• No results found

University of Groningen Insights into the transport mechanism of energy-coupling factor transporters Stanek, Weronika Karolina

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Insights into the transport mechanism of energy-coupling factor transporters Stanek, Weronika Karolina"

Copied!
17
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

University of Groningen

Insights into the transport mechanism of energy-coupling factor transporters

Stanek, Weronika Karolina

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date: 2018

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Stanek, W. K. (2018). Insights into the transport mechanism of energy-coupling factor transporters. University of Groningen.

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

CHAPTER 3

In vitro reconstitution of dynamically interacting integral membrane

subunits of Energy-Coupling Factor transporters

Weronika K. Stanek, Maria Majsnerowska, Dirk J. Slotboom

ABSTRACT

Energy-coupling factor (ECF) transporters are ATP-binding cassette (ABC) transporters that use a transmembrane protein (the S-component) to bind the transported substrate. The S-component is embedded in the bilayer either in a solitary state or in complex with the ECF module (EcfAA’T). The latter consists of the integral membrane subunit EcfT and two peripheral ABC-type ATPase subunits. In the solitary state the S-components are capable of scavenging the substrate from the surroundings, but translocation across the membrane usually requires the ECF module. Association and dissociation of the S-component from the ECF module is hypothesized to be part of the translocation cycle. In many ECF transporters multiple S-components specific for different substrates can interact with the same ECF module, leading to competition for association for substrate translocation. Here, we studied interaction of the S-components specific for pantothenate and folate from Lactobacillus delbrueckii with the same ECF module in a reconstituted system. The reconstituted ECF transporters maintained their transport activity and dynamic behavior in the proteoliposomes. Inhibition of pantothenate uptake by folate and vice versa indicated that association and dissociation of S-components must occur in the transport cycle. We postulate ECF transporters as a potential model system for studying interaction of membrane proteins in lipid bilayers.

(3)

INTRODUCTION

ATP-binding cassette (ABC) transporters are complexes of membrane proteins that mediate translocation of molecules across the bilayer and units able to hydrolyze ATP, which fuels transport. All ATPases of ABC transporters (also called Nucleotide-Binding Domains, NBDs) share highly conserved motifs. The membrane embedded part consists of two transmembrane subunits or domains (TMDs), each of which makes contact to a soluble nucleotide-binding domains via an α-helical element, called the coupling helix. ABC importers found in prokaryotes additionally make use of separate, soluble proteins for substrate binding and delivery to the transporter, called Substrate Binding Proteins or Domains (SBP or SBDs, shortly).

A notable subgroup of ABC transporters is formed by Energy Coupling Factor (ECF) transporters.1 The protein composition of ECF transporters is consistent with ABC transporters

architecture: two cytosolic ATPases (EcfA and EcfA’) and two transmembrane domains.2,3

Similar to ABC transporters, coupling helices transmit conformational changes caused by hydrolysis cycle of ATP from the NBDs to the transmembrane part of transporter.4,5 However,

in ECF type ABC transporters both coupling helices are part of only one transmembrane protein (EcfT). Together, the NBDs and EcfT form the so-called energy-coupling factor (also known as ECF module or energizing module). This module is the motor that drives accumulation of substrate with simultaneous ATP consumption, whereas ATP depletion restrained transport activity.6 ECF transporters owe their name to occurrence of this module.

In contrast to classical prokaryotic ABC importers, substrate binding in ECF transporters is mediated by the second integral membrane domain (the S-component). S-components are small proteins embedded in the cell membrane. S-components specific for different substrates have highly diverse amino acid sequences, but nonetheless share a similar fold of six α-helices.7,8 S-components have a double function, not only as scavengers of substrates

from environment, but also as transport domains alternating access of binding site by a toppling mechanism.9–12 All characterized S-components invariably exhibit high affinity for

their specific substrates in the absence of the ECF module.7,10,13 The high affinity allows them

to scavenge scarce substrates form the environment. Association with the ECF complex has two consequences: the protein topples over in the membrane so that the binding site becomes exposed to the cytoplasm, and the substrate binding site is disrupted leading to facilitated release into the cytosol.11 Type II ECF transporters, which are of main focus of this chapter,

have one single ECF module that can interact with multiple S-components. They have the S-components expressed upon the need for particular substrate from genes spread around the genome and not co-localized with operons coding for energizing module.3

The fascinating transport mechanism with a shared energizing module in ECF transporters was hypothesized for the first time by Henderson at al.1 His observations of transport in energized

cells of L. casei indicated that transporters needed a common component (the ECF module) in the complex with the substrate-binding protein to exhibit transport activity. 30 years later a comparative genomic study on ECF transporters revealed the identity of the ECF module and S-components. It was revealed that in organisms with type II ECF transporters genes of ECF module are usually not co-localized with S-component genes. Up to 12 different S-components were predicted to form complexes with one or two ECF modules.3 Further arguments in the

(4)

revealed a common surface area of S-components.14 Further investigations in a heterologous

expression system revealed that indeed different S-components can compete for a shared ECF module in a substrate-concentration dependent manner (Chapter 2).15 All these results

indicate an association and dissociation of S-components as a part of transport cycle. Here, we tested whether S-components competition can be detected in in vitro studies using a simple and defined liposome system. We chose to perform these studies in proteoliposomes instead of detergent because subunit association and dissociation in multimeric membrane transporters is expected to be dependent on a lipid bilayer since that is a native environment for the complex rearrangements in the transport cycle. Experimental results obtained with purified and solubilized protein may be burdened with artefacts of detergent micelle. The detergent micelle could either prevent S-component dissociation or the micelles might prevent their recurrent association.16 Therefore, we present here the functional reconstitution of multiple

component ECF transporters into the same proteoliposomes. Using the proteoliposomes system we show that, despite lack of sequence similarity, different S-components are able to interact with the same ECF module. Moreover, these integral membrane proteins compete for the interaction site on the ECF module. We conclude that dissociation of the S-component from ECF transporter complex and subsequent association of the same or different S-component is part of the transport cycle. Because of the recurring association and dissociation of S-components, ECF transporters are a great model for studying membrane protein interactions in the lipid bilayers.

MATERIAL AND METHODS

Mutagenesis

Mutations in EcfA and EcfA’ of ECF-FolT2 and ECF-PanT were introduced by two, consecutive rounds of QuikChange mutagenesis with primers listed below.

Primer name (mutation) Primer sequence (5’→3’)

Fw EcfA E177Q Ldb (E169Q in wild type) CATCATCCTGGATCAGTCGACCTCCATG Rev EcfA E177Q Ldb (E169Q in wild type) CATGGAGGTCGACTGATCCAGGATGATG

Fw EcfA’ E171Q Ldb TGTTTAGATCAGCCGGCAGCTGG Rev EcfA’ E171Q Ldb CCAGCTGCCGGCTGATCTAAACA Expression and membrane vesicles preparation

Whole complex ECF transporters from Lactobacillus delbrueckii subsp. bulgaricus were expressed in Escherichia coli MC1061 cells transformed with p2BAD vectors.17 The gene

organization in the expression vector was as follows: downstream the first arabinose promoter ECF module operon (10xHis-TEV-ecfAA’T) and further, downstream second arabinose promoter gene encoding PanT or FolT2 (panT-StrepII or folT2-StrepII, respectively). The expression from p2BAD plasmids was performed in a 5 L flask in LB Miller Broth with 100 µg/mL ampicillin. The E.coli culture was grown at 37°C with continuous shaking at 200 rpm. The expression of transporter was induced at OD600 between 0.6 and 0.8 with 10-3%

(5)

at 4°C.

Solitary S-components from L.delbrueckii, 10xHis-PanT or 10xHis-FolT2, were expressed from Lactococcus lactis NZ9000 cells transformed with pNZ8048 plasmid bearing the gene of either PanT or FolT2 protein after the nisin promoter. The expression from pNZ plasmid was performed in 1 L bottles with M17 media (Difco), 5 µg/mL chloramphenicol, and 2.0%

(w/v) glucose at 30°C. Overexpression of the solitary S-component was induced at OD600

of 0.8 with 0.1% (v/v) supernatant of nisin A producing strain. The cells were harvested by centrifugation (15 min, 6268 × g at 4°C) after the three-hour overexpression.

Membrane vesicles were prepared as described previously.2 Briefly, harvested cells were

diluted to OD600 around 100 with potassium phosphate buffer pH 7.5 and supplemented with

1 mM MgSO4 and DNase (~50 μg/mL). The cells were broken in Constant cell Disruption

System (Constant Systems Ltd) with 1 mM PMSF and 5 mM EDTA. For E.coli cells one passage at 25 kPsi and for L.lactis cells two passages at 39 kPsi were performed. Unbroken cell debris was separated by low-speed centrifugation (15 min, 27352 × g at 4°C). Subsequently, the membranes were concentrated by ultracentrifugation (120 min, 186010 × g at 4°C), homogenized in 50 mM potassium phosphate buffer pH 7.5, flash-frozen and stored at -80°C.

Protein purification and reconstitution into proteoliposomes

For the whole complex ECF transporter purification, membrane vesicles were thawed and incubated for 1 hour with 1% (w/v) n-dodecyl-β-D-maltopyranoside (DDM, Anatrace) in 50 mM potassium phosphate buffer pH 7.5 containing 300 mM NaCl, and 10% glycerol. Unsolubilized membrane fragments were removed by centrifugation (35 min, 286286 × g at 4°C). The solubilized protein solution was mixed with nickel-Sepharose resin equilibrated with solubilisation buffer and incubated for 1 hours with gentle rocking at 4°C. Proteins not bound to the resin were drained and subsequently washed away with 20 column volumes of 50 mM potassium phosphate buffer pH 7.5 supplemented with 300 mM NaCl, 50 mM imidazole pH 7.5 and 0.05% (w/v) DDM. The protein was eluted from the Ni-Sepharose column in three steps (fraction volumes 350 μL, 750 μL and 700 μL, respectively) in 50 mM potassium phosphate pH 7.5 supplemented with 300 mM NaCl, 500 mM imidazole pH 7.5 and 0.05% (w/v) DDM. The second fraction, the one with the highest protein content, was supplemented with 1 mM Na-EDTA and applied onto a size-exclusion chromatography column Sephadex200 10/300 (GE Healthcare) and eluted with 50 mM potassium phosphate buffer pH 7.5 supplemented with 150 mM NaCl and 0.05% DDM from a column previously equilibrated with that buffer. Protein fraction after the size-exclusion chromatography was used for the protein reconstitution into proteoliposomes according to a previously described method.18 Liposomes were composed of E.coli polar lipids supplemented with 1/3 (w/w) egg

phosphatidylcholine with final protein to lipid ratio in liposomes 1:1000 (w/w).

Solitary FolT2 was purified with the same approach as whole ECF transporters, but with different buffer compositions. In all buffers NaCl was replaced with KCl in corresponding concentrations. Membrane vesicles were solubilized with 1% DDM, and thereupon 0.38% n-nonyl-β-D-glucopyranoside (NG, Anatrace) was used instead of DDM in case of solitary FolT2 or a mixture of 0.05% DDM and solubilized with DDM E.coli polar lipids. Solitary S-components were reconstituted into the detergent-destabilized liposomes with protein to lipid ratios 1:250 or 1:1000 (w/w).

(6)

The co-reconstitution of multiple proteins was performed in the same manner as for individual reconstitution, always maintaining each protein to lipid ratio separately. The ECF transporters were treated as one complex and solitary S-components as independent proteins in the proteoliposomes.

Transport assay

In order to examine transport activity of proteins, radiolabeled substrate transport assay was performed as described previously with some modifications.11 Briefly, inclusion of 10

mM Mg2+-ATP or Mg2+-ADP into proteoliposomes was achieved with three consecutive

cycles of flash-freezing in liquid nitrogen and thawing at room temperature. Similar sized proteoliposomes were obtained after 11 passages through a polycarbonate filter (Avestin) with the pore size 400 nm. The remaining external nucleotides were removed by centrifugation in a final volume of 7 ml of 50 mM potassium phosphate (45 min, 286286 × g at 4°C). Subsequently, the proteoliposomes were resuspended in 50 mM potassium phosphate pH 7.5 to a protein concentration of 0.25 – 0.5 µg per time point (200 µL). The substrate uptake was performed at 30°C with stirring and initiated by adding the mixture of radiolabeled and non-radiolabeled substrates (5 nM and 95 nM respectively). For folate transport assays a folic acid [3,5,7,9 -3H] sodium salt (American Radiolabeled Chemicals) and for pantothenate transport

assays pantothenic acid, D-[2,3-3H] sodium salt (American Radiolabeled Chemicals) were

used. At given time intervals, 200 µL of the reaction mixture was withdrawn and diluted in the ice cold 50 mM potassium phosphate pH 7.5. Immediately, proteoliposomes were collected on pre-wetted cellulose nitrate filters. Subsequently, filters were washed with 2 mL of 50 mM potassium phosphate pH 7.5, dried for at least 1 hour at 80°C and dissolved in 5 mL of Filter Count scintillation liquid (Perkin Elmer). The radioactivity trapped inside the liposomes was determined with a Perkin Elmer Tri-carb 2800TR Scintillation counter.

RESULTS

To study the exchange of different S-components from the same ECF module, we chose the type II ECF transporter form Lactobacillus delbrueckii. This Gram-positive bacterium contains eight different S-components that share a common ECF module. We focused on the S-components for two different substrates: FolT2 and PanT, specific for folate and pantothenate, respectively. The two proteins share only 21.5% sequence identity, but comparison of crystal structures of FolT2 and a PanT homologue from Lactobacillus brevis shows that they adopt a common overall structure.11,14 We overproduced and purified the complete, four-subunit

complexes FolT2 and PanT, as well as the solitary S-components. Both ECF-FolT2 and ECF-PanT were stable as complexes in detergent solution and were reconstituted in active states into proteoliposomes (see below). The solitary S-component FolT2 could also be purified in a stable state in detergent solution, but PanT was more difficult to isolate. In the absence of the ECF module, the protein could not be expressed and purified in large quantities. Furthermore, isolated PanT in detergent solution was marginally stable in all tested conditions. Therefore, we designed our experiments in such a way that PanT was always purified in the complex with ECF module, but was allowed to dissociate from ECF module once reconstituted in the liposomes (see below). Only for a few crucial control experiments

(7)

we used purified, solitary PanT (see below, Supplementary Figure 1).

We produced the full four-subunit complexes in E. coli using the p2BAD vector. The solitary S-component genes were overexpressed from pNZ8048 vectors in Lactococcus lactis. The purified proteins were reconstituted into detergent-destabilized liposomes composed of E.coli polar lipids and egg PC to provide favorable membrane environment for protein activity (Chapter 6) and the “two-dimensional solvent” allowing the exchange of S-components.

Different S-components form a functional transport complex with the same ECF module

The full ECF transporter complexes ECF-FolT2 and ECF-PanT were both active when reconstituted into proteoliposomes (Figure 1A, B, Chapter 5), and mediated ATP-dependent uptake of the dedicated substrates (folate by ECF-FolT2 and pantothenate by ECF-PanT). The accumulation of radiolabeled substrate in proteoliposomes lumen was strictly dependent on the presence of luminal Mg2+-ATP. Mg2+-ADP did not support transport of pantothenate or

folate by ECF-PanT and ECF-FolT2, respectively (Figure 1, full circles). The ATP dependence is a defining feature of ABC transporters.6,11,13,19–21 Transport of each substrate was possible

only in the presence of a dedicated S-component. We were not able to detect transport of pantothenate by ECF-FolT2 or folate transport by ECF-PanT (Figure 1C, D, respectively), and conclude that the substrate specificity of the ECF transporters is determined entirely by the specific S-components (Figure 1A, B). In the absence of the ECF module, neither solitary FolT2 nor PanT from L. delbrueckii, was able to transport their dedicated substrates on their own (Figure 2B and Supplement 1B), showing that the ECF module is essential for transport. The liposomes used for the transport experiments shown in Figure 1, have a protein:lipid ratio of 1:1000 (w:w), which means on average a single ECF transporter per liposome (see calculation below). Our data show that the presence of a single ECF complex was enough to detect substrate accumulation in the assay. In case of pantothenate accumulation, the internal concentration of pantothenate in liposomes reached 3.76 µM in 25 min (37-fold accumulation) and folate was accumulated in 15 min to 4.55 µM (45-fold), assuming that the proteoliposomes have an average internal volume of 0.9 µL.18

Composition of the proteoliposomes

To study the competition of the S-components PanT and FolT2 for the same ECF module we reconstituted an excess of S-components relative to the ECF module in the liposomes, so that the amount of the latter would be limiting in transport assays. For the interpretation of the data, we took into account that membrane proteins can be incorporated in two orientations upon reconstitution in liposomes, with a mixture of inside-out and right-side-out oriented proteins (Chapter 5).22,23 Therefore, the pool of reconstituted transporters and S-components

oriented correctly for the activity assay was lower than the protein content. Additionally, the reconstitution efficiency into proteoliposomes was likely less than 100 %, further reducing the number of active proteins.

The absolute number of protein molecules in each proteoliposome was also relevant for the experiments. The size of proteoliposomes when extruded through 400 nm pore size filter is 350 ± 100 nm.18 Thus, for a full ECF transporter complex (Mw ~ 120 kDa) reconstituted using

(8)

each liposome. Consequently, the proteoliposome population in the ensemble is expected to be heterogeneous, with liposomes with a single transport complex in either the right-side-out or inside-out orientation present, as well as empty liposomes and liposomes containing more than one complex. The use of different chemical compositions (ATP, transported substrate) of the luminal and external solutions allowed us to probe the activity of proteins in a single orientation, with the proteins in the other orientation remaining “invisible”. For the solitary S-component FolT2 (Mw ~ 21 kDa) reconstituted in a ratio of 1:250 (w/w), we expected 26 protein molecules in each liposome. For comparison, when reconstitution ratio of FolT2 was decreased to 1:1000, then only ~ 6 protein molecules were expected in each vesicle

Figure 1Transport of [3H]pantothenate and [3H]folate into proteoliposomes with reconstituted ECF-PanT

and ECF-FolT2, respectively. A) Pantothenate transport by ECF-PanT in proteoliposomes loaded with Mg-ATP

(empty circles) and ADP (full circles). B) Folate transport by ECF-FolT2 in ATP (empty circles) and Mg-ADP (full circles) loaded proteoliposomes. Control of transporter substrate specificity: folate transport by ECF-PanT (C) and pantothenate transport by ECF-FolT2 (D).

Exchange of S-components in ECF module

We performed two types of experiments to show that the subunit composition of ECF transporter in membrane is dynamic. First, we co-reconstituted in proteoliposomes the whole complex ECF-PanT together with separately purified solitary FolT2. Because solitary FolT2 cannot transport folate alone (Figure 2B) uptake of radiolabeled folate in the liposomes containing FolT2, when co-reconstituted with full complex ECF-PanT, would be possible only if FolT2 could replace PanT in the full transporter. In the experiment, one full ECF-PanT transporter was present per liposome together with 26 molecules of FolT2. The excess of

A B

(9)

FolT2 over ECF-PanT was deliberately chosen for two reasons: (1) it facilitates the detection of competition between the S-components for the same ECF-module, as the amount of modules would be limiting; and (2) the pantothenate-specific S-component was delivered to the proteoliposome together with the ECF module, they were oriented in the liposome in the same way. The separately purified solitary FolT2 could insert in the liposomes either in the same orientation, which was desired for the competition experiments, or in the opposite orientation, which would not lead to functional association. The large excess of FolT2 ensured that there would be sufficient amount of molecules in the correct orientation for interaction with energy-coupling module.

The proteoliposomes with co-reconstituted ECF-PanT and solitary FolT2 showed transport activity for both folate and pantothenate (Figure 2A, C, open circles), indicating that exchange of the S-components had taken place, in which the solitary FolT2 had associated with the ECF module form ECF-PanT.

A B C

Figure 2 Transport assays showing exchange of S-components. A) [3H]pantothenate uptake in proteoliposomes

with co-reconstituted ECF-PanT and solitary FolT2. B) [3H]folate uptake in proteoliposomes loaded with Mg-ATP

and reconstituted with solitary FolT2. C) [3H]folate uptake in proteoliposomes with co-reconstituted ECF-PanT

and FolT2. Open circles represent proteoliposomes loaded with Mg-ATP, whereas full circles represent liposomes loaded with Mg-ADP. Uptake in Mg-ATP loaded proteoliposomes containing solitary FolT2 shown for comparison.

The exchange was not restricted to the combination of full ECF-PanT and solitary FolT2, reconstituted into proteoliposomes. We could also show that exchange is possible in a reverse system, where ECF-FolT2 was reconstituted as a whole complex together with solitary PanT, even though the yield and stability of the solitary PanT protein were low (Supplementary Figure 1). The proteoliposomes were able to take up pantothenate (Supplementary Figure 1A, empty circles). Notably, the uptake of pantothenate was observed only when ECF-FolT2 was co-reconstituted, and not observed when the same amount of solitary PanT was reconstituted alone (Supplementary Figure 1A, full circles).

As a second approach to demonstrate exchange of S-components in liposomes, we used a liposome system with two co-reconstituted full complexes: the wild type transporter complex for one substrate, and a complex with inactivated ATPase subunits for the other substrate. To inactivate the ATPases we created a double mutant, in which the glutamate residues from the Walker B motifs both in the EcfA and in the EcfA’ subunits were changed into glutamines (E169Q in EcfA and E171Q in EcfA’). The glutamates are necessary to coordinate a water

(10)

molecule for a nucleophilic attack on the bond between the γ- and ß-phosphate of ATP.24

Therefore, the glutamate-to-glutamine substitutions (EQ) are expected to be able to bind,25,26

but not hydrolyze ATP.27–30 Transport and ATPase activity of single mutants (E169Q or

E171Q) in ECF transporters was indeed significantly reduced (Chapter 5). With both NBDs mutated these activities were at the level of background (Chapter 5, also Chapter 7 in 30).

To test S-component exchange between two full complexes, wild type and mutant (EcfAE169Q; EcfA’E171Q) ECF-modules in complex with different S-components (FolT2 or PanT) were separately expressed and purified. They were subsequently co-reconstituted into liposomes to obtain approximately one native and one mutated complex in each liposome. By introducing ATP only in the lumen of the liposomes, and transported substrates only on the outside we could exclusively probe the activity of transporters in the right-side-out orientation. Co-reconstitution of the active ECF-PanT complex with the mutated and inactive ECF-FolT2 complex resulted in transport of both labeled substrates. The same result was obtained when active ECF-FolT2 was co-reconstituted with mutated ECF-PanT. These results show that S-component exchange had happened (Figure 3).

Figure 3Transport of folate and pantothenate into proteoliposomes co-reconstituted with two full complexes one of which contained the E to Q mutation in the Walker B motif. (A) Pantothenate transport and (B) folate

transport in proteoliposomes with co-reconstituted ECF-PanT and ECF(EQ)-FolT2. A) Full squares represents [3H]

pantothenate transport in proteoliposomes with two complexes, ECF-PanT and ECF(2EQ)-FolT2, empty circles represents [3H]pantothenate transport in ECF-PanT proteoliposomes, and full circles represents [3H]pantothenate

transport in ECF(2EQ)-FolT2 proteoliposomes. B) Full diamonds represents [3H]folate transport in proteoliposomes

with co-reconstituted ECF-PanT and ECF(2EQ)-FolT2, empty squares represents [3H]folate transport in

ECF-PanT proteoliposomes, and full circles [3H]folate transport in ECF(2EQ)-FolT2 proteoliposomes. C) Pantothenate

transport in proteoliposomes with co-reconstituted ECF-FolT2 and ECF(EQ)-PanT. Full circles represents [3H]

pantothenate transport in proteoliposomes with two complexes, ECF(2EQ)-PanT and ECF-FolT2, empty circles represents [3H]pantothenate transport in ECF-FolT2 proteoliposomes, and triangles represents uptake with Mg-ADP

loaded proteoliposomes with reconstituted ECF-PanT. Competition for a shared ECF module

In order to assay for competition of different S-components for the shared ECF module, we performed transport assays in proteoliposomes with co-reconstituted ECF-PanT and solitary FolT2. The full complex ECF-PanT and solitary FolT2 were expressed and purified separately. They were co-reconstituted to obtain approximately one full complex per liposome (one ECF module together with single PanT), and an excess of FolT2.

We found that radiolabeled pantothenate was taken up readily into the proteoliposomes A B C

(11)

containing ECF-PanT co-reconstituted with FolT2 (Figure 4A, opened circles). Addition of unlabeled folate revealed a dose-dependent reduction of [3H]pantothenate uptake (Figure

4A). It therefore appears that the substrate-loaded S-component FolT2 competes more effectively for use of the ECF module than the apo-protein. Using the same preparation of proteoliposomes we also followed the transport of radiolabeled folate. The highest folate transport rate was observed in the absence of unlabeled pantothenate as competing substrate (Figure 4B, open circles). Addition of unlabeled pantothenate caused a slight decrease in folate association with proteoliposomes (Figure 4B, full circles). However, increasing the amount of competing substrate caused only a marginal decrease of transport. The difference in sensitivity for the competing, substrate-bound S-component could be a reflection of the considerable excess of FolT2 over PanT, which was provided as subunit of the full complex ECF-PanT. In each liposome on average one full complex, ECF-PanT, and 26 molecules of FolT2 were present. To test whether the FolT2 excess may be the cause of the decreased sensitivity for added pantothenate in the folate uptake assay, we decreased the amount of co-reconstituted FolT2 by four folds to approximately 7 FolT2 molecules per liposome. Indeed, with this ratio of proteins the inhibitory effect of pantothanate was more pronounced (Figure 4D), although there was little dose dependence in the range of 0.5 to 50 µM pantothenate. The transport of radiolabeled pantothenate remained sensitive to the folate concentration (Figure 4C) in a dose dependent manner.

Figure 4 Transport of vitamins in proteoliposomes with co-reconstituted ECF-PanT and solitary FolT2. [3H]

pantothenate (A) and [3H]folate (B) transport into proteoliposomes with co-reconstituted ECF-PanT in protein to

lipid ratio 1:1000 (w:w) and solitary FolT2 in ratio 1:250 (w:w). (C) and (D) represents transport of pantothenate and folate, respectively; protein to lipid ratio for ECF-PanT is unchanged (1:1000) but ratio for FolT2 is decreased to 1:1000. The competing substrates were used at concentrations 0.5, 5 and 50 µM.

A B

(12)

We have not tested the system in which PanT was delivered as solitary component and FolT2 in complex with ECF module, because it was impossible to purify sufficient amounts of PanT, due to instability of the protein in detergent solution (see above). However, the data presented in Figures 2, 4, and Supplementary Figure 1 show that PanT and FolT2 are exchanged. Therefore, competition is likely to be an inherent feature of the system.

DISCUSSION

Cellular transport involves sophisticated machineries to allow movement of solutes through the lipid bilayer. Some transporters evolved to make a use of a single transmembrane translocator for many substrates. Examples are found in the family of ABC membrane transporters that include importers with more than one substrate binding protein (SBP) interacting with the same translocator complex. In some ABC importers different SBPs evolved as gene duplications followed by divergent evolution into two homologous periplasmic proteins with different substrate specificities31 or substrate affinities.32 Consequently, transport activity involves

interaction of multiple proteins with the shared transmembrane translocator. An alternative approach to increase efficiency of transport is observed in ABC importers with SBDs fused to TMDs. Each transmembrane domain may have up to three binding domains resulting in up to six SBDs connected to translocator.33,34 Recent single molecule FRET studies provide

inside into the competing behavior of SBDs to interact with shared translocator.35 In these

cases, competition of the substrate binding units for the shared, complementary part of the transporter resembles what is observed in ECF transporters, despite different localization and transport mechanism of the transporters. Extraordinarily, ECF transporters developed the ability to compete in the membrane environment.

We show here in an in vitro system that multi-transmembrane spanning proteins (S-components) not related in amino acid sequence can associate with a common partner forming a functional transporter. This study characterizes association and dissociation of S-components as a necessary part of transport mechanism in type II ECF transporters. We were able to find a pair of full complex ECF transporters from the same organism that were functional upon purification and reconstitution into the proteoliposomes. Furthermore, the results from experiments with reconstituted proteins showed that dissociation of S-component needs to occur to allow for transport of more than one substrate via the same ECF module. In the current model for the transport cycle of ECF transporters ATP is essential for the dissociation of S-components from the ECF module. The results from uptake assay with E to Q mutants of the Walker B motif of the ATPase subunits may suggest that only ATP binding, but not hydrolysis is necessary for the disassembly of S-component from the complex since these mutants are unable to hydrolyze ATP. However, it is possible that E to Q mutants are not completely inactive36 and residual ATPase activity is sufficient to release S-component from

the complex. Another possibility is that S-components have the ability to dissociate from the complex spontaneous possibly during the (harsh) reconstitution procedure.

We show here that FolT2 and PanT compete for the same ECF module. As already shown in in vivo experiments substrate bound S-components are out-competing those without substrate.1,15 We observe that increasing amounts of folate compete with [3H]pantothenate

uptake. With increasing saturation of FolT2 we push it towards the state where it was immediately binding with another substrate molecule, which subsequently associates again

(13)

with the ECF module.11 In case of [3H]folate uptake competition by PanT is less pronounced,

which can be a result of differences between S-components in their affinity for the ECF module or differences in the time needed for S-component to dock, release the substrate and be released from the complex.

The interaction between proteins is a foundation for virtually all processes in the cell. The study of protein-protein interactions in the lipid membrane is still challenging. Currently, there are few commonly used methods for identification of associated proteins in the bilayer: two-hybrid screen (Y2H),37 use of interactants bearing fluorescence labels forming a FRET

pair,38,39 proximity ligation assay,40 or a combination of more than one method to exclude

false results.41 These methods provide information about the growing number of interacting

complexes. However, many of the interactions have not been characterized in detail. This creates a need to select candidates as an experimental model. A single transmembrane helix model provides some information on the forces involved in dimerization.42,43 However, their

potential as experimental model system is limited due to limited size and complexity of interacting surfaces. A recently described model system with multi-transmembrane helices is the superfamily of ClC anion channels and antiporters.44,45 The studies of ClC-ec1 show that

protein interaction is dependent on its density in the membrane as well as high interacting surface complementarity and low affinity to the surrounding it lipids.

We propose ECF transporters as a natural protein complex for studying interactions between proteins not sharing amino acid sequence identity. Because of dissociation and association of different S-components as an indigenous part of the transport cycle, ECF transporters may serve as a novel model system for studying membrane protein interaction in the lipid bilayers.

ACKNOWLEDGMENTS

We thank Josy ter Beek, Raj Singh for constructing the wild type expression plasmids and Lotteke Swier for constructing ECF(EQ)-FolT2 expression plasmid.

REFERENCES

1. Henderson, G. B., Zevely, E. M. & Huennekens, F. M. Mechanism of folate transport in Lactobacillus casei: Evidence for a component shared with the thiamine and biotin transport systems. J. Bacteriol. 137, 1308–1314 (1979).

2. ter Beek, J., Duurkens, R. H., Erkens, G. B. & Slotboom, D. J. Quaternary structure and functional unit of energy coupling factor (ECF)-type transporters. J. Biol. Chem.

286, 5471–5475 (2011).

3. Rodionov, D. A. et al. A novel class of modular transporters for vitamins in

prokaryotes. J. Bacteriol. 91, 42–51 (2009).

4. Mourez, M., Hofnung, M. & Dassa, E. Subunit interactions in ABC transporters: A conserved sequence in hydrophobic membrane proteins of periplasmic permeases defines an important site of interaction with the ATPase subunits. EMBO J. 16, 3066– 3077 (1997).

5. Dawson, R. J. P., Hollenstein, K. & Locher, K. P. Uptake or extrusion: Crystal

structures of full ABC transporters suggest a common mechanism. Mol. Microbiol.

65, 250–257 (2007).

(14)

transport in Lactobacillus casei. J. Bacteriol. 139, 552–559 (1979).

7. Berntsson, R. P. -a. et al. Structural divergence of paralogous S components from

ECF-type ABC transporters. Proc. Natl. Acad. Sci. 109, 13990–13995 (2012).

8. Zhang, P., Wang, J. & Shi, Y. Structure and mechanism of the S component of a

bacterial ECF transporter. Nature 468, 717–720 (2010).

9. Finkenwirth, F. et al. ATP-dependent Conformational Changes Trigger Substrate

Capture and Release by an ECF-type Biotin Transporter. J. Biol. Chem. 290, 16929– 16942 (2015).

10. Slotboom, D. J. Structural and mechanistic insights into prokaryotic energy-coupling factor transporters. Nat. Rev. Microbiol. 12, 79–87 (2014).

11. Swier, L. J. Y. M., Guskov, A. & Slotboom, D. J. Structural insight in the toppling mechanism of an energy-coupling factor transporter. Nat. Commun. 7, 11072 (2016).

12. Henderson, G. B., Kojima, J. M. & Kumar, H. P. Kinetic Evidence for Two

Interconvertible Forms of the Folate Transport Protein from Lactobacillus caseit.

163, 1147–1152 (1985).

13. Erkens, G. B. et al. The structural basis of modularity in ECF-type ABC transporters. Nat. Struct. Mol. Biol. 18, 755–760 (2011).

14. Zhang, M. et al. Structure of a pantothenate transporter and implications for ECF

module sharing and energy coupling of group II ECF transporters. Proc. Natl. Acad. Sci. U. S. A. 111, 18560–5 (2014).

15. Majsnerowska, M., ter Beek, J., Stanek, W. K., Duurkens, R. H. & Slotboom, D.

J. Competition between Different S-Components for the Shared Energy Coupling Factor Module in Energy Coupling Factor Transporters. Biochemistry 54, 4763–4766 (2015).

16. Karpowich, N. K., Song, J. M., Cocco, N. & Wang, D.-N. ATP binding drives

substrate capture in an ECF transporter by a release-and-catch mechanism. Nat. Struct. Mol. Biol. 22, 565–571 (2015).

17. Birkner, J. P., Poolman, B. & Kocer, a. Hydrophobic gating of mechanosensitive channel of large conductance evidenced by single-subunit resolution. Proc. Natl. Acad. Sci. 109, 12944–12949 (2012).

18. Geertsma, E. R., Nik Mahmood, N. a B., Schuurman-Wolters, G. K. & Poolman, B.

Membrane reconstitution of ABC transporters and assays of translocator function. Nat. Protoc. 3, 256–266 (2008).

19. Xu, K. et al. Crystal structure of a folate energy-coupling factor transporter from Lactobacillus brevis. Nature 497, 268–71 (2013).

20. Karpowich, N. K. & Wang, D. Assembly and mechanism of a group II ECF

transporter. Proc. Natl. Acad. Sci. U. S. A. 110, 2534–9 (2013).

21. Jones, P. M. & George, A. M. The ABC transporter structure and mechanism:

Perspectives on recent research. Cell. Mol. Life Sci. 61, 682–699 (2004).

22. Rigaud, J. louis, Paternostre, M. therese & Bluzat, A. Mechanisms of Membrane

Protein Insertion into Liposomes during Reconstitution Procedures Involving the Use of Detergents. 2. Incorporation of the Light-Driven Proton Pump Bacteriorhodopsint. Biochemistry 27, 2677–2688 (1988).

23. Eytan, G. D. Use of liposomes for reconstitution of biological functions. BBA - Rev. Biomembr. 694, 185–202 (1982).

24. Hollenstein, K., Dawson, R. J. & Locher, K. P. Structure and mechanism of ABC

transporter proteins. Curr. Opin. Struct. Biol. 17, 412–418 (2007).

25. Schultz, K. M., Merten, J. A. & Klug, C. S. Characterization of the E506Q and

H537A dysfunctional mutants in the E. coli abc transporter MsbA. Biochemistry 50, 3599–3608 (2011).

(15)

26. Moody, J. E., Millen, L. & Binns, D. Cooperative, ATP-dependent association of the nucleotide binding cassettes during the catalytic cycle of ATP-binding cassette transporters. J. Biol. Chem. 277, 21111–21114 (2002).

27. Damas, J. M., Oliveira, a. S. F., Baptista, A. M. & Soares, C. M. Structural

consequences of ATP hydrolysis on the ABC transporter NBD dimer: Molecular dynamics studies of HlyB. Protein Sci. 20, 1220–1230 (2011).

28. Mehmood, S., Domene, C., Forest, E. & Jault, J.-M. Dynamics of a bacterial

multidrug ABC transporter in the inward- and outward-facing conformations. Proc. Natl. Acad. Sci. 109, 10832–10836 (2012).

29. Oldham, M. L., Khare, D., Quiocho, F. A., Davidson, A. L. & Chen, J. Crystal

structure of a catalytic intermediate of the maltose transporter. Nature 450, 515–521 (2007).

30. Swier, L. J. Y. M. On the transport mechanism of energy-coupling factor transporters. (University of Groningen, 2016).

31. Higgins, C. F. & Ames, G. F. Two periplasmic transport proteins which interact

with a common membrane receptor show extensive homology: complete nucleotide sequences. Proc. Natl. Acad. Sci. U. S. A. 78, 6038–6042 (1981).

32. Ghimire-Rijal, S., Lu, X., Myles, D. A. & Cuneo, M. J. Duplication of genes in an ATP-binding cassette transport system increases dynamic range while maintaining ligand specificity. J. Biol. Chem. 289, 30090–30100 (2014).

33. Heide, T. Van Der & Poolman, B. ABC transporters : one , two or four. 3, 938–943 (2002).

34. Fulyani, F. et al. Functional diversity of tandem substrate-binding domains in ABC transporters from pathogenic bacteria. Structure 21, 1879–1888 (2013).

35. Husada, F. et al. Watching conformational dynamics of ABC transporters with single-molecule tools. Biochem. Soc. Trans. 43, 1041–7 (2015).

36. Korkhov, V. M., Mireku, S. a. & Locher, K. P. Structure of AMP-PNP-bound vitamin B12 transporter BtuCD–F. Nature 490, 367–372 (2012).

37. Brückner, A., Polge, C., Lentze, N., Auerbach, D. & Schlattner, U. Yeast Two-Hybrid , a Powerful Tool for Systems Biology. 2763–2788 (2009). doi:10.3390/ijms10062763 38. Ll, D., Swift, S. & Lamond, A. I. Detecting Protein-Protein Interactions In Vivo

with FRET using Multiphoton Fluorescence Lifetime Imaging Microscopy ( FLIM ). 1–19 (2007). doi:10.1002/0471142956.cy1210s42

39. Kevin Truong and Mitsuhiko Ikura. The use of FRET imaging microscopy to detect

protein – protein interactions and protein conformational changes in vivo. Curr. Opin. Struct. Biol. 573–578 (2001).

40. Weibrecht, I. et al. Expert Review of Proteomics proteomics toolbox Proximity

ligation assays : a recent addition to the proteomics toolbox. 9450, (2014).

41. Ivanusic, D., Eschricht, M. & Denner, J. Investigation of membrane protein–protein interactions using correlative FRET-PLA. Biotechniques 57, 188–198 (2014). 42. Lemmon, M. A. et al. Glycophorin A dimerization is driven by specific interactions

between transmembrane α-helices. J. Biol. Chem. 267, 7683–7689 (1992).

43. Brosig, B. & Langosch, D. The dimerization motif of the glycophorin A transmembrane segment in membranes: Importance of glycine residues. Protein Sci. 7, 1052–1056 (2008).

44. Chadda, R. et al. The dimerization equilibrium of a ClC Cl À / H + antiporter in lipid bilayers. Elife 5, 1–24 (2016).

45. Last, N. B. & Miller, C. Functional Monomerization of a ClC-Type Fluoride

(16)

SUPPLEMENTARY INFORMATION

Supplementary Figure 1 Transport of [3H]pantothenate (A) and [3H]folate (B) into proteoliposomes with

co-reconstituted ECF-FolT2 and solitary PanT. In panel A, empty circles represent ECF-FolT2 co-co-reconstituted with

PanT, whereas full circles represent [3H]pantothenate uptake by solitary PanT. The folate transport exhibited rather

little activity. That could have been caused by poor quality ECF-FolT2 obtained during the purification together with the use of an old batch of radiolabeled folate (radio-damaged substrate).

(17)

Referenties

GERELATEERDE DOCUMENTEN

A commonly used vector for production of whole complex ECF transporters is the p2BAD plasmid with two identical arabinose promoters, one for the expression of the genes coding for

Rhodobacter capsulatus AMP-PCP - 5X40 1,45 18 Type II ECF transporters ECF-FolT Lactobacillus br evis Nucleotide free Inward-facing, no substrate bound 4HUQ 3 19 ECF-HmpT

Due to high sensitivity of studied transporters to the time spent outside the lipid bilayer as well as to gain a broader picture of transport mechanism we start investigation of lipid

Daarnaast hebben wij de uitwisseling getest tussen twee transport complexen, specifiek voor pantothenate en folate, co-gereconstitueerd met één ECF-module die de E naar Q mutatie

Insights into the transport mechanism of energy-coupling factor transporters Stanek, Weronika Karolina.. IMPORTANT NOTE: You are advised to consult the publisher's version

Dirk’s group: Alisa, Marysia, Sonja, Lotteke, Huifang (Hallie), Michael, Inda, Joris, Lisa, Rianne, Nynke, Rameez, Rajkumar (Raj), Nicola, Li (Iren), Stephan, Gianluca, Dorith,

The observed effects on the uptake rates of one vitamin in response to the presence of another vitamin are the result of differences in expression levels of specific S-component

Download date: 16-07-2021.. PROPOSITIONS belonging to the thesis Insights into the transport mechanism of Energy Coupling Factor transporters 1) S-components associate with