Sei sulla pagina 1di 6

Detergent-free isolation, characterization, and

functional reconstitution of a tetrameric K+ channel:


The power of native nanodiscs
Jonas M. Dörra,1, Martijn C. Koorengevela, Marre Schäfera, Alexander V. Prokofyevb,2, Stefan Scheidelaara,
Elwin A. W. van der Cruijsenb, Timothy R. Daffornc, Marc Baldusb, and J. Antoinette Killiana
a
Membrane Biochemistry and Biophysics, Bijvoet Center for Biomolecular Research, Utrecht University, 3584 CH Utrecht, The Netherlands; bNMR
Spectroscopy, Bijvoet Center for Biomolecular Research, Utrecht University, 3584 CH Utrecht, The Netherlands; and cSchool of Bio Sciences, University of
Birmingham, Edgbaston Birmingham B15 2TT, United Kingdom

Edited by Ramon Latorre, Centro Interdisciplinario de Neurociencias, Universidad de Valparaíso, Valparaíso, Chile, and approved November 21, 2014 (received
for review August 22, 2014)

A major obstacle in the study of membrane proteins is their sol- particles (10–13) (Fig. 1). The mechanism of action of SMA
ubilization in a stable and active conformation when using de- differs fundamentally from that of detergents: instead of dis-
tergents. Here, we explored a detergent-free approach to isolating rupting the lipid bilayer completely, SMA spontaneously self-
the tetrameric potassium channel KcsA directly from the mem- inserts and extracts intact membrane patches in the form of
brane of Escherichia coli, using a styrene-maleic acid copolymer. discoidal particles that are stabilized by a SMA annulus (14, 15).
This polymer self-inserts into membranes and is capable of extract- Because these nanodiscs conserve a spatially delimited native

COMPUTATIONAL BIOLOGY
ing membrane patches in the form of nanosize discoidal proteo- biomembrane including MPs, we term them “native nanodiscs.”
lipid particles or “native nanodiscs.” Using circular dichroism and

BIOPHYSICS AND
One of the main advantages of this system is the straightforward
tryptophan fluorescence spectroscopy, we show that the confor- extraction protocol without the need for detergent. It has been
mation of KcsA in native nanodiscs is very similar to that in de-
shown that the SMA polymer is capable of directly extracting
tergent micelles, but that the thermal stability of the protein is
native nanodiscs containing large functional protein complexes
higher in the nanodiscs. Furthermore, as a promising new applica-
from yeast (12), bacterial proteins involved in cell division (16)
tion, we show that quantitative analysis of the co-isolated lipids in
purified KcsA-containing nanodiscs allows determination of pref-
and photosynthesis (17), and several members of the ABC trans-
erential lipid–protein interactions. Thin-layer chromatography ex- porter family (13). The isolation of these proteins from a variety
periments revealed an enrichment of the anionic lipids cardiolipin of different organisms suggests a general applicability of SMA
and phosphatidylglycerol, indicating their close proximity to the solubilization for all MPs, irrespective of their expression host or
channel in biological membranes and supporting their functional native organism.
relevance. Finally, we demonstrate that KcsA can be reconstituted To further explore the potential of native nanodiscs, we used the
into planar lipid bilayers directly from native nanodiscs, which SMA polymer to isolate an oligomeric bacterial membrane protein:
enables functional characterization of the channel by electrophys- the tetrameric potassium channel from Streptomyces lividans (KcsA)
iology without first depriving the protein of its native environ- (18), expressed in Escherichia coli. KcsA is an ideal model protein
ment. Together, these findings highlight the potential of the use for such studies because it is well-characterized and because
of native nanodiscs as a tool in the study of ion channels, and of
membrane proteins in general. Significance

|
membrane–protein solubilization styrene-maleic acid copolymer | The study of membrane proteins is often hampered by their
| |
lipid–protein interactions nanodisc ion channels tendency to misfold when extracted by detergent. Here, we
explore a detergent-free approach to isolating membrane pro-

I ntegral membrane proteins (MPs) are an abundant class of


proteins that play key roles in a wide range of essential cellular
processes (1). To facilitate their study in vitro, detergent mole-
teins while retaining their native lipid environment, making use
of an amphipathic polymer that solubilizes intact membrane
patches in the form of nanodiscs. Using a potassium channel as
cules are commonly used to extract MPs out of their native lipid– a model protein, we show that these “native nanodiscs” are
bilayer environment (2). However, the use of detergents has highly thermostable particles that are suitable for spectroscopic
some inherent disadvantages. Most importantly, even though there studies, allowing structural characterization of the protein in its
are promising developments to improve their properties (3, 4), the native environment and direct analysis of the lipids in its im-
insufficient mimicking of a lipid bilayer by detergent micelles often mediate surroundings. We also demonstrate that the channel
leads to destabilization and rapid loss of function of the in- can be reconstituted from nanodiscs into planar lipid bilayers for
corporated protein (5). For many functional and structural studies, functional characterization, thus making native nanodiscs an
it is thus necessary to reconstitute the MP into a more stabilizing excellent alternative to detergent solubilization.
environment; for example, by replacing the detergent with am-
Author contributions: J.M.D., M.C.K., and J.A.K. designed research; J.M.D., M.C.K., M.S.,
phipathic polymers (amphipols) (6) or incorporating the MP into
S.S., and E.A.W.v.d.C. performed research; T.R.D. and M.B. contributed new reagents/
lipid nanodiscs with a surrounding protein scaffold (7). Both analytic tools; J.M.D., M.C.K., M.S., A.V.P., and J.A.K. analyzed data; and J.M.D. and J.A.K.
approaches have proven to be valuable tools for the study of wrote the paper.
structural and functional properties of MPs (8, 9); however, The authors declare no conflict of interest.
a limitation remains, as transfer of MPs into any of these systems This article is a PNAS Direct Submission.
requires initial solubilization by detergent. 1
To whom correspondence should be addressed. Email: j.m.dorr@uu.nl.
Recently, a detergent-free approach has been described using 2
Present address: The BIOS Lab on a Chip Group, MESA+ Institute of Nanotechnology,
amphipathic styrene-maleic acid copolymers (SMAs) as an al- University of Twente, 7500 AE Enschede, The Netherlands.
ternative to solubilize MPs directly from biological membranes This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.
in the form of nanodiscs, referred to as “Lipodisq” or SMA lipid 1073/pnas.1416205112/-/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.1416205112 PNAS | December 30, 2014 | vol. 111 | no. 52 | 18607–18612


Native Nanodiscs Surpass Detergent Micelles in Conserving Structural
A B Stability of KcsA. The purity of KcsA nanodiscs after size-exclu-
H H H H sion chromatography renders them directly suitable for studies
C C C C
H with circular dichroism (CD) and tryptophan fluorescence spec-
O O troscopy. Far-UV CD spectra of KcsA in native nanodiscs and
O O– DDM micelles at different temperatures are depicted in Fig. 3A.
H The shapes of the spectra acquired at room temperature are
n m virtually indistinguishable and show the features of a predom-
Styrene–Maleic Acid inantly α-helical protein with minima at around 208 and 222 nm,
in good agreement with previous studies (20, 26) and available
Fig. 1. (A) Chemical structure of SMA polymers at neutral pH. For this study, X-ray structures (27). Upon heat exposure, differences between
a polymer with an average SMA ratio of n:m = 2:1 was used. (B) Schematic
the spectra of KcsA in the different systems become visible. At
representation of a native nanodisc containing a KcsA tetramer (blue) and
native lipids (green). The outer hydrophobic surface of the lipids is shielded
95 °C, KcsA in DDM shows a loss of ∼50% in secondary structure,
by SMA (orange). with a well-defined transition at 72 °C, as measured by the ellip-
ticity at 222 nm (Fig. 3B). In contrast, KcsA in nanodiscs lacks
a genuine transition as a function of temperature, and the amount
reconstitution studies have shown that both its function and of helicity decreases only slightly with increasing temperature,
stability are strongly affected by lipid composition (19–21). In with a loss of ∼20% at 95 °C. In addition, the characteristic fea-
this work, we apply SMA to prepare and purify native nanodiscs tures of α-helical proteins are qualitatively conserved over the
with KcsA to compare the conformational properties and sta- whole temperature range for KcsA in nanodiscs, whereas they are
bility of the protein with those in detergent micelles. In addition, lost in DDM micelles above the transition.
we use native nanodiscs to investigate preferential lipid–protein Similar observations were made when the intrinsic fluores-
interactions by analyzing the composition of small patches of na- cence of KcsA in different environments was monitored. In-
tive membrane that are copurified with the protein. Finally, we creasing the temperature resulted in a strong decline in intensity
study the functional properties of KcsA on reconstitution from of the recorded emission spectra of KcsA that was more pro-
native nanodiscs into a planar lipid bilayer system. Our results nounced in detergent than in nanodiscs (Fig. 3C). At 20 °C, the
underscore the huge potential of SMA as a membrane-solubilizing maximum emission wavelengths on excitation at 295 nm were ∼329
agent, as well as the use of native nanodiscs as a membrane- and ∼332 nm for KcsA in nanodiscs and micelles, respectively. The
mimetic system for biophysical studies on ion channels, and MPs blue-shifted emission maximum in nanodiscs can be explained by
in general. the decreased solvent exposure of some tryptophans resulting from
better shielding by conserved lipid molecules than is seen in de-
Results tergent micelles. The thermal stability of the tertiary structure of
KcsA Can Be Solubilized and Purified in Native Nanodiscs. Prepara- KcsA was investigated by following the shift in the emission max-
tions of KcsA in native nanodiscs were obtained by the addition imum as a function of temperature. Consistent with the CD
of SMA polymer directly to lysed E. coli whole cells with over-
produced His-tagged KcsA, followed by isolation of the KcsA-
containing nanodiscs by Ni-affinity chromatography. SDS/PAGE A NN DDM B 50 * **
analysis of purified KcsA in native nanodiscs shows a pronounced – + – + heat
*

band at a molecular weight of ∼60 kDa (Fig. 2A, lane 2), similar to 250
40
KcsA
150
that observed with tetrameric KcsA purified in a standard pro- 100

tocol, using the nonionic detergent n-dodecyl-β-D-maltoside 75


T SlyD
(DDM) (lane 4) (22). Exposing the samples to 95 °C in the pres- 50 30
A280 (mAU)

ence of SDS results in a virtually complete transition to the mo- 37

nomeric form of KcsA, as evident from the appearance of a band 20


at ∼17 kDa (Fig. 2A, lanes 3 and 5). Thus, KcsA can be isolated in 25 **
native nanodiscs as a stable tetramer that is resistant to SDS at 20
10
room temperature, similar to what is described for KcsA in de-
tergent (23). In the isolations of KcsA in nanodiscs, we observed an M
additional pronounced double band of varying intensity at ∼25 kDa 0
SMA 5 10 15 20
that was heat-stable. Mass spectrometric analysis on a tryptic- Vel (mL)
digested sample of this band identified it as SlyD, a stress-up-
regulated, soluble 20.9-kDa protein endogenous to E. coli that C
contains multiple histidine residues in its C terminus (24). This
common contaminant of Ni-affinity purifications could be re-
moved completely by size-exclusion chromatography (Fig. 2B)
(25). The SMA polymer itself migrates at a low apparent mo-
lecular weight because of its high intrinsic negative charge and is
observed as a blue-stained front on the depicted gels (Fig. 2A, 50 nm
lanes 2 and 3).
The size and shape of purified KcsA-containing nanodiscs were Fig. 2. Purification of KcsA in native nanodiscs. (A) SDS/PAGE analysis of
further investigated, using negative-stain transmission electron purified KcsA in native nanodiscs (NN) and DDM micelles at room temper-
microscopy (Fig. 2C). The particles exhibit a round shape with a ature (−) and after incubation at 95 °C (+). Complete transitions from tet-
rameric (T) to monomeric (M) state are visible. (B) Size exclusion chromatogram
fairly homogeneous size distribution, with diameters of 10 ± 2 nm, showing effective separation of KcsA (*) from the soluble contaminant SlyD
which is in agreement with previous studies on native nanodiscs (**). (C) Negative stain transmission electron micrograph of native nanodiscs
(10, 12) and on protein-free nanodiscs obtained by SMA solubi- with KcsA after size exclusion chromatography. Round particles of an aver-
lization of synthetic liposomes (14, 15). age size of 10 ± 2 nm are visible. (Scale bar in the enlarged images, 10 nm.)

18608 | www.pnas.org/cgi/doi/10.1073/pnas.1416205112 Dörr et al.


A 5
Nanodisc 20 °C DDM 20 °C C Nanodisc 20°C thus directly yields particles that maintain a higher stability of the
1
Nanodisc 95 °C DDM 95 °C Nanodisc 95°C trapped protein than detergent micelles.
DDM 20°C
0 0.8 DDM 95°C
Biochemical Analysis of the Local Lipid Environment of KcsA Reveals
θ (mdeg)

Inorm (a. u.)


-5
0.6 Enrichment of Anionic Lipids. The isolation of intact nanopatches
of biological membranes in native nanodiscs provides a unique
0.4
opportunity to investigate the immediate local lipid environment
-10
0.2 of KcsA, and thus characterize preferential lipid–protein inter-
actions. Using TLC, the isolated lipids from native nanodiscs can
-15 0
200 210 220 230 240 250 300 325 350 375 400 be separated according to headgroup. Fig. 4A shows a represen-
λ (nm) λem (nm) tative chromatogram with lipid samples from the total cell lysate,
B D 345 the complete SMA-solubilized fraction, and purified KcsA nano-
-5 DDM DDM discs. In all samples, three pronounced bands are visible that can
θ222nm (mdeg)

λem, max (nm)

Nanodisc 339 Nanodisc


-10
be identified by comparison with synthetic reference lipids as
333 the zwitterionic phosphatidylethanolamine and the anionic lip-
-15 327
20 45 70 95 20 45 70 95 ids phosphatidylglycerol and cardiolipin. The lipid composition
T (°C) T (°C) of independent bacterial cultures was found to be very sensitive
Fig. 3. Comparison of the thermal stability of KcsA in different environ-
to the growth conditions and time, resulting in a variation of
ments. (A) Circular dichroism spectra of KcsA in native nanodiscs and DDM absolute values in the range of 5–10 mol%. For comparison,
micelles at 20 °C and 95 °C. Data are offset corrected averages of 8–10 scans. data from each independent nanodisc preparation were there-
(B) Corresponding thermal unfolding traces monitored by the ellipticity at fore normalized to the corresponding total cell lysate. Quanti-
222 nm. Data are averages of two experiments with errors too small to be tative analysis of the intensity of the chromatogram bands (Fig.
depicted. Solid lines are depicted to guide the eye. (C) Normalized fluores- 4B) shows that the lipid compositions of the total cell lysate and

COMPUTATIONAL BIOLOGY
cence intensity of the intrinsic tryptophans of KcsA at 20 °C and 95 °C. Data the soluble fraction are very similar and in good agreement with

BIOPHYSICS AND
are averages of three scans normalized to the intensity at 330 nm of the other studies on lipids of E. coli K 12 wild-type strains (28, 29).
respective spectra at 20 °C. (D) Corresponding thermal unfolding traces mon-
Thus, SMA does not preferentially solubilize any specific lipids
itored by the wavelength of maximum fluorescence emission. Solid lines are
depicted to guide the eye.
in the E. coli membrane. Strikingly, KcsA nanodiscs show higher
amounts of anionic lipids, with increases of 36% in phosphati-
dylglycerol and 61% in cardiolipin content compared with the
analysis, KcsA in detergent micelles shows a genuine transition at total extract. This isolation of enriched lipid species can hence
∼70 °C, whereas in nanodiscs, changes are less pronounced, be attributed to the preferential interaction of anionic lipids
without a clear transition for KcsA (Fig. 3D). The emission max- with KcsA.
imum of micellar KcsA at 95 °C is ∼343 nm, and is thus consid-
erably more red-shifted than in nanodiscs (∼337 nm). Virtually the
same results were obtained when the intensity of the fluorescence A 1 2 3
B 80 Total
at 330 nm was monitored as function of temperature. Thus, the
Relave amount (mol%)
Soluble
tryptophan residues of detergent-solubilized KcsA show larger CL 60 KcsA
conformational changes than those in nanodiscs, where the tertiary
structure is more stable even at high temperatures. PE 40
PG
The higher stability of KcsA in nanodiscs compared with
DDM micelles is further supported by the observation of a faint 20
monomer band of KcsA purified in DDM on SDS/PAGE at
room temperature that is absent for nanodiscs (Fig. 2A, lanes 2 Origin 0
CL PE PG
and 4). Upon storage at 4 °C, we observed an increase of this
difference over time. Tetrameric KcsA in nanodiscs was stable C 45
Total
over months, with a constant tetramer fraction above 95%,
Relave amount (weight%)

Soluble
whereas the tetramer fraction of DDM-solubilized protein de- KcsA
creased to ∼60% after 6 weeks. In line with this, we found that 30
neither freeze thawing nor lyophilization of KcsA in nanodiscs
affects the integrity of the quaternary structure, whereas DDM-
solubilized tetramers dissociate to some extent (Fig. S1). Thus, 15

using a complementary set of techniques, it is shown that native


nanodiscs serve as a good tool to conserve the structural stability
0
of the oligomeric KcsA. 12:0 14:0 16:0 16:1 17:c 18:0 18:1 19:c
A possible explanation for the lower stability of detergent- Fay acid
solubilized KcsA is the dynamic equilibrium between monomeric
Fig. 4. Analysis of coisolated lipids in native nanodiscs. (A) TLC on extracted
and micellar DDM (2), which may promote the dissociation of
lipids from lysed whole bacterial cells after addition of SMA (1), complete
the KcsA tetramer by segregation of monomers into separate soluble fraction separated by ultracentrifugation (2), and native nanodiscs
micelles, thus favoring further unfolding. In contrast, native with KcsA purified by Ni-affinity chromatography (3). (B) Molar composition
nanodiscs are stable particles, and no free SMA is needed in of the extracted lipid species according to headgroup of total lysate (or-
solution to maintain their stability. Thus, the more rigid structure ange), soluble fraction (blue), and KcsA nanodiscs (green). Data are shown as
of the particles could convey the higher stability of KcsA. How- average mol percentages with SDs (red error bars) of three independent
nanodisc isolations. Data for the soluble fraction and KcsA nanodiscs were
ever, it should also be noted that the mere presence of lipids can
normalized to the corresponding total lysate. Black error bars denote SDs of
increase the stability of KcsA, as suggested by studies on recon- the change compared with total lysate. (C) Gas chromatography analysis of
stituted protein (20, 22). Conserving and stabilizing a lipid en- the fatty acid composition of all isolated lipids. Depicted data are averages
vironment around MPs during their extraction by SMA polymers of two independent experiments. Errors are given as in B.

Dörr et al. PNAS | December 30, 2014 | vol. 111 | no. 52 | 18609
Preferential lipid–protein interactions were further analyzed Discussion
by investigating the composition of the acyl chains of all isolated We have shown that native nanodiscs with embedded KcsA are
lipids (Fig. 4C). Lipids from all samples predominantly contained formed spontaneously from bacterial cells upon addition of SMA
palmitic (16:0) and cis-vaccenic acid (18:1) chains with a combined polymer and that, using Ni-affinity and size-exclusion chro-
weight percentage of ∼70%, in agreement with other studies (28). matography, sufficient protein can be purified in nanodiscs for
In our analysis, no strong enrichment of specific acyl chains could extensive biophysical characterization. The isolation of KcsA in
be detected when comparing the lipids that were copurified with this manner is less cumbersome than standard detergent proto-
KcsA with the total extract or with the SMA-solubilized fraction. cols and has the additional advantage of conserving the native
Only the amount of short lauric (12:0) and myristic acid (14:0) conformation of the protein in a stabilizing environment com-
chains appeared to be slightly decreased in purified KcsA nano- prising native lipids. Our data suggest that native nanodiscs in
discs in favor of a slightly higher cis-vaccenic acid content. These general convey a higher stability of the incorporated protein than
results suggest a minor preference of KcsA for lipids with longer detergent micelles, in agreement with previous findings on the
chains, possibly to promote hydrophobic matching (30, 31). thermostability of an ABC transporter (13) and photosynthetic
reaction centers (17). In addition, as shown in this study, the
Reconstitution of KcsA into Planar Lipid Bilayers Enables Functional coisolation of MPs with a patch of biological membrane allows
Characterization. A final challenge remains in assessing the for the analysis of preferential interactions and further facilitates
functionality of KcsA in native nanodiscs. Because analysis of structural and functional studies on purified protein in a (near)
ion transport is not possible in nanodiscs, we attempted to re- native state, as discussed next.
constitute the channel into a compartment-forming bilayer sys- Insights into preferential lipid–protein interactions are im-
tem that enables investigations of protein-facilitated potassium portant to understand structural and functional properties of
conductivity across membranes. To this end, we used a planar MPs in a membrane environment. However, detailed infor-
lipid bilayer setup that has been used successfully for functional mation is not easily obtained with standard methods. Here, we
studies on KcsA delivered in protein-stabilized nanodiscs (32). exploited the extraction of intact nanopatches of biological
Seconds to minutes after the addition of native nanodiscs with membranes by the SMA polymer to directly identify preferential
KcsA, single-channel conductivity was observed as transient lipid–protein interactions that allow approximations of the situ-
discrete current changes with an amplitude of 10–15 pA (Fig. ation in vivo. Analysis of the lipid composition revealed an en-
5), implying spontaneous fusion events of single nanodiscs with richment of anionic lipids in close proximity to KcsA. This
the bilayer. The recorded traces show both high and low open finding is unlikely to be an artifact of the extraction protocol
probability states that are characteristic for KcsA monitored because we found that none of the three main E. coli lipid spe-
under similar conditions by single-channel recordings of KcsA cies is preferentially incorporated into nanodiscs. This is in-
fused from liposomes (33) or incorporated via a cell-free ex- triguing, as the SMA polymer has a high negative charge density
pression protocol (34). The results are also comparable with at pH 8.0 because of its many carboxylic acid moieties, which
those obtained by patch clamp studies of giant unilamellar could be expected to lead to electrostatic repulsion of anionic
vesicles with KcsA reconstituted from detergent (35). Together, lipids. The absence of a measurable effect of this repulsion hence
this suggests identical functional properties of KcsA recon- emphasizes the promiscuity of the polymer with respect to lipid
stituted from native nanodiscs compared with standard recon- species in biological membranes. Similarly, SMA did not pref-
stitution protocols. Upon addition of the K+-channel blocker erentially solubilize lipids containing specific fatty acids. This
tetraethylammonium, we observed a reversible complete de- indicates that lipid solubilization by SMA is applicable to all
pletion of all opening events (Fig. S2), further confirming the phospholipids in any membrane, irrespective of headgroup or
presence of reconstituted KcsA in the bilayer. Native nanodiscs acyl chains, as further supported by the successful extraction of
thus may serve as a powerful alternative to conventional ap- all major phospholipids of the mitochondrial membrane in
proaches for the functional reconstitution of ion channels. yeast (12), as well as the membrane of the photosynthetic
bacterium Rhodobacter sphaeroides (17).
The importance of anionic lipids for KcsA functionality and
A 10 pA
B 2 their close association with the channel has been well established
1s
by a wealth of studies using fluorescence quenching by bromi-
Events x 103

1.5
1 nated lipids (33), mass spectrometry (36), electrophysiology (33,
5 pA 0.5
37, 38), and molecular dynamic simulations (39, 40). In addition,
0.2 s
0
the diacylglycerol fragment found in crystal structures of KcsA
-5 0 5 10 15 20 was attributed to phosphatidylglycerol (19), and it has been
Current (pA)
10 pA 2 shown that anionic lipids strongly stabilize the KcsA tetramer
against dissociation by both heat (20, 22) and small fluorinated
Events x 103

1s
1.5
1 alcohols (21). In these studies, KcsA was purified in detergent,
0.5 and preferential interactions were either deduced from the co-
5 pA 0
purification of single tightly bound lipids or were observed in-
0.2 s -5 0 5 10 15 20 directly after reconstitution into synthetic lipid bilayer systems.
Current (pA)
In contrast, the method described in this work facilitates direct
Fig. 5. Functional characterization of KcsA reconstituted from native biochemical analysis of preferential lipid–protein interactions
nanodiscs into a planar lipid bilayer of E. coli polar lipid extract. (A) Typical that involve the native annular lipid environment. This paves the
single-channel current traces on addition of KcsA nanodiscs to a setup with way for a new application to study specific lipid–protein inter-
a symmetric 150-mM KCl solution at +100 mV. The top part represents actions by using the SMA polymer to isolate nanodiscs with
a channel in the high open probability mode that is characterized by long protein that has been conventionally reconstituted into bilayers
opening dwell times, as shown in the enlarged section indicated by a box of
dashed lines. Channels in the low open probability mode exhibit a distinctly
of well-defined composition. By systematically varying the com-
different pattern, with short opening times resulting in sequences of bursts position of these bilayers, one can then obtain very detailed in-
(Bottom). (B) All-point histograms for open/closed distribution of channels formation on the specificity of KcsA or other proteins for both
with high (Top) and low (Bottom) open probability calculated from the lipid headgroups and acyl chains. Aside from assessing lipid–
enlarged 2-s single-channel current traces from A. protein interactions, native nanodiscs offer the important

18610 | www.pnas.org/cgi/doi/10.1073/pnas.1416205112 Dörr et al.


additional advantage of copurifying proteins or other molecules containing 600 mM NaCl and 30 mM KCl (10 mL per 2 g wet cell weight) and
that weakly interact with the target MP, providing unique in- incubated with 10 μg/mL lysozyme and 5 μg/mL DNase for 45 min on ice. The
sights into the composition of its immediate membrane envi- suspension was aliquoted and mixed 1:1 with 6% (wt/vol) SMA in 50 mM Tris
at pH 8 for the preparation of native nanodiscs or 1:1 with 50 mM Tris buffer
ronment. This feature has been used very recently for the isolation of
at pH 8 for detergent solubilization, respectively. The cell suspension with
a detergent-labile complex of penicillin-binding proteins in Staphy- SMA was pushed through a disruptor at a pressure of 215 MPa and in-
lococcus aureus (16). Native nanodiscs thus are a powerful tool to cubated overnight with gentle agitation at room temperature, followed by
study preferential interactions of MPs both in vitro and in vivo. 45 min of centrifugation at 100,000 × g. SDS/PAGE analysis of supernatant
In addition to purification of MPs and analysis of their pref- and pellet after centrifugation revealed that 70–80% of the total amount of
erential interactions, native nanodiscs also allow structural and tetrameric KcsA was solubilized by SMA. The supernatant was then in-
functional characterization. Here, we used CD and fluorescence cubated for at least 4 h at 4 °C with 2.5 mL HisPur Ni-NTA agarose beads
spectroscopy to study the thermostability of KcsA in a straight- (Thermo Scientific) per liter of culture medium. The beads were transferred
forward way. Moreover, their small size makes native nanodiscs, to a gravity-flow column and washed three times each with buffers con-
taining 10 and 50 mM imidazole. Protein in native nanodiscs was then eluted
in principle, suitable for characterization by the full range of
with buffer containing 300 mM imidazole, yielding 1–2 mg protein per liter
biophysical techniques, including NMR spectroscopy, that have of culture medium on average, estimated by SDS/PAGE analysis and densi-
been successfully applied to MPs incorporated into similar lipid tometric comparison of the protein bands after Coomassie staining with
nanodiscs that are stabilized by scaffold proteins instead of SMA a gradient of BSA of known concentration, using the Quantity One software
(9). Native nanodiscs are also promising tools for studies on MPs (Biorad). For biophysical studies, the pooled eluted fractions were loaded on
in the rapidly developing field of single-particle cryo-electron a Superdex 200 10/300 GL size exclusion column (GE Healthcare) connected
microscopy. The use of amphipols to stabilize MPs has recently to an Äkta Prime Plus chromatography system (GE Healthcare) to separate
led to a high-resolution de novo structure determination of a the protein-containing nanodiscs from contaminating soluble proteins and free
cation channel at a resolution of 3.4 Å (41), breaking the side- polymer. Thereby, the buffer was exchanged to 10 mM Tris at pH 8, 100 mM
chain resolution barrier. Together with promising studies using NaCl, and 5 mM KCl, which was used in all subsequent experiments.
Detergent purification was performed on aliquots of lysed cells, as de-

COMPUTATIONAL BIOLOGY
scaffold-protein nanodiscs (42, 43), as well as initial studies with
scribed earlier (22), with the difference of using Tris buffer at pH 8.0 instead
native nanodiscs (44), these new developments suggest that

BIOPHYSICS AND
of Hepes. Eluted protein was extensively dialyzed against 1 mM DDM in the
native nanodiscs may become a powerful alternative to enable same buffer that was used for nanodiscs, using a dialysis membrane with a
high-resolution structural characterization of MPs in their na- 50-kDa molecular weight cutoff. Protein concentration was determined by
tive environment. the absorption at 280 nm, using a calculated extinction coefficient of 34,950
Several proteins have so far been functionally characterized in M−1 · cm−1 (45). For protein in native nanodiscs, this can only be considered
native nanodiscs by investigating their ligand-binding (13) and an estimate, as the phenyl groups of SMA contribute to the absorption in
optical (10–12, 17) properties, as well as their enzymatic activity this wavelength range.
(10, 12). Our data demonstrate that native nanodiscs can be also
used to directly reconstitute KcsA into planar lipid bilayers, Spectroscopy. Far-UV CD experiments were performed on a J-810 spec-
allowing electrophysiologic characterization of single channels in tropolarimeter (Jasco) with a Peltier thermo control element (Jasco), using Teflon-
sealed, polarimetrically checked quartz glass cuvettes with an optical pathlength
membranes with well-defined composition. To the best of our
of 1 mm and a volume of 350 μL (Hellma Analytics). Experimental parameters
knowledge, this represents the first evidence of purification and included a wavelength increment of 1 nm, a scan speed of 20 nm/min, a re-
transfer of an MP from the membrane of living cells to synthetic sponse time of 4 s, and a KcsA concentration of 0.06 mg/mL, as determined for
lipid bilayers without being deprived of its native lipid environ- DDM-solubilized protein. KcsA in native nanodiscs was diluted such that samples
ment during any stage in the procedure. The ability of native had the same signal intensity as protein in DDM, correcting for inaccuracies in
nanodiscs to fuse with bilayers can be considered a promising concentration determination. Samples were allowed to equilibrate for 30 min
first step toward enabling crystallization trials of MPs in a near- at the desired temperature before measurement. All resulting spectra are
native environment. Other prospects of successful reconstitution buffer- and offset-corrected averages of 8–10 scans in the range of 200–250 nm.
include a plethora of assays available for MPs in liposomes and Tryptophan fluorescence measurements were performed on a Cary Eclipse
supported bilayers in applications that require a well-defined spectrofluorometer (Varian) equipped with a thermo control unit (Varian) in
sealed quartz glass cuvettes with optical path lengths of 4 × 10 mm and an
lipid environment. With the inherent advantage of conserving
inner volume of 1.4 mL (Hellma). The excitation light beam had a wave-
the native lipid environment of MPs, native nanodiscs consti- length of 295 nm at a slit size of 5 nm, and the fluorescence emission was
tute a highly promising and convenient alternative to detergent recorded in the range of 300–400 nm, using a slit size of 5 nm, a wavelength
solubilization and may lead the way toward an in situ approach increment of 1 nm, a scan speed of 60 nm/min, and an integration time of
for structural and functional studies on MPs, including phar- 1 s. The KcsA concentration was 0.01 mg/mL. All samples were allowed to
macologic assays. equilibrate for 30 min at the desired temperature. Spectra were recorded in
triplicate and corrected for buffer contribution before analysis by nonlinear
Materials and Methods least squares fitting to a bimodal log-normal distribution, using the Excel
Materials and SMA Preparation. All chemicals and enzymes were purchased add-in Solver (Frontline Systems) (46).
from Sigma-Aldrich unless otherwise indicated. DDM was from Affymetrix,
and all reference lipids used were from Avanti Polar Lipids. The used polymer Lipid Analysis. Before lipid isolation, KcsA-containing native nanodiscs that
was SMA2000, a styrene-maleic anhydride copolymer with a molar styrene were eluted from Ni-NTA beads were washed with buffer on spin columns
maleic anhydride ratio of 2:1, a weight average molecular weight of 7.5 kDa, with a molecular weight cutoff of 30 kDa to remove imidazole. Lipids were
and a molar mass dispersity (polydispersity index) of 2.5 (Cray Valley). Con- then extracted according to a modified version of the method of Bligh and
version into SMA was achieved by hydrolysis in 1 M KOH and reflux for 2 h Dyer (47) and analyzed by quantitative TLC, as well as gas chromatography
while heating the suspension to 100 °C. Subsequently, the polymer was (for details, see SI Materials and Methods).
precipitated by the addition of HCl and washed 6 times with 100 mM HCl to
remove K+ ions. The sample was lyophilized, and hydrolysis was confirmed Electrophysiology. Single-channel recordings of KcsA were performed on a
by Fourier transform infrared spectroscopy (17). SMA stock solutions were Compact setup for planar lipid bilayer electrophysiology (Ionovation) con-
prepared by dissolving 6% (wt/vol) SMA powder in 50 mM unadjusted Tris nected to an EPC 10 amplifier (HEKA). Lipid bilayer formation was achieved by
buffer with gradual addition of NaOH solution until the pH reached the painting E. coli polar lipid extract dissolved in n-decane (50 mg/mL) over
neutral range. The solution was then stored at −20 °C and adjusted to pH 8.0 a 200-μm hole in a Teflon-septum separating two compartments. Both
and to the desired volume after thawing. compartments contained 150 mM KCl solution that was buffered with
10 mM Hepes to pH 7.0 in the cis compartment and 10 mM succinic acid to
Gene Expression and Protein Purification. KcsA was produced as described pH 4.0 in the trans compartment, respectively. After channel insertion, the
earlier (22). Cell pellets were resuspended in 50 mM Tris buffer at pH 8.0 conductivity at a constant voltage of +100 mV was recorded for several

Dörr et al. PNAS | December 30, 2014 | vol. 111 | no. 52 | 18611
minutes, using the PatchMaster software (HEKA). Data were sampled at 10 kHz the transmission electron microscopy measurements. We further thank Ruud
and digitally filtered at 1 kHz. All measurements were performed at 22 °C. Cox for help with the gas chromatography analysis, Hugo van den Hoek for
For a more detailed description of the method used for incorporation of assistance with size-exclusion chromatography, and Cray Valley for their
kind gift of SMA2000 polymer. Financial support received from the seventh
KcsA channels into planar bilayers as well as electron microscopy, see SI
framework program of the European Union (Initial Training Network “Mani-
Materials and Methods. Fold,” Grant 317371 to J.M.D.) and from the Netherlands Organisation for
Scientific Research (Grants 700.11.334 and 700.58.102 to E.A.W.v.d.C. and
ACKNOWLEDGMENTS. We thank Mohammed Jamshad and Rosemary M.B.), as well as via the research program of the Foundation for Funda-
Parslow for help with the initial trials of solubilization of KcsA in native mental Research on Matter (to S.S.), is gratefully acknowledged. T.R.D.
nanodiscs. We are indebted to Mirjam Damen for performing the mass acknowledges support from the Biotechnology and Biological Sciences Re-
spectrometric analysis on the SlyD protein and Hans Meeldijk for help with search Council (Grants BB/J017310/1, BB/I020349/1 and BB/G010412/1).

1. von Heijne G (2007) The membrane protein universe: What’s out there and why 24. Wülfing C, Lombardero J, Plückthun A (1994) An Escherichia coli protein consisting of
bother? J Intern Med 261(6):543–557. a domain homologous to FK506-binding proteins (FKBP) and a new metal binding
2. Garavito RM, Ferguson-Miller S (2001) Detergents as tools in membrane biochemistry. motif. J Biol Chem 269(4):2895–2901.
J Biol Chem 276(35):32403–32406. 25. Parsy CB, Chapman CJ, Barnes AC, Robertson JF, Murray A (2007) Two-step method to
3. McQuade DT, et al. (2000) Rigid amphiphiles for membrane protein manipulation. isolate target recombinant protein from co-purified bacterial contaminant SlyD after
Angew Chem Int Ed Engl 39(4):758–761. immobilised metal affinity chromatography. J Chromatogr B Analyt Technol Biomed
4. Chae PS, et al. (2010) Maltose-neopentyl glycol (MNG) amphiphiles for solubilization, Life Sci 853(1-2):314–319.
stabilization and crystallization of membrane proteins. Nat Methods 7(12):1003–1008. 26. van den Brink-van der Laan E, Chupin V, Killian JA, de Kruijff B (2004) Stability of KcsA
5. Bowie JU (2001) Stabilizing membrane proteins. Curr Opin Struct Biol 11(4):397–402. tetramer depends on membrane lateral pressure. Biochemistry 43(14):4240–4250.
6. Tribet C, Audebert R, Popot JL (1996) Amphipols: Polymers that keep membrane 27. Uysal S, et al. (2009) Crystal structure of full-length KcsA in its closed conformation.
proteins soluble in aqueous solutions. Proc Natl Acad Sci USA 93(26):15047–15050. Proc Natl Acad Sci USA 106(16):6644–6649.
7. Civjan NR, Bayburt TH, Schuler MA, Sligar SG (2003) Direct solubilization of heterol- 28. Raetz CR (1978) Enzymology, genetics, and regulation of membrane phospholipid
ogously expressed membrane proteins by incorporation into nanoscale lipid bilayers. synthesis in Escherichia coli. Microbiol Rev 42(3):614–659.
Biotechniques 35(3):556–560, 562–563. 29. Morein S, Henricson D, Rilfors L (1994) Separation of inner and outer membrane vesicles
8. Popot JL (2010) Amphipols, nanodiscs, and fluorinated surfactants: Three non- from Escherichia coli in self-generating Percoll gradients. Anal Biochem 216(1):47–51.
conventional approaches to studying membrane proteins in aqueous solutions. Annu 30. Williamson IM, Alvis SJ, East JM, Lee AG (2002) Interactions of phospholipids with the
Rev Biochem 79:737–775. potassium channel KcsA. Biophys J 83(4):2026–2038.
9. Bayburt TH, Sligar SG (2010) Membrane protein assembly into Nanodiscs. FEBS Lett 31. Killian JA (1998) Hydrophobic mismatch between proteins and lipids in membranes.
584(9):1721–1727. Biochim Biophys Acta 1376(3):401–415.
10. Knowles TJ, et al. (2009) Membrane proteins solubilized intact in lipid containing 32. Banerjee S, Nimigean CM (2011) Non-vesicular transfer of membrane proteins from
nanoparticles bounded by styrene maleic acid copolymer. J Am Chem Soc 131(22):
nanoparticles to lipid bilayers. J Gen Physiol 137(2):217–223.
7484–7485.
33. Marius P, et al. (2008) Binding of anionic lipids to at least three nonannular sites on the
11. Orwick-Rydmark M, et al. (2012) Detergent-free incorporation of a seven-trans-
potassium channel KcsA is required for channel opening. Biophys J 94(5):1689–1698.
membrane receptor protein into nanosized bilayer Lipodisq particles for functional
34. Friddin MS, et al. (2013) Single-channel electrophysiology of cell-free expressed ion
and biophysical studies. Nano Lett 12(9):4687–4692.
channels by direct incorporation in lipid bilayers. Analyst (Lond) 138(24):7294–7298.
12. Long AR, et al. (2013) A detergent-free strategy for the reconstitution of active en-
35. Chakrapani S, Cordero-Morales JF, Perozo E (2007) A quantitative description of KcsA
zyme complexes from native biological membranes into nanoscale discs. BMC Bio-
gating II: Single-channel currents. J Gen Physiol 130(5):479–496.
technol 13:41.
36. Demmers JAA, van Dalen A, de Kruijff B, Heck AJR, Killian JA (2003) Interaction of the
13. Gulati S, et al. (2014) Detergent-free purification of ABC (ATP-binding-cassette)
K+ channel KcsA with membrane phospholipids as studied by ESI mass spectrometry.
transporters. Biochem J 461(2):269–278.
FEBS Lett 541(1-3):28–32.
14. Orwick MC, et al. (2012) Detergent-free formation and physicochemical character-
37. Cuello LG, Romero JG, Cortes DM, Perozo E (1998) pH-dependent gating in the
ization of nanosized lipid-polymer complexes: Lipodisq. Angew Chem Int Ed Engl
Streptomyces lividans K+ channel. Biochemistry 37(10):3229–3236.
51(19):4653–4657.
38. Iwamoto M, Oiki S (2013) Amphipathic antenna of an inward rectifier K+ channel
15. Jamshad M, et al. (2014) Structural analysis of a nanoparticle containing a lipid bilayer
used for detergent-free extraction of membrane proteins. NanoResearch, in press. responds to changes in the inner membrane leaflet. Proc Natl Acad Sci USA 110(2):
16. Paulin S, et al. (2014) Surfactant-free purification of membrane protein complexes 749–754.
from bacteria: Application to the staphylococcal penicillin-binding protein complex 39. Deol SS, Domene C, Bond PJ, Sansom MSP (2006) Anionic phospholipid interactions
PBP2/PBP2a. Nanotechnology 25(28):285101. with the potassium channel KcsA: Simulation studies. Biophys J 90(3):822–830.
17. Swainsbury DJK, Scheidelaar S, van Grondelle R, Killian JA, Jones MR (2014) Bacterial 40. Weingarth M, et al. (2013) Structural determinants of specific lipid binding to po-
reaction centers purified with styrene maleic Acid copolymer retain native membrane tassium channels. J Am Chem Soc 135(10):3983–3988.
functional properties and display enhanced stability. Angew Chem Int Ed Engl 53(44): 41. Liao M, Cao E, Julius D, Cheng Y (2013) Structure of the TRPV1 ion channel de-
11803–11807. termined by electron cryo-microscopy. Nature 504(7478):107–112.
18. Schrempf H, et al. (1995) A prokaryotic potassium ion channel with two predicted 42. Gogol EP, et al. (2013) Three dimensional structure of the anthrax toxin translocon-
transmembrane segments from Streptomyces lividans. EMBO J 14(21):5170–5178. lethal factor complex by cryo-electron microscopy. Protein Sci 22(5):586–594.
19. Valiyaveetil FI, Zhou Y, MacKinnon R (2002) Lipids in the structure, folding, and 43. McCoy JG, et al. (2014) A KcsA/MloK1 chimeric ion channel has lipid-dependent
function of the KcsA K+ channel. Biochemistry 41(35):10771–10777. ligand-binding energetics. J Biol Chem 289(14):9535–9546.
20. Triano I, et al. (2010) Occupancy of nonannular lipid binding sites on KcsA greatly 44. Postis V, et al. (2014) The use of smalps as a novel membrane protein scaffold for
increases the stability of the tetrameric protein. Biochemistry 49(25):5397–5404. structure study by negative stain electron microscopy. Biochim Biophys Acta – Bio-
21. Raja M, Spelbrink REJ, de Kruijff B, Killian JA (2007) Phosphatidic acid plays a special membranes, in press.
role in stabilizing and folding of the tetrameric potassium channel KcsA. FEBS Lett 45. Gill SC, von Hippel PH (1989) Calculation of protein extinction coefficients from amino
581(29):5715–5722. acid sequence data. Anal Biochem 182(2):319–326.
22. van Dalen A, Hegger S, Killian JA, de Kruijff B (2002) Influence of lipids on membrane 46. Kemmer G, Keller S (2010) Nonlinear least-squares data fitting in Excel spreadsheets.
assembly and stability of the potassium channel KcsA. FEBS Lett 525(1-3):33–38. Nat Protoc 5(2):267–281.
23. Cortes DM, Perozo E (1997) Structural dynamics of the Streptomyces lividans K+ channel 47. Bligh EG, Dyer WJ (1959) A rapid method of total lipid extraction and purification.
(SKC1): Oligomeric stoichiometry and stability. Biochemistry 36(33):10343–10352. Can J Biochem Physiol 37(8):911–917.

18612 | www.pnas.org/cgi/doi/10.1073/pnas.1416205112 Dörr et al.

Potrebbero piacerti anche