You are on page 1of 10

Available online at www.sciencedirect.

com

Biochimica et Biophysica Acta 1778 (2008) 1601 – 1610


www.elsevier.com/locate/bbamem

Review
Formation of irreversibly bound annexin A1 protein domains on
POPC/POPS solid supported membranes
Simon Faiß a , Katja Kastl b,1 , Andreas Janshoff a , Claudia Steinem b,⁎
a
Institute of Physical Chemistry, University of Mainz, Welder-Weg 11, 55128 Mainz, Germany
b
Institute of Organic and Biomolecular Chemistry, University of Göttingen, Tammannstr. 2, 37077 Göttingen, Germany
Received 10 October 2007; received in revised form 9 December 2007; accepted 4 January 2008
Available online 12 January 2008

Abstract

The specific interaction of annexin A1 with phospholipid bilayers is scrutinized by means of scanning force and fluorescence microscopy,
quartz crystal microbalance, ellipsometry, and modeled by dynamic Monte Carlo simulations. It was found that POPC/POPS bilayers exhibit
phase separation in POPC- and POPS-enriched domains as a function of Ca2+ concentration. Annexin A1 interacts with POPC/POPS bilayers by
forming irreversibly bound protein domains with monolayer thickness on POPS-enriched nanodomains, while the attachment of proteins to the
POPC-enriched regions is fully reversible. A thorough kinetic analysis of the process reveals that both, the binding constant of annexin A1 at the
POPC-rich areas as well as the irreversible adsorption rate to the POPS-rich domains increases with calcium ion concentration. Based on the
thermodynamic and kinetic data, a possible mechanism of the annexin A1 membrane interaction can be proposed.
© 2008 Elsevier B.V. All rights reserved.

Keywords: Ellipsometry; Fluorescence microscopy; Micropatterned membranes; Monte Carlo simulation; QCM; SFM

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1602
2. Binding of annexin A1 to solid supported lipid bilayers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1602
2.1. Visualization of annexin A1 binding by fluorescence microscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1602
2.2. Thickness determination of annexin A1 by means of ellipsometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1603
2.3. Scanning force microscopy of annexin A1 bound to POPC/POPS membranes . . . . . . . . . . . . . . . . . . . . . . . 1604
3. Adsorption and desorption kinetics of annexin A1 at different Ca2+ concentrations . . . . . . . . . . . . . . . . . . . . . . . . . 1605
3.1. Theoretical modeling of the adsorption/desorption data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1606
3.2. Results of DMC simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1608
3.2.1. Impact of the calcium ion concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1608
3.2.2. Influence of incubation time on the parameter space. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1608
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1609
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1609

Abbreviations: ASF, available surface function; β-BODIPY 500/510 C12-HPC, 2-(4,4-Difluoro-5-methyl-4-bora-3a,4a-diaza-s-indacene-3-dodecanoyl)-1-


hexadecanoyl-sn-glycero-3-phosphocholine; DMC, Dynamic Monte Carlo simulation; FRET, Fluorescence recovery after photobleaching; POPC, 1-Palmitoyl-2-
oleoyl-sn-glycero-3-phosphocholine; POPS, 1-Palmitoyl-2-oleoyl-sn-glycero-3-phosphoserine; QCM, Quartz crystal microbalance; RSA, Random sequential
adsorption; SFM, Scanning force microscopy
⁎ Corresponding author. Tel.: +49 551 393294; fax: +49 551 393228.
E-mail address: Claudia.steinem@chemie.uni-goettingen.de (C. Steinem).
1
Current address: Toshiba Research Europe Ltd., Cambridge Research Laboratory, 260 Cambridge Science Park, Milton Road, Cambridge CB4 OWE, UK.

0005-2736/$ - see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.bbamem.2008.01.003
1602 S. Faiß et al. / Biochimica et Biophysica Acta 1778 (2008) 1601–1610

1. Introduction the concave side and is buried within the core in the absence of
Ca2+. The Ca2+ affinities are different for the various binding
Annexins comprise a family of membrane proteins that bind sites and increase by two orders of magnitude in the presence of
to negatively charged phospholipids in the presence of Ca2+ and negatively charged phospholipids. Based on the high-resolution
which are found in almost all organisms ranging from molds structural studies of annexin A1, Rosengarth and Luecke have
to mammals [1–3]. In 1990, a common nomenclature was proposed a two-step model for the annexin A1-membrane
proposed by Crumpton and Dedman based on the Greek term interaction leading to membrane aggregation. Starting with the
“annex” (= to connect), which refers to the unifying property protein in its inactive form, the first step would be the calcium
of annexins to be able to connect opposing membranes [4]. ion-mediated binding to anionic phospholipid headgroups of
Recently, this nomenclature has been fine tuned in such that one membrane. This process involves a change in conforma-
annexin genes expressed in vertebrates are termed “A” tion of the C-terminal core that results in the previously buried
annexins, those in invertebrates “B” annexins, while “C” an- N-terminal domain becoming solvent accessible. The second
nexins are found in fungi and molds, “D” annexins in plants step would be the binding of a second membrane via hydro-
and “E” annexins in protists. Mammals have 12 different an- phobic interactions with the now exposed amphipathic N-ter-
nexin genes (A1–A11 and A13) encoding for proteins with minal domain [12].
molecular masses of typically 30–40 kDa. All proteins are We were able to investigate both membrane binding pro-
structurally characterized by a highly conserved C-terminal cesses of annexin A1 separately by exploiting planar membra-
domain (core domain) that harbors the Ca2+- and membrane nes attached to a solid support. In a first step, the adsorption of
binding sites. Sequence analysis reveals that the core domain annexin A1 is monitored [14–17], while after removal of non-
is composed of four or eight (in case of annexin A6) annexin bound protein from solution, the second membrane binding
repeats, which are highly homologous among each other and can be followed by addition of unilamellar vesicles [18]. In this
among the annexin family. Within each annexin repeat a highly mini-review, we discuss the results obtained from quartz cry-
conserved sequence of 17 amino acids is found, which is called stal microbalance (QCM), ellipsometry, and scanning force
endonexin fold. In contrast to the core domain, the N-terminal microscopy measurements in conjunction with dynamic Monte
domains are rather variable. Their sizes range from a few amino Carlo simulations to elucidate the binding properties of annexin
acids (annexin A5) to more than hundred amino acids in case A1 and its ability to form protein domains as a function of
of annexin A7 and A11 and are thought to be responsible for phosphatidylserine and Ca2+ concentration.
the different functions of the annexins within cells.
Assessing the function of an annexin is still a difficult task, 2. Binding of annexin A1 to solid supported lipid bilayers
as some cell types express more than one annexin. This results
in a complex and overlapping expression pattern of a structu- The functionality of annexins, that is, the ability to bind to
rally conserved protein family, which renders the elucidation acidic phospholipids in the presence of calcium ions, is often
of their individual cellular functions very cumbersome. investigated in vitro by vesicle copelletation assays [19,20].
In this mini-review, we focus on annexin A1, which has In this method, however, annexin binding and vesicle aggre-
been identified as the cytosolic factor that aggregates isolated gation are intimately connected. In order to exclusively inves-
neutrophil specific granules and promotes exocytotic fusion tigate annexin binding to one membrane interface, planar
with the plasma membrane in a Ca2+ dependent fashion [5,6]. solid supported lipid bilayers are advantageous. These are
Since it has been localized at early endosomes, annexin A1 also available on various surfaces and can be investigated by fluo-
appears to be involved in endocytosis [7,8]. Moreover, it has rescence microscopy, scanning force microscopy, and quartz
been shown that annexin A1 is expressed at higher levels in cells crystal microbalance.
of the innate immune system such as neutrophils, monocytes/
macrophages, and mast cells than in cells of the adaptive im- 2.1. Visualization of annexin A1 binding by fluorescence
mune response such as T and B lymphocytes [9]. For example, microscopy
blood neutrophils and monocytes have abundant levels of an-
nexin A1, representing about 3–4% of total cytosolic proteins. By means of fluorescence microscopy, it is possible to in-
Results of the last 20 years have demonstrated that annexin vestigate the specificity of annexin A1 for acidic phospholipids.
A1 acts as glucocorticoid-regulated endogenous down-regu- Porcine annexin A1 was recombinantly expressed in Escheri-
lator of inflammation: if it is externalized on the cell surface chia coli and purified according to a procedure described by
of adherent neutrophils, it blocks their interaction with endo- Rosengarth et al. [21]. To allow for a direct comparison of
thelial cells [10]. Recently, the protein has also been suggested protein binding to membranes of different lipid composition
to participate in the plasma membrane repair mechanism [11]. within one experiment, individually addressable micropatterned
The crystal structure of annexin A1 was solved in the lipid bilayers on glass substrates were employed. Micropat-
presence [12] and absence [13] of Ca2+ ions showing the four terned membranes can be prepared by forming polymeric ca-
characteristic subdomains of the protein, each comprising five pillaries on a hydrophilic glass substrate that are filled with
α-helices. All eight Ca2+ binding sites are located at the convex unilamellar vesicle suspensions [22–25]. Capillaries are created
side of the slightly curved disk, while the N-terminus with an by placing an elastomeric polydimethylsiloxane stamp on the
intermediate length of 40 amino acids is presumably located at substrate forming a tight and sealed contact. Vesicle adsorption,
S. Faiß et al. / Biochimica et Biophysica Acta 1778 (2008) 1601–1610 1603

Fig. 1. Fluorescence micrographs of micropatterned lipid membranes on a glass surface after addition of 0.2 μM Texas Red labeled annexin A1. The membranes were
labeled with 0.5 mol% β-BODIPY 500/510 C12-HPC A. The fluorescence image was taken using an excitation wavelength of 488 nm to monitor the BODIPY
fluorescence. B. The fluorescence image was taken with excitation at 543 nm to monitor the Texas Red fluorescence of the protein.

spreading and fusion to form planar lipid bilayers is restricted ter at higher POPS content. Ca2+ ions are expected to bridge the
to the capillaries, which results in the formation of parallel negatively charged SiO2 surface with the negatively charged
membrane stripes. Fig. 1 shows fluorescence micrographs POPS headgroups leading to an increased adhesion of the
of four membrane stripes with a width of 15 μm compo- adsorbed vesicles, which is required for vesicle rupture on the
sed of POPC/POPS (4:1) and neat POPC doped with 0.5 mol% surface. The overshoot of the parameter Psi as found for POPC/
β-BODIPY 500/510 C12-HPC. To visualize membrane bound POPS (95:5) vesicles is indicative of the deposition of intact
annexin A1 by fluorescence microscopy, the protein was labe- vesicles followed by rupture and fusion on the surface. This
led with a Texas Red fluorescent dye covalently linked to
the amino groups of the protein. The membrane stripes were
incubated with 0.2 μM Texas Red labeled annexin A1 for 15 min
in the presence of 1 mM CaCl2 and afterwards rinsed with
protein-free buffer. Fig. 1B shows the corresponding fluore-
scence image obtained upon excitation of the Texas Red dye
of the protein at 543 nm. It is obvious that the membrane
stripes containing 20 mol% POPS show a red fluorescence,
indicating the specific binding of annexin A1 to POPS lipids.
The POPC stripes are, however, black demonstrating that PC
lipids fully prevent non-specific binding of the protein. This is
not the case on the pure glass substrate, where a faint red
fluorescence is discernable as a result of non-specific binding
of annexin A1. In Fig. 1A, the same membrane stripes were
visualized by excitation of the β-BODIPY dye of the lipid at
488 nm. The pure POPC bilayers show the characteristic green
fluorescence of the lipid dye, while the green fluorescence of
the membranes, composed of POPC/POPS, on which anne-
xin A1 has been bound to, is greatly diminished as a result of a
fluorescence resonance energy transfer (FRET) between β-BO-
DIPY and Texas Red. This FRET indicates that the protein is in
close proximity to the membrane.

2.2. Thickness determination of annexin A1 by means of


ellipsometry

Ellipsometry offers the unique possibility to study the change


in thickness on a SiO2/Si surface upon adsorption of lipids
and proteins [26]. Fig. 2A displays the change of the two Fig. 2. A. Formation of a solid supported lipid bilayer on SiO2/Si followed by
ellipsometric angles (Delta and Psi) of a SiO2/Si wafer exposed ellipsometry. Addition of 0.5 mg/mL vesicle suspension composed of either
to a vesicle suspension (0.5 mg/mL) consisting of either POPC/ POPC/POPS (4:1) (△,▲) or (95:5) (□,■) to a cleaned silica surface leads to a
decrease in Delta (open symbols) and increase in Psi (closed symbols). B. Time-
POPS (4:1) or POPC/POPS (95:5) in the presence of a buffer resolved thickness change after addition of 0.4 μM annexin A1 to a bilayer
containing 1 mM Ca2+ [26]. Consistent with the results of composed of POPC/POPS (4:1) in the presence of 1.0 mM Ca2+. After 20 min,
Reimhult et al. [27], bilayer formation occurs considerably fas- the sample was rinsed with a 1.0 mM CaCl2 containing buffer.
1604 S. Faiß et al. / Biochimica et Biophysica Acta 1778 (2008) 1601–1610

Fig. 3. Topographic scanning force microscopy images of POPC/POPS membranes with variable POPS content on silica in the presence of 0.05 mM CaCl2 after
addition of 0.4 μM annexin A1 and rinsing with protein-free buffer. A. 5 mol% POPS, B. 10 mol% POPS, C. 20 mol% POPS. The graph next to panel A shows the
corresponding height profile across two protein domains.

is reminiscent of comprehensive QCM studies put forward by thickness is reduced due to an incomplete coverage (ellipsometry)
Kasemo and coworkers [28,29], Richter et al. [30–32] and and indentation of the protein layer into the soft POPC/POPS
Janshoff and coworkers [17,33–36]. Eventually, a mean thickness bilayer (SFM), respectively.
of the supported membranes of 4–5 nm is achieved as calculated
from the change in Delta [26], which demonstrates that lipid 2.3. Scanning force microscopy of annexin A1 bound to
bilayers with a low number of defects have been formed. The POPC/POPS membranes
adsorption of annexin A1 on these bilayers was also followed by
ellipsometry. Fig. 2B shows the binding of annexin A1 to a bilayer SFM is highly suitable to investigate the lateral organization
composed of POPC/POPS (4:1) in the presence of 1 mM CaCl2. of membrane bound annexin A1 to POPC/POPS membranes.
An average thickness change of 2.5 nm is observed, which di- Usually, membranes are prepared on silicon surfaces by spread-
minishes to about 2.2 nm if rinsed with buffer containing 1 mM ing and fusion of unilamellar vesicles [37]. In previous studies
CaCl2 demonstrating that a substantial amount of annexin A1 is investigating the interaction of annexin A1 and A2 with solid
bound irreversibly to the solid supported bilayer as long as cal- supported bilayers, membranes were mainly immobilized on
cium ions are present. A similar thickness of about (2.2 ± 0.3) nm mica surfaces [38,39]. However, it turned out that the mica
was found for annexin A1 adsorbed on POPC/POPS membranes substrate influences the distribution of negatively charged PS
imaged by scanning force microscopy (SFM) [16]. Compared to lipids so that the content of PS in the leaflet that faces the
the height of annexin A1 protein domains observed on DPPC/ aqueous solution is diminished, which does not occur if silica
DPPS bilayers, where a mean thickness of (3.2 ± 0.3) nm was substrates such as silicon wafers are used [40]. Fig. 3 shows
monitored by means of SFM [14], the thickness of the annexin A1 intermittent contact mode SFM images of POPC/POPS mem-
layer on POPC/POPS is considerably smaller. The reduced branes on silica substrates containing 5, 10 and 20 mol% POPS,
thickness might be either attributed to a partial insertion of the respectively, after adsorption of annexin A1 in the presence of
protein in the soft matrix of the POPC/POPS bilayer or the 0.05 mM CaCl2 and subsequent rinsing with protein-free buffer.
S. Faiß et al. / Biochimica et Biophysica Acta 1778 (2008) 1601–1610 1605

Fig. 4. Topographic scanning force microscopy images of POPC/POPS membranes with variable POPS content on silica in the presence of 1.0 mM CaCl2 after
addition of 0.4 μM annexin A1 and rinsing with protein-free buffer. A. 5 mol% POPS, B. 10 mol% POPS, C. 20 mol% POPS. The graph next to panel A shows the
corresponding height profile across two protein domains.

From the topographic images, it can be deduced that annexin A1 the protein domains exhibit a mean radius of (145 ± 80) nm
is organized in laterally confined, more or less round protein and the coverage increases from 0.22 at 5 mol% POPS to one,
domains. In the absence of annexin A1, the POPC/POPS bila- where individual domains can no longer be distinguished. A
yers show in rare cases POPS-enriched domains with an in- comparison of the protein surface coverage at 5 mol% POPS
creased height of 0.5–1 nm but mostly these domains were not demonstrates that an increase in Ca2+ concentration results in
visible on silicon wafers. As shown in Fig. 3, the surface an increase in the surface coverage from 0.13 (0.05 mM CaCl2)
coverage increases as a function of the POPS content of the to 0.22 (1 mM CaCl2). From these experiments it was conclu-
membrane. At 5 mol% POPS, individual protein domains with ded that annexin A1 adsorbs irreversibly on POPC/POPS mem-
an approximate height of about 2 nm can be identified, while branes on silica and that the protein coverage increases with
at 20 mol% POPS protein domains merge so that individual both calcium ion concentration and POPS content. Lateral ag-
domains can no longer be distinguished. The observed height of gregates of monomolecular height (domains) are formed and
around (2.2 ± 0.3) nm [16] is lower than that of membrane bound the mean domain size decreases with increasing POPS content,
annexin A1 on gel-phase lipids [14] as discussed above. The but is independent of the calcium ion concentration.
mean domain radius decreases from (135 ± 70) nm at 5 mol%
POPS to (80 ± 45) nm at 10 mol% POPS, as inferred from grain 3. Adsorption and desorption kinetics of annexin A1 at
analysis. This might be discussed in terms of the presence of a different Ca2+ concentrations
larger number of nucleation sites of protein binding at higher
phosphatidylserine content. Fig. 4 shows SFM images of mem- Adsorption and desorption of annexin A1 on and from POPC/
brane bound annexin A1 at a 20 times larger CaCl2 concentration POPS membranes are well known to be a function of the Ca2+
(1 mM) as a function of the POPS content. Again, the protein concentration in solution. It is of great interest to elucidate to
coverage increases with the amount of POPS present in the what extend the amount of reversibly bound protein, the kinetic
membrane and individual annexin domains can only be obser- parameters and the lateral organization of the membrane are
ved at low phosphatidylserine concentration. At 5 mol% POPS, influenced by the Ca2+ concentration. To address these aspects,
1606 S. Faiß et al. / Biochimica et Biophysica Acta 1778 (2008) 1601–1610

A1 adsorption reaches saturation [15]. Annexin A1 binding to


the solid supported membrane is indicated by a decrease in
resonance frequency with maximum frequency shifts that are a
function of the Ca2+ concentration in solution (Fig. 5). Control
experiments, in which annexin A1 was either added to POPC/
POPS membranes in the absence of Ca2+-ions or added to POPC
membranes in the presence of 1 mM CaCl2, resulted in no
observable frequency shifts, which confirmed that the shifts in
resonance frequency can be attributed to a specific and Ca2+-
dependent binding of annexin A1 to acidic phospholipids.
Desorption of annexin from the POPC/POPS membranes was
initiated by rinsing the system with pure buffer solution con-
taining the corresponding CaCl2 concentration 10 min after
Fig. 5. Time-resolved frequency changes of 5 MHz quartz resonators during protein addition (Fig. 5). Independent of the Ca2+-concentration,
adsorption and desorption of annexin A1 on solid supported POPC/POPS (4:1) removal of annexin A1 from solution led to a small increase
membranes at different Ca2+ concentrations. QCM data (grey) and the results of
in resonance frequency, which can be attributed to annexin A1
computer simulations (black lines) are shown [14]. The arrows indicate the time
of buffer flushing. Simulation parameters are compiled in Table 1. desorption from the membrane. This result implies that the
majority of the protein is irreversibly bound on the membrane,
while a minor part is reversibly bound at a given Ca2+ con-
the quartz crystal microbalance (QCM) technique has been centration. These findings are the basis for a kinetic analysis
extensively used, which allows to follow the adsorption and based on computer simulations employing an off-lattice rever-
desorption of annexin A1 on and from membranes in a time- sible random sequential adsorption (RSA) algorithm on a he-
resolved and label free manner [34]. For these experiments, terogeneous surface.
solid supported monolayers composed of POPC/POPS (4:1)
were deposited on a gold electrode of a quartz plate, which was 3.1. Theoretical modeling of the adsorption/desorption data
functionalized with a self-assembly monolayer of octanethiol.
The quality of the chemisorbed monolayer and the hybrid bila- The experimental evidence suggested that POPC/POPS mem-
yer was monitored by impedance spectroscopy, which is very branes display two kinds of binding sites for annexin A1. One
sensitive to detection of small defects [41]. Protein binding to binding site is provided by POPS enriched domains, to which
the membrane was initiated by adding annexin A1 in the annexin A1 binds in an irreversible manner, while the binding
presence of different Ca2+ concentrations. A protein concentra- to the POPC enriched matrix, with single POPS molecules or
tion of 0.4 μM was used, which ensures that independent of small clusters, is reversible. Based on these findings a model
the Ca2+ concentration the frequency shift produced by annexin of annexin A1 binding to a heterogeneous surface has been

Fig. 6. Illustration of the RSA computer simulations based on the model of heterogeneous surfaces. Annexin A1 binds irreversibly to circular POPS-rich domains in the
membrane, while protein adsorption to the POPC rich matrix is reversible.
S. Faiß et al. / Biochimica et Biophysica Acta 1778 (2008) 1601–1610 1607

transferred into dynamic Monte Carlo (DMC) computer simula-


tions of the adsorption and desorption process. Considering a
linear flow of particles to the surface, the time-dependent surface
coverage can be expressed by two different rate equations (Eqs.
(1) and (2)) for reversible and irreversible binding:

dH1 kon pa2 qbulk U1 ðH1 Þ  koff H1


¼ ð1Þ
dt 1 þ kkon U1 ðH1 Þ
tr

dH2 kirr pa2 qbulk U2 ðH2 Þ


¼ ð2Þ
dt 1 þ kkirr U2 ðH2 Þ
tr

Θ1 and Θ2 are the reversible and irreversible surface covera-


ges, kon, koff, kirr, and ktr are the kinetic rate constants as described
in Fig. 6, and Φ1 and Φ2 are the available surface functions
(ASFs). The ASF describes the available surface for deposition
of particles from bulk solution as a function of coverage. For
instance, the classical Langmuir kinetics are represented by
an ASF of Φ(Θ) = 1 − Θ. These rate equations are solved by an
RSA algorithm, in which the proteins are modeled as disks with
radius of a = 3 nm based on the crystal structure of annexin
A1 [13]. The particle density in bulk solution is denoted as ρbulk.
The particles are randomly placed on a square that contains
circular domains representing the POPS rich phase (Fig. 6).
An average radius of 50 nm of the POPS domains was assu-
med independent of the Ca2+ concentration, which was deduced
from SFM images [16,39]. All simulations were carried out on
a simulation area of (1.5 × 1.5) μm2.
In order to compare the data produced by the computer
simulations with the results of the QCM technique, the sur-
face coverage as obtained from the simulations has to be
converted into a frequency shift. This is achieved by multi-
plying the calculated coverage with a constant negative va-
lue given by a measured frequency shift that arises from a
known protein surface coverage. The multiplying factor P
has been calculated by taking the surface coverage of ir-
reversibly bound protein to the membrane (Θirr), as derived
from SFM images, into account and corresponds to the re- Fig. 7. Frequency changes (grey) of 5 MHz quartz resonators and corresponding
results of computer simulations (black) after addition of 0.4 μM annexin A1 with
sonance frequency shift Δfirr from QCM readouts that has been various incubation times. A. c2+ Ca = 0.05 mM, Θdom = 0.15, B. cCa = 0.5 mM,
2+

Θdom = 0.255, C. c2+


Ca = 1.0 mM, Θ dom = 0.335. For all other parameters see Table 2.

Table 1
Results of computer simulations as shown in Fig. 5 monitored upon irreversible binding of annexin A1 to the
c2+
Ca / mM kon⁎ / 103 koff / s− 1 K / 105 M− 1 kirr⁎ / 105 Θdom Θsim
irr membrane (Eq. (3)):
(M s)− 1 (M s)− 1
Dfirr
0.01 0.51 0.007 0.73 0.34 0.065 0.035 P¼ ð3Þ
0.05 0.89 0.007 1.27 0.43 0.150 0.083 Hirr d Hjam
0.10 1.19 0.005 2.38 0.85 0.240 0.137
0.50 2.86 0.010 2.86 10.2 0.255 0.158 For the adsorption of 0.4 μM annexin A1 on a POPC/POPS
1.00 2.04 0.007 2.91 11.9 0.335 0.208 (4:1) membrane at 1 mM CaCl2 the resonance frequency shift
The parameters correspond to the binding of 0.4 μM annexin A1 to POPC/POPS is − 19.5 Hz and the corresponding surface coverage was
(4:1) membranes and an incubation time of 10 min. kon⁎ and koff are the rate determined to be 0.38 as deduced from SFM images. With a
constants of reversible adsorption and desorption on and from the POPC- theoretical maximal coverage of the surface referred to as the
enriched phase. kirr⁎ is the rate constant of the irreversible adsorption on the jamming limit of Θjam = 0.547, the absolute surface coverage is
POPS-enriched domains. K = kon⁎ / koff is the reversible binding constant, Θdom is
the fraction of the simulation area covered by POPS-enriched domains, and Θsim
0.21 and the multiplying factor is P = − 94 Hz. The conversion
irr
is the surface coverage of irreversibly adsorbed protein. The average domain factor P is slightly larger than that reported in an earlier study
radius was set to 50 nm. [16], since the SFM images allow for a more detailed estimation
1608 S. Faiß et al. / Biochimica et Biophysica Acta 1778 (2008) 1601–1610

Table 2 range from 0.005–0.010 s− 1 and show no obvious dependency


A. Kinetic parameters obtained from adapting dynamic Monte Carlo simulations on the calcium ion concentration.
to QCM measurements carried out in buffer containing different CaCl2
concentrations (see Fig. 7)
3.2.2. Influence of incubation time on the parameter space
0.05 mM CaCl2
tInc / s kon⁎ / 103 (M s)− 1 −1
koff / s 5
K / 10 M −1
kirr⁎ / 10 (M s)
5 −1
Coverage
Since it is well conceivable that the lateral organization of the
35 1.36 0.010 1.36 0.43 0.016 protein clusters as well as the lipid domains change over time, it
108 1.36 0.010 1.36 0.43 0.057 remained to be elucidated whether the kinetic parameters also
720 0.95 0.007 1.36 0.43 0.083 change as a function of incubation time. The incubation time is
1800 0.54 0.004 1.35 0.43 0.088 defined as the time between protein addition and flushing with
0.50 mM CaCl2
buffer. By terminating the adsorption process at a very early
tInc / s kon⁎ / 103 (M s)− 1 koff / s− 1 K / 105 M− 1 kirr⁎ / 105 (M s)− 1 Coverage time point, it should become possible to find evidence whether
34 2.8 0.010 2.9 10.2 0.06 lipid domains were present in full scale before adsorption takes
90 2.8 0.010 2.9 10.2 0.144 place or increase in size during adsorption [16]. Although it was
600 2.8 0.007 2.9 10.2 0.158

1.0 mM CaCl2
tInc / s kon⁎ / 103 (M s)− 1 koff / s− 1 K / 105 M− 1 kirr⁎ / 105 (M s)− 1 Coverage
35 2.04 0.007 2.91 11.9 0.066
75 2.04 0.007 2.91 11.9 0.181
600 2.04 0.007 2.91 11.9 0.208
1800 1.45 0.005 2.90 13.6 0.211
Adsorption kinetics of annexin A1 to POPC/POPS (4:1) membranes was
interrupted at time tInc by rinsing with buffer containing the given Ca2+
concentration.

of the protein surface coverage, which is slightly below the


theoretical jamming limit.

3.2. Results of DMC simulations

Prior to the kinetic analysis, the transport rate to the surface


ktr had to be quantified. Experimentally and theoretically a rate
of ktr = 10− 6 m/s was found and used throughout the simulations
[34,42].

3.2.1. Impact of the calcium ion concentration


Fig. 5 shows the simulated frequency changes for different
Ca2+ concentrations together with the QCM measurements.
Rate constants kon, koff, and kirr were stepwise adapted to ac-
hieve the best accordance between the QCM data and the
simulation. Noteworthy, the rate constants are defined in dif-
ferent parts of the curve, thus rendering the manual fitting
procedure very robust. The obtained values for the rate con-
stants are summarized in Table 1. The most important outcome
of the simulations and fitting procedure is that an increase in
Ca2+ concentration is accompanied by a substantial increase
of the overall domain area, on which irreversible adsorption
takes place. At 1 mM Ca2+, Θdom, the surface covered by the
POPS-enriched phase, was set to 0.335 to achieve a surface
coverage of 0.21 corresponding to the value obtained from the
analysis of SFM images, while Θdom decreases to 0.065 at
10 μM Ca2+ resulting in protein coverage of only 3.5%. Both
the reversible binding constant K ¼ kkonoff⁎ (k ⁎ ¼ kon pa2 NA with
the Avogadro number NA) and the rate constant for irreversi-

ble binding kirr ¼ kirr pa2 NA increase with the calcium ion con- Fig. 8. A. Course of the resonance frequency of adsorption Δfads (●) and
desorption Δfdes (▲) of annexin A1 binding to POPC/POPS (4:1) membranes as
centration. At 0.5 mM and 1 mM Ca2+, however, they are a function of Ca2+ concentration. B. Course of the binding constant K and C. the
almost constant and the change of the frequency course is rate constant of irreversible binding kirr⁎ as a function of Ca2+ concentration. The
mainly caused by the difference in domain area. The koff values time of protein incubation was 10 min for all measurements.
S. Faiß et al. / Biochimica et Biophysica Acta 1778 (2008) 1601–1610 1609

shown by SFM on pure bilayers that POPS-enriched domains a lower binding and unbinding dynamics is found by compari-
exist, it is unclear what the impact of protein binding on the son of the adsorption and desorption rate constants as a function
domain organization might be. Therefore, the time resolved of incubation time.
protein adsorption was monitored by QCM at different calcium Generally, the reversible binding constant K and the irre-
ion concentrations with incubation times ranging from about versible rate constant kirr⁎ increase both with the calcium ion
30 s to 30 min. Dynamic Monte Carlo simulations were per- concentration (Fig. 8 B,C). Plotting the parameters on a half
formed and the simulation parameters (kon⁎, koff, and kirr⁎) ad- logarithmic scale shows in both cases a sigmoidal curve with an
justed to match the experimental data. inflection point of 7·10− 5 M Ca2+ (K) and about 2.5·10− 4 M Ca2+
Fig. 7 shows the simulated frequency courses for different (kirr⁎). The considerable change of these two parameters at Ca2+
incubation times at three different Ca2+ concentrations together concentrations in the 10− 4 M range is consistent with the ob-
with the QCM results. The domain area was kept constant for a served change in resonance frequency, which shows for both,
fixed calcium ion concentration. The particle density for short Δfads and Δfdes a point of inflection at 7·10− 5 M Ca2+ (Fig. 8A).
incubation times was corrected for the inevitably existing Even though the results summarized in this article are all
protein concentration gradient at the beginning of the measure- obtained from in vitro experiments with well defined lipid com-
ment [37]. Table 2 summarizes the results obtained for three positions that mimic the inner leaflet of the plasma membrane,
different calcium ion concentrations and various incubation one might ask the question of how such observed POPS ag-
times. gregation and annexin A1 protein domain formation affects the
The simulated frequency courses agree well with the ex- protein's function within the cell. A conceivable scenario can be
perimental QCM data. While K and kirr remained constant, the envisioned as follows: A defined threshold of calcium ions is
rate constants of reversible binding, kon⁎ and koff, had to be needed to induce the formation of POPS clusters within a matrix
adjusted to lower values for longer incubation times [16,37]. At of POPC, which renders adsorption of annexin A1 irreversible as
high Ca2+ concentrations (0.5 and 1.0 mM) protein binding long as the calcium ion concentration is constant. If calcium ions
is virtually irreversible within the first 35 s and only a small are removed, demixing is reversed and annexin A1 is released.
amount of reversibly bound annexin can be released after 75 In fact, the presence of POPS domains entirely abolishes de-
and 90 s incubation time. The measurements at 0.05 mM Ca2+, sorption of annexin A1, thus creating small protein islands on
however, show a larger impact of the incubation time on the the surface, which might serve as new adsorption platforms
reversible binding dynamics. Already after 108 s, the same for other newly arriving proteins and membrane structures. The
amount of reversibly bound protein (3 Hz) is detectable as after binding properties of annexin A1 to POPS are thus modulated by
12 min (3 Hz) or even as at a later time point. This finding PS-clustering as a result of the Ca2+-concentration. This way, the
indicates that within the time scale of the experiment, a slight cell may avoid a high binding constant of annexin A1 per se,
rearrangement of proteins and lipids might occur at low cal- which would interfere with fast dynamics but rather allows
cium ion concentration, which changes the nature of binding controlling the binding affinity of annexin A1 by a modification
dynamics. A possible explanation of the apparently reduced of membrane organization as a function of calcium ions.
adsorption dynamics at low Ca2+ concentrations might be a
limited accessibility of free POPS molecules at later times, Acknowledgements
which is responsible for the reduction of kon⁎. At later times,
reversibly bound annexin A1 might recruit more POPS mo- A.J. gratefully acknowledges the DFG (SFB 625) for fi-
lecules, which eventually leads to a reduced desorption rate. nancial support.
Notably, a dynamic lateral rearrangement during the adsorp-
tion experiment is not captured by the RSA simulation and is References
inferred only indirectly from the change in the parameter space
necessary to describe the experimental data. Interestingly, the [1] V. Gerke, S.E. Moss, Annexins: from structure to function, Physiol. Rev.
binding constant K is preserved giving rise to an apparently 82 (2002) 331–371.
consistent adsorption isotherm, which misses to display subtle [2] U. Rescher, V. Gerke, Annexins — unique membrane binding proteins
with diverse functions, J. Cell Sci. 117 (2004) 2631–2639.
shifts from reversible to irreversible binding in the case of low [3] V. Gerke, C.E. Creutz, S.E. Moss, Annexins: linking Ca2+ signalling to
Ca2+ concentration over time. membrane dynamics, Nat. Rev., Mol. Cell Biol. 14 (2005) 206–213.
Dynamic Monte Carlo simulations in conjunction with ki- [4] M.J. Crumpton, J.R. Dedman, Protein terminology tangle, Nature 345
netic QCM measurements show that the reversible and ir- (1990) 212.
reversible binding as well as the lateral organization of the [5] P. Meers, T. Mealy, N. Pavlotsky, A.I. Tauber, Annexin I-mediated vesicular
aggregation: mechanism and role in human neutrophils, Biochemistry 31
membrane is only slightly dependent on the incubation time (1992) 6372–6382.
suggesting a rather stationary picture of the adsorption pro- [6] L. Oshry, P. Meers, T. Mealy, A.I. Tauber, Annexin-mediated membrane
cess. POPS aggregates are preformed due to the presence of fusion of human neutrophil plasma membranes and phospholipid vesicles,
calcium ions and are not affected by protein addition. This Biochim. Biophys. Acta 1066 (1991) 239–244.
is confirmed by SFM images of POPC/POPS bilayers immo- [7] J. Seemann, K. Weber, V. Gerke, Annexin I targets S100C to early endosomes,
FEBS Lett. 413 (1997) 185–190.
bilized on mica surfaces (no protein present) revealing small [8] J. Seemann, K. Weber, M. Osborn, R.G. Parton, V. Gerke, The association
domains of POPS in the presence of calcium ions [16,39]. of annexin I with early endosomes is regulated by Ca2+ and requires an
However, at low Ca2+ concentration a subtle tendency towards intact N-terminal domain, Mol. Biol. Cell 7 (1996) 1359–1374.
1610 S. Faiß et al. / Biochimica et Biophysica Acta 1778 (2008) 1601–1610

[9] F. D'Acquisto, A. Meghani, E. Lecona, G. Rosignoli, K. Raza, C.D. [26] S. Faiß, S. Schuy, D. Weiskopf, C. Steinem, A. Janshoff, Phase transition
Buckley, R.J. Flower, M. Perretti, Annexin-1 modulates T-cell activation of individually addressable microstructured membranes visualized by
and differentiation, Blood 109 (2007) 1095–1102. imaging ellipsometry, J. Phys. Chem., B 111 (2007) 13979–13986.
[10] M. Perretti, R.J. Flower, Annexin 1 and the biology of the neutrophil, [27] E. Reimhult, F. Höök, B. Kasemo, Intact vesicle adsorption and supported
J. Leukoc. Biol. 75 (2004) 25–29. biomembrane formation from vesicles in solution: influence of surface
[11] A. McNeil, U. Rescher, V. Gerke, P. McNeil, L, Requirement for annexin chemistry, vesicle size, temperature, and osmotic pressure, Langmuir 19
A1 in plasma membrane repair, J. Biol. Chem. 281 (2006) 35202–35207. (2003) 1681–1691.
[12] A. Rosengarth, H. Luecke, A calcium-driven conformational switch of the [28] C.A. Keller, K. Glasmästar, V.P. Zhdanov, B. Kasemo, Formation of
N-terminal and core domains of annexin A1, J. Mol. Biol. 326 (2003) supported membranes from vesicles, Phys. Rev. Lett. 84 (2000) 5443–5446.
1317–1325. [29] C.A. Keller, B. Kasemo, Surface specific kinetics of lipid vesicle
[13] A. Rosengarth, V. Gerke, H. Luecke, X-ray structure of full-length annexin adsorption measured with a quartz crystal microbalance, Biophys. J. 75
1 and implications for membrane aggregation, J. Mol. Biol. 306 (2001) (1998) 1397–1402.
489–498. [30] R.P. Richter, R. Bérat, A.R. Brisson, Formation of solid-supported lipid
[14] A. Janshoff, M. Ross, H.-J. Galla, V. Gerke, C. Steinem, Visualization of bilayers: an integrated view, Langmuir 22 (2006) 3497–3505.
annexin I binding to calcium-induced phosphatidylserine domains, Chem- [31] R.P. Richter, A.R. Brisson, Following the formation of supported lipid
BioChem 7/8 (2001) 587–590. bilayers on mica: a study combining AFM, QCM-D, and ellipsometry,
[15] K. Kastl, M. Ross, V. Gerke, C. Steinem, Kinetics and thermodynamics Biophys. J. 88 (2005) 3422–3433.
of annexin A1 binding to solid-supported membranes: A QCM study, [32] R.P. Richter, A. Mukhopadhyay, A.R. Brisson, Pathways of lipid vesicle
Biochemistry 41 (2002) 10087–10094. deposition on solid surfaces: A combined QCM-D and AFM study,
[16] K. Kastl, M. Menke, E. Lüthgens, S. Faiß, V. Gerke, A. Janshoff, C. Biophys. J. 85 (2003) 3035–3047.
Steinem, Partially reversible adsorption of annexin A1 on POPC/POPS [33] B. Pignataro, C. Steinem, H.-J. Galla, H. Fuchs, A. Janshoff, Specific
bilayers investigated by QCM measurements, SFM, and DMC simula- adhesion of vesicles monitored by scanning force microscopy and quartz
tions, ChemBioChem 77 (2006) 106–115. crystal microbalance, Biophys. J. 78 (2000) 487–498.
[17] C. Steinem, A. Janshoff, Specific adsorption of annexin A1 on solid [34] E. Lüthgens, A. Herrig, K. Kastl, C. Steinem, B. Reiss, J. Wegener, B.
supported membranes, in: C. Steinem, A. Janshoff (Eds.), Piezoelectric Pignataro, A. Janshoff, Adhesion of liposomes: a quartz crystal microbalance
sensors, Springer Verlag, Berlin, 2007, pp. 281–302. study, Meas. Sci. Technol. 14 (2003) 1865–1875.
[18] K. Kastl, A. Herrig, E. Lüthgens, A. Janshoff, C. Steinem, Scrutiny of annexin [35] B. Reiss, A. Janshoff, C. Steinem, J. Seebach, J. Wegener, Adhesion
A1 mediated membrane–membrane interaction by means of a thickness shear kinetics of functionalized vesicles and mammalian cells: a comparative
mode resonator and computer simulations, Langmuir 20 (2004) 7246–7253. study, Langmuir 19 (2003) 1816–1823.
[19] E. Bitto, W. Cho, Structural determinant of the vesicle aggregation activity [36] S. Faiß, E. Lüthgens, A. Janshoff, Adhesion and rupture of liposomes
of annexin I, Biochemistry 38 (1999) 14094–14100. mediated by electrostatic interaction monitored by thickness shear mode
[20] E. Bitto, M. Li, A.M. Tikhonov, M.L. Schlossman, W. Cho, Mechanism resonators, Eur. Biophys. J. 33 (2004) 555–561.
of annexin I-mediated membrane aggregation, Biochemistry 39 (2000) [37] S. Faiß, PhD, University of Mainz, Mainz 2007.
13469–13477. [38] A. Herrig, M. Janke, J. Austermann, V. Gerke, A. Janshoff, C. Steinem,
[21] A. Rosengarth, J. Rösgen, H.-J. Hinz, V. Gerke, A comparison of the Cooperative adsorption of ezrin on PIP2-containing membranes, Bio-
energetics of annexin I and annexin V, J. Mol. Biol. 288 (1999) 1013–1025. chemistry 45 (2006) 13025–13034.
[22] A. Janshoff, S. Künneke, Individually addressable solid supported lipid [39] M. Menke, V. Gerke, C. Steinem, Phosphatidylserine membrane domain
membranes formed by micromolding in capillaries, Eur. Biophys. J. 29 clustering induced by annexin A2/S100A10 heterotetramer, Biochemistry
(2000) 549–554. 44 (2005) 15296–15303.
[23] S. Künneke, A. Janshoff, Visualization of molecular recognition events on [40] R.P. Richter, N. Maury, A.R. Brisson, On the effect of the solid support on
microstructured lipid–membrane compartments by in situ scanning force the interleaflet distribution of lipids in supported lipid bilayers, Langmuir
microscopy, Angew. Chem. Int., Ed. 41 (2002) 314–317. 21 (2004) 299–304.
[24] S. Schuy, A. Janshoff, Thermal expansion of microstructured DMPC [41] C. Steinem, A. Janshoff, W.-P. Ulrich, M. Sieber, H.-J. Galla, Impedance
bilayers quantified by temperature controlled atomic force microscopy, analysis of supported lipid bilayer membranes: a scrutiny of different
ChemPhysChem 7 (2006) 1207–1210. preparation techniques, Biochim. Biophys. Acta 1279 (1996) 169–180.
[25] S. Schuy, A. Janshoff, Microstructuring of phospholipid bilayers on [42] Z. Adamczyk, B. Siwek, P. Waszynski, E. Musial, Kinetics of particle
gold surfaces by micromolding in capillaries, J. Coll. Int. Sci. 295 (2006) deposition in the radial impinging-jet cell, J. Coll. Int. Sci. 242 (2001)
93–99. 14–24.

You might also like