Protein quality in bacterial inclusion bodies

background image

Protein quality in bacterial
inclusion bodies

Salvador Ventura

1,3

and Antonio Villaverde

1,2

1

Institut de Biotecnologia i de Biomedicina, Universitat Auto`noma de Barcelona, Bellaterra, 08193 Barcelona, Spain

2

Departament de Gene`tica i de Microbiologia, Universitat Auto`noma de Barcelona, Bellaterra, 08193 Barcelona, Spain

3

Departament de Bioquı´mica i de Biologia Molecular, Universitat Auto`noma de Barcelona, Bellaterra, 08193 Barcelona, Spain

A common limitation of recombinant protein production
in bacteria is the formation of insoluble protein
aggregates known as inclusion bodies. The propensity
of a given protein to aggregate is unpredictable, and the
goal of a properly folded, soluble species has been
pursued using four main approaches: modification of
the protein sequence; increasing the availability of
folding assistant proteins; increasing the performance
of the translation machinery; and minimizing physico-
chemical conditions favoring conformational stress and
aggregation. From a molecular point of view, inclusion
bodies are considered to be formed by unspecific
hydrophobic interactions between disorderly deposited
polypeptides, and are observed as ‘molecular dust-balls’
in productive cells. However, recent data suggest that
these protein aggregates might be a reservoir of
alternative conformational states, their formation
being no less specific than the acquisition of the
native-state structure.

Introduction
Recombinant protein production is an essential tool for the
biotechnology industry and also supports expanding areas
of basic and biomedical research, including structural
genomics and proteomics. Although bacteria still rep-
resent a convenient production system, many recombi-
nant polypeptides produced in prokaryotic hosts undergo
irregular or incomplete folding processes that usually
result in their accumulation as insoluble, and usually
refractile, aggregates known as inclusion bodies (IBs)

[1,2]

. In fact, the solubility of bacterially produced proteins

is of major concern in production processes

[3,4]

because

IBs are commonly formed during overexpression of
heterologous genes, particularly of mammalian or viral
origin. Consequently, many biologically relevant protein
species are excluded from the market because they cannot
be harvested in the native form at economically con-
venient yields. Although some recombinant proteins do
occur in both the soluble and insoluble cell fractions, many
others are only produced as IBs. To date, the solubility of a
given gene product has not been anticipated before gene
expression. However, it is now clear that the extent of
protein aggregation is determined, at least partially, by a
combination of process parameters, including culture

media composition, growth temperature, production rate
(as result of diverse factors, such as gene dosage, promoter
strength, mRNA stability and codon usage)

[5,6]

, and the

availability of heat-shock chaperones

[7,8]

. All of these

factors can be manipulated to enhance solubility but the
operational range is more limited than that required for a
competent solubility control. Overexpression of chaper-
ones and other folding modulators along with the
recombinant gene has been the most successful approach
for the minimization of IB formation. During the past
decade, hundreds of articles have described particular
chaperone-assisted production experiments with poorly
concluding results, often because of inconsistencies when
considering different protein species, host cell strains or
expression systems

[8,9]

. Although still a matter of

speculation, the origin of such variability might lie in
the distinct requirements of different proteins when
folding in a prokaryotic environment.

In addition, despite the functional redundancy of the

quality control system, the activities of some chaperones
(such as DnaK) cannot be completely complemented by
others

[10]

, and their titration causes bottlenecks in the

folding process

[11]

. It is also true that an important part

of the bacterial protein quality-control system is organized
into partially overlapping sequential networks, in which
folding intermediates are delivered from one chaperone
(or chaperone set) to another

[12,13]

. This sequential

handling would prevent the proper folding of a misfolding-
prone species when one crucial folding element is not
available at the required concentrations; however, the
overexpression of this bottleneck chaperone would make
the next step of the folding process limiting.

Alternatively, IBs can be a source of relatively pure

protein because they can be easily purified from disrupted
cells. By using IBs as a starting material, and after
applying in vitro refolding procedures, native proteins can
be recovered ready for use

[14–20]

. The main concern

about using IBs as a source material for industrial
purposes is that in vitro refolding procedures are not
universal and need to be adapted for each specific protein.
In addition, the cost and speed of such refolding
procedures are not always convenient in the large-scale
formats needed in industry

[15,21]

.

The undesired aggregation of recombinant proteins has

been experienced since early recombinant DNA technol-
ogies were developed. However, the physiological and

Corresponding author: Villaverde, A. (

avillaverde@servet.uab.es

).

Available online 28 February 2006

Review

TRENDS in Biotechnology

Vol.24 No.4 April 2006

www.sciencedirect.com

0167-7799/$ - see front matter Q 2006 Elsevier Ltd. All rights reserved. doi:10.1016/j.tibtech.2006.02.007

background image

structural data that has been collected about IBs during
the past five years are now offering the first steps towards
an integrated model of protein aggregation in bacteria

[22]

. In addition, picturing how IB formation is connected

to the physiology of the cell during the conformational
stress imposed by protein overproduction is now
becoming possible.

Morphology and composition
In actively producing recombinant E. coli cells, IBs are
seen as refractile particles, usually occurring in the
cytoplasm

[23,24]

, although secretory proteins can also

form IBs in the periplasm

[25]

. Under electron microscopy,

IBs appear rather amorphous

[26]

but, after detergent-

based purification, scanning microscopy reveals them to
be rod-shaped particles

[24,27]

. In vitro protease digestion

of purified inclusion bodies occurs on IB-associated
proteins as a cascade process

[28,29]

in which target

sites are sequentially activated or exposed to the enzyme
in a defined manner. This in-order cleavage indicates both
conformational flexibility and accessibility of IB proteins.
Also, partially digested IBs have a granular architecture

[27]

that might be compatible with IBs being formed by

the clustering of protease-resistant, smaller aggregates.
Classical proteomics of IBs showed them to be relatively
homogeneous in composition and mainly formed by the
recombinant protein itself

[30–32]

. Although occurring in

variable proportions, the recombinant product can reach
more than 90% of the total embedded polypeptides

[2,22]

,

which is a convenient protein supply for further in vitro
refolding. The remaining material includes proteolytic
fragments of the recombinant protein

[33,34]

, traces of

membrane proteins

[30,35]

, phospholipids and nucleic

acids

[31]

, at least some of these being contaminants

retained during the IB purification procedures

[36]

. In E.

coli IBs, the small heat-shock proteins IbpA and IbpB
have been identified

[22,37,38]

in addition to the main

chaperones DnaK and GroEL

[22,35]

.

Molecular determinants
The large set of polypeptides forming bacterial IBs are not
related, either structurally or sequentially, and include
small, large, monomeric, multimeric, prokaryotic or
eukaryotic proteins. Thus, aggregation inside bacterial
factories has long been considered to be a nonspecific
process, resulting in the formation of disordered intra-
cellular precipitates. Accordingly, several general features
inherent to the particular molecular status of the protein
but irrespective of its nature have been suggested to
promote IB formation. These include: high local concen-
trations of the produced polypeptide; transient accumu-
lation of proteins in totally or partially unfolded
conformations, with reduced solubility related to that of
the native form

[3]

; the accumulation of unstructured

protein fragments as a result of proteolytic attack

[19]

; the

establishment of wrong interactions with the bacterial
folding machinery

[39]

; the lack of the post-translational

modifications needed for the solubility of some eukaryotic
polypeptides

[40]

; and the prevention of proper disulfide

pairing in the reducing cytoplasmic environment

[41]

.

Although such environmental factors are relevant for

IB formation, the intrinsic nature of a polypeptide and its
sequence also determine its partitioning between the
insoluble and soluble cell fractions. Several classical
observations, together with recent results, reinforce this
view. The high purity of the recombinant protein in IBs,
and

the

recurrent

observation

that

recombinant

expression results in the formation of a reduced number
of IBs (usually one)

[23]

, suggest that they might be

formed by the growth of a small number of initial founder
aggregates by a nucleation-like mechanism relying on
molecular recognition events. Several observations sup-
port this view. First, specificity of polypeptide association
during aggregation processes has been seen in in vitro
refolding studies of proteins in complex protein mixtures

[42]

. Second, the folding intermediates of different

proteins tend to self-associate, in vitro, instead of co-
aggregating, despite the fact they form IBs when
expressed individually in bacteria

[43]

. Finally, and more

interestingly, under certain conditions, co-expression of
two proteins from genes carried on the same plasmid
results in the formation of two types of cytoplasmic
aggregates, each enriched in one type of recombinant
protein

[44]

. This segregation of the protein aggregates is

not the result of a temporal dependence of deposition,
supporting the view that, seeing as it occurs in vitro,
aggregation of proteins into IBs is a selective process.

IBs have long been thought to be devoid of all molecular

architecture, according to the view that unspecific
hydrophobic interactions drive the deposition process.
However, pioneering studies in the early 1990s

[45–47]

,

together with more recent investigations

[48–50]

, run

against this view. The use of attenuated total reflectance
infra-red spectroscopy for IBs analysis has shown that,
irrespective of the native protein structure, formation of
IBs results in the acquisition of significant new b-sheet
structures compared with the native conformation, even
for b-sheet-rich proteins. The persistence of some native
conformation in addition to the presence of disordered
chain segments has been also described, the content
depending on the particular IB-forming protein

[51]

. The

structural data suggest that the newly formed b-sheet
architecture in IBs is stabilized by a network of hydrogen
bonds between different chains, resulting in tightly
packed,

extended

intermolecular

b

-sheets.

These

b

-sheet-rich polypeptides or polypeptide regions would

be resistant to proteolysis, and it is enticing to propose
that they might constitute the above mentioned multiple
protease-resistant nuclei within IBs, whereas proteins or
protein segments in native and specially disordered
conformations would constitute the protease-sensitive
part of IBs.

In this context, an obvious question arises: how do

specific interactions that occur during the nucleation
process result in a more or less common structure for all
IBs? Although only a few studies have addressed this topic
for IBs, it has been a key issue in the closely related area of
protein misfolding and aggregation into amyloid fibrils.
Independent of the forming protein, all amyloid fibrils
share a predominant b-sheet architecture

[52]

. This

conformation, as in the case of IBs, is stabilized mostly

Review

TRENDS in Biotechnology

Vol.24 No.4 April 2006

180

www.sciencedirect.com

background image

by the establishment of non-covalent interactions between
polypeptide backbones, which are common to all proteins

[53]

. For amyloids, it has been proven that the propen-

sities of protein backbones to aggregate are sharply
modulated by their amino acid sequences, with certain
stretches acting as ‘hot spots’ from which aggregation can
nucleate specifically

[54–56]

. This can be the case for IBs

too. Recently, it has been shown that a preformed IB can
act as an effective aggregation seed for the deposition of its
partially folded soluble protein counterpart in a dose-
dependent manner

[49]

. Moreover, the seeding process is

highly specific because IBs promote the deposition of
homologous but not heterologous polypeptides

[49]

.

Sequestering of homologous misfolded species into IBs
might be a refined mechanism to reduce the potential
toxicity of partially folded monomers or small oligomers

[57]

, of which the solvent-exposed hydrophobic surfaces

might interact, improperly, with a large number of cellular
components and/or exhaust the in vivo folding machinery,
thereby hampering the folding and function of the cell
proteins. Thus, the establishment of specific interactions
during aggregation might be a conserved strategy with a
role in cellular protection, which seems to be the case in
IB-forming recombinant bacteria

[58]

. In summary,

protein aggregation as bacterial IBs and as amyloid fibrils
shows more than one coincident trait (

Table 1

).

Sequence determinants
The impact of point mutations on IB formation in several
protein systems also suggests that the primary structure
of a polypeptide somehow determines its propensity to
aggregate into IBs, whereby specific changes have a huge
impact on solubility. However, to forecast the effect of
sequence changes on the aggregation propensity in E. coli
still constitutes a challenge because the structural and
thermodynamic context in which they occur must be taken
into account, and these parameters are not easily
predictable.

Furthermore,

consistently

identical

mutations in different protein systems have been shown
to result in dissimilar effects

[59–63]

. Nevertheless, the

increasing number of structural genomic initiatives, and
the concomitant need for soluble recombinant proteins,

has pushed several attempts to predict IB formation
directly from the primary structure

[64]

but still with

inconsistent results. Among the intrinsic factors proposed
to be related to the propensity of a polypeptide to be
incorporated into IBs are: the size of the polypeptide; its
phylogenetic origin; the protein family and/or fold; the
charge average; the proportion of aliphatic residues; the
in vivo half-life; the frequency of occurrence of certain
dipeptides and tripeptides within the sequence; the
proportion of residues with good b-sheet propensity; and
the fraction of turn-forming residues. The reasons behind
the discordance among approaches rely on the inherent
difficulty of the addressed problem, namely aggregation
propensity is the net result of several extrinsic and
intrinsic factors and many of them are important to
different extents depending on the protein and expression
contexts

[65]

. In addition, it is clear that the solubility of

recombinant heterologous proteins has nothing to do with
the forces that have shaped sequences during evolution.
Thus, it is implausible that particular polypeptide proper-
ties, which lead to increased solubility of a recombinant
protein, would dominate in any given group of proteins.
This hampers the detection of relevant patterns influen-
cing IB formation.

Protein quality and dynamics
Overall, recent data suggests that IBs might embrace
conformational states different to those observed in the
soluble cell fraction, ranging from enriched b-forms to
native or native-like structures

[45,48–50]

(

Figure 1

). The

heterogeneous conformational status of IB protein was
hinted by the modeling of in vitro IB proteolytic digestion,
where different species with distinctive proteolytic sensi-
tivity were detected

[27,28]

. Such heterogeneity is

probably supported by the fact that the volumetric IB
growth during gene overexpression is the result of
unbalanced protein deposition and simultaneous cell-
driven physiological removal. Interestingly, at least a
fraction of IB protein is in continuous dynamic transition
between soluble and insoluble cell fractions

[33]

and, in

the absence of protein synthesis, cytoplasmic IBs are
almost completely disintegrated in a few hours

[66]

.

Therefore, rather than being mere molecular ‘dust-balls’
of the folding machinery, IBs are protein reservoirs that
are profoundly integrated in the protein quality system of
the cell

[22]

, and the embedded protein is under

continuous quality surveillance. Disaggregating ATPase-
associated chaperones (AAA

C

), sharing conserved ATP

binding and hydrolysis motifs (essentially ClpB), are
probably key elements in IB protein release because
they are responsible for protein reactivation in thermally
stressed cells

[67–70]

. Small heat-shock proteins (IbpAB),

commonly associated with IB proteins

[38,71]

, are also

important contributors to the disintegration process,
acting in a chaperone team that includes ClpB and
DnaK

[72,73]

. Other cytoplasmic chaperones, such as

GroEL, GroES and ClpA, are probably assisting removal
of the IB protein because, upon arrest of protein
production, IBs are more stable in their respective absence

[11,26]

. Furthermore, in IB-forming recombinant E. coli

cells, DnaK, GroEL and IbpAB have been identified as IB

Table 1. Main functional and structural traits of bacterial
inclusion bodies resembling those of amyloids

Feature

Refs

High purity of the aggregate

[23]

Aggregation mainly from folding intermediates

[49,89]

Sequence-specific aggregation

[43,49]

Chaperon-modulated aggregation

[11,90]

Seeding-driven aggregation

[49]

Aggregation propensities strongly affected by point
mutations

[91–95]

Reduced aggregation by stabilization of the native
structure

[96,97]

Intermolecular, cross b-sheet organization or in
general, enrichment of b structure

[47,49]

Fibril-like organization (of soluble protein aggregates)

[86]

Amyloid-tropic dye binding

[49]

Enhanced proteolytic resistance (of a fraction of IB
protein species)

[27,28]

Protection from cytotoxicity

[58]

Review

TRENDS in Biotechnology

Vol.24 No.4 April 2006

181

www.sciencedirect.com

background image

components

[22,35,38]

. Intriguingly, most cellular DnaK

molecules have been observed at the IB interface

[26]

,

where this chaperone probably acts by refolding or
releasing IB polypeptides in cooperation with ClpB and
IbpAB

[67,72,74,75]

. Recent insights on the disaggrega-

tion process have provided fascinating details about its
molecular mechanics. The protein ClpB recognizes sub-
strates through the conserved Tyr251 residue sited at the
central pore of the first AAA domain. This fact suggests a
translocation event for ClpB-mediated protein removal

[76,77]

that acts on discrete protein molecules rather than

on aggregated sections

[78]

. Both DnaK and ClpB middle

domains might also contribute by providing an unfolding
force in a still unsolved mechanism, acting in coordination
with the translocation event

[79]

.

Conversely, it seems that proteases are secondary tools

for aggregate processing, acting on IB polypeptides once
released

[66]

or during disaggregation

[80]

; however,

in situ digestion of IB protein has been suspected, through
indirect in vivo and in vitro observations

[23,28,80,81]

. In

support of a direct proteolytic attack, the absence of either
Lon or ClpP proteases largely minimizes IB disintegration

[82]

. However, in a ClpP

K

background, IB proteins

released to the soluble cell fraction remain stable and
can refold to a functional form

[82]

, highlighting this

enzyme

as

a

controller

of

the

quality

of

disaggregated proteins.

The heterogeneous conformational nature of IB pro-

teins is, in addition, reflected by the relatively high
activity of IBs formed by enzymes such as galactosidases
and other glucanases

[6,10,83]

(

Table 2

). Recently, the

same has been observed for aggregating fluorescent
proteins that generate highly emitting IBs

[84]

. In fact,

when analyzing the specific activity of soluble and IB
forms of b-galactosidase fusions, such values are within
the same order of magnitude

[10]

. This similarity can be

partially attributed to the occurrence of ‘soluble aggre-
gates’

[85]

, namely clusters of soluble but biologically

inactive protein, organized as fibers, which might even-
tually be among IB precursors

[86]

. Such elements would

Translational apparatus

Translational apparatus

Deposition

Soluble fraction

Insoluble fraction

Refolding

and Proteolysis

TRENDS in Biotechnology

Figure 1. Recombinant proteins produced in distant translational factories within
the bacterial cytoplasm occur in either soluble or insoluble cell fractions. Such
entities are virtual cell compartments (indicated by a vertical dashed line) between
which proteins are distributed according to their fractionation under high-speed
centrifugation. A fraction of de novo synthesized polypeptides can immediately
reach the native conformation and are fully functional (yellow spheres). Other
molecules enter into incorrect, dead-end folding pathways, are non functional and
tend to aggregate because of the presence of solvent-exposed hydrophobic
surfaces (small brown boxes). Aberrant folding forms and folding intermediates
can have properly folded domains that, if embracing active sites, might be still fully
or partially functional, although tending to aggregate (orange boxes). The
backbones of these protein forms can interact in a sequence-dependent manner
and under second-order kinetics to form small, b-sheet-enriched, soluble
aggregates, organized as fibers or other cluster types. Soluble aggregates are
trapped, specifically, in larger aggregation nuclei, forming one or a few IBs (vertical
brown box in the insoluble cell fraction) according to first-order kinetics. Therefore,
IBs contain both inactive (unfolded) and active (partially folded or eventually
properly folded) protein species that might self-organize in a concentric manner.
Here, native-like species surround unfolded, densely packaged and proteolytically
stable polypeptide chains. Protein material is steadily transferred between these
virtual cell compartments by either deposition into IBs or refolding and/or
proteolysis of IB proteins, generating a conformational continuum between soluble
and insoluble cell fractions. Therefore, incorrect folding and aggregation, or proper
folding and solubility, are not perfectly pair-matched events because both active
and inactive protein forms can be found in either the soluble or the insoluble
fractions.

Table 2. Some structural and functional evidence that properly folded protein species are a significant component of bacterial IBs

IB protein

Structure (determination method)

Biological activity (% relative to the
soluble counterpart, when determined)

Refs

Green- and blue-fluorescent protein
fusions

High IB fluorescence emission in vivo
(between 20 and 30%)

[84]

b

-galactosidase and b-galactosidase

fusion proteins

High specific activity in purified IBs
(from around 30 up to more than 100%)

[6,10,84]

Di-hydropholate reductase

Low activity in purified IBs (6%)

[84]

Endoglucanase D

High activity in purified IBs (25%)

[83]

b

-lactamase

Detectable activity in purified IBs

[87]

HtrA1 serine protease

Detectable activity in purified IBs

[87]

Interleukin-1 b

Native-like secondary structure (FTIR)

a

[45]

Several a-helix-rich hyperthermophilic
proteins

Native-like secondary structure
(FTIR; NMR; CD)

b,c

[98]

TEM b-lactamase

Native-like secondary structure (FTIR)

[47]

Lipase

Native-like secondary structure (FTIR)

[50]

Human granulocyte-colony
stimulating factor

Native-like secondary structure (FTIR)

[99]

Human growth hormone

Native-like secondary structure (FTIR)

[100]

Human interferon a 2b

Native-like secondary structure (FTIR)

[100]

a

FTIR, Fourier transformed infrared spectroscopy.

b

NMR, nuclear magnetic resonance.

c

CD, circular dychroism.

Review

TRENDS in Biotechnology

Vol.24 No.4 April 2006

182

www.sciencedirect.com

background image

reduce the average specific activity of the recombinant
enzyme in the soluble cell fraction. Contrarily, an
important part of the IB protein population must be
properly folded and coexist with the background inter-
molecular b-sheet organization

[49]

(

Figure 1

). Again, this

might be indicative of conformational variability within
IBs as a result of either native-like and b-enriched
polypeptides, polypeptides trapped by b-enriched aggre-
gation determinants (but keeping properly folded active
site domains), or a combination of both. Although the
specific activity of IB enzymes relative to their soluble
versions is highly variable when comparing different
proteins (

Table 2

), IBs formed by enzymes seem to be

immediately useful in bioprocesses; they can skip any
refolding step because their porous nature would permit
substrate processing by the active enzyme molecules

[84]

.

Importantly, the availability of IbpAB and its occurrence
in enzyme IBs significantly enhances their biological
activities

[87]

. This observation confirms that these

small heat-shock proteins, believed to preserve the
folding-competent state of target proteins

[88]

and keep

them suitable for refolding

[67,72]

, are also efficient at

preserving their native structure within aggregates.

Conclusions and future prospects
Rather than being ‘scrambled eggs’, bacterial inclusion
bodies are dynamic and conformationally diverse struc-
tures, formed by a sequence-selective aggregation process
that is probably driven by certain ‘hot spots’ within the
protein sequence. Furthermore, neither are they the dead-
end of deficient folding processes but rather the transient
reservoirs of aggregated polypeptides that are still under
the quality control surveillance of cell chaperones and
proteases. Recent insights into IB structure reveal that
native or native-like proteins, or protein domains, coexist
with b-sheet-rich intermolecular assemblies that share
functional and architectural features with amyloid aggre-
gates. In addition, the biological activity of enzymes and
fluorescent proteins forming IBs is not dramatically lower
than their soluble counterparts. Deeper exploration of this
fact

will

open

intriguing

possibilities

for

the

biotechnological industry.

Acknowledgements

AV acknowledges the support for research on protein aggregation through
grants BIO2004–00700 (MEC;

http://www.mec.es/

) and 2005SGR-00956

(AGAUR;

http://agaur.gencat.net/

). SV is recipient of a ‘Ramo´n y Cajal’

contract awarded by the MCYT-Spain and co-financed by the Universitat
Auto`noma de Barcelona (UAB;

http://www.uab.es/

), and founded by

PNL2004–40 (UAB) and 2005SGR-00037(AGAUR).

References

1 Fahnert, B. et al. (2004) Inclusion bodies: formation and utilization.

Adv. Biochem. Eng. Biotechnol. 89, 93–142

2 Villaverde, A. and Carrio, M.M. (2003) Protein aggregation in

recombinant bacteria: biological role of inclusion bodies. Biotechnol.
Lett. 25, 1385–1395

3 Baneyx, F. and Mujacic, M. (2004) Recombinant protein folding and

misfolding in Escherichia coli. Nat. Biotechnol. 22, 1399–1408

4 Sorensen, H.P. and Mortensen, K.K. (2005) Soluble expression of

recombinant proteins in the cytoplasm of Escherichia coli. Microb.
Cell Fact. 4, 1

5 Strandberg, L. and Enfors, S.O. (1991) Factors influencing inclusion

body formation in the production of a fused protein in Escherichia
coli. Appl. Environ. Microbiol. 57, 1669–1674

6 Worrall, D.M. and Goss, N.H. (1989) The formation of biologically

active beta-galactosidase inclusion bodies in Escherichia coli. Aust.
J. Biotechnol. 3, 28–32

7 Hoffmann, F. and Rinas, U. (2004) Roles of heat-shock chaperones in

the production of recombinant proteins in Escherichia coli. Adv.
Biochem. Eng. Biotechnol. 89, 143–161

8 Thomas, J.G. and Baneyx, F. (1996) Protein misfolding and inclusion

body formation in recombinant Escherichia coli cells overexpressing
heat-shock proteins. J. Biol. Chem. 271, 11141–11147

9 Thomas, J.G. and Baneyx, F. (1997) Divergent effects of chaperone

overexpression and ethanol supplementation on inclusion body
formation in recombinant Escherichia coli. Protein Expr. Purif. 11,
289–296

10 Garcia-Fruitos, E. et al. (2005) Folding of a misfolding-prone beta-

galactosidase in absence of DnaK. Biotechnol. Bioeng. 90, 869–875

11 Carrio, M.M. and Villaverde, A. (2003) Role of molecular chaperones

in inclusion body formation. FEBS Lett. 537, 215–221

12 Buchberger, A. et al. (1996) Substrate shuttling between the DnaK

and GroEL systems indicates a chaperone network promoting
protein folding. J. Mol. Biol. 261, 328–333

13 Goloubinoff, P. et al. (1999) Sequential mechanism of solubilization

and refolding of stable protein aggregates by a bichaperone network.
Proc. Natl. Acad. Sci. U. S. A. 96, 13732–13737

14 Vallejo, L.F. and Rinas, U. (2004) Strategies for the recovery of active

proteins through refolding of bacterial inclusion body proteins.
Microb. Cell Fact. 3, 11

15 Middelberg, A.P. (2002) Preparative protein refolding. Trends

Biotechnol. 20, 437–443

16 Clark, E.D. (2001) Protein refolding for industrial processes. Curr.

Opin. Biotechnol. 12, 202–207

17 Panda, A.K. (2003) Bioprocessing of therapeutic proteins from the

inclusion bodies of Escherichia coli. Adv. Biochem. Eng. Biotechnol.
85, 43–93

18 Tsumoto, K. et al. (2003) Practical considerations in refolding

proteins from inclusion bodies. Protein Expr. Purif. 28, 1–8

19 Cabrita, L.D. and Bottomley, S.P. (2004) Protein expression and

refolding – a practical guide to getting the most out of inclusion
bodies. Biotechnol. Annu. Rev. 10, 31–50

20 Li, M. et al. (2004) In vitro protein refolding by chromatographic

procedures. Protein Expr. Purif. 33, 1–10

21 Schugerl, K. and Hubbuch, J. (2005) Integrated bioprocesses. Curr.

Opin. Microbiol. 8, 294–300

22 Carrio, M.M. and Villaverde, A. (2002) Construction and deconstruc-

tion of bacterial inclusion bodies. J. Biotechnol. 96, 3–12

23 Carrio, M.M. et al. (1998) Dynamics of in vivo protein aggregation:

building inclusion bodies in recombinant bacteria. FEMS Microbiol.
Lett. 169, 9–15

24 Bowden, G.A. et al. (1991) Structure and morphology of protein

inclusion bodies in Escherichia coli. Biotechnology 9, 725–730

25 Miot, M. and Betton, J.M. (2004) Protein quality control in the

bacterial periplasm. Microb. Cell Fact. 3, 4

26 Carrio, M.M. and Villaverde, A. (2005) Localization of chaperones

DnaK and GroEL in bacterial inclusion bodies. J. Bacteriol. 187,
3599–3601

27 Carrio, M.M. et al. (2000) Fine architecture of bacterial inclusion

bodies. FEBS Lett. 471, 7–11

28 Cubarsi, R. et al. (2001) In situ proteolytic digestion of inclusion body

polypeptides occurs as a cascade process. Biochem. Biophys. Res.
Commun. 282, 436–441

29 Cubarsi, R. et al. (2005) A mathematical approach to molecular

organization and proteolytic disintegration of bacterial inclusion
bodies. Math. Med. Biol. 22, 209–226

30 Rinas, U. and Bailey, J.E. (1992) Protein compositional analysis of

inclusion bodies produced in recombinant Escherichia coli. Appl.
Microbiol. Biotechnol. 37, 609–614

31 Valax, P. and Georgiou, G. (1993) Molecular characterization of beta-

lactamase inclusion bodies produced in Escherichia coli. 1. Compo-
sition. Biotechnol. Prog. 9, 539–547

Review

TRENDS in Biotechnology

Vol.24 No.4 April 2006

183

www.sciencedirect.com

background image

32 Rinas, U. et al. (1993) Characterization of inclusion bodies in

recombinant Escherichia coli producing high levels of porcine
somatotropin. J. Biotechnol. 28, 313–320

33 Carrio, M.M. et al. (1999) Proteolytic digestion of bacterial inclusion

body proteins during dynamic transition between soluble and
insoluble forms. Biochim. Biophys. Acta 1434, 170–176

34 Corchero, J.L. et al. (1996) The position of the heterologous domain

can influence the solubility and proteolysis of beta-galactosidase
fusion proteins in E. coli. J. Biotechnol. 48, 191–200

35 Jurgen, B. et al. (2000) Monitoring of genes that respond to

overproduction of an insoluble recombinant protein in Escherichia
coli glucose-limited fed-batch fermentations. Biotechnol. Bioeng. 70,
217–224

36 Georgiou, G. and Valax, P. (1999) Isolating inclusion bodies from

bacteria. Methods Enzymol. 309, 48–58

37 Hoffmann, F. and Rinas, U. (2000) Kinetics of heat-shock response

and inclusion body formation during temperature-induced pro-
duction of basic fibroblast growth factor in high-cell-density cultures
of recombinant Escherichia coli. Biotechnol. Prog. 16, 1000–1007

38 Allen, S.P. et al. (1992) Two novel heat shock genes encoding proteins

produced in response to heterologous protein expression in Escher-
ichia coli. J. Bacteriol. 174, 6938–6947

39 Mogk, A. et al. (2002) Mechanisms of protein folding: molecular

chaperones and their application in biotechnology. ChemBioChem 3,
807–814

40 Zhang, Y. et al. (1998) Expression of eukaryotic proteins in soluble

form in Escherichia coli. Protein Expr. Purif. 12, 159–165

41 Lilie, H. et al. (1998) Advances in refolding of proteins produced in E.

coli. Curr. Opin. Biotechnol. 9, 497–501

42 London, J. et al. (1974) Renaturation of Escherichia coli tryptopha-

nase after exposure to 8 M urea. Evidence for the existence of
nucleation centers. Eur. J. Biochem. 47, 409–415

43 Speed, M.A. et al. (1996) Specific aggregation of partially folded

polypeptide chains: the molecular basis of inclusion body compo-
sition. Nat. Biotechnol. 14, 1283–1287

44 Hart, R.A. et al. (1990) Protein composition of Vitreoscilla hemo-

globin inclusion bodies produced in Escherichia coli. J. Biol. Chem.
265, 12728–12733

45 Oberg, K. et al. (1994) Native-like secondary structure in interleukin-

1 beta inclusion bodies by attenuated total reflectance FTIR.
Biochemistry 33, 2628–2634

46 Georgiou, G. et al. (1994) Folding and aggregation of TEM beta-

lactamase: analogies with the formation of inclusion bodies in
Escherichia coli. Protein Sci. 3, 1953–1960

47 Przybycien, T.M. et al. (1994) Secondary structure characterization

of beta-lactamase inclusion bodies. Protein Eng. 7, 131–136

48 Ami, D. et al. (2003) FT–IR study of heterologous protein expression

in recombinant Escherichia coli strains. Biochim. Biophys. Acta
1624, 6–10

49 Carrio, M. et al. (2005) Amyloid-like properties of bacterial inclusion

bodies. J. Mol. Biol. 347, 1025–1037

50 Ami, D. et al. (2005) Kinetics of inclusion body formation studied in

intact cells by FT–IR spectroscopy. FEBS Lett. 579, 3433–3436

51 Fink, A.L. (1998) Protein aggregation: folding aggregates, inclusion

bodies and amyloid. Fold. Des. 3, R9–23

52 Rochet, J.C. and Lansbury, P.T., Jr. (2000) Amyloid fibrillogenesis:

themes and variations. Curr. Opin. Struct. Biol. 10, 60–68

53 Chiti, F. et al. (2003) Rationalization of the effects of mutations on

peptide and protein aggregation rates. Nature 424, 805–808

54 Sanchez de Groot, N. et al. (2005) Prediction of “hot spots” of

aggregation in disease-linked polypeptides. BMC Struct. Biol. 5,
18

55 Ivanova, M.I. et al. (2004) An amyloid-forming segment of beta2-

microglobulin suggests a molecular model for the fibril. Proc. Natl.
Acad. Sci. U. S. A. 101, 10584–10589

56 Ventura, S. et al. (2004) Short amino acid stretches can mediate

amyloid formation in globular proteins: the Src homology 3 (SH3)
case. Proc. Natl. Acad. Sci. U. S. A. 101, 7258–7263

57 Marianayagam, N.J. et al. (2004) The power of two: protein

dimerization in biology. Trends Biochem. Sci. 29, 618–625

58 Gonzalez-Montalban, N. et al. (2005) Bacterial inclusion bodies are

cytotoxic in vivo in absence of functional chaperones DnaK or GroEL.
J Biotechnol 118, 406–412

59 Malissard, M. and Berger, E.G. (2001) Improving solubility of

catalytic domain of human beta-1,4-galactosyltransferase 1 through
rationally designed amino acid replacements. Eur. J. Biochem. 268,
4352–4358

60 Luan, C.H. et al. (2004) High-throughput expression of C. elegans

proteins. Genome Res. 14, 2102–2110

61 Idicula-Thomas, S. and Balaji, P.V. (2005) Understanding the

relationship between the primary structure of proteins and its
propensity to be soluble on overexpression in Escherichia coli.
Protein Sci. 14, 582–592

62 Goh, C.S. et al. (2004) Mining the structural genomics pipeline:

identification of protein properties that affect high-throughput
experimental analysis. J. Mol. Biol. 336, 115–130

63 Bertone, P. et al. (2001) SPINE: an integrated tracking database

and data mining approach for identifying feasible targets in high-
throughput structural proteomics. Nucleic Acids Res. 29,
2884–2898

64 Koschorreck, M. et al. (2005) How to find soluble proteins: a

comprehensive analysis of alpha/beta hydrolases for recombinant
expression in E. coli. BMC Genomics 6, 49

65 King, J. et al. (1996) Thermolabile folding intermediates: inclusion

body precursors and chaperonin substrates. FASEB J. 10, 57–66

66 Carrio, M.M. and Villaverde, A. (2001) Protein aggregation as

bacterial inclusion bodies is reversible. FEBS Lett. 489, 29–33

67 Mogk, A. et al. (2003) Small heat shock proteins, ClpB and the DnaK

system form a functional triade in reversing protein aggregation.
Mol. Microbiol. 50, 585–595

68 Mogk, A. and Bukau, B. (2004) Molecular chaperones: structure of a

protein disaggregase. Curr. Biol. 14, R78–R80

69 Mogk, A. et al. (2003) Roles of individual domains and conserved

motifs of the AAA

C

chaperone ClpB in oligomerization, ATP

hydrolysis, and chaperone activity. J. Biol. Chem. 278, 17615–17624

70 Weibezahn, J. et al. (2004) Unscrambling an egg: protein disaggrega-

tion by AAA

C

proteins. Microb. Cell Fact. 3, 1

71 Lethanh, H. et al. (2005) The small heat-shock proteins IbpA and

IbpB reduce the stress load of recombinant Escherichia coli and delay
degradation of inclusion bodies. Microb. Cell Fact. 4, 6

72 Mogk, A. et al. (2003) Refolding of substrates bound to small Hsps

relies on a disaggregation reaction mediated most efficiently by
ClpB/DnaK. J. Biol. Chem. 278, 31033–31042

73 Mogk, A. et al. (2003) Small heat shock proteins, ClpB and the DnaK

system form a functional triade in reversing protein aggregation.
Mol. Microbiol. 50, 585–595

74 Mogk, A. et al. (1999) Identification of thermolabile Escherichia coli

proteins: prevention and reversion of aggregation by DnaK and ClpB.
EMBO J. 18, 6934–6949

75 Motohashi, K. et al. (1999) Heat-inactivated proteins are rescued by

the DnaK.J–GrpE set and ClpB chaperones. Proc. Natl. Acad. Sci. U.
S. A. 96, 7184–7189

76 Schlieker, C. et al. (2004) Substrate recognition by the AAA

C

chaperone ClpB. Nat. Struct. Mol. Biol. 11, 607–615

77 Weibezahn, J. et al. (2004) Thermotolerance requires refolding of

aggregated proteins by substrate translocation through the central
pore of ClpB. Cell 119, 653–665

78 Schlieker, C. et al. (2004) Solubilization of aggregated proteins by

ClpB/DnaK relies on the continuous extraction of unfolded polypep-
tides. FEBS Lett. 578, 351–356

79 Weibezahn, J. et al. (2005) Novel insights into the mechanism of

chaperone-assisted protein disaggregation. Biol. Chem. 386, 739–744

80 Corchero, J.L. et al. (1997) Limited in vivo proteolysis of aggregated

proteins. Biochem. Biophys. Res. Commun. 237, 325–330

81 Carbonell, X. and Villaverde, A. (2002) Protein aggregated into

bacterial inclusion bodies does not result in protection from
proteolytic digestion. Biotechnol. Lett. 24, 1939–1944

82 Vera, A. et al. (2005) Lon and ClpP proteases participate in the

physiological

disintegration

of

bacterial

inclusion

bodies.

J. Biotechnol. 119, 163–171

83 Tokatlidis, K. et al. (1991) High activity of inclusion bodies formed in

Escherichia coli overproducing Clostridium thermocellum endoglu-
canase D. FEBS Lett. 282, 205–208

84 Garcia-Fruitos, E. et al. (2005) Aggregation as bacterial inclusion

bodies does not imply inactivation of enzymes and fluorescent
proteins. Microb. Cell Fact. 4, 27

Review

TRENDS in Biotechnology

Vol.24 No.4 April 2006

184

www.sciencedirect.com

background image

85 Sorensen, H.P. and Mortensen, K.K. (2005) Advanced genetic

strategies for recombinant protein expression in Escherichia coli.
J. Biotechnol. 115, 113–128

86 de Marco, A. and Schroedel, A. (2005) Characterization of the

aggregates formed during recombinant protein expression in
bacteria. BMC Biochem. 6, 10

87 Kuczynska-Wisnik, D. et al. (2004) Escherichia coli small heat shock

proteins IbpA/B enhance activity of enzymes sequestered in inclusion
bodies. Acta Biochim. Pol. 51, 925–931

88 Kitagawa, M. et al. (2002) Escherichia coli small heat shock proteins,

IbpA and IbpB, protect enzymes from inactivation by heat and
oxidants. Eur. J. Biochem. 269, 2907–2917

89 Chrunyk, B.A. et al. (1993) Inclusion body formation and protein

stability in sequence variants of interleukin-1 beta. J. Biol. Chem.
268, 18053–18061

90 Thomas, J.G. et al. (1997) Molecular chaperones, folding catalysts,

and the recovery of active recombinant proteins from E. coli. To fold
or to refold. Appl. Biochem. Biotechnol. 66, 197–238

91 Izard, J. et al. (1994) A single amino acid substitution can restore the

solubility of aggregated colicin A mutants in Escherichia coli. Protein
Eng. 7, 1495–1500

92 Wetzel, R. et al. (1991) Mutations in human interferon gamma

affecting inclusion body formation identified by a general immuno-
chemical screen. Biotechnology (N. Y.) 9, 731–737

93 Krueger, J.K. et al. (1992) Evidence that the methylesterase of

bacterial chemotaxis may be a serine hydrolase. Biochim. Biophys.
Acta 1119, 322–326

94 Wetzel, R. and Chrunyk, B.A. (1994) Inclusion body formation by

interleukin-1 beta depends on the thermal sensitivity of a folding
intermediate. FEBS Lett. 350, 245–248

95 Nieba, L. et al. (1997) Disrupting the hydrophobic patches at the

antibody variable/constant domain interface: improved in vivo
folding and physical characterization of an engineered scFv
fragment. Protein Eng. 10, 435–444

96 Chan, W. et al. (1996) Mutational effects on inclusion body formation

in the periplasmic expression of the immunoglobulin VL domain REI.
Fold. Des. 1, 77–89

97 Yan, G. et al. (2003) A single residual replacement improves the

folding and stability of recombinant cassava hydroxynitrile lyase in
E. coil. Biotechnol. Lett. 25, 1041–1047

98 Umetsu, M. et al. (2004) Structural characteristics and refolding of

in vivo aggregated hyperthermophilic archaeon proteins. FEBS Lett.
557, 49–56

99 Jevsevar, S. et al. (2005) Production of nonclassical inclusion bodies

from which correctly folded protein can be extracted. Biotechnol.
Prog. 21, 632–639

100 Ami, D. et al. (2006) Structural analysis of protein inclusion bodies by

Fourier transform infrared microspectroscopy. Biochim. Biophys.
Acta

doi:10.1016/j.bbapap.2005.12.005

(

www.sciencedirect.com

)

Have you contributed to an Elsevier publication?

Did you know that you are entitled to a 30% discount on books?

A 30% discount is available to ALL Elsevier book and journal contributors when ordering books or stand-alone CD-ROMs directly
from us.

To take advantage of your discount:

1. Choose your book(s) from www.elsevier.com or www.books.elsevier.com

2. Place your order

Americas:
TEL: +1 800 782 4927 for US customers
TEL: +1 800 460 3110 for Canada, South & Central America customers
FAX: +1 314 453 4898
E-MAIL: author.contributor@elsevier.com

All other countries:
TEL: +44 1865 474 010
FAX: +44 1865 474 011
E-MAIL: directorders@elsevier.com

You’ll need to provide the name of the Elsevier book or journal to which you have contributed. Shipping is FREE on pre-paid
orders within the US, Canada, and the UK.

If you are faxing your order, please enclose a copy of this page.

3. Make your payment

This discount is only available on prepaid orders. Please note that this offer does not apply to multi-volume reference works or
Elsevier Health Sciences products.

For more information, visit www.books.elsevier.com

Review

TRENDS in Biotechnology

Vol.24 No.4 April 2006

185

www.sciencedirect.com


Document Outline


Wyszukiwarka

Podobne podstrony:
Learning about protein solubility from bacterial inclusion bodies
Formation of active inclusion bodies in E coli
Lipid domain in bacterial membr Nieznany
[13]Role of oxidative stress and protein oxidation in the aging process
Drugs Used in Bacterial & Viral Infections 2
BYT 2006 Quality in IT project
Advanced genetic strategies for recombinant protein expression in E coli
Rapid and efficient purification and refolding of a (His) tagged recombinant protein produced in E c
Strategies to maximize heterologous protein expression in E coli
Inclusion bodies formation and utilisation
Protein Degradation in the Large Intestine Relevance to Colorectal Cancer
Insanity Is An Attractive Quality In Woman
Quality in Tourism Standard Requirements
Molecular chaperones involved in heterologous protein expression in E coli

więcej podobnych podstron