• No results found

Membrane protein production in Escherichia coli cell-free lysates

N/A
N/A
Protected

Academic year: 2021

Share "Membrane protein production in Escherichia coli cell-free lysates"

Copied!
10
0
0

Loading.... (view fulltext now)

Full text

(1)

Review

Membrane protein production in Escherichia coli cell-free lysates

Erik Henrich, Christopher Hein, Volker Dötsch, Frank Bernhard

Institute of Biophysical Chemistry, Centre for Biomolecular Magnetic Resonance, J.W. Goethe-University, Frankfurt-am-Main, Germany

a r t i c l e

i n f o

Article history: Received 16 March 2015 Revised 17 April 2015 Accepted 21 April 2015 Available online 1 May 2015 Edited by Wilhelm Just Keywords:

Synthetic biology Nanodiscs

Membrane protein solubilization Expression modes

Escherichia coli lysates

a b s t r a c t

Cell-free protein production has become a core technology in the rapidly spreading field of synthetic biology. In particular the synthesis of membrane proteins, highly problematic proteins in conven-tional cellular production systems, is an ideal application for cell-free expression. A large variety of artificial as well as natural environments for the optimal co-translational folding and stabiliza-tion of membrane proteins can rastabiliza-tionally be designed. The high success rate of cell-free membrane protein production allows to focus on individually selected targets and to modulate their functional and structural properties with appropriate supplements. The efficiency and robustness of lysates from Escherichia coli strains allow a wide diversity of applications and we summarize current strate-gies for the successful production of high quality membrane protein samples.

Ó 2015 Federation of European Biochemical Societies. Published by Elsevier B.V. All rights reserved.

1. Introduction

Membrane proteins are one of the most important protein classes and they represent the majority of targets for modern drugs. Numerous limitations in their production by conventional in vivo expression systems largely retarded their functional and structural characterization. Toxicity, limited membrane space for their functional folding as well as inefficient transport and mem-brane insertion mechanisms pose significant barriers. Synthetic cell-free (CF) expression systems eliminate most of these classical problems in membrane protein production and thus rapidly became a new platform for their characterization[1–5]. In addi-tion, CF expression unifies a number of further beneficial proper-ties such as non-complicated operation, high success rates and large flexibility for the fast development of an array of diverse applications[6].

CF expression is a relatively old technique and known since more than 50 years after publication of the famous experiments of Nirenberg and co-workers on deciphering the genetic code with Escherichia coli lysates[7]. While the technique has been in use for decades as analytical scale protein production system, the overall synthesis efficiency was considerably improved after developing refined protocols and reaction designs [8,9]. Based on its easy

access and excellent advantages for the efficient labeling of proteins, CF expression systems became routine platforms for the production of soluble proteins for structural analysis[10–12]. In recent times, the variation of CF systems rapidly expanded and lysates of various organisms have been analyzed for their efficien-cies in protein production.

The available CF systems show tremendous differences in view of efficiencies and applications and selection of appropriate CF lysate sources can be crucial for the subsequent performance of a project. At the current state of the art, lysates from E. coli have been successfully used for the production of the by far highest variety of membrane proteins with regard to size, topology, function and ori-gin[13,14]. Documented applications include functional studies, X-ray crystallography, nuclear magnetic resonance (NMR) struc-tural analysis as well as the characterization of large membrane protein assemblies[15–18]. Key advantage of E. coli lysates is their reliable efficiency in protein synthesis allowing the routine pro-duction of even large membrane proteins in mg amounts per one mL of reaction. Detailed protocols for the preparation of E. coli lysates and for the set-up of corresponding CF reactions have been described[19–22]. CF membrane protein production was initiated approximately 10 years ago and applications are continuously expanding (Fig. 1). In parallel, the number of system variations and protocols is still increasing. The intention of this review is to give a guideline on what kind of decisions and considerations have to be made when approaching the CF expression of a membrane protein in E. coli lysates. Complementary information summarizing current approaches can be obtained in several recent reviews[1,3–

http://dx.doi.org/10.1016/j.febslet.2015.04.045

0014-5793/Ó 2015 Federation of European Biochemical Societies. Published by Elsevier B.V. All rights reserved. ⇑ Corresponding author at: Institute of Biophysical Chemistry, Center for

Biomolecular Magnetic Resonance, J.W. Goethe-University, Max-von-Laue Str. 9, Frankfurt-am-Main 60438, Germany. Fax: +49 69 798 29632.

E-mail address:fbern@bpc.uni-frankfurt.de(F. Bernhard).

FEBS Letters 589 (2015) 1713–1722

j o u r n a l h o m e p a g e : w w w . F E B S L e t t e r s . o r g

CORE Metadata, citation and similar papers at core.ac.uk

(2)

5,23]. E. coli lysates are used as coupled transcription/translation systems with linear or circular DNA as template. Transcription is usually controlled by supplied T7 RNA polymerase which can be isolated out of overexpressing strains according to routine proto-cols[24]. Alternatively, E. coli lysates with already adjusted con-centrations of T7 RNA polymerase are commercially available. Expression controlled by endogenous E. coli RNA polymerases could further be efficient[15,25]. Protocols are developed and opti-mized in small volumes of microplates and subsequent scaling up to even multiple liter volumes is possible[26]. Expression efficien-cies close to one mg per mL are already possible with single com-partment batch reactions and by using E. coli lysates from optimized strains as well as specific reaction conditions[27–29]. Crude E. coli lysates can even be replaced by reconstituted transla-tion machineries composed out of the purified individual compo-nents[30].

Protein production efficiencies considerably increase by provid-ing additional reservoirs of small molecule precursors in a second compartment[9]. Protein synthesis in such continuous exchange cell-free (CECF) configurations is ongoing for many hours. Container geometry, the ratios of reaction and reservoir volume as well as the exchange surface in between the two compartments are important parameters that determine the final protein produc-tion efficiencies. A number of alternative opproduc-tions comprising indi-vidual as well as commercially available CECF containers have been proposed[22]. In addition, recently developed dialyzer car-tridges in combination with standard deep-well microplates are well suitable for analytical and semi-preparative scale screening reactions in the CECF configuration.

2. Project strategies: approaching CF production of membrane proteins in E. coli lysates

CF expression can generally be considered as a complementary approach to expression in E. coli cells in cases when the cellular expression fails or only minute amounts of protein are synthesized

[13]. In addition it is a good choice for the quick labeling of a pro-tein as DNA templates can easily be shuttled in between expres-sion systems with E. coli cells or CF lysates. Initial considerations when approaching CF expression should address (i) whether the technique is generally appropriate for the production of the selected target, (ii) which lysate should be used, (iii) which reac-tion environments should be provided and (iv) how sample quality can be controlled and evaluated (Fig. 2). Completely integral mem-brane proteins are preferable and targets containing up to 12 transmembrane segments and exceeding 100kDa have been syn-thesized so far[14,31]. In contrast, membrane proteins containing large soluble domains such as anchored or single span membrane

proteins are more likely to experience folding problems during CF expression. Supplied detergents or other artificial hydrophobic compounds may affect the proper folding of the soluble domains. The presence of multiple disulfide bridges could furthermore cause problems and the screening of redox systems should then be con-sidered[32].

Common lysate sources are the E. coli K strain A19 or BL21 derivatives such as the strains C41 or C43 known as the gold stan-dards for membrane protein production in vivo[33]. In addition to different strains, also several modifications of lysate preparation can be considered. Standard S30 lysates (i.e. fractionated by cen-trifugation at 30 000g) still contain residual cellular membrane fragments in amounts of up to approximately 100

l

g lipids per one ml of lysate[18,34]. Inserted endogenous porins or ion chan-nels may thus cause background signals when working with highly sensitive detection methods such as electrophysiological measure-ments. This background can be avoided by using S60 extracts pre-pared at 60 000g[34]. On the other hand, S12 extracts prepared at 12 000g can be used alternatively to S30 extracts for other targets

[35]. Protein labeling might be performed in lysates of specifically designed E. coli knock out mutants or in chemically pre-treated lysates[36]. Lysates could further become enriched for chaperones that might improve the functional folding of synthesized proteins

[37]. Chaperone induction upon lysate production or supply of chaperones could alter the resulting lysate proteomes into more favorable environments.

CF expression is a highly versatile synthetic biology approach. Initial reaction designs can therefore be adjusted according to the individual requirements of a membrane protein or to the intended application (Fig. 2). Samples suitable for structural approaches can be obtained by the co- or post-translational solubilization of CF expressed membrane proteins in a number of detergents. Numerous membrane proteins strictly require lipids in order to support folding and stability[38,39]. Expression in presence of pre-formed nanodiscs or with supplied detergent micelles containing particular lipids will therefore be a good choice. The multitude of options in order to manipulate reaction conditions will usually generate a comprehensive library of samples containing the target protein in quite different environments and probably also with sig-nificant variations in quality (Fig. 3). The strategy resembles the classical ‘‘funnel’’ model of expression screens for membrane pro-tein crystallization[40,41]. In cellular expression systems, large target libraries are screened and selection by expression efficien-cies, membrane insertion and isolation procedures usually results into a subset of few promising targets (Fig. 3). In contrast, a num-ber of significant differences do exist with the CF-funnel model. First, the high success rate of CF expression allows focusing on few selected targets that are of prime interest. Second, a large sam-ple library of the desired target solubilized in a variety of different environments is generated. Third and most important is that screening tools mainly comprise additives affecting sample quality such as folding environments composed out of various combina-tions of hydrophobic compounds, stabilizers and ligands rather than parameters affecting production efficiencies as in the classical cellular funnel. The result of the CF-funnel model is therefore a subset of samples having the desired and functionally folded target in different environments such as detergent micelles or nanodiscs (Fig. 3).

The functional folding of a membrane protein usually shows more or less clear preferences for particular CF reaction environ-ments[39,42]. Efficient sample quality evaluation tools preferen-tially applicable directly in the crude CF reaction mixture such as specific ligand binding assays or enzymatic activity measurements are therefore valuable. The availability of such tools is frequently a major bottleneck in CF expression projects as only more advanced and time consuming assays such as size exclusion profiling or Fig. 1. PubMed referenced reports on the CF expression of membrane proteins in

between 2004 and 2015.

(3)

Fig. 2. Basic considerations important for the initial design of CF expression processes. Depending on the target properties and on the intended application, individual options and modules will be selected for the design of the final CF expression strategy.

Fig. 3. Schematic comparison of cellular versus CF screening strategies for membrane protein production. Both processes are compared as a ‘‘funnel’’ model. Symbols represent the individual membrane protein targets. The main differences in between the two strategies exist in the target list and in the screening strategies as discussed in the text. The CF funnel starts with one desired target represented by the star and generates a library of samples in a variety of different environments comprising hydrophobic solubilizing compounds, stabilizers or ligands. In contrast, the cellular funnel starts with a larger list of different targets that will all be synthesized in identical or very similar conditions. Quality control e.g. implementing functional assays will select a sample subset with the identical target solubilized in different environments in the cell-free funnel, whereas a subset of different targets solubilized in identical environments are usually selected in the cellular funnel.

(4)

analysis of transport function are available. However, an initial short list of promising expression conditions can be determined by using fusions with monitors such as green fluorescent protein (GFP) (Fig. 4) [43]. Reaction conditions resulting into aggregated membrane proteins could then instantly be identified by using fusions with C-terminally attached GFP derivatives [44,45]. Furthermore, concentration screens of compounds affecting yield and solubilization of a membrane protein can rapidly be monitored.

3. The first milestone: improving CF production efficiencies of membrane proteins

CF expression reactions have a largely reduced complexity if compared with living cells as environment for protein overproduc-tion. Cytotoxic effects are eliminated and the lysate preparation procedure enriches compounds relevant for the expression machinery. The open access furthermore facilitates the addition of e.g. protease inhibitors in order to prevent the degradation of synthesized proteins. As a result, expression problems such as low yields are usually only associated with inefficient translation and can rapidly be addressed by few standardized optimization steps. Adjusting the Mg2+ion concentration for each DNA template

is always mandatory as a first trouble shooting step and it has a clear impact on the final expression success[46,47]. Target frag-mentation could be caused by premature termination of transla-tion and may be addressed by using codon optimized synthetic genes.

The initial DNA template design usually comprises a number of alternative options that are either used for protocol development or for final preparative scale expressions (Fig. 4). Inefficient initia-tion of translainitia-tion is the most frequently encountered bottleneck when approaching high yield CF expression. Prime determinants of translation initiation are the first few codons of the mRNA and suboptimal nucleotide sequences could restrict their access by ribo-somal RNA subunits[48]. A rapid pragmatic approach in order to improve expression yields is the attachment of short sequence opti-mized N-terminal expression tags (Fig. 4)[49–51]. Screening a low number of nucleotide-optimized expression tags mostly results into significantly improved expression efficiencies within short time. If the modification of the target protein with the few amino acids of the expression tag causes non-desired altered protein characteris-tics, enrichment of the first six to seven natural mRNA codons for A or T nucleotides by directed mutagenesis will be an alternative strategy as demonstrated with a membrane-integrated domain of an ABC transporter[52]. Additional tags useful for rapid purification

(i.e. poly(His)n-tags or Strep-tags) could be attached to the

C-terminus of the target protein (Fig. 4).

Further major determinants for the expression yield are (i) the selected configuration, i.e. batch versus CECF, (ii) the selected energy resources and precursor concentrations and (iii) the precur-sor reservoir volume in the CECF configuration as well as its poten-tial refreshing during the reaction. Yields of even identical targets from different reports are therefore hard to compare as usually quite different reaction conditions have been used. Examples of detailed protocols and cost calculations for the preparative scale production of membrane proteins intended for structural projects have been published and could be taken as preliminary bench-marks[18,22].

4. Refining sample quality: designing synthetic environments with additives

If the above mentioned optimization strategies are considered, the success rate of membrane protein production in E. coli lysates is very high and independent from membrane protein topology or function. Based on the established production protocol, the sub-sequent refinement procedure will focus on the systematic screen-ing of expression environments suitable to support the functional folding and stability of the synthesized membrane protein (Fig. 5). Three distinct expression modes can serve as starting points for sample quality optimization. In the precipitate forming (P-CF) mode, the synthesized membrane proteins precipitate after translation and are solubilized post-translationally in detergent micelles. In contrast, in the detergent-based (D-CF) and the lipid-based (L-CF) modes, the synthesized membrane proteins will be co-translationally solubilized. Pure detergents, mixed micelles, micelles containing lipids, surfactants or amphipols could be sup-plied in the D-CF mode, while in the L-CF mode lipid bilayers are provided[3,53–58]. Overviews on the different types of artificial hydrophobic environments including their particular properties are given elsewhere[3,59].

Co-translational solubilization in either D-CF mode or L-CF mode has an obvious impact on the translation kinetics of mem-brane proteins. Whereas the synthesis of soluble proteins such as GFP is not affected by the presence of supplied detergents or lipids, the synthesis rate of membrane proteins is often notably reduced

[45]. Higher yields of membrane proteins can thus be obtained in the P-CF mode without supplied hydrophobic compounds. The pre-cipitates may retain folded structures and excessive refolding pro-tocols are usually not applied for their subsequent solubilization

[60]. Depending on the membrane protein, a number of detergents

Fig. 4. Improving the CF expression yield of a membrane protein target. The basic system efficiency, e.g. the E. coli lysate quality, is monitored by GFP expression benchmarking. Adjusting the Mg2+ion concentration for each target DNA template is furthermore crucial. Expression depends on efficient initiation of translation and can be optimized by either addition of small expression tags or by appropriate silent mutagenesis of the first natural 5’codons. Additional optional modifications include the attachment of C-terminal monitors such as GFP and of small standard purification-tags.

(5)

such as lyso-phosphoglycerol derivatives, phosphocholines or DDM are efficient in solubilization. The P-CF synthesized MraY translocase of Bacillus subtilis showed the highest specific activity after solubilization in DDM[46]. Further functional or even high resolution structural characterizations of P-CF generated mem-brane protein samples have been exemplified[15,31,47,61–64].

However, some membrane proteins may unfold irreversibly in the P-CF mode[46]. The next step would thus be to switch to the D-CF mode (Fig. 5). Systematic screens of detergents or surfactants with defined working concentrations that are tolerated by the CF system have been published[3,65–69]. Most commonly used are long chain Brij derivatives as well as peptide surfactants, digitonin, the micelle forming short chain lipid di-heptanoyl-phosphocholine or fluorinated surfactants[16,61,69,70–74]. Suitable environments are still hard to predict. The Agrobacterium curdlan synthase was active after D-CF expression in peptide surfactants but not in Brij derivatives [75], whereas numerous olfactory receptors showed similar activity in peptide surfactants or in Brij detergents

[67,74]. High quality samples of the mitochondrial carrier Ucp1 were obtained with fluorinated surfactants[69]. Screening mixed micelles or micelles containing some lipids for better stability of membrane proteins could further be advantageous [69]. The NMR structure of proteorhodopsin as well as the crystal structure of the eukaryotic Acetabularia rhodopsin II were obtained after their D-CF expression in E. coli lysates in presence of digitonin micelles containing some phosphocholines[16,17].

Membrane protein folding upon co-translational solubilization can further be supported by providing stabilizing additives. Addition of the all-trans retinal cofactor is required for the

functional folding of rhodopsin derivatives [16,17,76]. The con-trolled co-expression of individual subunits enabled and stabilized the assembly of functional ATP synthase [15,77]. Redox systems such as combinations of oxidized and reduced glutathione could be beneficial for the formation of disulfide bridges in eukaryotic membrane proteins. Tightly binding small molecule ligands could further help to support conformational stability e.g. during CF expression G protein-coupled receptors (GPCRs) as already shown in crystallization studies[78,79]. Efficient strategies for extensive post-translational modifications of membrane proteins in E. coli lysates have so far not been developed and are still future perspec-tives. The addition or co-expression of corresponding modifying enzymes such as glycosyl transferases generally appears to be fea-sible but still awaits further approval[80].

5. Implementing bilayers: production of lipid dependent membrane proteins

Lipids are the natural environment of membrane proteins and are sometimes essential as stabilizing or even structural elements

[38,81,82]. In addition, some membrane proteins do not fold in presence of detergents or do not even tolerate any contact with detergents [46]. The L-CF mode implementing supplied mem-branes as preformed liposomes, nanodiscs or as isolated micro-somes is mandatory for these types of proteins (Fig. 5). However, membrane protein insertion into bilayers is a different and less efficient process if compared with their solubilization in deter-gents. Natural insertion machineries could be used by providing Fig. 5. Sample quality evaluation and refinement of CF expression conditions. The three basic modes P-CF, D-CF and L-CF can be implemented either simultaneously or sequentially. For each mode, numerous combinations of additives will be supplied and the quality of the resulting samples will be screened in appropriate assays.

(6)

microsome fractions or by the directed synthesis of translocons

[2,83,84]. However, supplied artificial bilayers are devoid of any proteins which is in contrast to the typically high protein to lipid ratio of cellular membranes. This characteristic is most likely advantageous for spontaneous insertion of membrane proteins as artificial bilayers are then better accessible and provide more space for insertion [34,71,85–88]. In addition, the specific topology of artificial bilayers, e.g. the two-side access and the interface in between membrane and membrane scaffold protein in nanodiscs can further support membrane protein insertion. Providing some detergent together with liposomes may further weaken bilayers and promote insertion[89].

Well defined and preformed bilayers can be added into CF reac-tions either as liposomes or as nanodiscs (Fig. 6). In particular the nanodisc technology provides excellent synergies to CF expression

[90,91]. Nanodiscs and liposomes of identical membrane composi-tions differ significantly in a number of properties. Nanodiscs are stable and can be supplied in final concentrations of at least 100

l

M giving stoichiometric ratios even with highly expressing membrane proteins. In contrast, pre-sized liposomes tend to fuse during the reaction and might even precipitate [45]. Furthermore, nanodisc membranes and liposomes will have quite different topologies with regard to membrane curvature, lateral pressure and accessibility. Insertion of L-CF expressed GPCRs and other membrane proteins was in particular efficient with nan-odiscs [42,76,92,93]. L-CF produced nanodisc complexes with membrane proteins are highly soluble and suitable for many appli-cations such as ligand binding studies or the characterization of membrane protein oligomerization[42,94,95].

The selection of suitable membrane compositions is a crucial point and lipid size, charge and resulting membrane topology are important parameters for the membrane insertion and folding effi-ciency of membrane proteins (Fig. 6). The denaturation tempera-ture of bacteriorhodopsin was shown to change significantly with the membrane composition[96]. Furthermore, the sterol content of liposomes modulated the water permeability of inserted aqua-porins[97]. Transition temperatures of the designed membranes should always be below the usual CF reaction temperatures in between 25 °C and 35 °C in order to facilitate protein insertion

[98]. Lateral pressure in the membrane depends on the lipid head-groups as well as on the alkyl chain geometries and it is increased by double bonds in cis conformation[99,100]. Spontaneous inser-tion of membrane proteins into lipid bilayers is higher at low

membrane elastic stress. Consequently, phosphocholine lipids or even single chain lyso-phosphocholine lipids that reduce the cur-vature stress could be more successful in promoting membrane protein insertion in L-CF expression approaches. Negatively charged headgroups of phosphoglycerol derivatives could further improve membrane insertion and folding efficiency [101]. Accordingly, the functional folding of E. coli MraY translocase enzymes responsible for bacterial cell wall precursor formation was only observed after L-CF expression with nanodiscs assembled with lipids containing phosphoglycerol headgroups[39].

6. The core application: functional studies with CF expressed membrane proteins

Applications for the functional characterization of CF expressed membrane proteins appear to be hardly limited and already cover numerous examples of enzymes, pores, channels, transporters, receptors as well as of fully assembled multi protein complexes

[1,5]. However, despite the common use of E. coli lysates, the indi-vidual protocols for expression and processing differ significantly in view of the implemented expression mode, of the provided hydrophobic environments and of the selection of supplemented compounds. While high quality samples have been prepared out of all three expression modes, the preferred selection strictly depends on the requirements of the individual membrane protein targets. In general, the P-CF and D-CF modes are more interesting for structural studies, whereas the L-CF mode in combination with liposomes or nanodiscs becomes important for the characteriza-tion of lipid dependent activities of membrane proteins

[39,42,76,92,93].

Most current data are obtained from the characterization of CF expressed rhodopsin derivatives, GPCRs, porins and channels as well as of membrane integrated enzymes. Underlying reasons for this selection bias are the availability of fast or highly sensitive functional assays for these classes of membrane proteins such as the folding dependent light absorption of rhodopsin derivatives or the specific and high affinity ligand binding of GPCRs. Those assays are essential for the screening of suitable expression condi-tions in a reasonable period as pointed out in the previous chap-ters. A large variety of GPCRs synthesized in E. coli lysates have been characterized. Most GPCRs have been D-CF expressed in pres-ence of detergents[74,102–105], peptide surfactants[55,67,106]

or amphipol like substances[68]. Alternatively, P-CF expression

Fig. 6. Important lipid characteristics for the insertion and stabilization of membrane proteins exemplified on the basis of a nanodisc complex. Lipid properties such as type of headgroup, the length of the alkyl chain or saturation affect the fluidity, thickness, charge and lateral pressure of the bilayer. Direct lipid interaction with membrane proteins can have a further impact on protein folding, stabilization or activity.

(7)

with subsequent detergent solubilization was implemented

[103,107]. Even synthetic membranes built by diblock copolymers could become a suitable expression environment for GPCRs[108]. Recent studies with the human endothelin receptors indicated that L-CF expression into preformed nanodiscs could be a fast tool for the subsequent characterization of GPCR/ligand interactions by surface plasmon resonance techniques even in crude CF reaction mixtures[42]. Alternatively, the direct insertion of CF expressed GPCRs as well as of other membrane proteins into tethered lipid bilayer membranes immobilized on chip surfaces could enable future drug screening attempts[109–111].

7. Arising perspectives: structural studies with membrane proteins

High success rates, rapid access to protein samples and efficient productivities make CF expression systems with E. coli lysates suit-able for structural studies[112]. Seleno-methionine labeled sam-ples of the multidrug transporter EmrE were a first example of a CF synthesized and crystallized membrane protein [113]. Only low diffracting crystals were obtained with a human voltage-dependent anion channel[114], whereas high resolution crystal structures of D-CF expressed Acetabularia rhodopsin II

[17] and the P-CF expressed bacterial diacylglycerokinase DgkA

[18]were solved at 3.2 Å and at 2.28 Å, respectively. The structure of DgkA was a proof-of principle study with a detailed comparison of samples obtained either after P-CF expression or after inclusion body production in E. coli cells. A comparable X-ray diffraction was finally obtained with both samples[18].

NMR structural approaches profit most from the labeling oppor-tunities enabled by CF expression. Solution NMR spectra of CF syn-thesized membrane proteins were obtained in different hydrophobic environments such as detergents[52,115], amphipols

[116], nanodiscs[76]or bicelles[117]. The NMR structure of prote-orhodopsin was solved with D-CF synthesized protein in presence of optimized detergent mixtures[16]. Structures of a subunit of the Alzheimer’s disease related protease presenilin-1, of membrane integrated domains of various signal kinases and of several small human integral membrane proteins were obtained with P-CF syn-thesized samples [60,62,63]. Future drug screening approaches based on combinations of NMR analysis of P-CF expressed mem-brane proteins and molecular simulation could thus become feasi-ble[118,119].

Structural dynamics is often crucial for protein–protein or pro-tein–ligand interactions as well as for enzymatic catalysis or trans-port cycles. Spectroscopic methods like NMR and EPR can provide closer insights into these processes. Both techniques require label-ing with spin probes or isotopes and pulsed EPR experiments like PELDOR became the method of choice for the determination of structural dynamics. Several dynamic features of membrane pro-teins have been investigated by these methods in the last years

[120]. Protocols for labeling are available [120,121] and some examples of in vivo expressed membrane proteins are the analysis of the heterodimeric ABC transporter BmrCD[122]and the vitamin transporter ButB [123]. Spin labeling is often performed with lysine or cysteine specific coupling chemistry. First data of direct spin label insertion via CF expression were obtained with the incorporation of 57Fe into a soluble FeFe hydrogenase [124].

Protocols developed for the direct insertion of spin labels into CF expressed soluble proteins could open the door for similar future studies with membrane proteins[125].

Structural characterization by advanced single particle electron microscopy techniques will come into focus in the near future

[126]. Well-resolved structures of larger assemblies such as the CF synthesized 540kDa ATP–synthase complex of C. thermarum can be obtained[15]. In addition, much lower amounts of samples

are required for electron microscopy and whole virus particles syn-thesized in even mammalian lysates have been visualized [127]. The continuously increasing resolution of electron microscopy could allow to analyze even smaller particles such as complexes of membrane proteins in nanodiscs. Protein inserted into nan-odiscs can be analyzed by atomic force microscopy as well as force spectroscopy, revealing their native conformation [39,94]. The evolving field of native mass spectrometry of membrane proteins and membrane protein complexes[128]especially in combination with the nanodisc technology[129,130]gives a further promising perspective for the characterization of CF expressed membrane proteins.

8. Outlook and perspectives

E. coli lysates have already become the workhorse in the field of CF systems and the versatility of applications will continuously grow. Evaluation of lysate blends combining benefits of different organisms is among the future developments and it will further boost the production of difficult biomolecules. Based on valuable intrinsic features such as the cost-effective preparation of lysates, the robustness and reliability of the translation machinery and the immense accumulated knowledge in recombinant protein expression, CF systems with E. coli lysates will further strengthen their role as standard platform for preparative scale protein pro-duction. Industrial scale applications such as the production of antibody-drug conjugates and the design of hybrid biosynthetic pathways with new and advanced properties in CF environments are already emerging. CF expression completely re-designed mem-brane protein production and the elimination of many restrictions caused by the need to maintain cell viability opened numerous new avenues of applications. Multitudes of chemically synthesized compounds can now directly be used for the co-translational stabi-lization of membrane proteins. The systematic screening of lipid effects on membrane proteins by the combination of nanodiscs with CF reactions is just one example for the excellent synergies in this new field of synthetic biology.

Acknowledgements

This study was funded by the Collaborative Research Center (SFB) 807 of the German Research Foundation (DFG). Support was further provided by Instruct, part of the European Strategy Forum on Research Infrastructures (ESFRI) and funded by national member subscriptions.

References

[1] Junge, F., Haberstock, S., Roos, C., Stefer, S., Proverbio, D., Dötsch, V. and Bernhard, F. (2011) Advances in cell-free protein synthesis for the functional and structural analysis of membrane proteins. N. Biotechnol. 28 (3), 262–271. [2] Fenz, S.F., Sachse, R., Schmidt, T. and Kubick, S. (2014) Cell-free synthesis of membrane proteins: tailored cell models out of microsomes. Biochim. Biophys. Acta 1838 (5), 1382–1388.

[3] Hein, C., Henrich, E., Orbán, E., Dötsch, V. and Bernhard, F. (2014) Hydrophobic supplements in cell-free systems: designing artificial environments for membrane proteins. Eng. Life Sci. 14 (4), 365–379. [4] Proverbio, D., Henrich, E., Main, G.U.F.A., Orbán, E., Dötsch, V. and Bernhard, F.

(2014) Membrane protein quality control in cell-free expression systems: tools, strategies and case studies in: Membrane Proteins Production for Structural Analysis (Mus-Veteau, Isabelle, Ed.), pp. 45–70, Springer, Heidelberg.

[5] Sachse, R., Dondapati, S.K., Fenz, S.F., Schmidt, T. and Kubick, S. (2014) Membrane protein synthesis in cell-free systems: from bio-mimetic systems to bio-membranes. FEBS Lett. 588 (17), 2774–2781.

[6] Smith, M.T., Wilding, K.M., Hunt, J.M., Bennett, A.M. and Bundy, B.C. (2014) The emerging age of cell-free synthetic biology. FEBS Lett. 588 (17), 2755– 2761.

[7] Eiserling, F., Levin, J.G., Byrne, R., Karlsson, U., Nirenberg, M.W. and Sjoestrand, F.S. (1964) Polyribosomes and DANN-dependent amino acid incorporation in Escherichia coli extracts. J. Mol. Biol. 10 (3), 536–540.

(8)

[8] Spirin, A.S., Baranov, V.I., Ryabova, L.A., Ovodov, S.Y. and Alakhov, Y.B. (1988) A continuous cell-free translation system capable of producing polypeptides in high yield. Science 242, 1162–1164.

[9] Kigawa, T., Yabuki, T., Yoshida, Y., Tsutsui, M., Ito, Y., Shibata, T. and Yokoyama, S. (1999) Cell-free production and stable-isotope labeling of milligram quantities of proteins. FEBS Lett. 442 (1), 15–19.

[10] Sawasaki, T., Ogasawara, T., Morishita, R. and Endo, Y. (2002) A cellfree protein synthesis system for high-throughput proteomics. Proc. Natl. Acad. Sci. U.S.A. 99, 14652–14657.

[11]Yokoyama, S. (2003) Protein expression systems for structural genomics and proteomics. Curr. Opin. Chem. Biol. 7, 39–43.

[12]Vinarov, D.A., Lytle, B.L., Peterson, F.C., Tyler, E.M., Volkman, B.F. and Markley, J.L. (2004) Cell-free protein production and labeling protocol for NMR based structural proteomics. Nat. Methods 1, 149–153.

[13]Savage, D.F., Anderson, C.L., Robles-Colmenares, Y., Newby, Z.E. and Stroud, R.M. (2007) Cell-free complements in vivo expression of the E. coli membrane proteome. Protein Sci. 16 (5), 966–976.

[14]Schwarz, D., Daley, D., Beckhaus, T., Dotsch, V. and Bernhard, F. (2010) Cell-free expression profiling of E. coli inner membrane proteins. Proteomics 10 (9), 1762–1779.

[15]Matthies, D., Haberstock, S., Joos, F., Dötsch, V., Vonck, J., Bernhard, F. and Meier, T. (2011) Cell-free expression and assembly of ATP synthase. J. Mol. Biol. 413 (3), 593–603.

[16]Reckel, S., Gottstein, D., Stehle, J., Löhr, F., Verhoefen, M.K., Takeda, M., Silvers, R., Kainosho, M., Glaubitz, C., Wachtveitl, J., Bernhard, F., Schwalbe, H., Güntert, P. and Dötsch, V. (2011) Solution NMR structure of proteorhodopsin. Angew. Chem. Int. Ed. Engl. 50 (50), 11942–11946.

[17]Wada, T., Shimono, K., Kikukawa, T., Hato, M., Shinya, N., Kim, S.Y., Kimura-Someya, T., Shirouzu, M., Tamogami, J., Miyauchi, S., Jung, K.H., Kamo, N. and Yokoyama, S. (2011) Crystal structure of the eukaryotic light-driven proton-pumping rhodopsin, Acetabularia rhodopsin II, from marine alga. J. Mol. Biol. 411 (5), 986–998.

[18]Boland, C., Li, D., Shah, S.T., Haberstock, S., Dötsch, V., Bernhard, F. and Caffrey, M. (2014) Cell-free expression and in meso crystallisation of an integral membrane kinase for structure determination. Cell. Mol. Life Sci. 71 (24), 4895–4910.

[19]Kigawa, T., Yabuki, T., Matsuda, N., Matsuda, T., Nakajima, R., Tanaka, A. and Yokoyama, S. (2004) Preparation of Escherichia coli cell extract for highly productive cell-free protein expression. J. Struct. Funct. Genomics 1–2, 63– 68.

[20] Swartz, J.R., Jewett, M.C. and Woodrow, K.A. (2004) Cell-free protein synthesis with prokaryotic combined transcription–translation. Methods Mol. Biol. 267, 169–182.

[21]Schwarz, D., Junge, F., Durst, F., Frölich, N., Schneider, B., Reckel, S., Sobhanifar, S., Dötsch, V. and Bernhard, F. (2007) Preparative scale expression of membrane proteins in Escherichia coli-based continuous exchange cell-free systems. Nat. Protoc. 2 (11), 2945–2957.

[22]Schneider, B., Junge, F., Shirokov, V.A., Durst, F., Schwarz, D., Dötsch, V. and Bernhard, F. (2010) Membrane protein expression in cell-free systems. Methods Mol. Biol. 601, 165–186.

[23]Junge, F., Schneider, B., Reckel, S., Schwarz, D., Dotsch, V. and Bernhard, F. (2008) Large-scale production of functional membrane proteins. Cell. Mol. Life Sci. 65 (11), 1729–1755.

[24]Li, Y., Wang, E. and Wang, Y. (1999) A modified procedure for fast purification of T7 RNA polymerase. Protein Expr. Purif. 16 (2), 355–358.

[25]Shin, J. and Noireaux, V. (2010) Efficient cell-free expression with the endogenous E. Coli RNA polymerase and sigma factor 70. J. Biol. Eng. 4 (8), 1– 9.

[26]Zawada, J.F., Yin, G., Steiner, A.R., Yang, J., Naresh, A., Roy, S.M., Gold, D.S., Heinsohn, H.G. and Murray, C.J. (2011) Microscale to manufacturing scale-up of cell-free cytokine production – a new approach for shortening protein production development timelines. Biotechnol. Bioeng. 108 (7), 1570–1578. [27]Kim, T.W., Kim, D.M. and Choi, C.Y. (2006) Rapid production of milligram quantities of proteins in a batch cell-free protein synthesis system. J. Biotechnol. 124 (2), 373–380.

[28]Jewett, M.C., Calhoun, K.A., Voloshin, A., Wuu, J.J. and Swartz, J.R. (2008) An integrated cell-free metabolic platform for protein production and synthetic biology. Mol. Syst. Biol. 4 (220), 1–10.

[29]Kim, H.C. and Kim, D.M. (2009) Methods for energizing cell-free protein synthesis. J. Biosci. Bioeng. 108 (1), 1–4.

[30] Shimizu, Y., Inoue, A., Tomari, Y., Suzuki, T., Yokogawa, T., Nishikawa, K. and Ueda, T. (2001) Cell-free translation reconstituted with purified components. Nat. Biotechnol. 19 (8), 751–755.

[31]Keller, T., Egenberger, B., Gorboulev, V., Bernhard, F., Uzelac, Z., Gorbunov, D., Wirth, C., Koppatz, S., Dötsch, V., Hunte, C., Sitte, H.H. and Koepsell, H. (2011) The large extracellular loop of organic cation transporter 1 influences substrate affinity and is pivotal for oligomerization. J. Biol. Chem. 286 (43), 37874–37886.

[32]Goerke, A.R. and Swartz, J.R. (2008) Development of cell-free protein synthesis platforms for disulfide bonded proteins. Biotechnol. Bioeng. 99 (2), 351–367.

[33]Miroux, B. and Walker, J.E. (1996) Over-production of proteins in Escherichia coli: mutant hosts that allow synthesis of some membrane proteins and globular proteins at high levels. J. Mol. Biol. 260 (3), 289–298.

[34]Berrier, C., Guilvout, I., Bayan, N., Park, K.H., Mesneau, A., Chami, M., Pugsley, A.P. and Ghazi, A. (2011) Coupled cell-free synthesis and lipid vesicle

insertion of a functional oligomeric channel MscL MscL does not need the insertase YidC for insertion in vitro. Biochim. Biophys. Acta 1808 (1), 41–46. [35] Kim, T.W., Keum, J.W., Oh, I.S., Choi, C.Y., Park, C.G. and Kim, D.M. (2006) Simple procedures for the construction of a robust and cost-effective cell-free protein synthesis system. J. Biotechnol. 126 (4), 554–561.

[36] Hong, S.H., Ntai, I., Haimovich, A.D., Kelleher, N.L., Isaacs, F.J. and Jewett, M.C. (2014) Cell-free protein synthesis from a release factor 1 deficient Escherichia coli activates efficient and multiple site-specific nonstandard amino acid incorporation. ACS Synthetic Biol. 3 (6), 398–409.

[37] Niwa, T., Kanamori, T., Ueda, T. and Taguchi, H. (2012) Global analysis of chaperone effects using a reconstituted cell-free translation system. Proc. Natl. Acad. Sci. U.S.A. 109 (23), 8937–8942.

[38] Hunte, C. and Richers, S. (2008) Lipids and membrane protein structures. Curr. Opin. Struct. Biol. 18 (4), 406–411.

[39] Roos, C., Zocher, M., Müller, D., Münch, D., Schneider, T., Sahl, H.G., Scholz, F., Wachtveitl, J., Ma, Y., Proverbio, D., Henrich, E., Dötsch, V. and Bernhard, F. (2012) Characterization of co-translationally formed nanodisc complexes with small multidrug transporters, proteorhodopsin and with the E. coli MraY translocase. Biochim. Biophys. Acta 1818 (12), 3098–3106.

[40] Lewinson, O., Lee, A.T. and Rees, D.C. (2008) The funnel approach to the precrystallization production of membrane proteins. J. Mol. Biol. 377 (1), 62– 73.

[41] Hammon, J., Palanivelu, D.V., Chen, J., Patel, C. and Minor Jr., D.L. (2009) A green fluorescent protein screen for identification of well-expressed membrane proteins from a cohort of extremophilic organisms. Protein Sci. 18 (1), 121–133.

[42] Proverbio, D., Roos, C., Beyermann, M., Orbán, E., Dötsch, V. and Bernhard, F. (2013) Functional properties of cell-free expressed human endothelin A and endothelin B receptors in artificial membrane environments. Biochim. Biophys. Acta 1828 (9), 2182–2192.

[43] Drew, D., Slotboom, D.J., Friso, G., Reda, T., Genevaux, P., Rapp, M., Meindl-Beinker, N.M., Lambert, W., Lerch, M., Daley, D.O., Van Wijk, K.J., Hirst, J., Kunji, E. and De Gier, J.W. (2005) A scalable, GFP-based pipeline for membrane protein overexpression screening and purification. Protein Sci. 14 (8), 2011–2017.

[44] Muller-Lucks, A., Bock, S., Wu, B. and Beitz, E. (2012) Fluorescent in situ folding control for rapid optimization of cell-free membrane protein synthesis. PLoS One 7 (7), e42186.

[45] Roos, C., Kai, L., Proverbio, D., Ghoshdastider, U., Filipek, S., Dötsch, V. and Bernhard, F. (2013) Co-translational association of cell-free expressed membrane proteins with supplied lipid bilayers. Mol. Membr. Biol. 30 (1), 75–89.

[46] Ma, Y., Münch, D., Schneider, T., Sahl, H.G., Bouhss, A., Ghoshdastider, U., Wang, J., Dötsch, V., Wang, X. and Bernhard, F. (2011) Preparative scale cell-free production and quality optimization of MraY homologues in different expression modes. J. Biol. Chem. 286 (45), 38844–38853.

[47] Rath, P., Demange, P., Saurel, O., Tropis, M., Daffe, M., Dotsch, V., Ghazi, A., Bernhard, F. and Milon, A. (2011) Functional expression of the PorAH channel from Corynebacterium glutamicum in cell-free expression systems: implications for the role of the naturally occurring mycolic acid modification. J. Biol. Chem. 286 (37), 32525–32532.

[48] Voges, D., Watzele, M., Nemetz, C., Wizemann, S. and Buchberger, B. (2004) Analyzing and enhancing mRNA translational efficiency in an Escherichia coli in vitro expression system. Biochem. Biophys. Res. Commun. 318 (2), 601–614.

[49] Ahn, J.H., Hwang, M.Y., Lee, K.H., Choi, C.Y. and Kim, D.M. (2007) Use of signal sequences as an in situ removable sequence element to stimulate protein synthesis in cell-free extracts. Nucleic Acids Res. 35 (4), e21.

[50] Kralicek, A.V., Radjainia, M., Mohamad Ali, N.A., Carraher, C., Newcomb, R.D. and Mitra, A.K. (2011) A PCR-directed cell-free approach to optimize protein expression using diverse fusion tags. Protein Expr. Purif. 80 (1), 117–124.

[51] Haberstock, S., Roos, C., Hoevels, Y., Dotsch, V., Schnapp, G., Pautsch, A. and Bernhard, F. (2012) A systematic approach to increase the efficiency of membrane protein production in cell-free expression systems. Protein Expr. Purif. 82 (2), 308–316.

[52] Tumulka, F., Roos, C., Lohr, F., Bock, C., Bernhard, F., Dotsch, V. and Abele, R. (2013) Conformational stabilization of the membrane embedded targeting domain of the lysosomal peptide transporter TAPL for solution NMR. J. Biomol. NMR 57 (2), 141–154.

[53] Park, K.H., Berrier, C., Lebaupain, F., Pucci, B., Popot, J.L., Ghazi, A. and Zito, F. (2007) Fluorinated and hemifluorinated surfactants as alternatives to detergents for membrane protein cell-free synthesis. Biochem. J. 403 (1), 183–187.

[54] Park, K.H., Billon-Denis, E., Dahmane, T., Lebaupain, F., Pucci, B., Breyton, C. and Zito, F. (2011) In the cauldron of cell-free synthesis of membrane proteins: playing with new surfactants. N. Biotechnol. 28 (3), 255–261. [55] Wang, X., Corin, K., Baaske, P., Wienken, C.J., Jerabek-Willemsen, M., Duhr, S.,

Braun, D. and Zhang, S. (2011) Peptide surfactants for cell-free production of functional G protein-coupled receptors. Proc. Natl. Acad. Sci. U.S.A. 108 (22), 9049–9054.

[56] Bazzacco, P., Billon-Denis, E., Sharma, K.S., Catoire, L.J., Mary, S., Le Bon, C., Point, E., Baneres, J.L., Durand, G., Zito, F., Pucci, B. and Popot, J.L. (2012) Nonionic homopolymeric amphipols: application to membrane protein folding, cell-free synthesis, and solution nuclear magnetic resonance. Biochemistry 51 (7), 1416–1430.

(9)

[57]Koutsopoulos, S., Kaiser, L., Eriksson, H.M. and Zhang, S. (2012) Designer peptide surfactants stabilize diverse functional membrane proteins. Chem. Soc. Rev. 41 (5), 1721–1728.

[58]Zheng, X., Dong, S., Zheng, J., Li, D., Li, F. and Luo, Z. (2014) Expression, stabilization and purification of membrane proteins via diverse protein synthesis systems and detergents involving cell-free associated with self-assembly peptide surfactants. Biotechnol. Adv. 32 (3), 564–574.

[59]Seddon, A.M., Curnow, P. and Booth, P.J. (2004) Membrane proteins, lipids and detergents: not just a soap opera. Biochim. Biophys. Acta 1666 (1–2), 105–117.

[60] Maslennikov, I., Klammt, C., Hwang, E., Kefala, G., Okamura, M., Esquivies, L., Mörs, K., Glaubitz, C., Kwiatkowski, W., Jeon, Y.H. and Choe, S. (2010) Membrane domain structures of three classes of histidine kinase receptors by cell-free expression and rapid NMR analysis. Proc. Natl. Acad. Sci. U.S.A. 107 (24), 10902–10907.

[61]Kai, L., Kaldenhoff, R., Lian, J., Zhu, X., Dötsch, V., Bernhard, F., Cen, P. and Xu, Z. (2010) Preparative scale production of functional mouse aquaporin 4 using different cell-free expression modes. PLoS One 5 (9), e12972.

[62]Sobhanifar, S., Schneider, B., Lohr, F., Gottstein, D., Ikeya, T., Mlynarczyk, K., Pulawski, W., Ghoshdastider, U., Kolinski, M., Filipek, S., Guntert, P., Bernhard, F. and Dotsch, V. (2010) Structural investigation of the C-terminal catalytic fragment of presenilin 1. Proc. Natl. Acad. Sci. U.S.A. 107 (21), 9644–9649.

[63]Klammt, C., Maslennikov, I., Bayrhuber, M., Eichmann, C., Vajpai, N., Chiu, E.J., Blain, K.Y., Esquivies, L., Kwon, J.H., Balana, B., Pieper, U., Sali, A., Slesinger, P.A., Kwiatkowski, W., Riek, R. and Choe, S. (2012) Facile backbone structure determination of human membrane proteins by NMR spectroscopy. Nat. Methods 9 (8), 834–839.

[64]Shenkarev, Z.O., Lyukmanova, E.N., Butenko, I.O., Petrovskaya, L.E., Paramonov, A.S., Shulepko, M.A., Nekrasova, O.V., Kirpichnikov, M.P. and Arseniev, A.S. (2013) Lipid-protein nanodiscs promote in vitro folding of transmembrane domains of multi-helical and multimeric membrane proteins. Biochim. Biophys. Acta 1828 (2), 776–784.

[65]Klammt, C., Schwarz, D., Fendler, K., Haase, W., Dötsch, V. and Bernhard, F. (2005) Evaluation of detergents for the soluble expression of alpha-helical and beta-barrel-type integral membrane proteins by a preparative scale individual cell-free expression system. FEBS J. 272 (23), 6024–6038. [66]Gourdon, P., Alfredsson, A., Pedersen, A., Malmerberg, E., Nyblom, M., Widell,

M., Berntsson, R., Pinhassi, J., Braiman, M., Hansson, O., Bonander, N., Karlsson, G. and Neutze, R. (2008) Optimized in vitro and in vivo expression of proteorhodopsin: a seven-transmembrane proton pump. Protein Expr. Purif. 58 (1), 103–113.

[67]Corin, K., Baaske, P., Ravel, D.B., Song, J., Brown, E., Wang, X., Geissler, S., Wienken, C.J., Jerabek-Willemsen, M., Duhr, S., Braun, D. and Zhang, S. (2011) A robust and rapid method of producing soluble, stable, and functional G-protein coupled receptors. PLoS One 6 (10), e23036.

[68]Klammt, C., Perrin, M.H., Maslennikov, I., Renault, L., Krupa, M., Kwiatkowski, W., Stahlberg, H., Vale, W. and Choe, S. (2011) Polymer-based cell-free expression of ligand-binding family B G-protein coupled receptors without detergents. Protein Sci. 20 (6), 1030–1041.

[69]Blesneac, I., Ravaud, S., Juillan-Binard, C., Barret, L.A., Zoonens, M., Polidori, A., Miroux, B., Pucci, B. and Pebay-Peyroula, E. (2012) Production of UCP1 a membrane protein from the inner mitochondrial membrane using the cell free expression system in the presence of a fluorinated surfactant. Biochim. Biophys. Acta 1818 (3), 798–805.

[70] Klammt, C., Schwarz, D., Dötsch, V. and Bernhard, F. (2007) Cell-free production of integral membrane proteins on a preparative scale. Methods Mol. Biol. 375, 57–78.

[71]Martin, M., Albanesi, D., Alzari, P.M. and de Mendoza, D. (2009) Functional in vitro assembly of the integral membrane bacterial thermosensor DesK. Protein Expr. Purif. 66 (1), 39–45.

[72]Corin, K., Baaske, P., Ravel, D.B., Song, J., Brown, E., Wang, X., Wienken, C.J., Jerabek-Willemsen, M., Duhr, S., Luo, Y., Braun, D. and Zhang, S. (2011) Designer lipid-like peptides: a class of detergents for studying functional olfactory receptors using commercial cell-free systems. PLoS One 6 (11), e25067.

[73]Miot, M. and Betton, J.M. (2011) Reconstitution of the Cpx signaling system from cell-free synthesized proteins. N. Biotechnol. 28 (3), 277–281. [74]Tegler, L.T., Corin, K., Hillger, J., Wassie, B., Yu, Y. and Zhang, S. (2015)

Cell-free expression, purification, and ligand-binding analysis of Drosophila melanogaster olfactory receptors DmOR67a, DmOR85b and DmORCO. Sci. Rep. 5 (7867), 1–7.

[75]Periasamy, A., Shadiac, N., Amalraj, A., Garajova, S., Nagarajan, Y., Waters, S., Mertens, H.D. and Hrmova, M. (2013) Cell-free protein synthesis of membrane (1,3)-beta-D-glucan (curdlan) synthase: co-translational insertion in liposomes and reconstitution in nanodiscs. Biochim. Biophys. Acta 1828 (2), 743–757.

[76]Lyukmanova, E.N., Shenkarev, Z.O., Khabibullina, N.F., Kopeina, G.S., Shulepko, M.A., Paramonov, A.S., Mineev, K.S., Tikhonov, R.V., Shingarova, L.N., Petrovskaya, L.E., Dolgikh, D.A., Arseniev, A.S. and Kirpichnikov, M.P. (2012) Lipid-protein nanodiscs for cell-free production of integral membrane proteins in a soluble and folded state: comparison with detergent micelles, bicelles and liposomes. Biochim. Biophys. Acta 1818 (3), 349–358. [77]Kuruma, Y., Suzuki, T. and Ueda, T. (2010) Production of multi-subunit

complexes on liposome through an E. coli cell-free expression system. Methods Mol. Biol. 607, 161–171.

[78] Rasmussen, S.G., Choi, H.J., Rosenbaum, D.M., Kobilka, T.S., Thian, F.S., Edwards, P.C., Burghammer, M., Ratnala, V.R., Sanishvili, R., Fischetti, R.F., Schertler, G.F., Weis, W.I. and Kobilka, B.K. (2007) Crystal structure of the human beta2 adrenergic G-protein-coupled receptor. Nature 450 (7168), 383–387.

[79] Rasmussen, S.G., Choi, H.J., Fung, J.J., Pardon, E., Casarosa, P., Chae, P.S., Devree, B.T., Rosenbaum, D.M., Thian, F.S., Kobilka, T.S., Schnapp, A., Konetzki, I., Sunahara, R.K., Gellman, S.H., Pautsch, A., Steyaert, J., Weis, W.I. and Kobilka, B.K. (2011) Structure of a nanobody-stabilized active state of the beta(2) adrenoceptor. Nature 469 (7329), 175–180.

[80] Guarino, C. and DeLisa, M.P. (2012) A prokaryote-based cell-free translation system that efficiently synthesizes glycoproteins. Glycobiology 22 (5), 596– 601.

[81] Opekarova, M. and Tanner, W. (2003) Specific lipid requirements of membrane proteins – a putative bottleneck in heterologous expression. Biochim. Biophys. Acta 1610 (1), 11–22.

[82] Schwall, C.T., Greenwood, V.L. and Alder, N.N. (2012) The stability and activity of respiratory complex II is cardiolipin-dependent. Biochim. Biophys. Acta 1817 (9), 1588–1596.

[83] Wuu, J.J. and Swartz, J.R. (2008) High yield cell-free production of integral membrane proteins without refolding or detergents. Biochim. Biophys. Acta 1778 (5), 1237–1250.

[84] Matsubayashi, H., Kuruma, Y. and Ueda, T. (2015) Cell-free synthesis of SecYEG translocon as the fundamental protein transport machinery. Orig. Life Evol. Biosp.. Epub ahead of print.

[85] Hovijitra, N.T., Wuu, J.J., Peaker, B. and Swartz, J.R. (2009) Cell-free synthesis of functional aquaporin Z in synthetic liposomes. Biotechnol. Bioeng. 104 (1), 40–49.

[86] Liguori, L. and Lenormand, J.L. (2009) Production of recombinant proteoliposomes for therapeutic uses. Methods Enzymol. 465, 209–223. [87] Chalmeau, J., Monina, N., Shin, J., Vieu, C. and Noireaux, V. (2011)

F-Hemolysin pore formation into a supported phospholipid bilayer using cell-free expression. Biochim. Biophys. Acta 1808 (1), 271–278.

[88] Berrier, C., Peyronnet, R., Betton, J.M., Ephritikhine, G., Barbier-Brygoo, H., Frachisse, J.M. and Ghazi, A. (2015) Channel characteristics of VDAC-3 from Arabidopsis thaliana. Biochem. Biophys. Res. Commun. (in press).

[89] Shimono, K., Goto, M., Kikukawa, T., Miyauchi, S., Shirouzu, M., Kamo, N. and Yokoyama, S. (2009) Production of functional bacteriorhodopsin by an Escherichia coli cell-free protein synthesis system supplemented with steroid detergent and lipid. Protein Sci. 18 (10), 2160–2171.

[90] Shadiac, N., Nagarajan, Y., Waters, S. and Hrmova, M. (2013) Close allies in membrane protein research: cell-free synthesis and nanotechnology. Mol. Membr. Biol. 30 (3), 229–245.

[91] Roos, C., Kai, L., Haberstock, S., Proverbio, D., Ghoshdastider, U., Ma, Y., Filipek, S., Wang, X., Dotsch, V. and Bernhard, F. (2014) High-level cell-free production of membrane proteins with nanodiscs. Methods Mol. Biol. 1118, 109–130.

[92] Katzen, F., Fletcher, J.E., Yang, J.P., Kang, D., Peterson, T.C., Cappuccio, J.A., Blanchette, C.D., Sulchek, T., Chromy, B.A., Hoeprich, P.D., Coleman, M.A. and Kudlicki, W. (2008) Insertion of membrane proteins into discoidal membranes using a cell-free protein expression approach. J. Proteome Res. 7 (8), 3535–3542.

[93] Gao, T., Petrlova, J., He, W., Huser, T., Kudlick, W., Voss, J. and Coleman, M.A. (2012) Characterization of de novo synthesized GPCRs supported in nanolipoprotein discs. PLoS One 7 (9), e44911.

[94] Zocher, M., Roos, C., Wegmann, S., Bosshart, P.D., Dotsch, V., Bernhard, F. and Muller, D.J. (2012) Single-molecule force spectroscopy from nanodiscs: an assay to quantify folding, stability, and interactions of native membrane proteins. ACS Nano 6 (1), 961–971.

[95] Lai, G. and Renthal, R. (2013) Integral membrane protein fragment recombination after transfer from nanolipoprotein particles to bicelles. Biochemistry 52 (52), 9405–9412.

[96] Allen, S.J., Curran, A.R., Templer, R.H., Meijberg, W. and Booth, P.J. (2004) Controlling the folding efficiency of an integral membrane protein. J. Mol. Biol. 342 (4), 1293–1304.

[97] Kai, L. and Kaldenhoff, R. (2014) A refined model of water and CO(2) membrane diffusion: effects and contribution of sterols and proteins. Sci. Rep. 4 (6665), 1–6.

[98] Meijberg, W. and Booth, P.J. (2002) The activation energy for insertion of transmembrane alpha-helices is dependent on membrane composition. J. Mol. Biol. 319 (3), 839–853.

[99] Booth, P.J. (2005) Sane in the membrane: designing systems to modulate membrane proteins. Curr. Opin. Struct. Biol. 15 (4), 435–440.

[100] Booth, P.J. and Curnow, P. (2009) Folding scene investigation: membrane proteins. Curr. Opin. Struct. Biol. 19 (1), 8–13.

[101] Seddon, A.M., Lorch, M., Ces, O., Templer, R.H., Macrae, F. and Booth, P.J. (2008) Phosphatidylglycerol lipids enhance folding of an alpha helical membrane protein. J. Mol. Biol. 380 (3), 548–556.

[102] Ishihara, G., Goto, M., Saeki, M., Ito, K., Hori, T., Kigawa, T., Shirouzu, M. and Yokoyama, S. (2005) Expression of G protein coupled receptors in a cell-free translational system using detergents and thioredoxin-fusion vectors. Protein Expr. Purif. 41 (1), 27–37.

[103] Junge, F., Luh, L.M., Proverbio, D., Schäfer, B., Abele, R., Beyermann, M., Dötsch, V. and Bernhard, F. (2010) Modulation of G-protein coupled receptor sample quality by modified cell-free expression protocols: a case study of the human endothelin A receptor. J. Struct. Biol. 172 (1), 94–106.

(10)

[104] Basu, D., Castellano, J.M., Thomas, N. and Mishra, R.K. (2013) Cell-free protein synthesis and purification of human dopamine D2 receptor long isoform. Biotechnol. Prog. 29 (3), 601–608.

[105] Rues, R.B., Orban, E., Dotsch, V. and Bernhard, F. (2014) Cell-free expression of G-protein coupled receptors: new pipelines for challenging targets. Biol. Chem. 395 (12), 1425–1434.

[106] Wang, X., Cui, Y. and Wang, J. (2013) Efficient expression and immunoaffinity purification of human trace amine-associated receptor 5 from E. coli cell-free system. Protein Pept. Lett. 20 (4), 473–480.

[107] Sansuk, K., Balog, C.I., van der Does, A.M., Booth, R., de Grip, W.J., Deelder, A.M., Bakker, R.A., Leurs, R. and Hensbergen, P.J. (2008) GPCR proteomics: mass spectrometric and functional analysis of histamine H1 receptor after baculovirus-driven and in vitro cell free expression. J. Proteome Res. 7 (2), 621–629.

[108] de Hoog, H.P., Lin JieRong, E.M., Banerjee, S., Decaillot, F.M. and Nallani, M. (2014) Conformational antibody binding to a native, cell-free expressed GPCR in block copolymer membranes. PLoS One 9 (10), e110847. [109] Robelek, R., Lemker, E.S., Wiltschi, B., Kirste, V., Naumann, R., Oesterhelt, D.

and Sinner, E.K. (2007) Incorporation of in vitro synthesized GPCR into a tethered artificial lipid membrane system. Angew. Chem. Int. Ed. Engl. 46 (4), 605–608.

[110] Leutenegger, M., Lasser, T., Sinner, E.K. and Robelek, R. (2008) Imaging of G protein-coupled receptors in solid-supported planar lipid membranes. Biointerphases 3 (2), Fa136.

[111] Coutable, A., Thibault, C., Chalmeau, J., Francois, J.M., Vieu, C., Noireaux, V. and Trevisiol, E. (2014) Preparation of tethered-lipid bilayers on gold surfaces for the incorporation of integral membrane proteins synthesized by cell-free expression. Langmuir 30 (11), 3132–3141.

[112] Bernhard, F. and Tozawa, Y. (2013) Cell-free expression – making a mark. Curr. Opin. Struct. Biol. 23 (3), 374–380.

[113] Chen, Y.J., Pornillos, O., Lieu, S., Ma, C., Chen, A.P. and Chang, G. (2007) X-ray structure of EmrE supports dual topology model. Proc. Natl. Acad. Sci. U.S.A. 104 (48), 18999–19004.

[114] Deniaud, A., Liguori, L., Blesneac, I., Lenormand, J.L. and Pebay-Peyroula, E. (2010) Crystallization of the membrane protein hVDAC1 produced in cell-free system. Biochim. Biophys. Acta 1798 (8), 1540–1546.

[115] Abdine, A., Verhoeven, M.A., Park, K.H., Ghazi, A., Guittet, E., Berrier, C., Van Heijenoort, C. and Warschawski, D.E. (2010) Structural study of the membrane protein MscL using cell-free expression and solid-state NMR. J. Magn. Reson. 204 (1), 155–159.

[116] Elter, S., Raschle, T., Arens, S., Viegas, A., Gelev, V., Etzkorn, M. and Wagner, G. (2014) The use of amphipols for NMR structural characterization of 7-TM proteins. J. Membr. Biol. 247 (9–10), 957–964.

[117] Uhlemann, E.M., Pierson, H.E., Fillingame, R.H. and Dmitriev, O.Y. (2012) Cell-free synthesis of membrane subunits of ATP synthase in phospholipid

bicelles: NMR shows subunit a fold similar to the protein in the cell membrane. Protein Sci. 21 (2), 279–288.

[118] Maslennikov, I. and Choe, S. (2013) Advances in NMR structures of integral membrane proteins. Curr. Opin. Struct. Biol. 23 (4), 555–562.

[119] Lindert, S., Maslennikov, I., Chiu, E.J., Pierce, L.C., McCammon, J.A. and Choe, S. (2014) Drug screening strategy for human membrane proteins: from NMR protein backbone structure to in silica- and NMR-screened hits. Biochem. Biophys. Res. Commun. 445 (4), 724–733.

[120] Bordignon, E. (2012) Site-directed spin labeling of membrane proteins. Top. Curr. Chem. 321, 121–157.

[121] Schmidt, M.J., Borbas, J., Drescher, M. and Summerer, D. (2014) A genetically encoded spin label for electron paramagnetic resonance distance measurements. J. Am. Chem. Soc. 136 (4), 1238–1241.

[122] Mishra, S., Verhalen, B., Stein, R.A., Wen, P.C., Tajkhorshid, E. and McHaourab, H.S. (2014) Conformational dynamics of the nucleotide binding domains and the power stroke of a heterodimeric ABC transporter. ELife 3, e02740. [123] Cafiso, D.S. (2014) Identifying and quantitating conformational exchange in

membrane proteins using site-directed spin labeling. Acc. Chem. Res. 47 (10), 3102–3109.

[124] Kuchenreuther, J.M., Guo, Y., Wang, H., Myers, W.K., George, S.J., Boyke, C.A., Yoda, Y., Alp, E.E., Zhao, J., Britt, R.D., Swartz, J.R. and Cramer, S.P. (2013) Nuclear resonance vibrational spectroscopy and electron paramagnetic resonance spectroscopy of 57Fe-enriched [FeFe] hydrogenase indicate stepwise assembly of the H-cluster. Biochemistry 52 (5), 818–826. [125] Kuchenreuther, J.M., Shiigi, S.A. and Swartz, J.R. (2014) Cell-free synthesis of

the H-cluster: a model for the in vitro assembly of metalloprotein metal centers. Methods Mol. Biol. 1122, 49–72.

[126] Kühlbrandt, W. (2014) The resolution revolution. Science 343 (6178), 1443– 1444.

[127] Kobayashi, T., Yukigai, J., Ueda, K., Machida, K., Masutani, M., Nishino, Y., Miyazawa, A. and Imataka, H. (2013) Purification and visualization of encephalomyocarditis virus synthesized by an in vitro protein expression system derived from mammalian cell extract. Biotechnol. Lett. 35 (3), 309– 314.

[128] Politis, A., Stengel, F., Hall, Z., Hernandez, H., Leitner, A., Walzthoeni, T., Robinson, C.V. and Aebersold, R. (2014) A mass spectrometry-based hybrid method for structural modeling of protein complexes. Nat. Methods 11 (4), 403–406.

[129] Marty, M.T., Zhang, H., Cui, W., Blankenship, R.E., Gross, M.L. and Sligar, S.G. (2012) Native mass spectrometry characterization of intact nanodisc lipoprotein complexes. Anal. Chem. 84 (21), 8957–8960.

[130] Hopper, J.T., Yu, Y.T., Li, D., Raymond, A., Bostock, M., Liko, I., Mikhailov, V., Laganowsky, A., Benesch, J.L., Caffrey, M., Nietlispach, D. and Robinson, C.V. (2013) Detergent-free mass spectrometry of membrane protein complexes. Nat. Methods 10 (12), 1206–1208.

References

Related documents

The objective of this study is to identify successes and challenges in the implementation of the HEP in Ethiopia, find out initiatives undertaken to improve program challenges,

So, for classification artificial intelligence tools like artificial neural network (ANN) and Support vector machine (SVM) are applied form which we can assess

Firewall parser architecture for a give protocol, such as an IP-telephony protocol, comprising a parser component (PS) which has interfaces (FAI, DI) allowing the parser component

The present study investigated the effect of processing regular and irregular word morphology on L2 vocabulary learning amongst low-level native Arabic learners

This is astounding, as numerous co-benefits would, in principle, be available from such policies: off-grid solar energy could contribute to electrification in those regions not

It is becoming increasingly apparent that all aspects of people’s lives including the way education is taught and delivered are greatly influenced by

Scores indicated significant improvement in functional physical status (activities of daily living), balance, distance walked in 6 min, and quality of life in the physical