You are on page 1of 11

BBA - General Subjects 1864 (2020) 129330

Contents lists available at ScienceDirect

BBA - General Subjects


journal homepage: www.elsevier.com/locate/bbagen

Microsystem for the single molecule analysis of membrane transport T


proteins
Rikiya Watanabe
Molecular Physiology Laboratory, RIKEN, Saitama, Japan

A R T I C LE I N FO A B S T R A C T

Keywords: Micro-chamber arrays enable highly sensitive and quantitative bioassays at the single-molecule level.
Biomembrane Accordingly, they are widely used for ultra-sensitive biomedical applications, e.g., digital PCR and digital ELISA.
Membrane protein However, the versatility of micro-chambers is generally limited to reactions in aqueous solutions, although
Single molecule biophysics various functions of membrane proteins are extremely important. To address this issue, microsystems using
bioMEMS
arrayed micro-sized chambers sealed with lipid bilayers, referred to here as a “biomembrane microsystems”,
have been developed by many research groups for the analysis of membrane proteins. In this review, I would like
to introduce recent progress on the single molecule analysis of membrane transport proteins using a biomem-
brane microsystem, and discuss the future prospects for its use in analytical and pharmacological applications.

1. Introduction with lipid bilayers have been developed by many research groups for
the analysis of membrane transport; in these systems, transport activity
The maintenance of an appropriate intracellular environment is a is measured optically based on the accumulation or consumption of a
constant challenge for all living organisms, from prokaryotes to mul- substrate present in the chambers [9–15]. Microsystems enhance both
ticellular eukaryotes. Intracellular homeostasis is in general maintained sensitivity and throughput, thereby allowing the highly sensitive ana-
by membrane proteins which mediate the translocation of various lysis of various transporter proteins in a high throughput manner. No-
compounds across biomembranes [1]. Therefore, the analysis of tably, our “biomembrane microsystem”, which contains > 100,000
transport mechanisms is crucial to understand cell physiology, as well arrayed atto- to femto-liter chambers sealed with stable lipid bilayers,
as the action of drugs [2,3]. The transport of various substrates is has been used to perform the most sensitive and parallel analysis of
mainly conducted by transport proteins embedded in the cell mem- membrane transport, enabling the single molecule analysis of trans-
brane. For several decades, the functional dynamics of transporters porters with extremely low transport activities (< 10 molecules s−1)
have been studied using a variety of single-molecule techniques [4,5], [5,16]. Moreover, we have successfully demonstrated some physiolo-
which offer key benefits over macroscopic assay methods, as they un- gical aspects of the membrane system, such as an asymmetric transbi-
lock the ability to quantify transport events [6]. layer phospholipid distribution [16,17], and the modulation of mem-
One of the most robust systems used to analyze membrane transport brane potential across lipid bilayers [18]. Thus, our system paves the
at the single molecule level is patch clamp recording, which measures way for understanding the mechanism of membrane transport under
the flux of ions as an electric current under constant electrical voltage semi-physiologic conditions, as well as allowing for further analytical
across a membrane. Recent developments have resulted in the auto- and pharmacological applications. In this review, I would like to in-
mation of patch clamp recordings, thus enabling massively parallel troduce our current microsystem, and discuss the future prospects for
analysis of transport activities, e.g. pores and channels [7,8]; however, membrane transport analysis at the single molecule level.
transport rates of most transporters (< 102 molecules s−1) are much
smaller than those of ion channels (> 107 molecules s−1), and there- 2. Part I. Biomembrane microsystems
fore, it is difficult to detect their activities as an electric current.
Moreover, patch clamp recording cannot detect the flux of electrically Over the past few years, we have developed various “biomembrane
neutral substrates, and therefore, the development of a more versatile microsystems” in order to carry out the single molecule analysis of
system has been long desired. membrane transport proteins under semi-physiological conditions
Recently, microsystems arrayed with micro-sized chambers sealed [5,16–21]. The microsystem consists of a glass substrate containing the

E-mail address: rikiya.watanabe@riken.jp.

https://doi.org/10.1016/j.bbagen.2019.03.016
Received 9 December 2018; Received in revised form 19 March 2019; Accepted 22 March 2019
Available online 26 March 2019
0304-4165/ © 2019 Elsevier B.V. All rights reserved.
R. Watanabe BBA - General Subjects 1864 (2020) 129330

Fig. 1. Biomembrane microsystem.


(a) Components of the system. Top, a glass block
with a sample injection port (l: w: h = 20 mm:
20 mm: 5 mm); middle, a spacer sheet with one side
open (Frame-Seal, BIO-RAD); bottom, a glass slide
with > 10,000 microchambers printed. (b)
Photograph (top) and illustration (bottom) of the
assembled system. The access ports for sample in-
jection are indicated using red circles. (c) Schematic
illustration of the generation of phospholipid mem-
branes. Membrane bilayers were formed over mi-
crochambers via sequential injection of liquids from
the access port. Gray, blue, and yellow are the first
aqueous solution, second aqueous solution, and
chloroform solution containing lipids, respectively.
(d) Fluorescent image of Alexa 647 encapsulated into
microchambers sealed by lipid membranes.

microarray, a spacer sheet, and a glass block with an access port for chamber via the sequential injection of liquids from the access port
sample injection (Fig. 1a). Microarrays with > 10,000 micro-chambers (Fig. 1c). This technique is similar to the conventional technique known
(ɸ = 2.4–8.2 μm and h = 0.03–0.5 μm: V = 0.2–71 fL) were fabricated as the “painting technique,” which forms lipid bilayers through the
using soft lithography on a hydrophilic glass substrate coated with deposition of a lipid solution across a small aperture in a hydrophobic
CYTOP (Asahi-glass), a carbon‑fluorine hydrophobic polymer (Fig. 1b). material [12,23]. The experimental procedures used to form the lipid
Notably, in previous studies fluorine polymers have frequently been membrane are comprised of three steps. First, the micro-chambers are
used as a support medium for lipid membranes [22]. The fabrication filled with an aqueous solution. A chloroform solution containing
procedure is comprised of three steps. First, CYTOP is spin-coated onto phospholipids, e.g., phosphatidylcholine (PC), phosphatidylethanola-
a glass substrate of a specified thickness. Photolithography is then used mine (PE), and phosphatidylserine (PS), is then infused via the access
to pattern mask the structures onto the photo-resist cover on the entire port. Due to the hydrophobicity of the top orifice, the interface between
surface of the CYTOP layer. The resist-patterned substrate is then dry- water and chloroform is formed over the orifice, where a lipid mono-
etched with O2 plasma using a reactive-ion etching system (RIE) to layer is then deposited. Finally, a second aqueous solution is infused to
expose the glass substrate, resulting in the fabrication of the micro- form another lipid monolayer at the interface with the chloroform so-
chamber. Owing to the treatment with O2 plasma, the bottom and walls lution, which is then deposited over the entire surface of the device.
of the micro-chamber become hydrophilic, while the top orifice remains Thus, bilayers are formed over the micro-chambers, whereas mono-
very hydrophobic. These different surfaces are crucial for forming lipid layers are formed elsewhere. It is worth noting that we form lipid bi-
bilayers in a high throughput manner (described later). layers using chloroform as a solvent for the lipid, whereas decane or
Lipid membranes are formed on the top orifice of the micro- hexadecane have typically been used in conventional approaches [22].

2
R. Watanabe BBA - General Subjects 1864 (2020) 129330

A prominent feature of our approach is that the lipid membranes


spontaneously thin down to form bilayers without mechanical pertur-
bation, e.g., hydraulic pressure, which is required for bilayer formation
using decane or hexadecane. This lack of mechanical perturbation is
critically important, because typically more than half of the lipid
membranes are broken by the mechanical perturbation, resulting in an
efficiency of < 30% for bilayer formation using these approaches [22].
Thinning of the lipid membranes is driven by drainage of the organic
solvent over the hydrophobic support. Because chloroform is more
water soluble than decane or hexadecane, and so can be more easily
dissolved in an aqueous solution, it is more easily drained over a hy-
drophobic support, and thus, our approach does not require any me-
chanical perturbation, which contributes to increasing the efficiency of
bilayer formation to over 97% (Fig. 1d).

2.1.1. Membrane potential


Membrane potential, one of the main driving forces for transport by
membrane proteins, is indispensable in cell physiology. For example,
numerous types of ion channels in neurons are controlled by membrane
potential, allowing for signal transmission via intra- or inter-cellular
interfaces. To mimic physiological conditions, a microsystem with a
nano-sized electrode has been developed to modulate membrane po-
tential in a highly quantitative manner [18] (Fig. 2).
The system is fabricated using conventional vacuum metal deposi-
tion and photolithography. First, 500-nm-thick Au layers and 20-nm-
thick chromium adhesion layers are coated onto the glass substrate
using a vacuum evaporator. Following this, CYTOP is spin-coated onto
the Au layer at a thickness of 500 nm. Photolithography is then con-
ducted to pattern-mask structures onto the photo-resist covering the
entire surface of the CYTOP layer. The resist-patterned substrate is then
dry-etched with O2 plasma using a reactive-ion etching system to ex-
pose the Au surface, and regions bare of Au and Cr are then etched
using Au and Cr etchants, respectively. A schematic illustration of the
system is shown in Fig. 2a. The Au and fluororesin layers are used as a
nano-sized electrode to modulate membrane voltage, and to support the
lipid bilayer membrane, respectively. The through-hole structures on
these layers are used as micro-chambers to detect biological reactions
with high sensitivity.
To evaluate membrane voltage, the micro-chambers are filled with a
buffer solution containing DiBAC4 [24], a fluorescent indicator which
increases its fluorescence intensity in proportion to the amplitude of the
membrane voltage, and the lipid bilayers are then formed over the top
orifices as described above (Fig. 2b). As shown in Fig. 2c, the fluores-
cence intensity of DiBAC4 changes following voltage modulation Fig. 2. Membrane protential.
through the nano-sized electrodes. In addition, the change in intensity (a) Photograph (top) and illustration (bottom) of the fabricated microsystem to
of DiBAC4 is highly stable for > 10 min. Thus, a long-term quantitative modulate membrane potential. Microchambers (ϕ = 3 μm) were fabricated on a
modulation of membrane voltage using a nano-electrode has been de- double layer of fluororesin (h = 500 nm) and Au (h = 500 nm). (b) Membrane
monstrated in this microsystem. potential monitoring. A membrane potential indicator, DiBAC4, was en-
capsulated in the chamber. Membrane potential was modulated using elec-
2.1.2. Asymmetric phospholipid distribution trodes on a chamber and a top glass block. (c) Fluorescence intensity of DiBAC4
Most biological membranes possess an asymmetric transbilayer (green) against applied membrane potential using nano-sized electrodes
(black).
phospholipid distribution, the maintenance of which requires en-
dogenous enzymes to expend energy by promoting phospholipid
translocation [25]. In particular, the plasma membrane of most eu- First, an aqueous solution is infused into the flow channel of the system.
karyotes maintains a high degree of asymmetry. For example, PS and PE After this infusion, the micro-chamber is filled with an aqueous solu-
are strictly confined to the inner leaflet, which controls various cellular tion. Second, a lipid solution is then infused to flush away the first
functions, such as signal transduction, membrane fusion, and cell aqueous solution. The chloroform solution used for this process con-
apoptosis [26–28]. Two different methods have been developed to tains a small amount of lipid, and therefore, lipid monolayers, but not
achieve asymmetric phospholipid distributions in our system in order to multilayers, are formed at the top orifice of chambers. Third, a second
reproduce the physiological environment (Fig. 3) [16,17]. chloroform solution containing a different lipid composition, and a
First, a method was developed whereby two lipid monolayers are second aqueous solution are sequentially infused to form a second lipid
independently prepared, and then assembled on the top orifice of the monolayer at their interface, which is then deposited over the entire
micro-chamber to create an asymmetric distribution (top, Fig. 3a) [17]. surface of the system. In this process, the hydrocarbon tails in the
The experimental procedure used for this is comprised of three steps. second lipid layer are zipped together with those of the first lipid layer

3
R. Watanabe BBA - General Subjects 1864 (2020) 129330

Fig. 3. Assymtery of phospholipid distribution.


(a) Asymmetric lipid bilayer formation using two
compositions of phospholipids. First, an aqueous
solution is injected into the microsystem through an
access port. Second, the first lipid solution con-
taining fluorescent-labeled lipids (fluorescein-DHPE)
is injected. Third, another lipid solution without
fluorescein-DHPE is injected. Finally, a second aqu-
eous solution is injected to flush the second lipid
solution. During this process, asymmetric lipid bi-
layers are formed on the orifices of the micro-
chambers. The bottom panel is the fluorescence
image of fluorescein-DHPE on the asymmetric lipid
bilayers. (b) Asymmetric lipid bilayer formation by
irradiating a high-energy laser. 1) Membrane bi-
layers formed over microchambers were exposed a
laser to 2) photobleach fluorescently labeled phos-
pholipids (red). 3) Fluorescent lipids diffused into
the bilayer from the surrounding area, but were then
restricted to the outer layer. (c) The fluorescence
images of microchambers, TopFluor-TMR-PS (red)
before and after photo-bleaching.

to form an asymmetric bilayer over the micro-chamber, whereas The experimental procedures for this approach are as follows: First,
monolayers are formed elsewhere. Notably, this method allows for the lipid membranes are formed in the microsystem using a fatty-acid la-
formation of > 100,000 asymmetric lipid bilayer membranes with an beled fluorescent phospholipid, e.g., 18:1–6:0 TopFluor-TMR-PC, -PS,
efficiency of over 97% (bottom, Fig. 3a). -PE or 18:1-6:0 NBD-PS (Avanti, USA). The fluorescence of the bilayer
A second method for creating an asymmetric distribution of phos- membrane in the micro-chamber can be clearly seen to be higher than
pholipids involves irradiation with a high-energy laser (Fig. 3b) [16]. that of the surrounding monolayers (Fig. 3b, top Fig. 3c). Following

4
R. Watanabe BBA - General Subjects 1864 (2020) 129330

this, the fluorescent lipids in the bilayer membranes over the micro- 3. Part II. Single molecule analysis of membrane transport
chambers are photobleached by exposure to a high-energy laser for a proteins
few seconds in order to establish an asymmetric lipid distribution As
shown in the bottom panel of Fig. 3c, the fluorescence in the micro- Over the past few years, single molecule analysis has been demon-
chambers gradually recovers over a period of several hundred seconds strated for a variety of membrane transport proteins, e.g., pores,
to the same level as seen on the surrounding surfaces, confirming that channels, ion-pumps, and phospholipid scramblases, using “biomem-
the membrane over the micro-chambers is contiguous with the mono- brane microsystems” [5,16–21]. In this section, I would like to in-
layer in other parts of the system. The recovery of the fluorescence over troduce the achievements made for the single molecule analysis of
the micro-chambers do not exceed the fluorescence of surrounding transport proteins.
surfaces (Fig. 3b), indicating that the vertical translocation of phos-
pholipids between the two layers of the membrane does not occur
3.1. Pore: α-hemolysin
during measurement or is very slow, as has been observed in plasma
membranes [29]. In other words, this result shows that photobleaching
A transmembrane pore is a membrane protein that passively
generates an asymmetric membrane bilayer in which unbleached
transport substrates along their concentration gradient. The activities of
fluorescent-phospholipids are present only in the upper layer. Notably,
pores are relatively higher than other transport proteins, e.g., ion-
the asymmetry is maintained over 2 h after photo-bleaching, and is
pumps, because both sides of the permeation pathway are always open,
moreover reproduced at the same bilayers for many times by laser ir-
and as a result, they have been well characterized mechanistically using
radiations. These technical advantages are useful for single molecule
patch clamp recordings.
analysis of phospholipid transport proteins, such as flippase, floppase,
A single molecule transport assay was carried out on our micro-
and scramblase (see below).
system using α-hemolysin (Fig. 5) [30], a toxic membrane protein that
binds to lipid bilayer membranes and forms transmembrane nanopores
2.1.3. Generation of concentration gradient
(ϕ = 1–2 nm) that have been used as various biomedical sensors, e.g.,
Owing to the small volume of the chambers in our system, a minute
in nanopore DNA sequencing [31]. While monomers of α-hemolysin are
quantity of product is detectable, allowing for the direct monitoring of
soluble in water, once bound to a lipid bilayer, α-hemolysin assembles
the individual activities of isolated single membrane proteins in a
into a heptameric ring and forms a nanopore at the center, resulting in
highly sensitive and quantitative manner. However, even though the
the passive transport of small molecules as small as 1 nm through its
chambers are highly integrated in the system, it is difficult to conduct
pore.
multiple bioassays under different conditions using integrated cham-
For the assay, a fluorescent dye, Alexa 488, was encapsulated in the
bers in parallel, because the composition of the reaction solution en-
chambers as the transport substrate, and a low concentration of α-he-
capsulated in the chambers is uniform over the entire area of the
molysin solution (i.e., 1 μg mL−1) was injected into the flow channel
system. Recently, this technical issue was addressed by developing a
after lipid bilayer formation (Fig. 5a). In this setup, the passive trans-
novel system capable of forming a variety of concentration gradients of
port activity of α-hemolysin was detectable by analyzing the fluor-
target molecules encapsulated in the chambers (Fig. 4) [21].
escent intensity of the chamber, because Alexa 488 passively diffuses
In this system, based on the advection diffusion model, a con-
out of the chamber through the α-hemolysin pores, resulting in a de-
centration gradient of target molecules is formed along the flow
crease in chamber fluorescent intensity. Notably, a simple physico-
channel via the sequential injection of several liquids from the access
chemical model has been established for the passive transport by α-
port (Fig. 4a). First, an aqueous solution containing the target mole-
hemolysin pores in the microsystem, where diffusion of the dye mole-
cules is infused into the flow channel to fill the individual chambers, as
cule into the α-hemolysin pore is the kinetic bottleneck, and obeys
described above. Second, a specified amount of a second aqueous so-
Fick's law [5,19]. In this model, the fluorescence intensity, F(t), of Alexa
lution lacking target molecules (7–12 μL) is infused at a defined flow
488 encapsulated in the chamber is expressed by Eq. (1):
rate (0.5–1.5 μL/s) using an electric pipette. In this step, the con-
centration gradients are generated based on the advection-diffusion N ∙D∙d 2∙π ⎞
process. The aqueous solution first filled in the flow channel is gradu- F (t ) = Fo ∙exp ⎛−
⎜ ⎟t + F1
⎝ 4∙L∙V ⎠ (1)
ally diluted from the inlet as the second solution for dilution is infused.
Then, the concentration gradients are generated along the flow direc- where N is the number of α-hemolysin pores reconstituted in the lipid
tion, i.e., the first solution is not completely diluted because the volume membrane, D is the diffusion coefficient of Alexa 488, d is the diameter
of second solution is smaller than that of the flow channel. Finally, to of the α-haemolysin pore (~1 nm), L is the length of the α-hemolysin
encapsulate the target molecules into the chambers, a chloroform so- pore (~10 nm), and V is the volume of the chamber. According to Eq.
lution containing phospholipids and a third aqueous solution are suc- (1), the fluorescence intensity of the chamber exponentially decreases
cessively infused. After infusion, the lipid-bilayer membrane is formed due to the passive transport activity of α-hemolysin with a rate constant
on the orifice of the individual chambers, as described above, resulting proportional to the number of α-hemolysin pores (N).
in the encapsulation of the target molecules. Fig. 5b, c show typical time courses of the fluorescent signals of
To examine the feasibility of this approach, a fluorescent dye (Alexa 1 μM Alexa 488 encapsulated in 7 fL chambers, where the decay of
488) was used as an indicator of the concentration gradient. The con- fluorescent intensity represents passive transport through the α-hemo-
centration gradient was formed using the first or second aqueous so- lysin pore. The response of each individual chamber is not homo-
lutions, with or without 1 μM Alexa 488, respectively. As expected, the geneous (Fig. 5b), representing the stochastic formation of α-hemolysin
fluorescence intensity of Alexa 488 encapsulated in the chambers in- pores in lipid bilayers. The distribution of the rate constant for fluor-
creased along the flow channel (Fig. 4b, c), confirming the formation of escent decay, determined by fitting with an exponential function (black
a concentration gradient of target molecules in our system. Further- line in Fig. 5c), exhibits four distinctive peaks (Fig. 5d). The intervals
more, the concentration gradient of the target molecule, i.e., Alexa 488, between the peaks are essentially constant, a typical feature of a single-
became steeper as the flow rate increased, and/or the volume of the molecule digital assay [32,33], and each peak corresponds to the ac-
second aqueous solution decreased (Fig. 4b). Accordingly, the con- tivity of 0, 1, 2, or 3 α-hemolysin pores.
centration gradient in the micro-chamber array can be quantitatively The rate constant of the passive diffusion through a single α-he-
modulated in proportion to the distance from the access port, L (Fig. 4b, molysin pore was determined from the peak interval to be
c). So far, using this system, multiple single-molecule bioassays have 5.5 ± 1.5 × 10−4 s−1. The initial transport flux of a single α-hemo-
been performed on the system under various conditions. lysin pore (v) was thus estimated as v = n · k, where n is the number of

5
R. Watanabe BBA - General Subjects 1864 (2020) 129330

Fig. 4. Concentration gradient.


(a) Schematic illustration of the concentration gradient of target molecules. The concentration gradients were formed on individual micro-chambers depending on
the distance (L) from the access port for sample injection. The individual micro-chambers were sealed with lipid-bilayer membranes. (b) Concentrations of Alexa 488
in micro-chambers are plotted against the flow rate (top) or volume of the second buffer (bottom) used to generate the concentration gradients. The solid lines
represent linear regressions. (c) Fluorescence image of Alexa 488 (green) encapsulated in micro-chambers (h = 0.5 μm) at each L. The flow rate of liquid was 1.5 μL/s
and the volume of the second buffer solution was 7 μL.

fluorescent dye molecules encapsulated in each chamber, and extensive biochemical studies have been conducted for ion-pumps, a
n = 7 fL × 1 μM × NA = 4.2 × 103 molecules. The initial transport single-molecule analysis has not achieved due to the technical diffi-
flux per α-hemolysin pore was 2.3 molecules s−1, which is much lower culties in detecting slow transport events in highly sensitive manner.
than the detection limit of conventional patch clump recordings (~107 For the assay, RhP-M, a fluorescent pH indicator which increases its
ions s−1). Thus, through the use of the biomembrane microsystem the fluorescent intensity under acidic conditions [35], is encapsulated in
highly sensitive detection of passive transport by a transmembrane pore the chamber, and a buffer solution containing a low concentration of
can be achieved down to the single molecule level. FoF1 is injected into the flow channel after lipid bilayer formation.
Following this, ATP is injected to initiate proton pumping. In this setup,
3.1.1. Ion pump: F-type ATPase FoF1 molecules with an outward orientation of the catalytic core do-
An ion pump is a membrane protein that actively transports ions by main can hydrolyze ATP to pump protons into the chamber, decreasing
consuming electro-chemical energy, e.g., ATP hydrolysis or ion-motive the pH, and thereby increasing the fluorescent intensity of RhP-M
force (imf). The activities of ion pumps are relatively low (> 100 mo- (Fig. 6a). Fig. 6b, c show typical time courses of fluorescent images
lecules s−1) because one side of the ion permeation pathway is always obtained after ATP injection. Notably, some RhP-M molecules bound
closed, i.e., they have to induce conformational changes to open and nonspecifically to the chamber wall due to its hydrophobicity, ex-
close the outward or inward gate on the pathway, and therefore, it is hibiting ring-like shapes in fluorescent images (Fig. 6b). Similarly to the
technically challenging to measure the low activity of ion pumps at the fluorescence signals obtained from the single-molecule α-hemolysin
single molecule level. assay, the fluorescent signals in this assay show heterogeneity. As ex-
A single molecule analysis of an ion pump was carried out on our pected, Fig. 6b (left) exhibited discrete signal levels, representing the
microsystem using the F-type ATPase (FoF1) (Fig. 6a), which mediates stochastic reconstitution of the FoF1 molecules. The active chambers
proton transport by coupling with ATP hydrolysis [34]. Although reached a plateau at approximately 1600, which corresponds to a pH of

6
R. Watanabe BBA - General Subjects 1864 (2020) 129330

Fig. 5. Single molecule analysis of pore, α-hemolysin.


(a) Schematic illustration of passive transport of α-hemolysin using the biomembrane microsytem. The fluorescent dye (Alexa 488) diffused out from the chamber via
the pore on the membrane formed by α-hemolysin. (b) Fluorescent images of the passive transport activity in the chambers. The images were recorded just after α-
hemolysin injection (left panel) and 3000 s later (middle panel). The right panel, diff., shows the intensity difference between these two images in the form of a color
gradient. (c) Continuous recording of the passive transport activity of 1 μg mL−1 α-hemolysin. The chambers containing 0 α-hemolysin pores, 1 α-hemolysin pore, 2
α-hemolysin pores, and 3 α-hemolysin pores were plotted as gray, blue, red, and green, respectively. Solid lines represent the fittings with the single exponential
decay: y = C1 · exp.(−k · t) + C2, where k is the rate constant of passive transport. (d) In the histogram, the number of chambers is plotted versus the rate constant of
passive transport, k. The four peaks can be attributed to occupancies of 0, 1, 2, or 3 α-hemolysin pores per chamber. They were fitted to a sum of Gaussians.

approximately 5.4. The transmembrane proton gradient (ΔpH) gener- 3.1.2. Phospholipid scramblase: TMEM16F
ated was approximately 1.7, which is at a level of equilibrium with the Phospholipid scramblase is a membrane protein which translocates
free energy of ATP hydrolysis. At the end of the observation period, a phospholipids between the inner and outer leaflets of cell membranes in
H+ ionophore, nigericin (arrow in Fig. 6c), was injected into the flow order to disrupt their asymmetric lipid-distribution [36]. The bio-
channel. The active chambers uniformly recovered their fluorescence chemical features of scramblase have not been well characterized be-
toward the original level. This nigericin-induced fluorescence recovery cause the purification of intact scramblase is difficult, and moreover,
ensures that the fluorescence increase represents the ATP-driven with a few exceptions, membrane bilayers containing asymmetrically
proton-pumping activity of FoF1. distributed phospholipids are challenging to construct [17,37,38].
To estimate the proton transport rate for FoF1 during the initial A single molecule analysis of phospholipid scramblase is carried out
phase, the initial rate is measured from the linear portion of the time- using our microsystem [16] and the transmembrane protein 16F
course (typically from 1500 s to 4000 s) where the fluorescence in- (TMEM16F), a Ca2+-dependent phospholipid scramblase [39] that
creases at a constant rate (Fig. 6c). Similarly to the α-hemolysin assay, mediates PS exposure in activated platelets during blood clotting, and
the distribution of the proton-pumping rate exhibites three distinct, regulates hydroxyapatite release from osteoblasts during bone miner-
regularly spaced peaks, indicating that each peak represented chambers alization (Fig. 7a) [40,41]. Despite its physiological importance, a
with 0, 1, or 2 molecules of FoF1 (Fig. 6d). The rate of proton pumping single-molecule analysis has not been achieved so far due to the tech-
by FoF1 is estimated from the intervals between the peaks to be nical difficulties in preparing asymmetric lipid bilayers and in detecting
27.5 s−1. Thus, the biomembrane microsystem is the first to achieve the lipid translocation in a highly sensitive manner. For the assay, a buffer
highly sensitive detection of active transport by an ion pump at the solution containing a low concentration of TMEM16F is injected into
single molecule level. the flow channel after lipid bilayer formation. Following this, an
asymmetric distribution of the fluorescent lipid is formed on membrane

7
R. Watanabe BBA - General Subjects 1864 (2020) 129330

Fig. 6. Single molecule analysis of ion-pump, F-type ATPase.


(a) Schematic illustration of F-type ATPase (FoF1) active transport. FoF1 pumps the proton from outside to inside the microchambers by hydrolyzing ATP. The
fluorescent pH indicator (RhP-M) is encapsulated in the microchamber for pH monitoring. (b) Fluorescence images of the proton pumping of FoF1 in the chambers.
The images were recorded just after the injection of 200 μM ATP (left panel) and 6000 s later (middle panel). The right panel, diff., shows the intensity difference
between the left and middle panels in the form of a color gradient. (c) Continuous recording of proton pumping. The chambers containing 0 molecules, 1 molecule,
and 2 molecules of active FoF1 were plotted as gray, blue, and red, respectively. (d) In the histogram, the number of chambers is plotted versus the slope of
fluorescence increments from 1500 to 4000 s. The three peaks can be attributed to occupancies of 0, 1, or 2 active FoF1 molecules per chamber. They were fitted with
a sum of Gaussians.

bilayers by irradiation with a high-energy laser, as described above. The fluorescence up to their original level in the presence of 100 μM Ca2+,
scrambling assay is initiated by injecting 100 μM Ca2+ into the channel but not in the presence of 100 μM Ca2+ plus 1.0 μM of the inhibitor
to activate the phospholipid scrambling activity of TMEM16F. Using EGCg [42] (Fig. 7c), demonstrating that phospholipid scrambling by
this setup, the phospholipid translocation of TMEM16F is detected by TMEM16F can be triggered more than once in our microsystem.
analyzing the fluorescent intensity of the bilayers, because TMEM16F A kinetic analysis is carried out by conducting the scrambling assay
can transports fluorescent lipids from the outer leaflet to the inner under various experimental conditions, e.g., different membrane sizes
leaflet of bilayers, resulting in an increase in their fluorescent intensity. or at different temperatures. The scrambling catalyzed by TMEM16F
Fig. 7b, c display typical time courses of fluorescent images obtained slowes down (Fig. 7d) as the area of the bilayer membrane increased.
after Ca2+ injection, where the fluorescent increase in the bilayers re- Because TMEM16F-mediated scrambling is reversible [39], the reaction
presents the transport of a fluorescent lipid by TMEM16F. The response should follow the reaction
of each individual bilayer is not homogeneous (Fig. 7b), showing the k
stochastic reconstitution of TMEM16F molecules. [Lipidout ] ⇄ [Lipidin]
k (3)
The fluorescent intensity of the bilayers gradually increases to reach
the original intensity level before photobleaching (Fig. 7c), irrespective where k is the rate constant, and [Lipidout] and [Lipidin] are the con-
of the phospholipid composition of the membranes. Notably, when centrations of fluorescent lipids at the outer and inner layers, respec-
active bilayers are laser-photobleached for a second time to re-create an tively. Notably, [Lipidout] is considered to be constant because it is
asymmetric distribution, the bilayers gradually regain their contiguous with a large lipid monolayer in the surrounding areas. Since

8
R. Watanabe BBA - General Subjects 1864 (2020) 129330

(caption on next page)

9
R. Watanabe BBA - General Subjects 1864 (2020) 129330

Fig. 7. Single molecule analysis of phospholipid scramblase, TMEM16F.


(a) Schematic illustration of the phospholipid scrambling by TMEM16F. 1) TMEM16F was incorporated into the asymmetric membrane bilayers with fluorescent
phospholipids (red). 2) Addition of CaCl2 initiated phospholipid scrambling, followed by lateral diffusion of unbleached phospholipids from surrounding areas,
increasing the fluorescence intensity in microchambers. (b) Purified TMEM16F was loaded at 3.4 μg mL−1, and scrambling was initiated with 100 μM CaCl2.
Subsequently, the asymmetric membrane bilayers were again formed by laser irradiation with injection of 1 μM EGCg. Maximum intensity projections of confocal
images of TopFluor-TMR-PS (red) obtained from the same microchambers before (top) and after addition of 100 μM CaCl2 (middle) or 100 μM CaCl2 + 1 μM EGCg
(bottom). Scale bar: 5 μm. (c) Time course of TMEM16F-mediated phospholipid scrambling. Lipid bilayers with (red) or without (gray) TMEM16F were treated with
100 μM CaCl2 (black arrowheads), and fluorescence from TopFluor-TMR-PS was followed over time. Subsequently, the microchambers were re-exposed for a few
seconds to a laser at time points marked in orange, with (lower panel) or without (upper panel) prior injection of 1 μM EGCg (red arrowhead). (d) Time course of
TMEM16F-mediated phospholipid scrambling at different area (S) of the bilayer membrane, 8.6 or 71 μm2. Black curves are fits to the single-exponential function
y = C1 + C2 × (1 – exp.[−k × (t – 1000)]). (e) The rate constant of fluorescence increase (k*) of TopFluor-TMR-PS was plotted against the inverse of membrane area
(1/S). Data in blue, red, and orange were obtained using TopFluor-TMR-PS at 16 °C, 25 °C, and 31 °C, respectively. Black lines are linear regressions. (f) An Arrhenius
plot for TMEM16F-mediated phospholipid scrambling, with k values from Fig. 7e. The solid line represents linear regression, and the inset contains thermodynamic
parameters at 25 °C, as determined from the plot.

the fluorescent intensity over the chamber (FI) is proportional to the size of micro-chambers (ɸ < 100 μm) drastically affect the efficiency of
concentration of fluorescent lipids and the membrane area, FI is thus bilayer formation. In addition, the experimental conditions for recon-
determined as a function of time t; stituting membrane proteins into lipid bilayers are widely varied (not
universal) among protein species and ways of purification. These points
FI = FIout + FIin⋅[1 − exp(−k ∗⋅t )] (4)
represent remaining technical challenges for single molecule analysis
using the biomembrane microsystems.
k 1
k∗ = ⋅ Single-molecule analysis using micro-chamber arrays is an emerging
η S (5)
approach to detect various bio-reactions, e.g., hydrolysis, protein
where FIout and FIin are the fluorescent intensities of the outer and inner synthesis, and membrane transport, with a high sensitivity down to the
layers, respectively, S is the area of the membrane over the chambers, single-molecule level, enabling biomedical applications, such as digital
and η is the number of lipid molecules (~2.0 × 106/μm2) per unit area PCR and digital ELISA. The biomembrane microsystem described here
of the monolayer, as determined previously [43]. Fitting the time has broadened the versatility of micro-chambers to a single molecule
course in Fig. 7d to Eq. (4) demonstrates k*, the rate constant of analysis for membrane proteins, which are promising drug targets for
fluorescent increase at 25 °C. As expected from Eq. (5), k* is inversely the treatment of various diseases. Thus, this microsystem will enable
proportional to S (Fig. 7e), demonstrating that k, the rate of TMEM16F- further analytical and pharmacological applications, such as use in high
mediated lipid transport, is 4.5 × 104 lipids/s at 25 °C (Fig. 7f); this throughput drug screening, and as an early diagnosis tool.
value is ~108-fold faster than the spontaneous vertical lipid flip-flop
across membrane bilayers [29]. Importantly, the k value does not differ Acknowledgements
between fluorescence substrates, such as TopFluor-TMR-PS, -PC, and
-PE, indicating that TMEM16F does not distinguish between the head This work was supported by a Grant-in-Aid for Scientific Research
group moieties of phospholipids, and confirming that TMEM16F non- (JP17H03660, JP15H05591) from the Japan Society for the Promotion
specifically scrambles phospholipids. of Science, a PRESTO Grant (JPMJPR13LC) from the Japan Science and
The TMEM16F-mediated scrambling is temperature dependent Technology Agency, and a PRIME Grant (JP17gm0910020) from the
(Fig. 7e), with k values of 1.4 × 104 and 7.1 × 104 lipids/s at 16 °C and Japan Agency for Medical Research and Development (to R.W.).
35 °C, respectively. The corresponding Arrhenius plot fitted well to a
linear function (Fig. 7f), indicating that TMEM16F-mediated lipid References
transport is governed by a single rate-limiting step, at least from 16 °C
to 31 °C. The thermodynamic parameters of this rate-limiting step are as [1] M.H. Saier Jr., A functional-phylogenetic classification system for transmembrane
solute transporters, Microbiol. Mol. Biol. Rev. 64 (2000) 354–411.
follows: ΔG‡ = 47 kJ/mol, ΔH‡ = 82 kJ/mol, and TΔS‡ = 35 kJ/mol at [2] C. International Transporter, et al., Membrane transporters in drug development,
25 °C (inset, Fig. 7f). The ΔG‡ value for spontaneous vertical lipid flip- Nat. Rev. Drug Disc. 9 (2010) 215–236.
flop is approximately 100 kJ/mol [29], showing that TMEM16F sig- [3] K. Sugano, et al., Coexistence of passive and carrier-mediated processes in drug
transport, Nat. Rev. Drug Disc. 9 (2010) 597–614.
nificantly reduces the activation free energy. Thus, this biomembrane
[4] S. Veshaguri, et al., Direct observation of proton pumping by a eukaryotic P-type
microsystem is the first to achieve a highly sensitive detection of ATPase, Science 351 (2016) 1469–1473.
phospholipid transport by phospholipid scramblase at the single mo- [5] R. Watanabe, et al., Arrayed lipid bilayer chambers allow single-molecule analysis
lecule level. of membrane transporter activity, Nat. Commun. 5 (2014) 4519.
[6] A.J. Garcia-Saez, P. Schwille, Single molecule techniques for the study of membrane
proteins, App. Microbiol. Biotechnol. 76 (2007) 257–266.
[7] S.B. Kodandaramaiah, et al., Automated whole-cell patch-clamp electrophysiology
3.1.3. Future prospects
of neurons in vivo, Nat. Methods 9 (2012) 585–587.
A high-throughput single-molecule analysis of the membrane [8] J. Dunlop, et al., High-throughput electrophysiology: an emerging paradigm for ion-
transport proteins, α-hemolysin, F-type ATPase, and TMEM16F has channel screening and physiology, Nat. Rev. Drug Discov. 7 (2008) 358–368.
been achieved using the “biomembrane microsystem”. The micro- [9] O. Keminer, et al., Optical recording of signal-mediated protein transport through
single nuclear pore complexes, Proc. Natl. Acad. Sci. U. S. A. 96 (1999)
system has the following advantages for single molecule analysis: i) a 11842–11847.
high sensitivity using ultra-small chambers, ii) high throughput using [10] H. Bayley, P.S. Cremer, Stochastic sensors inspired by biology, Nature 413 (2001)
chambers in parallel, and iii) a high compatibility for membrane pro- 226–230.
[11] K. Sumitomo, et al., Ca2+ ion transport through channels formed by alpha-hemo-
tein analysis. These features may extend the use of our microsystem for lysin analyzed using a microwell array on a Si substrate, Biosens. Bioelectron. 31
single molecule analysis toward other membrane transport proteins, (2012) 445–450.
e.g., ion-channels, secondary active transporters, and phospholipid [12] S. Ota, et al., Microfluidic lipid membrane formation on microchamber arrays, Lab
Chip 11 (2011) 2485–2487.
flippases/floppases. Such studies will reveal the generality and un- [13] T. Tonooka, et al., Lipid bilayers on a picoliter microdroplet array for rapid fluor-
iqueness of the operating principles of membrane transport proteins, escence detection of membrane transport, Small 10 (2014) 3275–3282.
thereby promoting in vitro membrane protein studies. [14] M. Urban, et al., Highly parallel transport recordings on a membrane-on-nanopore
chip at single molecule resolution, Nano Lett. 14 (2014) 1674–1680.
The currently developed microsystem has several limitations. First, [15] A. Kleefen, et al., Multiplexed parallel single transport recordings on nanopore
the concentration of bivalent cations (< 10 mM Ca2+ or Mg2+) and the

10
R. Watanabe BBA - General Subjects 1864 (2020) 129330

arrays, Nano Lett. 10 (2010) 5080–5087. [29] T.C. Anglin, J.C. Conboy, Kinetics and thermodynamics of flip-flop in binary
[16] R. Watanabe, et al., Single-molecule analysis of phospholipid scrambling by phospholipid membranes measured by sum-frequency vibrational spectroscopy,
TMEM16F, Proc. Natl. Acad. Sci. U. S. A. 115 (2018) 3066–3071. Biochemistry 48 (2009) 10220–10234.
[17] R. Watanabe, et al., High-throughput formation of lipid bilayer membrane arrays [30] L. Song, et al., Structure of staphylococcal alpha-hemolysin, a heptameric trans-
with an asymmetric lipid composition, Sci. Rep. 4 (2014) 7076. membrane pore, Science 274 (1996) 1859–1866.
[18] R. Watanabe, et al., Novel nano-device to measure voltage-driven membrane [31] D. Branton, et al., The potential and challenges of nanopore sequencing, Nat.
transporter activity, IEEE Trans. Nanotechnol. 15 (2016) 70–73. Biotechnol. 26 (2008) 1146–1153.
[19] N. Soga, et al., Attolitre-sized lipid bilayer chamber array for rapid detection of [32] Y. Zhang, H. Noji, Digital bioassays: theory, applications, and perspectives, Anal.
single transporters, Sci. Rep. 5 (2015) 11025. Chem. 89 (2017) 92–101.
[20] R. Watanabe, et al., Arrayed water-in-oil droplet bilayers for membrane transport [33] D. Kim, et al., Research highlights: digital assays on chip, Lab Chip 15 (2015)
analysis, Lab Chip 16 (2016) 3043–3048. 17–22.
[21] R. Watanabe, et al., High-throughput single-molecule bioassay using micro- [34] W. Junge, et al., Torque generation and elastic power transmission in the rotary
chamber arrays with a concentration gradient of target molecules, Lab Chip 18 F0F1-ATPase, Nature 459 (2009) 364–370.
(2018) 2849–2853. [35] D. Asanuma, et al., Acidic-pH-activatable fluorescence probes for visualizing exo-
[22] M. Zagnoni, Miniaturised technologies for the development of artificial lipid bilayer cytosis dynamics, Angew. Chem. Int. Ed. Engl. 53 (2014) 6085–6089.
systems, Lab Chip 12 (2012) 1026–1039. [36] K. Segawa, S. Nagata, An apoptotic 'eat me' signal: phosphatidylserine exposure,
[23] R. Iino, et al., A single-cell drug efflux assay in bacteria by using a directly acces- Trends Cell Biol. 25 (2015) 649–650.
sible femtoliter droplet array, Lab Chip 12 (2012) 3923–3929. [37] S. Pautot, et al., Engineering asymmetric vesicles, Proc. Natl. Acad. Sci. U. S. A. 100
[24] D.E. Epps, et al., Characterization of the steady-state and dynamic fluorescence (2003) 10718–10721.
properties of the potential-sensitive dye bis-(1,3-dibutylbarbituric acid)trimethine [38] M. Iwamoto, S. Oiki, Contact bubble bilayers with flush drainage, Sci. Rep. 5 (2015)
oxonol (Dibac4(3)) in model systems and cells, Chem. Phys. Lipids 69 (1994) 9110.
137–150. [39] J. Suzuki, et al., Calcium-dependent phospholipid scrambling by TMEM16F, Nature
[25] G. van Meer, et al., Membrane lipids: where they are and how they behave, Nat. 468 (2010) 834–838.
Rev. Mol. Cell. Biol. 9 (2008) 112–124. [40] N. Pedemonte, L.J.V. Galietta, Structure and function of TMEM16 proteins (anoc-
[26] J.M. Boon, B.D. Smith, Chemical control of phospholipid distribution across bilayer tamins), Physiol. Rev. 94 (2014) 419–459.
membranes, Med. Res. Rev. 22 (2002) 251–281. [41] S. Nagata, et al., Exposure of phosphatidylserine on the cell surface, Cell Death
[27] D.L. Bratton, et al., Appearance of phosphatidylserine on apoptotic cells requires Differ. 23 (2016) 952–961.
calcium-mediated nonspecific flip-flop and is enhanced by loss of the aminopho- [42] T. Suzuki, et al., Functional swapping between transmembrane proteins TMEM16A
spholipid translocase, J. Biol. Chem. 272 (1997) 26159–26165. and TMEM16F, J. Biol. Chem. 289 (2014) 7438–7447.
[28] V.A. Fadok, et al., Loss of phospholipid asymmetry and surface exposure of phos- [43] J.F. Nagle, S. Tristram-Nagle, Structure of lipid bilayers, Biochim. Biophys. Acta
phatidylserine is required for phagocytosis of apoptotic cells by macrophages and 1469 (2000) 159–195.
fibroblasts, J. Biol. Chem. 276 (2001) 1071–1077.

11

You might also like