wang et al. 2014). Additionally,
GPx8 has recently been shown to rapidly clear hydrogen peroxide produced by
Ero1α, hence it prevents leakage of hydrogen peroxide from the ER (Ramming et
al. 2014). GPx8 was identified as a cellular substrate of the hepatitis C virus
(HCV) NS3-4A protease, and it is cleaved by the HCV NS3-4A protease at
Cysteine 11 (Morikawa et al. 2014). This study also shows that GPx8 is involved
in HCV particle production, but not in HCV entry or RNA replication.
Recently, human peroxiredoxin IV (Prx IV), an ER localized protein was
shown to protect cells by removing peroxide produced during disulfide bond
formation (Tavender & Bulleid 2010b, Zito et al. 2010b, Tavender et al. 2010)
with data suggesting that Prx IV may provide a parallel pathway to GPx7 and
53
GPx8 for peroxide utilization. However, the oxidation of a substrate protein to the
native state by Prx IV + PDI + peroxide in vitro (Zito et al. 2010b) is significantly
slower than that reported in this study (Paper II) for GPx7 or GPx8 + PDI (II Fig.
5d and f). Even though the assay systems were not same, this suggests that the
PDI peroxidases are at least as efficient as Prx IV at converting peroxide into
disulfide bond formation and their association with Ero1 suggests that they may
provide the major route for peroxide utilization.
All together strongly suggest that PDI peroxidases and Prx IV effectively
utilize any peroxide produced during disulfide bond formation for disulfide bond
formation in the ER.
6.3
Applications of disulfide bond formation
Industrial recombinant protein production is a multi-billion dollar a year market.
Many of the FDA approved biologics are disulfide bonded proteins, including
anticancer drugs, insulin, interleukins, interferons, growth factors, antibody and
antibody fragments etc and the number of such products is growing. The most
common route for the industrial scale production of proteins that contain disulfide
bonds was to express these proteins in the cytoplasm of E. coli as inclusion
bodies, since it is low cost and high yield. The insoluble proteins generated are
then refolded in vitro to get correct folded products. However this step is often
inefficient, time consuming and costly. E. coli can naturally make disulfide
bonded proteins in the periplasm (Collet et al. 2002). However, the volume of
periplasm is rather small compared to the volume of cytoplasm, typically
comprising approximately 10% of the total cell volume (Ghuysen & Hakenbeck
1994). Furthermore, the secretion and folding machinery can be easily saturated.
This means that yields of proteins produced in the periplasm are often 10-fold or
more lower than equivalent yields in the cytoplasm. The problems with the high
level production of disulfide bond containing proteins in E. coli has led to a shift
in the past decade towards the use of eukaryotic production systems such as CHO
cells (Palomares et al. 2004). While the ease of use has increased and associated
costs of using eukaryotic cells has decreased significantly over recent years, they
are still more costly, more difficult to scale up and less regulatory authority
friendly than the use of E. coli.
To get around the limitations of the use of E. coli, researchers have been
trying to create E. coli strains that can produce native disulfide bonded proteins in
the cytoplasm. The cytoplasm of E. coli has been commonly used to produce
54
recombinant proteins at industrial and laboratory scales, but it is not suitable for
the direct production of disulfide bonded proteins as it lacks enzyme systems to
generate disulfide bonds. While strains have been produced which lack the
reducing pathways normally found in the cytoplasm and which can produce much
higher levels of disulfide bonded proteins than wild-type strains (Bessette et al.
1999), the yields and/or quality are often not optimal even when a disulfide
isomerase is co-expressed. In contrast, eukaryotic cells have organelles in which
disulfide bond formation occurs, such as the ER (Hatahet & Ruddock 2009) and
mitochondrial inter-membrane space (Herrmann et al. 2007). These both have
catalytic systems that use sulfhydryl oxidases to make disulfide bonds. The
sulfhydryl oxidases use molecular oxygen to catalyze the oxidation of a dithiol to
a disulfide. Hence we thought, it might be possible to create an efficient system to
produce disulfide bonded recombinant proteins in the cytoplasm of E. coli by
introducing sulfhydryl oxidases into this compartment. For proteins that contain a
few disulfide bonds such as PhoA and AppA, co-expression of the sulfhydryl
oxidase Erv1p gave yields of active disulfide-bonded protein produced up to
1000-fold over the wild-type E. coli control. Moreover, the production of active
PhoA and AppA is more effective than the production in ∆gor ∆trxB background
strain (paper III).
For proteins that are more difficult to be folded correctly because they
contain many disulfide bonds or which contain non-sequential disulfide bonds,
such as vtPA and AppA, may require a catalyst of the other step of native
disulfide bond formation, disulfide isomerization e.g. the use of PDI or DsbC.
In addition to the requirement for catalysts of disulfide bond formation and
isomerization, proteins like vtPA which contain a large number of disulfide bonds
may require longer for the system to catalyze the correct folding of those proteins.
This in turn may lead to the expressed proteins forming inclusion bodies rather
than soluble natively folded states due to the aggregation process being faster than
the folding process. To avoid the issues associated with folding intermediate
aggregation, highly soluble fusion proteins, e.g. MBP, have been widely use to
increase the solubility of the folding intermediates. The use of a MBP fusion
protein, in conjunction with catalysts of disulfide bond formation and
isomerization, significantly increases the yields of disulfide bonded proteins in
our system (paper IV). Furthermore, pre-expression of the folding factors so they
are available before the protein of interest is expressed, and the correct choice of
media also result in higher yields of active complex disulfide bonded proteins. In
combination, yields of active vtPA about 800 fold higher than previously reported
55
using DsbC co-expression in a Δgor ΔtrxB strain were obtained in our system
(paper IV). However, homogeneous natively folded vtPA is not observed.
Other disulfide bonded proteins such as resistin, BPTI, CSF3 and BMP4 have
also been produced by this system (paper IV). Those proteins were expressed in
high yields of soluble, and/or active proteins. However, only proteins with a few
disulfide bonds were produced homogenously in high yields in this system. The
most complex protein we were able to produce to near homogeneity during this
work was Resistin that contains 5 intramolecular disulfide bonds per monomer
plus an inter-molecular disulfide bond between subunits in the dimer. Coexpression of Erv1 in E. coli cytoplasm also significantly increases the yield and
quality of expressed Camelidae single domain antibodies (Veggiani et al. 2011).
Even though our system is significantly better than Δgor ΔtrxB strain in the
production of disulfide bonded proteins, it still needs to be optimized for proteins
which contain a high number of disulfide bonds, especially a high number of non
-sequential disulfide bonds. Subsequent optimization with the group has
significantly improved the processes and now the system is able to generate g/L
amounts of homogeneously folded disulfide bond containing eukaryotic proteins
in the cytoplasm of E. coli in fermentation (HGH, scFv, interleukins; unpublished
results) and more than 100mg/L of correctly folded scFv and Fab antibody
fragments from shake flasks (unpublished results). These optimized systems are
all based on the work undertaken as part of this study (papers III and IV).
56
7
Conclusions
Human PDI is a key enzyme involved in disulfide bond formation in the ER.
Structural information will be very useful to understand the mechanisms of action
of the PDI. Prior to this work the structure of three domain (a, b, a’) were solved,
but the structure of the substrate binding b’x domain was not. It was shown that
the b’x is not suitable for structural studies due to the flexibility of the x linker.
Using a screen of 75 mutants, this study has identified mutants of b’x that
exclusively adopt one of these alternative conformations. The crystal structure of
one of these mutants, I272A, was determined. This structure reveals a
hydrophobic pocket in the b’ domain which is capped by the x region. This study
also showed that the two conformers are also significant in full-length PDI. These
alternative conformations are linked to the ability of PDI to interact with folding
substrates and hence to the mechanisms of action of this folding catalyst.
Pathways for disulfide bond formation in the ER are not yet fully understood.
Until recently the Ero1 family has been considered to be the main source of
oxidation in the ER. These proteins produce disulfide bonds and hydrogen
peroxide as a byproduct. This raises the question: what pathways the ER uses to
remove hydrogen peroxide efficiently, as it will cause cell toxicity if it
accumulates. This thesis has characterized two ER resident proteins, GPx7 and
GPx8, that are involved in eliminating hydrogen peroxide. Human GPx7 and
GPx8 were noted as secreted GPxs. This study showed that they are actually
located in the ER and associate with the peroxide-producing Ero1α. Furthermore,
neither has significant GPx activity, but instead have PDI peroxidase activity. It is
shown in this thesis that the addition of PDI and GPx7 or GPx8 and peroxide to a
reduced unfolded protein results in the very rapid achievement of the folded
disulfide-bonded state. Furthermore, the addition of GPx7 to a mixture of purified
Ero1α, PDI, and GSH resulted in a significant increase in oxygen consumption by
Ero1α. Hence, these proteins, together with PDI, form an efficient way to prevent
hydrogen peroxide accumulation in the ER by utilizing hydrogen peroxide to
form a disulfide bond in substrate proteins.
In parallel to these studies on understanding the mechanisms of disulfide
bond formation in vivo, knowledge of these pathways was applied to the
production of disulfide bond containing proteins. The cytoplasm of E. coli
contains disulfide reducing pathways. Hence production of disulfide bond
containing proteins in the cytoplasm of E. coli results in non-disulfide bonded
proteins that normally form insoluble aggregates. It has been noted previously
57
that the disruption of at least one of the reducing pathways in cytoplasm of E. coli
is essential for the efficient generation of disulfide bonds in proteins (Bessette et
al. 1999). The study reported here has developed a system that showed efficient
disulfide bond formation can be gained by exogenous expression of Erv1p in E.
coli cytoplasm. This study refutes the current paradigm in the field that disruption
of at least one of the reducing pathways is essential for the efficient production of
disulfide bond containing proteins in the cytoplasm of E. coli and open up new
possibilities for the use of E. coli as a microbial cell factory. The addition of PDI
helps to increase the efficiency of production of multi-disulfide bonds containing
proteins, in particular those containing non-sequential disulfide bonds. The yield
of disulfide bond containing protein production is further increased significantly
by fusion of the targeted protein to a highly soluble fusion protein, e.g. MBP, and
by manipulating growth conditions. Various disulfide bond containing proteins
that contain from one to multi-disulfide bonds have been successfully produced
using this technology. To date the system is able to generate g/L amounts of
homogeneously folded disulfide bond containing commercially interesting human
proteins in the cytoplasm of E. coli in fermentation (HGH, scFv, interleukins) and
more than 100mg/L of correctly folded scFv and Fab antibody fragments from
shake flasks.
58
References
Alberts B, Johnson A, Lewis J, Walter P, Raff M, Roberts K (2002) Molecular Biology of
the Cell 4th Edition: International Student Edition: Routledge.
Adler KB, Tuvim MJ & Dickey BF (2013) Regulated mucin secretion from airway
epithelial cells. Front Endocrinol (Lausanne) 4: 129
Alanen HI, Raykhel IB, Luukas MJ, Salo KE & Ruddock LW (2011) Beyond KDEL: the
role of positions 5 and 6 in determining ER localization. J Mol Biol 409: 291-297
Alanen HI, Salo KE, Pekkala M, Siekkinen HM, Pirneskoski A & Ruddock LW (2003a)
Defining the domain boundaries of the human protein disulfide isomerases. Antioxid
Redox Signal 5: 367-374
Alanen HI, Williamson RA, Howard MJ, Lappi AK, Jantti HP, Rautio SM, Kellokumpu S
& Ruddock LW (2003b) Functional characterization of ERp18, a new endoplasmic
reticulum-located thioredoxin superfamily member. J Biol Chem 278: 28912-28920
Alanen HI, Williamson RA, Howard MJ, Hatahet FS, Salo KE, Kauppila A, Kellokumpu S
& Ruddock LW (2006) ERp27, a new non-catalytic endoplasmic reticulum-located
human protein disulfide isomerase family member, interacts with ERp57. J Biol Chem
281: 33727-33738
Allen S, Balabanidou V, Sideris DP, Lisowsky T & Tokatlidis K (2005) Erv1 mediates the
Mia40-dependent protein import pathway and provides a functional link to the
respiratory chain by shuttling electrons to cytochrome c. J Mol Biol 353: 937-944
Alon A, Grossman I, Gat Y, Kodali VK, DiMaio F, Mehlman T, Haran G, Baker D,
Thorpe C & Fass D (2012) The dynamic disulphide relay of quiescin sulphydryl
oxidase. Nature 488: 414-418
Anelli T, Alessio M, Bachi A, Bergamelli L, Bertoli G, Camerini S, Mezghrani A, Ruffato
E, Simmen T & Sitia R (2003) Thiol-mediated protein retention in the endoplasmic
reticulum: the role of ERp44. EMBO J 22: 5015-5022
Anelli T, Alessio M, Mezghrani A, Simmen T, Talamo F, Bachi A & Sitia R (2002)
ERp44, a novel endoplasmic reticulum folding assistant of the thioredoxin family.
EMBO J 21: 835-844
Anfinsen CB (1973) Principles that govern the folding of protein chains. Science 181: 223230
Appenzeller-Herzog C (2011) Glutathione- and non-glutathione-based oxidant control in
the endoplasmic reticulum. J Cell Sci 124: 847-855
Appenzeller-Herzog C, Riemer J, Christensen B, Sorensen ES & Ellgaard L (2008) A
novel disulphide switch mechanism in Ero1alpha balances ER oxidation in human
cells. EMBO J 27: 2977-2987
Appenzeller-Herzog C, Riemer J, Zito E, Chin KT, Ron D, Spiess M & Ellgaard L (2010)
Disulphide production by Ero1alpha-PDI relay is rapid and effectively regulated.
EMBO J 29: 3318-3329
59
Baker KM, Chakravarthi S, Langton KP, Sheppard AM, Lu H & Bulleid NJ (2008) Low
reduction potential of Ero1alpha regulatory disulphides ensures tight control of
substrate oxidation. Embo J 27: 2988-2997
Banci L, Bertini I, Cefaro C, Ciofi-Baffoni S, Gallo A, Martinelli M, Sideris DP, Katrakili
N & Tokatlidis K (2009) MIA40 is an oxidoreductase that catalyzes oxidative protein
folding in mitochondria. Nat Struct Mol Biol 16: 198-206
Banhegyi G, Lusini L, Puskas F, Rossi R, Fulceri R, Braun L, Mile V, di Simplicio P,
Mandl J & Benedetti A (1999) Preferential transport of glutathione versus glutathione
disulfide in rat liver microsomal vesicles. J Biol Chem 274: 12213-12216
Barak NN, Neumann P, Sevvana M, Schutkowski M, Naumann K, Malesevic M,
Reichardt H, Fischer G, Stubbs MT & Ferrari DM (2009) Crystal structure and
functional analysis of the protein disulfide isomerase-related protein ERp29. J Mol
Biol 385: 1630-1642
Bass R, Ruddock LW, Klappa P & Freedman RB (2004) A major fraction of endoplasmic
reticulum-located glutathione is present as mixed disulfides with protein. J Biol Chem
279: 5257-5262
Benham AM, Cabibbo A, Fassio A, Bulleid N, Sitia R & Braakman I (2000) The
CXXCXXC motif determines the folding, structure and stability of human Ero1Lalpha. EMBO J 19: 4493-4502
Bessette PH, Aslund F, Beckwith J & Georgiou G (1999) Efficient folding of proteins with
multiple disulfide bonds in the Escherichia coli cytoplasm. Proc Natl Acad Sci U S A
96: 13703-13708
Bihlmaier K, Mesecke N, Terziyska N, Bien M, Hell K & Herrmann JM (2007) The
disulfide relay system of mitochondria is connected to the respiratory chain. J Cell
Biol 179: 389-395
Burny A (1997) Prions and the problems they raise. Bull Mem Acad R Med Belg 152:
247-263
Byrne LJ, Sidhu A, Wallis AK, Ruddock LW, Freedman RB, Howard MJ & Williamson
RA (2009) Mapping of the ligand-binding site on the b' domain of human PDI:
interaction with peptide ligands and the x-linker region. Biochem J 423: 209-217
Cabibbo A, Pagani M, Fabbri M, Rocchi M, Farmery MR, Bulleid NJ & Sitia R (2000)
ERO1-L, a human protein that favors disulfide bond formation in the endoplasmic
reticulum. J Biol Chem 275: 4827-4833
Cenci S, van Anken E & Sitia R (2011) Proteostenosis and plasma cell pathophysiology.
Curr Opin Cell Biol 23: 216-222
Chakravarthi S & Bulleid NJ (2004) Glutathione is required to regulate the formation of
native disulfide bonds within proteins entering the secretory pathway. J Biol Chem
279: 39872-39879
Chakravarthi S, Jessop CE & Bulleid NJ (2006) The role of glutathione in disulphide bond
formation and endoplasmic-reticulum-generated oxidative stress. EMBO Rep 7: 271275
60
Chen W, Shang WH, Adachi Y, Hirose K, Ferrari DM & Kamata T (2008) A possible
biochemical link between NADPH oxidase (Nox) 1 redox-signalling and ERp72.
Biochem J 416: 55-63
Chen Y, Zhang Y, Yin Y, Gao G, Li S, Jiang Y, Gu X & Luo J (2005) SPD--a web-based
secreted protein database. Nucleic Acids Res 33: D169-173
Chivers PT, Prehoda KE & Raines RT (1997) The CXXC motif: a rheostat in the active
site. Biochemistry 36: 4061-4066
Cho SH & Beckwith J (2009) Two snapshots of electron transport across the membrane:
insights into the structure and function of DsbD. J Biol Chem 284: 11416-11424
Cumming RC, Andon NL, Haynes PA, Park M, Fischer WH & Schubert D (2004) Protein
disulfide bond formation in the cytoplasm during oxidative stress. J Biol Chem 279:
21749-21758
Daithankar VN, Schaefer SA, Dong M, Bahnson BJ & Thorpe C (2010) Structure of the
human sulfhydryl oxidase augmenter of liver regeneration and characterization of a
human mutation causing an autosomal recessive myopathy. Biochemistry 49: 67376745
Demmer J, Zhou C & Hubbard MJ (1997) Molecular cloning of ERp29, a novel and
widely expressed resident of the endoplasmic reticulum. FEBS Lett 402: 145-150
Desilva MG, Lu J, Donadel G, Modi WS, Xie H, Notkins AL & Lan MS (1996)
Characterization and chromosomal localization of a new protein disulfide isomerase,
PDIp, highly expressed in human pancreas. DNA Cell Biol 15: 9-16
Desilva MG, Notkins AL & Lan MS (1997) Molecular characterization of a pancreasspecific protein disulfide isomerase, PDIp. DNA Cell Biol 16: 269-274
Dong G, Wearsch PA, Peaper DR, Cresswell P & Reinisch KM (2009) Insights into MHC
class I peptide loading from the structure of the tapasin-ERp57 thiol oxidoreductase
heterodimer. Immunity 30: 21-32
Dutton RJ, Boyd D, Berkmen M & Beckwith J (2008) Bacterial species exhibit diversity in
their mechanisms and capacity for protein disulfide bond formation. Proc Natl Acad
Sci U S A 105: 11933-11938
Dutton RJ, Wayman A, Wei JR, Rubin EJ, Beckwith J & Boyd D (2010) Inhibition of
bacterial disulfide bond formation by the anticoagulant warfarin. Proc Natl Acad Sci
U S A 107: 297-301
Elledge SJ, Zhou Z & Allen JB (1992) Ribonucleotide reductase: regulation, regulation,
regulation. Trends Biochem Sci 17: 119-123
Ellgaard L & Ruddock LW (2005) The human protein disulphide isomerase family:
substrate interactions and functional properties. EMBO Rep 6: 28-32
Ferrari DM, Nguyen Van P, Kratzin HD & Soling HD (1998) ERp28, a human
endoplasmic-reticulum-lumenal protein, is a member of the protein disulfide
isomerase family but lacks a CXXC thioredoxin-box motif. Eur J Biochem 255: 570579
Fewell SW, Travers KJ, Weissman JS & Brodsky JL (2001) The action of molecular
chaperones in the early secretory pathway. Annu Rev Genet 35: 149-191
61
Frand AR, Cuozzo JW & Kaiser CA (2000) Pathways for protein disulphide bond
formation. Trends Cell Biol 10: 203-210
Frand AR & Kaiser CA (1998) The ERO1 gene of yeast is required for oxidation of
protein dithiols in the endoplasmic reticulum. Mol Cell 1: 161-170
Frand AR & Kaiser CA (1999) Ero1p oxidizes protein disulfide isomerase in a pathway for
disulfide bond formation in the endoplasmic reticulum. Mol Cell 4: 469-477
Frand AR & Kaiser CA (2000) Two pairs of conserved cysteines are required for the
oxidative activity of Ero1p in protein disulfide bond formation in the endoplasmic
reticulum. Mol Biol Cell 11: 2833-2843
Freedman RB (1995) The formation of protein disulphide bonds. Curr Opin Struct Biol 5:
85-91
Freedman RB, Klappa P & Ruddock LW (2002) Protein disulfide isomerases exploit
synergy between catalytic and specific binding domains. EMBO Rep 3: 136-140
Fu X, Wang P, Fukui M, Long C, Yin L, Choi HJ & Zhu BT (2012) PDIp is a major
intracellular oestrogen-storage protein that modulates tissue levels of oestrogen in the
pancreas. Biochem J 447: 115-123
Fu XM, Dai X, Ding J & Zhu BT (2009) Pancreas-specific protein disulfide isomerase has
a cell type-specific expression in various mouse tissues and is absent in human
pancreatic adenocarcinoma cells: implications for its functions. J Mol Histol 40: 189199
Gane PJ, Freedman RB & Warwicker J (1995) A molecular model for the redox potential
difference between thioredoxin and DsbA, based on electrostatics calculations. J Mol
Biol 249: 376-387
Garbi N, Hammerling G & Tanaka S (2007) Interaction of ERp57 and tapasin in the
generation of MHC class I-peptide complexes. Curr Opin Immunol 19: 99-105
Garbi N, Tanaka S, Momburg F & Hammerling GJ (2006) Impaired assembly of the major
histocompatibility complex class I peptide-loading complex in mice deficient in the
oxidoreductase ERp57. Nat Immunol 7: 93-102
Goldberger RF, Epstein CJ & Anfinsen CB (1964) Purification and Properties of a
Microsomal Enzyme System Catalyzing the Reactivation of Reduced Ribonuclease
and Lysozyme. J Biol Chem 239: 1406-1410
Goldstone D, Haebel PW, Katzen F, Bader MW, Bardwell JC, Beckwith J & Metcalf P
(2001) DsbC activation by the N-terminal domain of DsbD. Proc Natl Acad Sci U S A
98: 9551-9556
Goodstadt L & Ponting CP (2004) Vitamin K epoxide reductase: homology, active site and
catalytic mechanism. Trends Biochem Sci 29: 289-292
Goplen D, Wang J, Enger PO, Tysnes BB, Terzis AJ, Laerum OD & Bjerkvig R (2006)
Protein disulfide isomerase expression is related to the invasive properties of
malignant glioma. Cancer Res 66: 9895-9902
Grantham J, Ruddock LW, Roobol A & Carden MJ (2002) Eukaryotic chaperonin
containing T-complex polypeptide 1 interacts with filamentous actin and reduces the
initial rate of actin polymerization in vitro. Cell Stress Chaperones 7: 235-242
62
Gross E, Sevier CS, Heldman N, Vitu E, Bentzur M, Kaiser CA, Thorpe C & Fass D (2006)
Generating disulfides enzymatically: reaction products and electron acceptors of the
endoplasmic reticulum thiol oxidase Ero1p. Proc Natl Acad Sci U S A 103: 299-304
Gross E, Sevier CS, Vala A, Kaiser CA & Fass D (2002) A new FAD-binding fold and
intersubunit disulfide shuttle in the thiol oxidase Erv2p. Nat Struct Biol 9: 61-67
Gutscher M, Pauleau AL, Marty L, Brach T, Wabnitz GH, Samstag Y, Meyer AJ & Dick
TP (2008) Real-time imaging of the intracellular glutathione redox potential. Nat
Methods 5: 553-559
Ghuysen JM, Hackenbeck R (1994) Bacterial Cell Wall: Elsevier.
Hansen HG, Soltoft CL, Schmidt JD, Birk J, Appenzeller-Herzog C & Ellgaard L (2014)
Biochemical evidence that regulation of Ero1beta activity in human cells does not
involve the isoform-specific cysteine 262. Biosci Rep
Harju TH, Peltoniemi MJ, Rytila PH, Soini Y, Salmenkivi KM, Board PG, Ruddock LW
& Kinnula VL (2007) Glutathione S-transferase omega in the lung and sputum
supernatants of COPD patients. Respir Res 8: 48
Harper JD & Lansbury PT, Jr. (1997) Models of amyloid seeding in Alzheimer's disease
and scrapie: mechanistic truths and physiological consequences of the time-dependent
solubility of amyloid proteins. Annu Rev Biochem 66: 385-407
Hartl FU & Hayer-Hartl M (2009) Converging concepts of protein folding in vitro and in
vivo. Nat Struct Mol Biol 16: 574-581
Hatahet F & Nguyen VD, Salo KE, Ruddock LW (2010) Disruption of reducing pathways
is not essential for efficient disulfide bond formation in the cytoplasm of E. coli.
Microb Cell Fact 9: 1475-2859
Hatahet F & Ruddock LW (2007) Substrate recognition by the protein disulfide isomerases.
Febs J 274: 5223-5234
Hatahet F & Ruddock LW (2009) Protein disulfide isomerase: a critical evaluation of its
function in disulfide bond formation. Antioxid Redox Signal 11: 2807-2850
Hatahet F & Ruddock LW (2013) Topological plasticity of enzymes involved in disulfide
bond formation allows catalysis in either the periplasm or the cytoplasm. J Mol Biol
425: 3268-3276
Hayano T & Kikuchi M (1995) Molecular cloning of the cDNA encoding a novel protein
disulfide isomerase-related protein (PDIR). FEBS Lett 372: 210-214
Heckler EJ, Rancy PC, Kodali VK & Thorpe C (2008) Generating disulfides with the
Quiescin-sulfhydryl oxidases. Biochim Biophys Acta 4: 567-577
Helenius A & Aebi M (2004) Roles of N-linked glycans in the endoplasmic reticulum.
Annu Rev Biochem 73: 1019-1049
Heras B, Edeling MA, Schirra HJ, Raina S & Martin JL (2004) Crystal structures of the
DsbG disulfide isomerase reveal an unstable disulfide. Proc Natl Acad Sci U S A 101:
8876-8881
Herrmann JM & Kohl R (2007) Catch me if you can! Oxidative protein trapping in the
intermembrane space of mitochondria. J Cell Biol 176: 559-563
Herrmann JM & Neupert W (2000) Protein transport into mitochondria. Curr Opin
Microbiol 3: 210-214
63
Higo T, Hattori M, Nakamura T, Natsume T, Michikawa T & Mikoshiba K (2005)
Subtype-specific and ER lumenal environment-dependent regulation of inositol 1,4,5trisphosphate receptor type 1 by ERp44. Cell 120: 85-98
Hirsch I, Weiwad M, Prell E & Ferrari DM (2014) ERp29 deficiency affects sensitivity to
apoptosis via impairment of the ATF6-CHOP pathway of stress response. Apoptosis
19: 801-815
Hollak CE & Wijburg FA (2014) Treatment of lysosomal storage disorders: successes and
challenges. J Inherit Metab Dis
Hoober KL, Sheasley SL, Gilbert HF & Thorpe C (1999) Sulfhydryl oxidase from egg
white. A facile catalyst for disulfide bond formation in proteins and peptides. J Biol
Chem 274: 22147-22150
Horibe T, Gomi M, Iguchi D, Ito H, Kitamura Y, Masuoka T, Tsujimoto I, Kimura T &
Kikuchi M (2004) Different contributions of the three CXXC motifs of human
protein-disulfide isomerase-related protein to isomerase activity and oxidative
refolding. J Biol Chem 279: 4604-4611
Hund E, Singer R, Linke RP, Willig F & Grau A (2002) [Hereditary amyloidoses
associated with transthyretin mutations]. Nervenarzt 73: 930-936
Hurtley SM & Helenius A (1989) Protein oligomerization in the endoplasmic reticulum.
Annu Rev Cell Biol 5: 277-307
Hwang C, Sinskey AJ & Lodish HF (1992) Oxidized redox state of glutathione in the
endoplasmic reticulum. Science 257: 1496-1502
Ikawa M, Wada I, Kominami K, Watanabe D, Toshimori K, Nishimune Y & Okabe M
(1997) The putative chaperone calmegin is required for sperm fertility. Nature 387:
607-611
Inaba K, Masui S, Iida H, Vavassori S, Sitia R & Suzuki M (2010) Crystal structures of
human Ero1alpha reveal the mechanisms of regulated and targeted oxidation of PDI.
EMBO J 29: 3330-3343
Irvine AG, Wallis AK, Sanghera N, Rowe ML, Ruddock LW, Howard MJ, Williamson
RA, Blindauer CA & Freedman RB (2014) Protein disulfide-isomerase interacts with
a substrate protein at all stages along its folding pathway. PLoS One 9
Ito K & Inaba K (2008) The disulfide bond formation (Dsb) system. Curr Opin Struct Biol
18: 450-458
Janiszewski M, Lopes LR, Carmo AO, Pedro MA, Brandes RP, Santos CX & Laurindo FR
(2005) Regulation of NAD(P)H oxidase by associated protein disulfide isomerase in
vascular smooth muscle cells. J Biol Chem 280: 40813-40819
Jeong W, Lee DY, Park S & Rhee SG (2008) ERp16, an endoplasmic reticulum-resident
thiol-disulfide oxidoreductase: biochemical properties and role in apoptosis induced
by endoplasmic reticulum stress. J Biol Chem 283: 25557-25566
Jessop CE, Watkins RH, Simmons JJ, Tasab M & Bulleid NJ (2009) Protein disulphide
isomerase family members show distinct substrate specificity: P5 is targeted to BiP
client proteins. J Cell Sci 122: 4287-4295
64
Jin DY, Tie JK & Stafford DW (2007) The conversion of vitamin K epoxide to vitamin K
quinone and vitamin K quinone to vitamin K hydroquinone uses the same active site
cysteines. Biochemistry 46: 7279-7283
Jordan PA, Stevens JM, Hubbard GP, Barrett NE, Sage T, Authi KS & Gibbins JM (2005)
A role for the thiol isomerase protein ERP5 in platelet function. Blood 105: 15001507
Jucker M & Walker LC (2011) Pathogenic protein seeding in Alzheimer disease and other
neurodegenerative disorders. Ann Neurol 70: 532-540
Kadokura H & Beckwith J (2002) Four cysteines of the membrane protein DsbB act in
concert to oxidize its substrate DsbA. EMBO J 21: 2354-2363
Kadokura H & Beckwith J (2010) Mechanisms of oxidative protein folding in the bacterial
cell envelope. Antioxid Redox Signal 13: 1231-1246
Karala AR, Lappi AK & Ruddock LW (2010) Modulation of an active-site cysteine pKa
allows PDI to act as a catalyst of both disulfide bond formation and isomerization. J
Mol Biol 396: 883-892
Karala AR, Lappi AK, Saaranen MJ & Ruddock LW (2009) Efficient peroxide-mediated
oxidative refolding of a protein at physiological pH and implications for oxidative
folding in the endoplasmic reticulum. Antioxid Redox Signal 11: 963-970
Karala AR, Psarrakos P, Ruddock LW & Klappa P (2007) Protein disulfide isomerases
from C. elegans are equally efficient at thiol-disulfide exchange in simple peptidebased systems but show differences in reactivity towards protein substrates. Antioxid
Redox Signal 9: 1815-1823
Karala AR & Ruddock LW (2010) Bacitracin is not a specific inhibitor of protein disulfide
isomerase. Febs J 277: 2454-2462
Katzen F & Beckwith J (2000) Transmembrane electron transfer by the membrane protein
DsbD occurs via a disulfide bond cascade. Cell 103: 769-779
Kemmink J, Darby NJ, Dijkstra K, Nilges M & Creighton TE (1996) Structure
determination of the N-terminal thioredoxin-like domain of protein disulfide
isomerase using multidimensional heteronuclear 13C/15N NMR spectroscopy.
Biochemistry 35: 7684-7691
Kemmink J, Dijkstra K, Mariani M, Scheek RM, Penka E, Nilges M & Darby NJ (1999)
The structure in solution of the b domain of protein disulfide isomerase. J Biomol
NMR 13: 357-368
Kettner K, Blomberg A & Rodel G (2004) Schizosaccharomyces pombe ER
oxidoreductin-like proteins SpEro1a p and SpEro1b p. Yeast 21: 1035-1044
Klappa P, Ruddock LW, Darby NJ & Freedman RB (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
Kleizen B, van Vlijmen T, de Jonge HR & Braakman I (2005) Folding of CFTR is
predominantly cotranslational. Mol Cell 20: 277-287
Knowles TP, Vendruscolo M & Dobson CM (2014) The amyloid state and its association
with protein misfolding diseases. Nat Rev Mol Cell Biol 15: 384-396
65
Kobayashi T, Takahashi Y & Ito K (2001) Identification of a segment of DsbB essential
for its respiration-coupled oxidation. Mol Microbiol 39: 158-165
Kober FX, Koelmel W, Kuper J, Drechsler J, Mais C, Hermanns HM & Schindelin H
(2013) The crystal structure of the protein-disulfide isomerase family member ERp27
provides insights into its substrate binding capabilities. J Biol Chem 288: 2029-2039
Kodali VK & Thorpe C (2010) Oxidative protein folding and the Quiescin-sulfhydryl
oxidase family of flavoproteins. Antioxid Redox Signal 13: 1217-1230
Koehler CM (2004) New developments in mitochondrial assembly. Annu Rev Cell Dev
Biol 20: 309-335
Koivunen P, Salo KE, Myllyharju J & Ruddock LW (2005) Three binding sites in proteindisulfide isomerase cooperate in collagen prolyl 4-hydroxylase tetramer assembly. J
Biol Chem 280: 5227-5235
Kojima R, Okumura M, Masui S, Kanemura S, Inoue M, Saiki M, Yamaguchi H, Hikima
T, Suzuki M, Akiyama S & Inaba K (2014) Radically different thioredoxin domain
arrangement of ERp46, an efficient disulfide bond introducer of the mammalian PDI
family. Structure 22: 431-443
Kozlov G, Maattanen P, Schrag JD, Hura GL, Gabrielli L, Cygler M, Thomas DY &
Gehring K (2009) Structure of the noncatalytic domains and global fold of the protein
disulfide isomerase ERp72. Structure 17: 651-659
Lappi AK, Lensink MF, Alanen HI, Salo KE, Lobell M, Juffer AH & Ruddock LW (2004)
A conserved arginine plays a role in the catalytic cycle of the protein disulphide
isomerases. J Mol Biol 335: 283-295
Lappi AK & Ruddock LW (2011) Reexamination of the role of interplay between
glutathione and protein disulfide isomerase. J Mol Biol 409: 238-249
Leach MR, Cohen-Doyle MF, Thomas DY & Williams DB (2002) Localization of the
lectin, ERp57 binding, and polypeptide binding sites of calnexin and calreticulin. J
Biol Chem 277: 29686-29697
Lee S, Park B, Kang K & Ahn K (2009) Redox-regulated export of the major
histocompatibility complex class I-peptide complexes from the endoplasmic reticulum.
Mol Biol Cell 20: 3285-3294
Li W, Schulman S, Dutton RJ, Boyd D, Beckwith J & Rapoport TA (2010) Structure of a
bacterial homologue of vitamin K epoxide reductase. Nature 463: 507-512
Liberman M, Bassi E, Martinatti MK, Lario FC, Wosniak J, Jr., Pomerantzeff PM &
Laurindo FR (2008) Oxidant generation predominates around calcifying foci and
enhances progression of aortic valve calcification. Arterioscler Thromb Vasc Biol 28:
463-470
Limauro D, De Simone G, Pirone L, Bartolucci S, D'Ambrosio K & Pedone E (2014)
Sulfolobus solfataricus thiol redox puzzle: characterization of an atypical protein
disulfide oxidoreductase. Extremophiles 18: 219-228
Madonna S, Papa R, Birolo L, Autore F, Doti N, Marino G, Quemeneur E, Sannia G,
Tutino ML & Duilio A (2006) The thiol-disulfide oxidoreductase system in the coldadapted bacterium Pseudoalteromonas haloplanktis TAC 125: discovery of a novel
disulfide oxidoreductase enzyme. Extremophiles 10: 41-51
66
Malhotra JD, Miao H, Zhang K, Wolfson A, Pennathur S, Pipe SW & Kaufman RJ (2008)
Antioxidants reduce endoplasmic reticulum stress and improve protein secretion. Proc
Natl Acad Sci U S A 105: 18525-18530
Matos CF, Robinson C, Alanen HI, Prus P, Uchida Y, Ruddock LW, Freedman RB &
Keshavarz-Moore E (2014) Efficient export of prefolded, disulfide-bonded
recombinant proteins to the periplasm by the Tat pathway in Escherichia coli
CyDisCo strains. Biotechnol Prog 30: 281-290
Matsuo Y, Masutani H, Son A, Kizaka-Kondoh S & Yodoi J (2009) Physical and
functional interaction of transmembrane thioredoxin-related protein with major
histocompatibility complex class I heavy chain: redox-based protein quality control
and its potential relevance to immune responses. Mol Biol Cell 20: 4552-4562
May JM (2012) Vitamin C transport and its role in the central nervous system. Subcell
Biochem 56: 85-103
McCarthy AA, Haebel PW, Torronen A, Rybin V, Baker EN & Metcalf P (2000) Crystal
structure of the protein disulfide bond isomerase, DsbC, from Escherichia coli. Nat
Struct Biol 7: 196-199
Mesecke N, Terziyska N, Kozany C, Baumann F, Neupert W, Hell K & Herrmann JM
(2005) A disulfide relay system in the intermembrane space of mitochondria that
mediates protein import. Cell 121: 1059-1069
Molinari M, Galli C, Piccaluga V, Pieren M & Paganetti P (2002) Sequential assistance of
molecular chaperones and transient formation of covalent complexes during protein
degradation from the ER. J Cell Biol 158: 247-257
Morikawa K, Gouttenoire J, Hernandez C, Dao Thi VL, Tran HT, Lange CM, Dill MT,
Heim MH, Donze O, Penin F, Quadroni M & Moradpour D (2014) Quantitative
proteomics identifies the membrane-associated peroxidase GPx8 as a cellular
substrate of the hepatitis C virus NS3-4A protease. Hepatology 59: 423-433
Nguyen VD, Hatahet F, Salo KE, Enlund E, Zhang C & Ruddock LW (2011) Preexpression of a sulfhydryl oxidase significantly increases the yields of eukaryotic
disulfide bond containing proteins expressed in the cytoplasm of E.coli. Microb Cell
Fact 10: 1475-2859
Nguyen VD, Saaranen MJ, Karala AR, Lappi AK, Wang L, Raykhel IB, Alanen HI, Salo
KE, Wang CC & Ruddock LW (2011) Two endoplasmic reticulum PDI peroxidases
increase the efficiency of the use of peroxide during disulfide bond formation. J Mol
Biol 406: 503-515
Nguyen VD, Wallis K, Howard MJ, Haapalainen AM, Salo KE, Saaranen MJ, Sidhu A,
Wierenga RK, Freedman RB, Ruddock LW & Williamson RA (2008) Alternative
conformations of the x region of human protein disulphide-isomerase modulate
exposure of the substrate binding b' domain. J Mol Biol 383: 1144-1155
Oka OB & Bulleid NJ (2013) Forming disulfides in the endoplasmic reticulum. Biochim
Biophys Acta 1833: 2425-2429
Oka OB, Pringle MA, Schopp IM, Braakman I & Bulleid NJ (2013) ERdj5 is the ER
reductase that catalyzes the removal of non-native disulfides and correct folding of the
LDL receptor. Mol Cell 50: 793-804
67
Olivari S, Cali T, Salo KE, Paganetti P, Ruddock LW & Molinari M (2006) EDEM1
regulates ER-associated degradation by accelerating de-mannosylation of foldingdefective polypeptides and by inhibiting their covalent aggregation. Biochem Biophys
Res Commun 349: 1278-1284
Olivari S, Galli C, Alanen H, Ruddock L & Molinari M (2005) A novel stress-induced
EDEM variant regulating endoplasmic reticulum-associated glycoprotein degradation.
J Biol Chem 280: 2424-2428
Pace CN, Shirley BA, McNutt M & Gajiwala K (1996) Forces contributing to the
conformational stability of proteins. FASEB J 10: 75-83
Pagani M, Fabbri M, Benedetti C, Fassio A, Pilati S, Bulleid NJ, Cabibbo A & Sitia R
(2000) Endoplasmic reticulum oxidoreductin 1-lbeta (ERO1-Lbeta), a human gene
induced in the course of the unfolded protein response. J Biol Chem 275: 2368523692
Palomares LA, Estrada-Mondaca S & Ramirez OT (2004) Production of recombinant
proteins: challenges and solutions. Methods Mol Biol 267: 15-52
Park SW, Zhen G, Verhaeghe C, Nakagami Y, Nguyenvu LT, Barczak AJ, Killeen N &
Erle DJ (2009) The protein disulfide isomerase AGR2 is essential for production of
intestinal mucus. Proc Natl Acad Sci U S A 106: 6950-6955
Peaper DR, Wearsch PA & Cresswell P (2005) Tapasin and ERp57 form a stable disulfidelinked dimer within the MHC class I peptide-loading complex. Embo J 24: 3613-3623
Peltoniemi MJ, Karala AR, Jurvansuu JK, Kinnula VL & Ruddock LW (2006) Insights
into deglutathionylation reactions. Different intermediates in the glutaredoxin and
protein disulfide isomerase catalyzed reactions are defined by the gamma-linkage
present in glutathione. J Biol Chem 281: 33107-33114
Peltoniemi MJ, Rytila PH, Harju TH, Soini YM, Salmenkivi KM, Ruddock LW & Kinnula
VL (2006) Modulation of glutaredoxin in the lung and sputum of cigarette smokers
and chronic obstructive pulmonary disease. Respir Res 7: 133
Pennisi E (2012) Society for Integrative and Comparative Biology. Robotic fish point to
schooling gene. Science 335: 276-277
Pihlajaniemi T, Helaakoski T, Tasanen K, Myllyla R, Huhtala ML, Koivu J & Kivirikko
KI (1987) Molecular cloning of the beta-subunit of human prolyl 4-hydroxylase. This
subunit and protein disulphide isomerase are products of the same gene. EMBO J 6:
643-649
Pirneskoski A, Klappa P, Lobell M, Williamson RA, Byrne L, Alanen HI, Salo KE,
Kivirikko KI, Freedman RB & Ruddock LW (2004) Molecular characterization of the
principal substrate binding site of the ubiquitous folding catalyst protein disulfide
isomerase. J Biol Chem 279: 10374-10381
Pollard MG, Travers KJ & Weissman JS (1998) Ero1p: a novel and ubiquitous protein
with an essential role in oxidative protein folding in the endoplasmic reticulum. Mol
Cell 1: 171-182
68
Pons V, Rolland C, Nauze M, Danjoux M, Gaibelet G, Durandy A, Sassolas A, Levy E,
Terce F, Collet X & Mas E (2011) A severe form of abetalipoproteinemia caused by
new splicing mutations of microsomal triglyceride transfer protein (MTTP). Hum
Mutat 32: 751-759
Prinz WA, Aslund F, Holmgren A & Beckwith J (1997) The role of the thioredoxin and
glutaredoxin pathways in reducing protein disulfide bonds in the Escherichia coli
cytoplasm. J Biol Chem 272: 15661-15667
Rainey-Barger EK, Mkrtchian S & Tsai B (2007) Dimerization of ERp29, a PDI-like
protein, is essential for its diverse functions. Mol Biol Cell 18: 1253-1260
Ramming T, Hansen HG, Nagata K, Ellgaard L & Appenzeller-Herzog C (2014) GPx8
peroxidase prevents leakage of H2O2 from the endoplasmic reticulum. Free Radic
Biol Med 70: 106-116
Rancy PC & Thorpe C (2008) Oxidative protein folding in vitro: a study of the cooperation
between quiescin-sulfhydryl oxidase and protein disulfide isomerase. Biochemistry 47:
12047-12056
Ratjen F & Doring G (2003) Cystic fibrosis. Lancet 361: 681-689
Raykhel I, Alanen H, Salo K, Jurvansuu J, Nguyen VD, Latva-Ranta M & Ruddock L
(2007) A molecular specificity code for the three mammalian KDEL receptors. J Cell
Biol 179: 1193-1204
Reinhardt C, von Bruhl ML, Manukyan D, Grahl L, Lorenz M, Altmann B, Dlugai S, Hess
S, Konrad I, Orschiedt L, Mackman N, Ruddock L, Massberg S & Engelmann B
(2008) Protein disulfide isomerase acts as an injury response signal that enhances
fibrin generation via tissue factor activation. J Clin Invest 118: 1110-1122
Riemer J, Bulleid N & Herrmann JM (2009) Disulfide formation in the ER and
mitochondria: two solutions to a common process. Science 324: 1284-1287
Ritz D & Beckwith J (2001) Roles of thiol-redox pathways in bacteria. Annu Rev
Microbiol 55: 21-48
Rowe ML, Ruddock LW, Kelly G, Schmidt JM, Williamson RA & Howard MJ (2009)
Solution structure and dynamics of ERp18, a small endoplasmic reticulum resident
oxidoreductase. Biochemistry 48: 4596-4606
Ruddock LW (2006) Gaining access to ERp57 function. Structure. 14(8):1209-1210.
Ruddock LW (2012) Low-molecular-weight oxidants involved in disulfide bond formation.
Antioxid Redox Signal 16: 1129-1138
Ruddock LW & Molinari M (2006) N-glycan processing in ER quality control. J Cell Sci
119: 4373-4380
Russell SJ, Ruddock LW, Salo KE, Oliver JD, Roebuck QP, Llewellyn DH, Roderick HL,
Koivunen P, Myllyharju J & High S (2004) The primary substrate binding site in the
b' domain of ERp57 is adapted for endoplasmic reticulum lectin association. J Biol
Chem 279: 18861-18869
Rutkevich LA & Williams DB (2012) Vitamin K epoxide reductase contributes to protein
disulfide formation and redox homeostasis within the endoplasmic reticulum. Mol
Biol Cell 23: 2017-2027
69
Ryser HJ & Fluckiger R (2005) Progress in targeting HIV-1 entry. Drug Discov Today 10:
1085-1094
Saaranen MJ, Karala AR, Lappi AK & Ruddock LW (2010) The role of dehydroascorbate
in disulfide bond formation. Antioxid Redox Signal 12: 15-25
Saaranen MJ & Ruddock LW (2013) Disulfide bond formation in the cytoplasm. Antioxid
Redox Signal 19: 46-53
Saaranen MJ, Salo KE, Latva-Ranta MK, Kinnula VL & Ruddock LW (2009) The Cterminal active site cysteine of Escherichia coli glutaredoxin 1 determines the
glutathione specificity of the second step of peptide deglutathionylation. Antioxid
Redox Signal 11: 1819-1828
Sadler JE (2009) von Willebrand factor assembly and secretion. J Thromb Haemost 7
Suppl 1: 24-27
Sato Y, Kojima R, Okumura M, Hagiwara M, Masui S, Maegawa K, Saiki M, Horibe T,
Suzuki M & Inaba K (2013) Synergistic cooperation of PDI family members in
peroxiredoxin 4-driven oxidative protein folding. Sci Rep 3: 2456
Schroder M & Kaufman RJ (2005) The mammalian unfolded protein response. Annu Rev
Biochem 74: 739-789
Schulman S, Wang B, Li W & Rapoport TA (2010) Vitamin K epoxide reductase prefers
ER membrane-anchored thioredoxin-like redox partners. Proc Natl Acad Sci U S A
107: 15027-15032
Schwaller M, Wilkinson B & Gilbert HF (2003) Reduction-reoxidation cycles contribute
to catalysis of disulfide isomerization by protein-disulfide isomerase. J Biol Chem 278:
7154-7159
Sevier CS (2010) New insights into oxidative folding. J Cell Biol 188: 757-758
Sevier CS & Kaiser CA (2006) Disulfide transfer between two conserved cysteine pairs
imparts selectivity to protein oxidation by Ero1. Mol Biol Cell 17: 2256-2266
Sevier CS, Qu H, Heldman N, Gross E, Fass D & Kaiser CA (2007) Modulation of cellular
disulfide-bond formation and the ER redox environment by feedback regulation of
Ero1. Cell 129: 333-344
Silverman EK & Sandhaus RA (2009) Clinical practice. Alpha1-antitrypsin deficiency. N
Engl J Med 360: 2749-2757
Stockton JD, Merkert MC & Kellaris KV (2003) A complex of chaperones and disulfide
isomerases occludes the cytosolic face of the translocation protein Sec61p and affects
translocation of the prion protein. Biochemistry 42: 12821-12834
Sullivan DC, Huminiecki L, Moore JW, Boyle JJ, Poulsom R, Creamer D, Barker J &
Bicknell R (2003) EndoPDI, a novel protein-disulfide isomerase-like protein that is
preferentially expressed in endothelial cells acts as a stress survival factor. J Biol
Chem 278: 47079-47088
Takahashi YH, Inaba K & Ito K (2004) Characterization of the menaquinone-dependent
disulfide bond formation pathway of Escherichia coli. J Biol Chem 279: 47057-47065
Tavender TJ & Bulleid NJ (2010a) Molecular mechanisms regulating oxidative activity of
the Ero1 family in the endoplasmic reticulum. Antioxid Redox Signal 13: 1177-1187
70
Tavender TJ & Bulleid NJ (2010b) Peroxiredoxin IV protects cells from oxidative stress
by removing H2O2 produced during disulphide formation. J Cell Sci 123: 2672-2679
Tavender TJ, Springate JJ & Bulleid NJ (2010) Recycling of peroxiredoxin IV provides a
novel pathway for disulphide formation in the endoplasmic reticulum. EMBO J 29:
4185-4197
Taylor RW & Turnbull DM (2005) Mitochondrial DNA mutations in human disease. Nat
Rev Genet 6: 389-402
Taylor SW, Fahy E & Ghosh SS (2003) Global organellar proteomics. Trends Biotechnol
21: 82-88
Taylor SW, Fahy E, Zhang B, Glenn GM, Warnock DE, Wiley S, Murphy AN, Gaucher
SP, Capaldi RA, Gibson BW & Ghosh SS (2003) Characterization of the human heart
mitochondrial proteome. Nat Biotechnol 21: 281-286
Thorpe C, Hoober KL, Raje S, Glynn NM, Burnside J, Turi GK & Coppock DL (2002)
Sulfhydryl oxidases: emerging catalysts of protein disulfide bond formation in
eukaryotes. Arch Biochem Biophys 405: 1-12
Tian G, Xiang S, Noiva R, Lennarz WJ & Schindelin H (2006) The crystal structure of
yeast protein disulfide isomerase suggests cooperativity between its active sites. Cell
124: 61-73
Tokatlidis K (2005) A disulfide relay system in mitochondria. Cell 121: 965-967
Tokuhiro K, Ikawa M, Benham AM & Okabe M (2012) Protein disulfide isomerase
homolog PDILT is required for quality control of sperm membrane protein ADAM3
and male fertility [corrected]. Proc Natl Acad Sci U S A 109: 3850-3855
Toppo S, Vanin S, Bosello V & Tosatto SC (2008) Evolutionary and structural insights
into the multifaceted glutathione peroxidase (Gpx) superfamily. Antioxid Redox
Signal 10: 1501-1514
Tu BP, Ho-Schleyer SC, Travers KJ & Weissman JS (2000) Biochemical basis of
oxidative protein folding in the endoplasmic reticulum. Science 290: 1571-1574
Urade R, Okudo H, Kato H, Moriyama T & Arakaki Y (2004) ER-60 domains responsible
for interaction with calnexin and calreticulin. Biochemistry 43: 8858-8868
Ushioda R, Hoseki J, Araki K, Jansen G, Thomas DY & Nagata K (2008) ERdj5 is
required as a disulfide reductase for degradation of misfolded proteins in the ER.
Science 321: 569-572
Wajih N, Hutson SM & Wallin R (2007) Disulfide-dependent protein folding is linked to
operation of the vitamin K cycle in the endoplasmic reticulum. A protein disulfide
isomerase-VKORC1 redox enzyme complex appears to be responsible for vitamin K1
2,3-epoxide reduction. J Biol Chem 282: 2626-2635
Wajih N, Sane DC, Hutson SM & Wallin R (2005) Engineering of a recombinant vitamin
K-dependent gamma-carboxylation system with enhanced gamma-carboxyglutamic
acid forming capacity: evidence for a functional CXXC redox center in the system. J
Biol Chem 280: 10540-10547
Wallin R, Wajih N & Hutson SM (2008) VKORC1: a warfarin-sensitive enzyme in
vitamin K metabolism and biosynthesis of vitamin K-dependent blood coagulation
factors. Vitam Horm 78: 227-246
71
Wallis AK, Sidhu A, Byrne LJ, Howard MJ, Ruddock LW, Williamson RA & Freedman
RB (2009) The ligand-binding b' domain of human protein disulphide-isomerase
mediates homodimerization. Protein Sci 18: 2569-2577
van Lith M, Hartigan N, Hatch J & Benham AM (2005) PDILT, a divergent testis-specific
protein disulfide isomerase with a non-classical SXXC motif that engages in disulfidedependent interactions in the endoplasmic reticulum. J Biol Chem 280: 1376-1383
van Lith M, Karala AR, Bown D, Gatehouse JA, Ruddock LW, Saunders PT & Benham
AM (2007) A developmentally regulated chaperone complex for the endoplasmic
reticulum of male haploid germ cells. Mol Biol Cell 18: 2795-2804
Van PN, Rupp K, Lampen A & Soling HD (1993) CaBP2 is a rat homolog of ERp72 with
proteindisulfide isomerase activity. Eur J Biochem 213: 789-795
Wang C, Li W, Ren J, Fang J, Ke H, Gong W, Feng W & Wang CC (2013) Structural
insights into the redox-regulated dynamic conformations of human protein disulfide
isomerase. Antioxid Redox Signal 19: 36-45
Wang L, Vavassori S, Li S, Ke H, Anelli T, Degano M, Ronzoni R, Sitia R, Sun F & Wang
CC (2008) Crystal structure of human ERp44 shows a dynamic functional modulation
by its carboxy-terminal tail. EMBO Rep 9: 642-647
Wang L, Zhang L, Niu Y, Sitia R & Wang CC (2014) Glutathione peroxidase 7 utilizes
hydrogen peroxide generated by Ero1alpha to promote oxidative protein folding.
Antioxid Redox Signal 20: 545-556
Wang ZV, Schraw TD, Kim JY, Khan T, Rajala MW, Follenzi A & Scherer PE (2007)
Secretion of the adipocyte-specific secretory protein adiponectin critically depends on
thiol-mediated protein retention. Mol Cell Biol 27: 3716-3731
Webb CT, Gorman MA, Lazarou M, Ryan MT & Gulbis JM (2006) Crystal structure of
the mitochondrial chaperone TIM9.10 reveals a six-bladed alpha-propeller. Mol Cell
21: 123-133
Veggiani G & de Marco A (2011) Improved quantitative and qualitative production of
single-domain intrabodies mediated by the co-expression of Erv1p sulfhydryl oxidase.
Protein Expr Purif 79: 111-114
Welch RW, Wang Y, Crossman A, Jr., Park JB, Kirk KL & Levine M (1995)
Accumulation of vitamin C (ascorbate) and its oxidized metabolite dehydroascorbic
acid occurs by separate mechanisms. J Biol Chem 270: 12584-12592
Wickner W & Schekman R (2005) Protein translocation across biological membranes.
Science 310: 1452-1456
Vinaik R, Kozlov G & Gehring K (2013) Structure of the non-catalytic domain of the
protein disulfide isomerase-related protein (PDIR) reveals function in protein binding.
PLoS One 8: e62021
Winterbourn CC & Metodiewa D (1999) Reactivity of biologically important thiol
compounds with superoxide and hydrogen peroxide. Free Radic Biol Med 27: 322328
Winther JR & Thorpe C (2014) Quantification of thiols and disulfides. Biochim Biophys
Acta 1840: 838-846
72
Zeskind BJ, Jordan CD, Timp W, Trapani L, Waller G, Horodincu V, Ehrlich DJ &
Matsudaira P (2007) Nucleic acid and protein mass mapping by live-cell deepultraviolet microscopy. Nat Methods 4: 567-569
Zhou Y, Cierpicki T, Jimenez RH, Lukasik SM, Ellena JF, Cafiso DS, Kadokura H,
Beckwith J & Bushweller JH (2008) NMR solution structure of the integral membrane
enzyme DsbB: functional insights into DsbB-catalyzed disulfide bond formation. Mol
Cell 31: 896-908
Zito E, Chin KT, Blais J, Harding HP & Ron D (2010a) ERO1-beta, a pancreas-specific
disulfide oxidase, promotes insulin biogenesis and glucose homeostasis. J Cell Biol
188: 821-832
Zito E, Melo EP, Yang Y, Wahlander A, Neubert TA & Ron D (2010b) Oxidative protein
folding by an endoplasmic reticulum-localized peroxiredoxin. Mol Cell 40: 787-797
Zito E, Hansen HG, Yeo GS, Fujii J & Ron D (2012) Endoplasmic reticulum thiol oxidase
deficiency leads to ascorbic acid depletion and noncanonical scurvy in mice. Mol Cell
48: 39-51
73
74
Original articles
I
Nguyen VD, Wallis K, Howard MJ, Haapalainen AM, Salo KE, Saaranen MJ, Sidhu
A, Wierenga RK, Freedman RB, Ruddock LW, Williamson RA (2008) Alternative
conformations of the x region of human protein disulphide-isomerase modulate
exposure of the substrate binding b' domain. J Mol Biol 383(5):1144-55.
II Nguyen VD, Saaranen MJ, Karala AR, Lappi AK, Wang L, Raykhel IB, Alanen HI,
Salo KE, Wang CC, Ruddock LW (2011) Two endoplasmic reticulum PDI
peroxidases increase the efficiency of the use of peroxide during disulfide bond
formation. J Mol Biol 406(3):503-15.
III Hatahet F, Nguyen VD, Salo KE, Ruddock LW (2010) Disruption of reducing
pathways is not essential for efficient disulfide bond formation in the cytoplasm of E.
coli. Microb Cell Fact 9(67):1-9.
IV Nguyen VD, Hatahet F, Salo KE, Enlund E, Zhang C, Ruddock LW (2011) Preexpression of a sulfhydryl oxidase significantly increases the yields of eukaryotic
disulfide bond containing proteins expressed in the cytoplasm of E. coli. Microb Cell
Fact 10(1):1-13.
Reprinted with permission from Elsevier (I, II) and BioMed Central (III, IV).
Original publications are not included in the electronic version of the dissertation.
75
76
ACTA UNIVERSITATIS OULUENSIS
SERIES D MEDICA
1267. Kiviniemi, Marjo (2014) Mortality, disability, psychiatric treatment and medication
in first-onset schizophrenia in Finland : the register linkage study
1268. Kallio, Merja (2014) Neurally adjusted ventilatory assist in pediatric intensive care
1269. Väyrynen, Juha (2014) Immune cell infiltration and inflammatory biomarkers in
colorectal cancer
1270. Silvola, Anna-Sofia (2014) Effect of treatment of severe malocclusion and related
factors on oral health-related quality of life
1271. Mäkelä, Tuomas (2014) Systemic transplantation of bone marrow stromal cells :
an experimental animal study of biodistribution and tissue targeting
1272. Junttila, Sanna (2014) Studies of kidney induction in vitro using gene expression
profiling and novel tissue manipulation technique
1273. Häkli, Sanna (2014) Childhood hearing impairment in northern Finland :
prevalence, aetiology and additional disabilities
1274. Lehto, Liisa (2014) Interactive two-step training and management strategy for
improvement of the quality of point-of-care testing by nurses : implementation of
the strategy in blood glucose measurement
1275. Harjunen, Vanessa (2014) Skin stem cells and tumor growth : functions of
collagen XVIII in hair follicle cycling and skin cancer, and Bmx tyrosine kinase in
tumor angiogenesis
1276. Savela, Salla (2014) Physical activity in midlife and health-related quality of life,
frailty, telomere length and mortality in old age
1277. Laitala, Anu (2014) Hypoxia-inducible factor prolyl 4-hydroxylases regulating
erythropoiesis, and hypoxia-inducible lysyl oxidase regulating skeletal muscle
development during embryogenesis
1278. Persson, Maria (2014) 3D woven scaffolds for bone tissue engineering
1279. Kallankari, Hanna (2014) Perinatal factors as predictors of brain damage and
neurodevelopmental outcome : study of children born very preterm
1280. Pietilä, Ilkka (2015) The role of Dkk1 and Wnt5a in mammalian kidney
development and disease
1281. Komulainen, Tuomas (2015) Disturbances in mitochondrial DNA maintenance in
neuromuscular disorders and valproate-induced liver toxicity
Book orders:
Granum: Virtual book store
http://granum.uta.fi/granum/
D 1282
OULU 2015
UNIVERSITY OF OUL U P.O. Box 8000 FI-90014 UNIVERSITY OF OULU FINLA ND
U N I V E R S I TAT I S
O U L U E N S I S
ACTA
A C TA
D 1282
ACTA
U N I V E R S I T AT I S O U L U E N S I S
Van Dat Nguyen
University Lecturer Santeri Palviainen
Postdoctoral research fellow Sanna Taskila
Van Dat Nguyen
Professor Esa Hohtola
MECHANISMS AND
APPLICATIONS OF DISULFIDE
BOND FORMATION
Professor Olli Vuolteenaho
University Lecturer Veli-Matti Ulvinen
Director Sinikka Eskelinen
Professor Jari Juga
University Lecturer Anu Soikkeli
Professor Olli Vuolteenaho
Publications Editor Kirsti Nurkkala
ISBN 978-952-62-0724-7 (Paperback)
ISBN 978-952-62-0725-4 (PDF)
ISSN 0355-3221 (Print)
ISSN 1796-2234 (Online)
UNIVERSITY OF OULU GRADUATE SCHOOL;
UNIVERSITY OF OULU,
FACULTY OF BIOCHEMISTRY AND MOLECULAR MEDICINE;
BIOCENTER OULU
D
MEDICA