• No results found

Lipid Bilayer-like Mixed Self-Assembled Monolayers with Strong Mobility and Clustering-Dependent Lectin Affinity

N/A
N/A
Protected

Academic year: 2021

Share "Lipid Bilayer-like Mixed Self-Assembled Monolayers with Strong Mobility and Clustering-Dependent Lectin Affinity"

Copied!
25
0
0

Loading.... (view fulltext now)

Full text

(1)

http://www.diva-portal.org

Postprint

This is the accepted version of a paper published in Langmuir. This paper has been peer-reviewed but does not include the final publisher proof-corrections or journal pagination.

Citation for the original published paper (version of record):

Yeung, S Y., Sergeeva, Y., Dam, T., Jönsson, P., Pan, G. et al. (2019)

Lipid Bilayer-like Mixed Self-Assembled Monolayers with Strong Mobility and Clustering-Dependent Lectin Affinity

Langmuir, 35(24): 8174-8181

https://doi.org/10.1021/acs.langmuir.9b01452

Access to the published version may require subscription. N.B. When citing this work, cite the original published paper.

Permanent link to this version:

(2)

Lipid bilayer-like mixed self-assembled monolayers with

1

strong mobility and clustering dependent lectin affinity

2

3

Sing Yee Yeunga, Yulia Sergeeva, Tommy Damb, Peter Jönssonb, Guoqing Pana, Vivek 4

Chaturvedia and Börje Sellergrena* 5

a Department of Biomedical Sciences and Biofilms-Research Center for Biointerfaces 6

(BRCB), Faculty of Health and Society, Malmö University, 205 06 Malmö, Sweden.

7

b Department of Physical Chemistry, Lund University, Box 124, 22100 Lund, Sweden 8 * borje.sellergren@mau.se 9 10 Abstract 11

Glycans at the surface of cellular membranes modulate biological activity via multivalent

12

association with extracellular messengers. The lack of tuneable simplified models mimicking

13

this dynamic environment complicates basic studies of these phenomena. We here present a

14

series of mixed reversible self-assembled monolayers (rSAMs) that addresses this deficiency.

15

Mixed rSAMs were prepared in water by simple immersion of a negatively charged surface in

16

a mixture of sialic acid- and hydroxy-terminated benzamidine amphiphiles. Surface

17

compositions derived from infrared reflection absorption spectroscopy (IRAS) and film

18

thickness information (atomic force microscopy, ellipsometry) suggest the latter to be

19

statistically incorporated in the monolayer. These surfaces affinity for the lectin hemagglutinin

20

revealed a strong dependence of the affinity on the presentation, density and mobility of the

21

sialic acid ligands. Hence, a spacer length of 2 ethylene glycol and a surface density of 15%

22

resulted in a dissociation constant Kd,multi of 1.3 x 10-13 M, on a par with the best di- or tri-23

(3)

saccharide based binders reported to date, whereas a density of 20% demonstrated complete

24

resistance to hemagglutinin binding. These results correlated with ligand mobility measured by

25

fluorescence recovery after photobleaching (FRAP) which showed a dramatic drop in the same

26

interval. The results have a direct bearing on biological cell surface multivalent recognition

27

involving lipid bilayers and we believe will impact the design of model surfaces and sensors

28

for both fundamental and applied studies.

(4)

Introduction

30

Glycans, covering the surface of cellular membranes, play crucial roles in a wide range of cell

31

surface processes, such as pathogen recognition and binding, antibody recognition and cell-cell

32

interaction and signalling.1-3 As a singular glycan-protein interaction is weak, multiple 33

association between the two entities is necessary to achieve high avidity. Elucidating their

34

biological roles require understanding of the glycan’s topologies in relation to the binding

35

affinity of its corresponding ligand. The complexity and diverse nature of glycans and their

36

native environment have hampered progress in the field. In order to decode glycan’s

37

functionality for improving therapeutics and diagnostics, researchers use simplified and

38

tuneable two or three-dimensional models.1-5 39

Strategies for immobilizing glycans onto scaffolds can be either static (polymers, dendrimers,

40

self-assembled monolayers (SAMs)) or dynamic (supported lipid bilayers (SLBs), host-guest

41

based, poly rotaxanes). The latter is particularly attractive due their responsive nature and

42

biomimetic characteristics.1, 6-9 Surfaces with lateral adaptability such as SLBs or liposomes 43

spatially rearranges the glycans during the binding processes. These unique properties as

44

compared to their static counterparts enhance binding affinity and aid in the discovery of

45

secondary mechanisms.10-13 Despite their importance, apart from lipid assemblies, 2D interfaces 46

with long-range mobility resembling biological membranes suitable for molecular recognition

47

are rare.

48

We have recently reported on an air-stable and adaptable biosensing platform, reversible

self-49

assembled monolayers (rSAMs), featuring strongly enhanced affinity and sensitivity towards

50

viruses and lectins.13 This new sensing system utilizes noncovalent amidinium-carboxylate ion 51

pairs for building up stable two-dimensional assemblies, akin to lipid bilayers but with a simple

52

preparation process and fast on/off rates. Thus, benzamidine-terminated amphiphiles

(5)

thiol SAMs to form ordered monolayers with a tunable pH responsiveness. Layer thicknesses

55

and order correlate with the molecular length of the amphiphile, which beyond a certain length

56

feature liquid crystalline-like order and an odd even chain length related tendency to form

57

bilayers. These layers are stable towards rinsing and resist exchange by common plasma

58

proteins and charged surfactants, while resisting non-specific protein adsorption.13-17 59

Here we demonstrate the use of mixed rSAMs as a tunable platform to control surface density

60

and presentation of glycans using different mole fractions of sialic acid amphiphiles (E2-SA

61

and E4-SA) in ω-(ethylene glycol)2,α-(4-amidinophenoxy)decanes (E2-OH) solutions (Fig. 1). 62

The mixed rSAMs surfaces demonstrated tunable surface density, while retaining fluidity with

63

the formation of distinct saccharide clusters akin to gangliosides in lipid bilayers.12 Glycan 64

surface density thus modulates the dynamics and affinity of these surfaces towards its

65

corresponding lectins.

66

67

Figure 1. Schematic illustration of preparation and evaluation of sialic acid-functionalized mixed reversible self-assembled monolayers (rSAMs). i. 16-mercaptohexadecanoic acid

(MHA) modified surface placed in 50 µM mixed sialic acid amphiphiles, E2-SA, E4-SA or E6-SA mixed with E2-OH in pH 8 HEPES buffer solution. ii. Ordered mixed rE6-SAMs formed after 18 hrs of incubation and rinsing with pH 8 HEPES buffer. iii. Hemagglutinin binding isotherms on mixed rSAMs obtained. Images are not drawn to scale.

(6)

Results and discussion

69 70

Design of ligands and rSAMs: Optimization of ligand decorated SAMs demands attention to

71

multiple factors governing the multivalent interactions with the receptor. Two key parameters

72

to consider are the nature and length of the linker group and the surface density of ligands. For

73

instance, accessibility to lectin binding is enhanced by inserting linkers containing 2-3

74

ethyleneglycol repeat units between the glycan and the mesogen unit (Fig. 1) whereas lateral

75

spacing requires mixed SAMs to contain less than 20 % of ligand decorated amphiphile to

76

promote strong multivalent interactions.18-24 We have here compared sialic acid terminated 77

amphiphiles with two, four and six EG repeats in the sialic acid tether (E2-SA, E4-SA, E6-SA)

78

in combination with an OH terminated amidine with 2 EG repeats (E2-OH). The amphiphiles

79

were synthesised as described in the Supporting Information with the final step being the click

80

coupling of alkyne-modified sialic acid and azide-terminated amidine.

81

Mixed rSAM formation and stability: Our first objective was to study the formation and nature

82

of mixed rSAMs of ligand and OH terminated amphiphiles. For this purpose, we first chose the

83

shorter OH terminated amphiphiles E2-OH and the medium size ligand terminated amphiphile

84

E4-SA. Based on previous reports demonstrating optimal sialic acid coverages with respect to

85

hemagglutinin binding of 20 % or below 19-21, 23-24 a series of mixed rSAMs was prepared by 86

immersing 16-mercaptohexadecanoic acid (MHA) modified gold surfaces (MHA-SAMs) in pH

87

8 HEPES buffer containing varying mole fractions of sialic acid amphiphile, E4-SA in filler

88

E2-OH,

c

E4-SA = 0 to 0.2 (50 µM). Layer composition and molecular order and orientation of 89

the rSAM amphiphiles were evaluated by in situ (ISE) and ex situ (ESE) ellipsometry, quartz

90

crystal microbalance (QCM-D), infrared reflection-adsorption spectroscopy (IRAS) and atomic

91

force microscopy (AFM) (Fig. 2 and 3).

(7)

A comparison of the assembly kinetics by ISE and limiting equilibrium thickness for the

93

homogenous, E4-SA and E2-OH and mixed,

c

E4-SA = 0.2 systems suggested incorporation of 94

both amphiphiles in the latter layer (Fig. 2A).

95 96 97 98 99 100

A

B

C

0 0.05 0.1 0.15 0.2 1 0 20 40 60 80 χE4-SA Thickness (Å) E4-SA E2 0 100 200 300 400 500 0 20 40 60 80 Time (s) Thickness (Å) 0 0.2 1 E4-SA E2 2920 cm-1 C-H, asym 2851 cm-1 C-H, sym 2800 3000 3200 3400 3600 3800 Wavenumbers (cm-1) Absorbance 3350 cm-1 OH and NH2 0 0.05 0.1 0.15 0.2 χE4-SA = 1 0.002 842 cm-1 C-H 1511 cm-1 (C=C)1,4 800 1000 1200 1400 1600 Wavenumbers (cm-1) 1400 cm-1 C-O-C 1190 cm-1 C-C-O 1609 cm-1 (C=C)1,4

Figure 2. Surface characterization of mixed rSAMs. A. Real-time change in film thickness

upon exposure of an MHA-SAM to pH 8 HEPES buffer containing 50 µM cE4-SA = 0, 0.2 or 1.

B. Ex situ ellipsometric thickness of rinsed MHA-SAMs surfaces modified with cE4-SA = 0 - 1.

C. Baseline-corrected IRAS spectra of MHA-SAMs modified with cE4-SA = 0 – 1. Dotted lines

in A and B indicate calculated length of the corresponding amphiphiles. Assignments of the IR bands were based on positions reported in literature.13-17

(8)

All three systems demonstrated a rapid initial adsorption upon amphiphile exposure with a close

101

to monolayer thickness attained after ca. 100 s. This was followed by a considerably slower

102

phase reflecting continued amphiphile incorporation or ordering of the existing layer

103

components. Interestingly, the mixed rSAM with

c

E4-SA = 0.2 displayed the largest thickness as 104

compared to the homogenous systems of pure E2-OH or E4-SA.

105

Ex-situ air ellipsometry was then used to estimate the layer thickness of rinsed homogenous

106

and mixed rSAMs after incubation in the amphiphilic solutions overnight (Fig. 2B). Starting

107

with the rSAM of E2 a thickness slightly beneath the molecular length of E2 (Fig. S1) was

108

measured suggesting a nearly complete layer of E2 oriented perpendicular to the surface.

109

Monitoring the self-assembly by QCM-D confirmed this result, indicating the formation of a

110

relatively rigid film of a mass agreeing with the ISE results (ca 2.3 mg/m2) (Fig. S13). With 111

increasing E4-SA the thickness then continuously increased to reach a value of nearly 70 Å at

112

a ligand density of

c

E4-SA = 0.15. Beyond this ligand density the thickness dropped to reach a 113

value corresponding to sub-monolayer at

c

E4-SA = 1. As we previously argued13, this suggests 114

that the steric repulsion of the bulky sialic acid head groups (10 Å of the sialic acid head group

115

as compared to 4 Å of the benzene ring (see Fig. S1) precludes close packing of the layers in

116

the homogenous system. With the inclusion of filler E2-OH in the mixed systems (3 Å across

117

the ethylene glycol head group) the sialic acid head groups are brought further apart leading to

118

an improved lateral packing.25 Although qualitatively the behaviour above makes sense, the 119

quantitative changes in thickness are puzzling. The two-fold increase in thickness observed by

120

increasing the ligand density from

c

E4-SA = 0 to

c

E4-SA = 0.15 way exceeds the theoretical value 121

expected for statistically mixed monolayers. Formation of bilayered structures is one possible

122

explanation for this behaviour that we put forward in our first report. Although refractive index

123

effects could be another contributing factor, the increased IRAS signal intensity of the C=C

(9)

stretch band at 1609 cm-1 (Fig. 2C) supports the former explanation. To give further clues 125

regarding these structural features IRAS, goniometry and AFM were used to study the layers.

126

rSAM characterisation by IRAS and goniometry: Although the layer formed by homogenous

127

sialic acid amphiphiles,

c

E4-SA = 1 appeared partly unstable to rinsing (Fig. 2C), the layers 128

formed using mixed sialic acid and filler amphiphiles,

c

E4-SA = 0.05 – 0.2 exhibit relative band 129

intensities and bandwidths corresponding to well-ordered layers. Important structural detail is

130

revealed considering the intensities of the (C=C)1,4 stretch at 1609 cm-1 and 1511 cm-1 that have 131

transition dipole vectors oriented along the 1,4-axis of the benzene ring relative to the intensities

132

of the aromatic C–H out-of-plane bending mode at 841 cm-1 and the amidine N-C-N asymmetric 133

stretch at 1650 cm-1, that have transition dipole vectors perpendicular to the 1,4-axis. The 134

pronounced increase of the former and the concomitant decrease of the latter signals with

135

reference to the transmission spectra of the neat compounds point to a near upright position of

136

the layer amphiphiles.

137

In the high frequency region, the CH2 stretch vibrations at 2918 cm-1 (asym) and 2850 cm-1 138

(sym) and the sharpness of these bands of the layer spectra indicate the presence of trans

139

extended closely-packed amphiphiles. It is interesting to note that layers with

c

E4-SA = 0.2 and 140

0.1 feature a slight shift to lower wavenumbers (2919 cm-1 (asym); 2850 cm-1 (sym) and 2918 141

cm-1 (asym); 2849 cm-1 (sym) respectively) as compared to

c

E4-SA = 0.15 (2920 cm-1 (asym); 142

2852 cm-1 (sym)). This suggests an enhanced molecular order at the former sialic acid 143

densitiesand coincides with the decrease in amphiphile mobility (vide infra).

144

To obtain information regarding the percentage of sialic acid amphiphiles in the mixed rSAMs,

145

we compared the IRAS spectra of the homogenous filler and sialic acid layers, focusing on the

146

prominent signals from sialic acid at ca. 3350 cm-1 (bonded OH and mono substituted amide), 147

(10)

1400 cm-1 (C-O-C) and 1190 cm-1 (C-OH) in the E4-SA spectra. These signals corresponded 148

well with literature observations.27 Although a visual comparison of the intensities of the signals 149

characteristic for sialic acids correlated with an increased mole fraction of E4-SA in the

150

adsorption solution, a quantitative analysis was avoided due to overlapping signals and

151

inconsistency in the fingerprint regions between replicas.

152

Water contact angle measurements were performed to gain insight into surface hydrophilicity

153

(see Table S1). The highest value (Q = 50°) was measured for cE4-SA=0, showing that the rSAM 154

of neat E2-OH featured a less hydrophilic character. This value is higher than those measured

155

for PEG-SAMs on gold (Q ≈ 36°)31 which can be explained by the larger number of EG repeats 156

(> 3) in the latter compared to E2-OH. Lower values were measured for the mixed rSAMs

157

reflecting the introduction of E4-SA with its more hydrophilic head group. The most

158

hydrophilic surface (Q = 22°) was measured for cE4-SA=1, confirming the partial destabilization 159

of the rSAM and exposure of the MHA thiol SAM (vide supra).

160

rSAM characterisation by AFM: Atomic force microscopy (AFM) of the rSAMs was

161

performed in the peak force tapping mode (Fig. 3A, Fig. S2). The image of the rSAM of pure

162

OH-terminated amphiphile (E2-OH), cE4-SA = 0 was relatively featureless as expected from the 163

monolayer thickness recorded by ISE, with a roughness RRMS = 0.41 nm. Comparing this 164

surface with the rSAM formed from the lowest E4-SA concentration in the adsorption solution,

165

cE4-SA = 0.05, the roughness value had increased to RRMS = 0.54 nm. rSAMs formed from 166

higher E4-SA concentrations, cE4-SA = 0.1 – 0.2 featured distinct nano-sized domains protruding 167

ca. 20 Å from the shorter domains. This height difference roughly equates to the calculated

168

length difference (17 Å) between E4-SA and filler E2-OH suggesting that the tall domains are

169

primarily E4-SA and the shorter domains are primarily the filler. Similar phase separations were

170

observed in mixed carbohydrate self-assembled monolayers over a 28 day period28 and 171

(11)

gangliosides, GM3 in SLB membranes29 supporting our observations with long-range lateral 172

mobility in the rSAMs (vide infra).

173 174 175 176 177 178 179 180 Χ = 0 Χ = 0.1 Χ = 0.15 Χ = 0.2

Increasing mole fraction of E4-SA, χE4-SA

A

B

C

Χ = 0.05

D

0.00 0.05 0.10 0.15 0.20 0.25 0 5 10 15 20 χE4-SA Percentage Coverage by T all Domains (%) R2 = 0.9623 0.00 0.05 0.10 0.15 0.20 0.0 0.5 1.0 1.5 χE4-SA D ( µm 2/s) 0 100 200 300 400 0 5 10 15 x (nm) z (nm) 0 0.05 0.1 0.15 0.2 Δ = 2.4 nm Δ = 1.3 nm Δ = 2.3 nm Δ = 2.0 nm χE4-SA =

Figure 3. Atomic force microscopy (AFM) topography and profile and fluorescence recovery after photobleaching (FRAP) of layers formed by cE4-SA = 0-0.2 in the adsorption

solution. A. AFM topographic image of layers formed using cE4-SA = 0-0.2. B. Height profile

of AFM topographic images of layers in A. C. Plot of area covered by taller domains as a function of cE4-SA . D. Diffusivity of E2-FAM in different rSAMs measured with FRAP. The

rSAMs were doped with 1 % fluorescein-tagged amidine E2-FAM in the adsorption solution as reported in the experimental section. In the case of cE4-SA=0.2 no recovery was discernable

(12)

Both the mean size and the percent area coverage by the taller domains increased as a function

181

of mole fraction of E4-SA in the adsorption solution (Fig. 3A and C). Whereas the domain size

182

increased from ca. 25 to 100 nm in the range cE4-SA = 0.1-0.2 the coverage of the tall domains 183

correlated linearly with cE4-SA. Based on the slope of the line in Fig. 3C we conclude that at 184

least 70% of E4-SA should reside in the tall regions whereas 30% are either not incorporated

185

or reside in the thin regions. The nearly linear increase in average ISE layer thickness (Fig. 2B)

186

and the intensities of the E4-SA characteristic bands (Fig. 2C) in the interval cE4-SA=0-0.15 187

support the latter explanation. This moreover agrees with studies of mixed carbohydrate-based

188

self-assembled monolayers28 and indicate that mixed rSAMs of E4-SA and E2-OH form in a 189

near stochiometric manner. We then turned to investigate their affinity for the lectin

190

hemagglutinin.

191

rSAM binding affinity for hemagglutinin: The influence of the rSAM parameters (e.g. ligand

192

density, presentation, adsorption time) on the binding affinity of a trimeric sialic acid binding

193

protein, hemagglutinin (HA, dimensions (Å): 135 x (15-40) x (15-40)) was evaluated. After

194

formation and rinsing of the rSAMs in pH 8 buffer, HA was added and the film thickness

195

followed in real time by ISE until a stable reading was obtained followed by another rinsing

196

step. We first studied the adsorption of HA on mixed rSAMs of E2-SA and E2-OH at different

197

ligand densities (Fig. S3, Table S2). In agreement with our previous report,13 all curves display 198

positive cooperativity (h > 1) with a steep initial portion followed by clear saturation at

199

concentrations exceeding 20 nM. The cooperative Hill equation gave in all cases the highest

200

goodness of fit with the corresponding binding parameters listed in Table S2.

201

Highest affinity and capacity were measured for the rSAM with the lowest ligand density (c E2-202

SA=0.01) followed by slightly lower affinities at higher densities with exception of cE2-SA=0.2 203

(vide infra) (Fig. S3). This general trend agrees with previous literature reports on glycolipid

204

doped SLBs.11,12 Moreover, we observed that the EG terminated E2-monolayer generally 205

(13)

displayed a poor protein resistance and that also the 100% E2-SA rSAM bound HA with high

206

affinity. These results are supported by our previous findings.13,17 The latter rSAM was then 207

shown to display weaker affinity for the competing lectin (Concanavalin A) and human serum

208

albumin whereas the complementary lectin HA could be inhibited by the heavily sialylated

209

protein mucin. This showed that the recognition was driven by sialic acid-lecin interactions.

210

In view of the 2-phase kinetic profiles of the amphiphile adsorption (vide supra) we suspected

211

the time of rSAM adsorption to be a critical parameter. A comparison of two rSAMs (c E4-212

SA=0.1) prepared using a short (5000 s) and long (12 h) adsorption time confirmed this 213

assumption (Fig. 4A). Fig. 4 and Table 1 show that an extended adsorption time leads to a

214

dramatic boost of the rSAM binding affinity for HA (see also Supporting Information Figs

S4-215

S10). In addition, contrasting with the behaviour in Fig. S3, the binding affinity depended now

216

critically on the ligand density peaking at cE4-SA=0.15. 217

218

219

A. B.

220

Figure 4. HA binding isotherms on mixed rSAMs of E2-OH and E4-SA (A) for different

221

adsorption times (cE4-SA=0.1) and (B) for different mole fractions E4-SA (cE4-SA) after 12h 222

rSAM equilibration.

223 224

Fitting the curve assuming a cooperative Hill equation resulted in a dissociation constant

225

Kdmulti of 1.3 (± 0.1) x 10-13 M, on a par with the best binders reported to date based on higher 226 10-2 10-1 100 101 102 0.0 0.5 1.0 [HA] (nM) Γ (mg m -2) 5000 s 12 h 10-810-710-610-510-410-310-210-1 100 101 0.0 0.5 1.0 [HA] (nM) Γ (mg m -2) 0.05 0.1 0.2 0.15 χE4-SA

(14)

saccharide ligands. The optimal surface coverage matches well with literature observations of

227

both adaptable and non-adaptable sialic acid systems.19-21,23 Questions remained concerning the 228

origin of the strongly enhanced affinity observed.

229

230

Table 1. Dissociation constant (Kd) and surface saturation capacity (Γmax) of HA binding

231

to mixed rSAMs of E4-SA and E2 with 12h adsorption time

232

c

E4-SA Kd (M) Γmax (mg m-2) h

0.05 1.3 (± 0.9) x 10-12 0.22 ± 0.06 0.3 ± 0.1

0.1 3.1 (± 1.4) x 10-11 0.59 ± 0.09 0.6 ± 0.1

0.15 1.3 (± 0.1) x 10-13 0.58 ± 0.15 0.5 ± 0.03

a) Data for cE4-SA=0 and 0.2 have been left out due to weak curvature and poor adherence to the Hill

233

equation.

234

b) The large errors of the data in Fig. 4B are due to a systematic variability of Γmax. This is clearly seen

235

from plots of normalized values (see Fig. S5).

236 237 238

The low amount of adsorbed protein at cE4-SA = 0.2 for rSAMs prepared for both short and long 239

adsorption times (Fig. 4, S3-S4), suggests the lateral distribution of sialic acid head groups to

240

be an important factor for modulating the binding affinity for hemagglutinin (Fig. 5). The

241

drastic drop in binding affinity coincides with the increase in cluster size from 50 nm to 100

242

nm, as seen in the AFM image of the layer formed from cE4-SA = 0.2 (Fig. 3B). It has been 243

reported that formation of larger clusters in ganglioside GM1 containing supported lipid

244

bilayers decreases binding affinity of cholera toxin.12 Hydrogen bonding between the sialic acid 245

head groups was inferred as a likely cause for the segregation of ganglioside, GM3 or GM1 in

246

those membranes.29-30 However, the binding affinity was highest for the lowest ligand density 247

(as low as 0.02% GM1) and then dropped gradually in the interval 0.02% - 10%. This contrasts

248

with the results in Fig.4 and Table 1 where the affinity peaked at an intermediate ligand density

249

whereas it agrees with the gradual decrease in affinity observed for rSAMs prepared with

(15)

shorter adsorption times (Fig. S3, Table S2). Hence, in spite of the similar clustering behaviour

251

in the mixed rSAMs in the former case, this is unlikely to be the main cause for the lack of

252

binding of hemagglutinin at cE4-SA = 0.2. 253

254

255

Study of rSAM dynamics by FRAP: We then turned to investigate the ligand lateral mobility

256

as this was inferred as a key factor accounting for the enhanced rSAM affinity.17 To gain insight 257

into this phenomenon, fluorescence recovery after photobleaching (FRAP) was performed on

258

the layers. Dye doped rSAMs were made by including 1 mol% fluorescein-tagged amidine

E2-259

FAM (Fig. S11) in the mixed amidine solution prior to rSAM formation.

260

After rinsing in HEPES buffer pH 8, the rSAMs were examined by fluorescence

TIRF-261

microscopy in order to follow the recovery process in the bleached regions. First, we observed

262

that a significant portion of the total fluorescence (ca 10-35%) stemmed from immobile

263

fluorophores (Fig. S11-S12). In view of the lack of obvious correlation between the latter and

264

the type of sample (Fig. S12) we attribute this to substrate variability defects in the anchor layer

265 cE4-SA = 0.10 cE4-SA= 0.15 cE4-SA = 0.2 Kd= 3.1 x 10-11M Kd= 1.3 x 10-13M Kd> 10-11M Gmax= 0.1 mg m-2 Gmax= 0.6 mg m-2 Gmax= 0.6 mg m-2

Figure 5. Sialic acid density effect on binding affinity of HA. Schematic illustration of

(16)

leaving room for amidines (including the dye tagged amidine) to enter and bind irreversibly. In

266

this context, we have previously shown that benzamidines have a strong affinity for bare gold.15 267

Assuming this explanation, we decided to calculate the diffusion constant while correcting for

268

the immobile fraction using Eq. 1 (see experimental section) (Fig. 3D). The lowest diffusivity

269

was displayed for the pure E2 rSAM (cE4-SA=0) with a D = 0.51 µm2/s was lower than that of 270

E4-SA rSAMs which appeared more mobile. The highest diffusivity was observed for rSAMs

271

with cE4-SA=0.05 and cE4-SA=0.15. With a D ≈ 1 µm2/s these surfaces display diffusivities in the 272

order of supported lipid bilayers (Fig. S11).33 The results in Table S3 show that the diffusivities 273

of the rSAMs withcE4-SA=0.05-0.15 are significantly higher than for the pure E2 rSAM (c E4-274

SA=0). Most striking is the lack of mobility at cE4-SA=0.20. At this SA density, no recovery 275

was observed indicating all fluorophores to be immobile. The larger clustering observed at this

276

ligand density apparently stiffens the rSAM and inhibits ligand diffusion (Fig. 5). This lack of

277

mobility we believe can explain the concomitant drop in HA affinity and capacity when

278

comparing rSAMs with cE4-SA=0.15 and cE4-SA=0.20. Moreover, the binding affinities (Table 279

1) and rSAM order, inferred from the CH-stretch frequencies (Table S1, vide supra), seem in

280

part to correlate with the ligand mobility. The rSAMs featuring the lowest ligand mobility (c E4-281

SA=0.10 and cE4-SA=0.20) also display the lowest affinities. This explanation contrasts with that 282

of Cremer et al. where the decreasing affinity was attributed to ligand clustering alone.12 283

284

Conclusions

285

We have demonstrated that mixed rSAMs featuring ligand terminated amhiphiles form in a near

286

statistical manner resulting in adaptable surfaces that are optimal for multivalent receptor

287

binding. This leads to a strong boost in overall affinity as here, exemplified by the sialic

acid-288

hemagluttinine ligand receptor pair. The layers are simple to prepare, they demonstrate lateral

(17)

mobility and can be reversibly assembled in response to pH. These features we believe render

290

these systems attractive as models for studies of lipid bilayer membrane dynamics and

291

multivalent molecular recognition.

292

293

Experimental section

294 295

Preparation of amphiphile and protein stock solutions. E4-SA was synthesised as described

296

in the Supporting Information whereas E2-OH, E2-FAM and E2-SA were synthesized as

297

previously reported.13,17 2.5 mM amphipihile stock solutions were prepared in 0 - 5 % ethanolic 298

pH 8 HEPES buffer. Stock solutions of 42 µM hemagglutinin were prepared in Milli-Q water

299

and unused fractions were stored at −80 °C.

300

rSAM preparation. SAMs of mercaptohexadecanoic acid (MHA) on gold were prepared as

301

previously reported17 and was fully immersed into a pH 8 HEPES buffer solution (0.01 M) 302

containing the amidines E4-SA or E2-SA and filler E2, (total concentration=50 µM)at ambient

303

conditions. After 5000 s or 12-18 hrs, the rSAM- modified surfaces were removed from

304

solution, rinsed with pH 8 HEPES buffer and dried under a stream of nitrogen before

305

characterisation by spectroscopic ellipsometry, infrared reflection-adsorption specotroscopy

306

(IRAS) and atomic force microscopy (AFM) measurements. Samples for in situ ellipsometry

307

(ISE) were not dried under a stream of nitrogen and kept wet throughout the entire process.

308

Spectroscopic ellipsometry. Ex situ ellipsometric measurements were taken using a UVISEL

309

HORIBA spectroscopic ellipsometer covering a wavelength range of 200 – 820 nm and

310

incidence angle of 70° at room temperature in air. Optical constants of substrates (MHA) were

311

determined before adsorption of the amphiphiles and each surface was sampled randomly at

3-312

4 points. rSAM thicknesses were modelled based on a homogenous 3-layer model

(18)

rSAM-air) using a Cauchy layer, with an extinction coefficient, k = 0 and an assumed refractive

314

index, n = 1.45 to allow systematic comparison. To verify the accuracy of the ellipsometric

315

measurements, the thickness of MHA on gold was determined. An experimental thickness of

316

19±1 Å corresponded well with literature values.32 317

IRAS. The measurements were made using a Nicolet 6400 instrument, equipped with a liquid

318

nitrogen-cooled MCT-A detector, operating at a resolution of 4 cm-1. Data was collected with 319

a Smart SAGA™ accessory operating at an angle of incidence of 80°. The instrument was

320

purged with nitrogen before and during measurements. Each spectrum is the sum of 512 scans

321

on the modified surfaces using an unreacted, cleaned gold substrate as reference. Each spectrum

322

was processed using OMNIC software and baseline corrected.

323

Atomic Force Microscopy (AFM) Measurements. The surfaces were modified as described

324

above using freshly deposited gold on mica as reported in supporting information. The surfaces

325

were examined with a commercial Atomic Force Microscope (AFM) (MultiMode 8 SPM with

326

a NanoScope V control unit, Bruker AXS) in air at room temperature in PeakForce Tapping®

327

mode. Cantilevers with nominal spring constant 0.5819 N m−1 were employed. Analysis and 328

processing of AFM images were performed using the WSxN 5.0 Develop 8.2.5 Each substrate

329

was scanned randomly at min. 3 points.

330

Fluorescence recovery after photobleaching (FRAP). The rSAMs were prepared as

331

described above with the exception that 1 mol% of fluorescein-tagged amidine (E2-FAM) was

332

added to the amidine solution. After rinsing with HEPES buffer of pH 8, the surfaces were dried

333

using nitrogen gas and concealed in the dark. The rSAMs were again washed in HEPES buffer

334

pH 8 before examination with a fluorescence microscope. Samples were observed using

TIRF-335

microscopy in a Nikon Eclipse Ti fluorescence microscope with a 100x oil objective for all

336

FRAP measurements. The FRAP measurements were done using a 15-mW laser emitting at

(19)

The recovery process in the bleached region was observed by taking images of the sample with

339

five seconds apart. The diffusion constant (D) and immobile fractions (g0) were derived using 340

a MATLAB-programme specially developed for this purpose (Fig. S11-S12, Tables S3).34 341

Compensating for the influence of the immobile fraction the diffusion is given by35: 342

(1) 𝐷 = 𝐷#$%∙'()'+)*

*

343

Where D = diffusion constant corrected for the immobile fraction and Dexp = experimentally 344

measured diffusion constant with FRAP. Applying expression (1) on the above-presented

345

diffusion constants yielded the results shown in Fig. 3D. Performing the Students t-test now

346

gave the values presented in Table S3 showing that there is a significant difference between the

347

diffusion constants of all rSAMs except between cE4-SA = 0.05 and cE4-SA = 0.15. 348

In situ ellipsometry. Saturation plots were obtained by sequential addition of proteins as

349

previously reported.17 Additional amounts of protein were added once the limiting steady state 350

had been attained where adsorbed amount change is <0.035 mg m-2 per 30 data points. 351

Statistical methods. Error bars are standard deviation of observed values unless stated

352

otherwise. Molecular lengths of the compounds were estimated from energy minimized models

353

obtained by molecular mechanics using the MM2 force field (ChemDraw 3D, CambridgeSoft).

354

355

ACKNOWLEDGMENTS

356

This work was supported by grants from the Swedish Research Council (contract number

357

C0296601) and Marie Skłodowska-Curie Actions (H2020-MSCA-IF-2017, 794999 – Y.

358

Sergeeva and H2020-MSCA-IF-2014-EF, 658953 – G. Pan). PJ was supported by a grant from

359

the Swedish Research Council (number: 2018-03872). Giuliana Grasso (Malmö University) is

360

acknowledged for experimental assistance. We are grateful to Javier Soutres and Juan Francisco

(20)

Gonzales (Malmö University) for assistance and helpful discussion concerning the AFM 362 experiments. 363 364 Supporting Information 365

Experimental procedures and data from goniometry, IRAS, ISE, AFM, QCM and FRAP

366

supporting the main findings of the paper.

367

368

Keywords: self-assembled monolayers, carbohydrates, lectins, lipid bilayer, adaptable.

369

370

References

371

1. Delbianco, M.; Bharate, P.; Varela-Aramburu, S.; Seeberger, P. H., Carbohydrates in

372

Supramolecular Chemistry. Chemical Reviews 2016, 116 (4), 1693-1752.

373

2. Kiessling, L. L.; Gestwicki, J. E.; Strong, L. E., Synthetic multivalent ligands in the

374

exploration of cell-surface interactions. Current Opinion in Chemical Biology 2000, 4 (6),

375

696-703.

376

3. Bertozzi, C. R.; Kiessling, L. L., Chemical Glycobiology. Science 2001, 291 (5512),

377

2357-2364.

378

4. Horlacher, T.; Seeberger, P. H., Carbohydrate arrays as tools for research and

379

diagnostics. Chemical Society Reviews 2008, 37 (7), 1414-1422.

380

5. Hushegyi, A.; Tkac, J., Are glycan biosensors an alternative to glycan microarrays?

381

Analytical Methods 2014, 6 (17), 6610-6620.

382

6. Muller, C.; Despras, G.; Lindhorst, T. K., Organizing multivalency in carbohydrate

383

recognition. Chemical Society Reviews 2016, 45 (11), 3275-3302.

(21)

7. Evans, S. V.; MacKenzie, C. R., Characterization of protein–glycolipid recognition at

385

the membrane bilayer. Journal of Molecular Recognition 1999, 12 (3), 155-168.

386

8. Boullanger, P., Amphiphilk carbohydrates as a tool for molecular recognition in

387

organized systems. In Glycoscience Synthesis of Substrate Analogs and Mimetics, Driguez,

388

H.; Thiem, J., Eds. Springer Berlin Heidelberg: Berlin, Heidelberg, 1998; pp 275-312.

389

9. Hyun, J. Y.; Pai, J.; Shin, I., The Glycan Microarray Story from Construction to

390

Applications. Accounts of Chemical Research 2017, 50 (4), 1069-1078.

391

10. Zhu, X. Y.; Holtz, B.; Wang, Y.; Wang, L.-X.; Orndorff, P. E.; Guo, A., Quantitative

392

Glycomics from Fluidic Glycan Microarrays. Journal of the American Chemical Society

393

2009, 131 (38), 13646-13650.

394

11. Jung, H.; Robison, A. D.; Cremer, P. S., Multivalent ligand–receptor binding on

395

supported lipid bilayers. J Struct Biol 2009, 168 (1), 90-94.

396

12. Shi, J.; Yang, T.; Kataoka, S.; Zhang, Y.; Diaz, A. J.; Cremer, P. S., GM1 Clustering

397

Inhibits Cholera Toxin Binding in Supported Phospholipid Membranes. Journal of the

398

American Chemical Society 2007, 129 (18), 5954-5961.

399

13. Yeung, S. Y.; Mucha, A.; Deshmukh, R.; Boutrus, M.; Arnebrant, T.; Sellergren, B.,

400

Reversible Self-Assembled Monolayers (rSAMs): Adaptable surfaces for enhanced

401

multivalent interactions and ultrasensitive virus detection. ACS Central Science 2017, 3 (11),

402

1198-1207.

403

14. Auer, F.; Nelles, G.; Sellergren, B., Odd-Even Chain Length-Dependent Order in

pH-404

Switchable Self-Assembled Layers. Chem. Eur. J. 2004, 10 (13), 3232-3240.

405

15. Auer, F.; Sellergren, B.; Swietlow, A.; Offenhäuser, A. Self-Assembled Layers of

406

Bisbenzamidines on Gold. Langmuir 2000, 16, 5936-5944.

407

16. Sellergren, B.; Swietlow, A.; Arnebrant, T.; Unger, K., Consecutive Selective

408

Adsorption of Pentamidine and Phosphate Biomolecules on a Self-Assembled Layer: 

(22)

Reversible Formation of a Chemically Selective Coating. Analytical Chemistry 1996, 68 (2),

410

402-407.

411

17. Yeung, S. Y.; Ederth, T.; Pan, G.; Cicėnaitė, J.; Cárdenas, M.; Arnebrant, T.;

412

Sellergren, B., Reversible Self-Assembled Monolayers (rSAMs) as Robust and Fluidic Lipid

413

Bilayer Mimics. Langmuir 2018, 34 (13), 4107-4115.

414

18. Bhatia, S.; Camacho, L. C.; Haag, R., Pathogen Inhibition by Multivalent Ligand

415

Architectures. Journal of the American Chemical Society 2016, 138 (28), 8654-8666.

416

19. Kingery-Wood, J. E.; Williams, K. W.; Sigal, G. B.; Whitesides, G. M., The

417

agglutination of erythrocytes by influenza virus is strongly inhibited by liposomes

418

incorporating an analog of sialyl gangliosides. Journal of the American Chemical Society

419

1992, 114 (18), 7303-7305.

420

20. Papp, I.; Sieben, C.; Sisson, A. L.; Kostka, J.; Böttcher, C.; Ludwig, K.; Herrmann,

421

A.; Haag, R., Inhibition of Influenza Virus Activity by Multivalent Glycoarchitectures with

422

Matched Sizes. ChemBioChem 2011, 12 (6), 887-895.

423

21. Spevak, W.; Nagy, J. O.; Charych, D. H.; Schaefer, M. E.; Gilbert, J. H.; Bednarski,

424

M. D., Polymerized liposomes containing C-glycosides of sialic acid: Potent inhibitors of

425

influenza virus in vitro infectivity. Journal of the American Chemical Society 1993, 115 (3),

426

1146-1147.

427

22. Kwon, S.-J.; Na, D. H.; Kwak, J. H.; Douaisi, M.; Zhang, F.; Park, E. J.; Park, J.-H.;

428

Youn, H.; Song, C.-S.; Kane, R. S.; Dordick, J. S.; Lee, K. B.; Linhardt, R. J., Nanostructured

429

glycan architecture is important in the inhibition of influenza A virus infection. Nature

430

Nanotechnology 2016, 12, 48-54.

431

23. Hushegyi, A.; Pihíková, D.; Bertók, T.; Adam, V.; Kizek, R.; Tkac, J., Ultrasensitive

432

detection of influenza viruses with a glycan-based impedimetric biosensor. Biosensors &

433

bioelectronics 2016, 79, 644-649.

(23)

24. Charych, D.; Nagy, J.; Spevak, W.; Bednarski, M., Direct colorimetric detection of a

435

receptor-ligand interaction by a polymerized bilayer assembly. Science 1993, 261 (5121),

436

585-588.

437

25. Fyrner, T.; Lee, H.-H.; Mangone, A.; Ekblad, T.; Pettitt, M. E.; Callow, M. E.;

438

Callow, J. A.; Conlan, S. L.; Mutton, R.; Clare, A. S.; Konradsson, P.; Liedberg, B.; Ederth,

439

T., Saccharide-Functionalized Alkanethiols for Fouling-Resistant Self-Assembled

440

Monolayers: Synthesis, Monolayer Properties, and Antifouling Behavior. Langmuir 2011, 27

441

(24), 15034-15047.

442

26. Lin-Vien, D.; Colthup, N. B.; Fateley, W. G.; Grasselli, J. G., APPENDIX 3 - A

443

Summary of Characteristic Raman and Infrared Frequencies A2 - Lin-Vien, Daimay. In The

444

Handbook of Infrared and Raman Characteristic Frequencies of Organic Molecules,

445

Academic Press: San Diego, 1991; pp 477-490.

446

27. Khajehpour, M.; Dashnau, J. L.; Vanderkooi, J. M., Infrared spectroscopy used to

447

evaluate glycosylation of proteins. Analytical Biochemistry 2006, 348 (1), 40-48.

448

28. Tantakitti, F.; Burk-Rafel, J.; Cheng, F.; Egnatchik, R.; Owen, T.; Hoffman, M.;

449

Weiss, D. N.; Ratner, D. M., Nanoscale clustering of carbohydrate thiols in mixed SAMs on

450

gold. Langmuir 2012, 28 (17), 6950-6959.

451

29. Cantu, L.; Corti, M.; Brocca, P.; Del Favero, E., Structural aspects of

ganglioside-452

containing membranes. Biochimica et Biophysica Acta (BBA) - Biomembranes 2009, 1788

453

(1), 202-208.

454

30. Gu, R.-X.; Ingólfsson, H. I.; de Vries, A. H.; Marrink, S. J.; Tieleman, D. P.,

455

Ganglioside-Lipid and Ganglioside-Protein Interactions Revealed by Coarse-Grained and

456

Atomistic Molecular Dynamics Simulations. The Journal of Physical Chemistry B 2017, 121

457

(15), 3262-3275.

(24)

31. Cerruti, M.; Fissolo, S.; Carraro, C.; Ricciardi, C.; Majumdar, A.; Maboudian, R.

459

Poly(ethylene glycol) Monolayer Formation and Stability on Gold and Silicon Nitride

460

Substrates. Langmuir 2008, 24, 10646-10653.

461

32. Bain, C. D.; Troughton, E. B.; Tao, Y. T.; Evall, J.; Whitesides, G. M.; Nuzzo, R. G.,

462

Journal of the American Chemical Society 1989, 111 (1), 321-335.

463

33. Deng, Y.; Wang, Y.; Holtz, B.; Li, J.; Traaseth, N.; Veglia, G.; Stottrup, B. J.; Elde, R.;

464

Pei, D.; Guo, A.; Zhu, X. Y. Fluidic and Air-Stable Supported Lipid Bilayer and

Cell-465

Mimicking Microarrays. Journal of the American Chemical Society 2008, 130, 6267-6271

466

34. Jönsson, P., M. P. Jonsson, J. O. Tegenfeldt, and F. Hk. 2008. A Method Improving the

467

Accuracy of Fluorescence Recovery After Photobleaching Analysis. Biophys J

95:5334-468

5348." (doi: 10.1529/biophysj.108.134874).

469

35. Jönsson, P.; Jonsson, M. P.; Höök, F. Sealing of Submicrometer Wells by a Shear-Driven

470

Lipid Bilayer. Nano Letters 2010, 10, 1900-1906.

471 472 473

(25)

Table of Contents Graphics

475 476 477 E2-OH MHA O O HS

Sialic Acid Amphiphiles n = 4, E4-SA O O H2N H2N 10 2 O OH N N N O O OH OH HO NH O OH O HO O O H2N H2N 10 n O Hemagglutinin pH 8 Dissociation constant, Kd,multi ΧE4-SA = 0.15 ΧE4-SA = 0.2 1.3 x 10-13 M Minimal binding

Sialic acid-functionalized reversible self-assembled monolayers (rSAMs) with lipid bilayer like fluidity exhibit pM range binding affinity towards hemagglutinin. A pronounced dependence of the affinity on ligand density, clustering and mobility was observed.

Figure

Figure 1. Schematic illustration of preparation and evaluation of sialic acid-functionalized  mixed  reversible  self-assembled  monolayers  (rSAMs)
Figure 2. Surface characterization of mixed rSAMs. A. Real-time change in film thickness  upon exposure of an MHA-SAM to pH 8 HEPES buffer containing 50 µM c E4-SA  = 0, 0.2 or 1
Figure  3.  Atomic  force  microscopy  (AFM)  topography  and  profile  and  fluorescence  recovery after photobleaching (FRAP) of layers formed by  c E4-SA  = 0-0.2 in the adsorption  solution
Figure 4.  HA binding isotherms on mixed rSAMs of E2-OH and E4-SA (A) for different 221
+2

References

Related documents

Diese zweite Inkonsequenz besteht in dem (vor dem Hintergrund dessen, was man vor der Lektüre von Cooks Buch schon über das Märtyrertum weiß, wenig überraschenden)

Denna kvantitativa studie hade som syfte att få svar på hur föräldrar till elever med autism och/eller ADHD i mellanstadieåldern uppfattar att skolan fungerar överlag för deras

Kunskap om språkstörning saknas till stor del bland lärare och dessutom är forskningen kring anpassningar för elever med språkstörning mycket knapphändig.. Därför är det

Two different remediation techniques were evaluated: bioremediation to degrade the organic contaminants with simultaneous monitoring of the inorganic co-contaminants, and soil

Instead of a mixed real and virtual world, the simulator environment and robot are joined to form a combined real world where every move of the user is matched by

Identified requirements on a simulator relating to physical and psychological user aspects are support for unobtrusive and wireless use, high field of view, high

Higher magnification images of samples grown at condition iv, with or without a magnet present (see Fig. 8 for an example), reveal that the individual parti- cles have changed both

The aim of this study was to describe and explore potential consequences for health-related quality of life, well-being and activity level, of having a certified service or