You are on page 1of 14

Article

Cite This: ACS Nano 2018, 12, 6429−6442 www.acsnano.org

Multivalent Flexible Nanogels Exhibit Broad-


Spectrum Antiviral Activity by Blocking Virus
Entry
Pradip Dey,†,§,⊥ Tobias Bergmann,‡,⊥ Jose Luis Cuellar-Camacho,† Svenja Ehrmann,†
Mohammad Suman Chowdhury,† Minze Zhang,‡ Ismail Dahmani,∥ Rainer Haag,*,† and Walid Azab*,‡

Institut für Chemie und Biochemie, Freie Universität Berlin, Takustrasse 3, 14195 Berlin, Germany
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.


Institut für Virologie, Robert von Ostertag-Haus, Zentrum für Infektionsmedizin, Freie Universität Berlin, Robert-von-Ostertag-Str.
7-13, 14163 Berlin, Germany
Downloaded via HUMBOLDT UNIV ZU BERLIN on May 14, 2019 at 14:00:35 (UTC).

§
Polymer Science Unit, Indian Association for the Cultivation of Science, 2A and 2B Raja S.C. Mullick Road, Kolkata 700032, India

Institute of Biochemistry and Biology, University of Potsdam, Karl-Liebknecht-Str. 24-25, 14476 Potsdam, Germany
*
S Supporting Information

ABSTRACT: The entry process of viruses into host cells is


complex and involves stable but transient multivalent
interactions with different cell surface receptors. The initial
contact of several viruses begins with attachment to
heparan sulfate (HS) proteoglycans on the cell surface,
which results in a cascade of events that end up with virus
entry. The development of antiviral agents based on
multivalent interactions to shield virus particles and
block initial interactions with cellular receptors has
attracted attention in antiviral research. Here, we designed
nanogels with different degrees of flexibility based on
dendritic polyglycerol sulfate to mimic cellular HS. The designed nanogels are nontoxic and broad-spectrum, can
multivalently interact with viral glycoproteins, shield virus surfaces, and efficiently block infection. We also visualized
virus−nanogel interactions as well as the uptake of nanogels by the cells through clathrin-mediated endocytosis using
confocal microscopy. As many human viruses attach to the cells through HS moieties, we introduce our flexible nanogels
as robust inhibitors for these viruses.
KEYWORDS: multivalent, herpes simplex virus, heparan sulfate, nanoparticles, click chemistry, polyglycerol

T he parasitic nature and high diversity of viruses pose a


great hurdle in designing antiviral compounds that
could inhibit viral replication across virus families.
Antiviral drugs available to date mainly target specific viral
proteins in order to interrupt the replication cycle, which
Heparan sulfate (HS) moieties are cell surface molecules
that serve as primary receptors or co-receptors for several
microbial pathogens, including bacteria, viruses, and parasites.
It has been shown that HS is involved in the infection of many
viruses, including herpesviruses, arteriviruses, papillomaviruses,
results in reduced replication, giving the immune system time and flaviviruses, through facilitating their internalization or
to control or even clear the infection.1 Targeting early steps of interaction with secondary receptors.5,6 Herpesviruses con-
infection like entry of viral particles into host cells could stitute a large family of viruses that can infect a wide variety of
provide a broader inhibition approach. In general, viruses host species, including humans. Herpes simplex virus type 1
attach to and enter host cells using multivalent interactions of
(HSV-1) is a human alphaherpesvirus with a high seropreva-
viral ligands with receptors and co-receptors present on the cell
lence of 70−80% in adults.7 It primarily infects epithelial cells
membrane.2 The use of monovalent drugs to inhibit viral entry
and consequent spread is not very effective due to the strong and spreads to neurons, where it establishes life-long latent
multivalent virus−cell interactions.3,4 Consequently, there is a infections to evade immune detection. Symptomatic reac-
need for the development of antiviral agents based on tivations can present as orolabial and ocular lesions, herpes
multivalent interactions to efficiently shield virus particles
and inhibit virus−cell interactions and subsequent entry and Received: March 1, 2018
infection. It is challenging, however, to maintain host cell Accepted: June 12, 2018
viability while targeting the virus. Published: June 12, 2018

© 2018 American Chemical Society 6429 DOI: 10.1021/acsnano.8b01616


ACS Nano 2018, 12, 6429−6442
ACS Nano Article

Figure 1. Schematic depiction of flexible and rigid dPGS-based nanogels. Preparation was done using linPG and dPG as cross-linkers,
respectively, by strain-promoted azide alkyne cycloaddition via an inverse nanoprecipitation technique. The scheme shows the structure of
dPG and models of rigid and flexible nanogels. A denotes “upregulation in concentration, leading to cross-linking in the droplets”. B
indicates “addition of water, removal of acetone, and dialysis”. C denotes “nanogels dispersed in water”.

gladiatorum, and eczema herpeticum as well as central nervous carbon nanotubes, carbon dots,34 or nanodiamonds33)
complications such as neonatal herpes and herpetic encepha- functionalized with negatively charged functional groups such
litis.8,9 Several glycoproteins spike the viral lipid envelope and as sulfonate (SO3−),35 sulfate (−OSO3−),33 or hydroxyl
are responsible for attachment, specific receptor binding, and/ (−OH) groups have been used as multivalent inhibitors of
or immune modulation. Herpesvirus particles initially attach to viral entry.31,36,37 To effectively shield and block virus entry,
the cell surface through a charge-based interaction of we have recently shown that sulfated nanoparticles of the same
glycoprotein (gC) and/or gB to heparan sulfate moieties of size as virus particles significantly established multivalent
the extracellular matrix.10 Although relatively nonspecific and interactions and were capable of potently inhibiting virus
reversible, this first interaction is important to set the stage for entry.33,38 However, using carbon nanomaterials in vivo is of
all subsequent processes including stable binding, fusion, and concern considering their toxicity due to reactive oxygen
penetration.11 Absence of HS,12−19 blocking with hepa- species (ROS) generation in the target cells and oxidative
rin,15,18,20−22 or deletion of the gC encoding genes results in stress.39,40 Polymeric nanogels are cross-linked hydrogel
virions that are impaired in their ability to attach to the particles with a 3D network composed of water-soluble/
cell.16,23,24 Currently, most of the herpesvirus therapeutics swellable polymers. They represent a better and promising
include nucleoside, nucleotide, and pyrophosphate analogues choice for the design of virus entry inhibitors41 due to the fact
that target DNA replication. Acyclovir, Valacyclovir, Famcy- that they can be degraded to smaller sized fragments, which
clovir, and other relative drugs have been shown to reduce the can be removed by renal clearance.42 In this context, polymeric
morbidity and mortality associated with encephalitis. However, nanogels functionalized with bioactive ligands may play a
the development of resistance is a major obstacle.11 Previous crucial role in terms of interfering with viral infection.
trials to block virus entry through targeting HS−virus In this report, sulfated nanogels (NGs) based on dendritic
interaction showed promising results.25,26 However, the polyglycerol sulfate (dPGS) are prepared in the size range of
developed compounds stuck with their weak antiviral activity 100−200 nm to match virus size.38 For nanogelation, we used
in vivo.25 No drugs inhibiting viral entry are commercially the inverse nanoprecipitation technique based on a bio-
available so far. orthogonal click reaction. Furthermore, the flexibility (defor-
Another HS-binding virus is the unrelated porcine mation) of these NGs is tuned by using two different types of
reproductive and respiratory syndrome virus (PRRSV) that spacers, that is, dendritic and linear polyglycerols (Figure 1).
has a major impact on the swine industry. It is a single- We hypothesize that flexible NGs (F-NGs) can efficiently
stranded RNA virus that is classified within the family shield the virus particle and consequently are more potent
Arteriviridae.27 The major envelope proteins M and GP5 inhibitors for HS-binding viruses as compared to the rigid ones
form a complex that has been shown to bind to heparin, which (R-NGs). The rationale to use NGs with different deformation
decreases virus infectivity in a dose-dependent manner.28 potential is that flexible NGs can adapt to the virus surface
Until now, mainly dendritic and linear polymers, silver during the binding process, which leads to higher valent
(Ag)29,30 and gold (Au)31 nanoparticles, and functionalized 2D interactions and hence reduces the probability of detachment
nanomaterials (such as thermally reduced graphene oxide,32,33 from the viral surface. To this end, we investigated the
6430 DOI: 10.1021/acsnano.8b01616
ACS Nano 2018, 12, 6429−6442
ACS Nano Article

efficiency and the mechanism of inhibition of virus infection as

−34.3 ± 4.9

−16.5 ± 1.5

−5.1 ± 0.9
ζ-potential
well as the biological fate of NGs in cells.

(mV)
RESULTS AND DISCUSSION
Synthesis and Formation of Sulfated Nanogels Using

176 ± 47

174 ± 52

192 ± 71
Inverse Nanoprecipitation. Dendritic polyglycerol sulfate

(d, nm)
size by
NTA
azide (dPGS-N3), dPG-cyclooctyne (dPG-Oct), and linear
polyglycerol−cyclooctyne (linPG-Oct) were used as macro-
monomers for cross-linking by strain-promoted azide alkyne

0.043

0.107

0.079
from
DLS
cycloaddition reaction to form dPGS nanogels. The synthesis

PDI
of the macromonomers is discussed in the Supporting
Information. Inverse nanoprecipitation was introduced by

202.8 ± 4.1

211.5 ± 2.8

203.4 ± 2.3
our group for the preparation of hydrophilic nanogels as it has

intensity
(d, nm)
size by
several advantages: it is a surfactant-free technique, no
hydrophobic organic solvents are required, and size can be
controlled very easily by varying the good solvent to
nonsolvent ratio.43,44 The R-NGs (sulfated rigid NGs; dPGS-

Z average size

191.3 ± 3.9

191.4 ± 1.4

186.3 ± 0.8
dPG), F-NGs (sulfated flexible NGs; dPGS-linPG), and C-

in water
(d, nm)
NGs (nonsulfated control NGs, dPG-dPG) were formed by
the inverse nanoprecipitation technique,43 where the polymers
were dissolved in water and precipitated in acetone (Figure 1).
Here, R-NGs and F-NGs were the nanogels containing dPGS,

Z average size

not measured
156.2 ± 1.3

98.8 ± 0.06
in acetone
(d, nm)
where R-NG is the rigid particle composed of dendritic
polyglycerol (dPG), and F-NG is the flexible particle
composed of linPG. The dPGS content (Table 1, column 6)
in the polysulfated nanogels was determined from the
elemental analysis of freeze-dried nanogels measuring sulfur
(from CHNS, wt %)
dPGS content

not applicable
(S) content. They were found to be in the range of 40 to 28%
for R-NGs and F-NGs, respectively. Neutral C-NGs (control 40.4

28.4
nanogels) are solely composed of dendritic polyglycerol (dPG-
OH) and prepared by cross-linking of dPG-N3 and dPG-Oct.
Characterization of Nanogels. The particles were
characterized by dynamic light scattering (DLS), ζ-potential,
nanosight tracking analysis (NTA), 1H NMR spectroscopy,
volume of
acetone

200.0

200.0

320.0
(mL)

cryogenic transmission electron microscopy (cryo-TEM), and


atomic force microscopy (AFM). After nanoprecipitation, the
sizes were measured in acetone using DLS. The particle sizes
volume of

are shown in Table 1 (column 7). Afterward, the nanogels


(mL)
water

5.0

5.0

1.0

were transferred into water to quench them by adding excess


water and removing the acetone in vacuo (Figure 1). R-NGs
and F-NGs had different sizes (around 156 and 98 nm,
cyclooctyne/linPG-
cyclooctyne (mg)

respectively) in acetone. However, the hydrodynamic diame-


15.0 (16.0 μmol of

15.0 (6.64 μmol of


wt of dPG-

ters of both nanogels were found to be 191 nm in water with a


cyclooctyne)

cyclooctyne)

cyclooctyne)
2.0 (2.3 μmol

monomodal distribution (Table 1 and Figure 2a). As expected,


the nanogels swelled more in water compared to acetone. The
F-NGs swelled more (from 98 to 191 nm) compared to R-
NGs (from 156 to 191 nm). This can be explained due to the
inherent difference in cross-linkers used in the preparation of
10.0 (5.2 μmol)

10.0 (5.2 μmol)

3.0 (3.52 μmol)


wt of dPGS-N3

R-NGs and F-NGs. The presence of dPGS (in R-NGs and F-


NGs) and linPG (in F-NGs) was confirmed by 1H NMR of the
(mg)

dried nanogels (Supplementary Figures 1 and 2) and elemental


Table 1. Description of Nanogels

analysis. In addition, the size of the particles was measured by


NTA (Figure 2b). They were similar in size as those obtained
from DLS with a monomodal distribution (Figure 2a and
(dPGS-N3−linPG-cyclooctyne)
(dPGS-N3−dPG-cyclooctyne)

(dPG-N3−dPG-cyclooctyne)

Table 1, column 11). As expected, R-NGs and F-NGs both


have shown negative ζ-potential in the range from −15 to −35
mV compared to the control C-NGs (ζ-potential −5 mV;
nanogels

Table 1, column 12). Furthermore, the morphology of


nanogels was visualized by cryo-TEM (Supplementary Figure
3a,b). The diameters of individual nanogels were estimated by
analyzing the images taken by cryo-TEM using ImageJ
C-NGs
R-NGs

F-NGs

software, and histograms were plotted (Supplementary Figure


3c,d). The average sizes of dry R-NGs and F-NGs were found
6431 DOI: 10.1021/acsnano.8b01616
ACS Nano 2018, 12, 6429−6442
ACS Nano Article

technique is that changes in material properties can be


followed while imaging in liquid or buffer conditions even at
desired temperatures. The cationic nature of the lysine layer
allowed the nanogels to adhere on the surface, where spreading
of nanogels was different according to their flexibility, as seen
in the morphology maps (Supplementary Figure 4b,c) and
cross section profiles (Figure 2e,f) when imaged in water. For
R-NGs, the height image shows particles with a quasi-spherical
shape with a clear bumpy morphology. R-NGs had a mean
height of h = 65 ± 18 nm and a mean width of wm = 173 ± 56
nm (n = 74) (Supplementary Figure 4h). Surface topography
and cross-sectional height profiles of F-NGs are shown in
Figure 2d,f. A significant difference in morphology compared
to that of R-NGs (Supplementary Figure 4b,c) is observed for
F-NGs as they clearly spread out more on the substrate, and
their surface presented a more fibrous texture with clear distal
extensions from the center. A zoomed region for these with a
3D reconstruction is given in Supplementary Figure 5, where a
thread-like texture is resolved on their surface, likely generated
by a mesh of linear connecting polymers. Because of this more
enhanced spreading, the height of F-NGs was drastically
reduced to a mean value of h = 35 ± 12 nm with an increase in
measured width of wm = 228 ± 47 nm (Supplementary Figure
4h, n = 74). The pronounced spreading of these flexible
multivalent F-NGs driven by electrostatic interactions
resembles the adherence of cells on flat substrates upon first
contact with a substrate, a process that is mainly controlled by
their elasticity.45,46 This obvious difference in NG spreading
reflects their different degree of flexibility. We showed that the
size of rigid (R-NGs) and flexible (F-NGs) nanogels in
solution determined by DLS and NTA corroborated with each
Figure 2. Characterization of nanogels by DLS, NTA, and QNM other, whereas their height attained after spreading on the
imaging. (a) Size distribution by intensity of all nanogels using substrate was very different and dependent on their flexibility.
DLS. (b) Comparative size distributions obtained using NTA. The extent of spreading of a homogeneous soft particle on a
Morphology map (3D construction) of (c) rigid nanogels (R-NGs) particular substrate will depend on its surface interactions,
and (d) flexible nanogels (F-NGs). Cross sections for single NGs
as shown in the blue dashed lines (Supplementary Figure 4b,c) for interfacial surface energy, and elastic contributions from the
(e) R-NGs and (f) F-NGs. Arrows indicate the edges of NGs after deformed particle. The atomic force microscopy results
spreading on the substrate. (g) Simplified schematic representa- revealed that NGs’ elasticity has a distinctive influence on
tion of AFM photo detection system and tip applying repetitive how spreading of NGs occurred on similar positively charged
Nanoindentations across the scanned region. Here, the external surfaces. To quantify the degree of NG spreading, we
load applied by the AFM tip induces deformation onto the soft considered the ratio A = h/w for the determined values of
nanogels, which can be determined as the difference in separation height h and width w. From the mean values for the height h
distance from the tip initial contact point with the nanogels and its and width w for rigid and flexible NGs, ratios of w = 2.6h for
separation distance at maximal loading force. The nanogels height rigid (R-NGs) and w = 6.2h for flexible NGs (F-NGs) are
is obtained from the difference in separation distance from the
obtained. Imaging of soft nanoparticles with PeakForce QNM
initial contact point on the nanogels surface and the lower region
corresponding in this case to the substrate. (h) Average mode (Supplementary Figure 4a)45−47 allows controlling and
deformation obtained by applying same force on the nanogels. minimizing the maximal applied normal force on the NGs’
Deformations were calculated by measuring 72 NGs, and the bars surface by the AFM tip. This contributes to obtaining accurate
indicate the mean with standard deviation of 72 NGs counted; determination of the NGs’ height but still introduces an
***p < 0.0001 (Kolmogorov−Smirnov normality test followed by overestimation of their width induced by the finite size of the
two-tailed Mann−Whitney test). AFM tip. This contribution to the “true” NGs width (w) is
discussed below in the Methods section.
to be 109 ± 17 and 97 ± 18 nm, respectively. On the contrary, The process of adhesion to substrates induces deformation
the diameters obtained from DLS and NTA, which were in soft nanoparticles. Flexibility of different nanoparticles can
measured in solution (water), were ∼200 nm. These be compared through their degree of deformation induced by
differences can be attributed to the fact that DLS and NTA the AFM tip (Figure 2g). Maps of deformation for R-NGs and
measure the hydrodynamic diameter, whereas the hydration F-NGs are provided (Supplementary Figure 4d,e). The maps
shell is not visualized in cryo-TEM experiments. show regions where the AFM tip penetrated the NGs to
Morphology and Flexibility of Nanogels Using different depths. In the image, brighter regions represent areas
Atomic Force Microscopy. Sulfated nanogels (R-NGs and where the tip penetrated deeper while indenting on the NGs
F-NGs) were attached and imaged on poly-L-lysine-coated (Supplementary Figure 4d,e). On the other hand, the dark
mica substrates using AFM (3D reconstruction, Figure 2c,d; backgrounds are regions scanned on the substrate where the
2D image, Supplementary Figure 4b,c). An advantage of this tip could not easily penetrate because the substrate is harder.
6432 DOI: 10.1021/acsnano.8b01616
ACS Nano 2018, 12, 6429−6442
ACS Nano Article

Figure 3. Cytotoxicity and inhibition capacity of nanogels in Vero E6 cells. (a) Viability assay (WST-1) in Vero E6 cells for mock and
hydrogen peroxide (as negative controls) as well as C-NGs, F-NGs, and R-NGs (each 1 mg/mL). (b) Inhibition of HSV-1 infection at NGs
concentrations of 50, 100, and 200 μg/mL determined by flow cytometry. (c) Inhibition of HSV-1 infection with NGs at 200 μg/mL
determined by flow cytometry (black bars, postinfection; white bars, preinfection) or (d) plaque reduction assay. Postinfection refers to NGs
added to cells that had been infected with HSV-1 1 h before, and preinfection refers to NGs mixed with virus before being applied to the
cells. Virus was used at a multiplicity of infection (MOI) of 1. Normalized mean and standard deviation of three independent experiments
each in duplicates for (a−c). Circles represent individual plaque numbers from three independent experiments normalized to the mean
(with standard deviation) of the control (HSV) (d); *p < 0.01, **p < 0.001, ***p < 0.0001 (Kolmogorov−Smirnov normality test followed
by OneWay ANOVA with Bonferroni post-test for (a−c). OneWay ANOVA with Bonferroni post-test for (d). In (b), stars indicate a
significant difference of F-NGs and R-NGs to the contol C-NGs and between each other).

The profile of penetration on a single particle is given for R- 6). To investigate the effect of the nanogels on virus entry, cells
NGs and F-NGs (Supplementary Figure 4f,g, respectively). were either treated with a mixture of NGs at 200 μg/mL and
Because both samples were imaged with the same type of AFM virus at an MOI of 1 (preinfection) or first infected with HSV-
tips and using the same maximal loading force (500 pN), it was 1 at an MOI of 1 for 1 h, then washed before nanogels were
possible to directly compare their differences in deformability. added (postinfection). In the preinfection setup, the infection
An average penetration depth of 6.3 ± 3.6 nm was obtained for was reduced to less than 5% and around 30% for F-NGs and R-
R-NGs, whereas a 22 ± 7.8 nm was found for F-NGs (Figure NGs, respectively. However, when the cells were infected first,
2h). This represents a 10 and 62% deformation of their height NGs were less or not effective in reducing the infection (Figure
for R-NGs and F-NGs, respectively. This is a significant 3c and Supplementary Figure 6). After 1 h of infection, HSV-1
difference in NGs’ deformation, which validates our chemical is expected to have shuttled its genome into the nucleus of the
approach for the preparation of rigid and flexible NGs with infected cell and started gene expression.49 Therefore, it is
controllable material properties. likely that the NGs exhibit their inhibiting effects at the level of
Nanogels Are Potent Inhibitors of HS-Dependent virus attachment, receptor binding, and/or penetration rather
Viruses. At first, cytotoxicity of the prepared NGs was than at downstream events. We clearly showed that both
assessed on Vero E6 and MARC-145 cells using the WST-1 functionalized NGs inhibited infection of cells with HSV-1
viability assay.48 At a concentration of 1 mg/mL and 200 μg/ with high efficiency at the stage of viral attachment. Flexibility
mL, respectively, none of the NGs significantly reduced cell of the nanogels has a significant impact on inhibition as only
viability (Figure 3a and Supplementary Figure 7a). Next, the around half of the concentration of F-NGs as compared to R-
efficiency of the NGs to inhibit infection of Vero E6 cells with NGs is needed to achieve 50% inhibition. The fact that the
HSV-1 was evaluated. Nanogels in different concentrations infection was strongly reduced when NGs and virus particles
(ranging from 50 to 200 μg/mL) were incubated with HSV-1 can freely interact (preinfection), but less inhibition is seen
[at a multiplicity of infection (MOI) of 1] on a monolayer of when the cells are already infected (postinfection), shows that
Vero E6 cells for 24 h. Infected cells producing virus-encoded inhibition occurs early in the infection cycle. The 25%
green fluorescent protein (GFP)33 were quantified using flow reduction in infection observed in the postinfection experi-
cytometry (Figure 3b). Although the control nanogels (C- ments (Figure 3c) is likely due to the inhibition of egressed
NGs) did not show any inhibition, the two sulfated NGs (F- virus particles that are newly synthesized in the infected cells.
NGs and R-NGs: at a concentration of 200 μg/mL) were Infection of cells with virus at an MOI of 1 statistically yields
active against HSV-1 and have inhibitory effect up to 90 and an infection rate of around 70%. In a 20 h time frame, HSV-1
60% in the case of F-NGs and R-NGs, respectively. The can well undergo an entire replication cycle, producing new
effective concentrations (EC) with 50% inhibition were virions that can infect previously uninfected cells.50 The
calculated to be 90 and 164 μg/mL for F-NGs and R-NGs, presence of NGs in the supernatant inhibits this second round
respectively. Here, the F-NGs were more potent compared to of infection, leading to a smaller reduction in infection
the R-NGs though the concentration of dPGS present in both compared to the controls. Accordingly, when the infection was
nanogel preparations was in a similar range (Table 1, column allowed to proceed for only 6 h, which is an enough time to
6433 DOI: 10.1021/acsnano.8b01616
ACS Nano 2018, 12, 6429−6442
ACS Nano Article

produce GFP but not infectious progenies, no inhibition was (positive-NGs; 214 nm in size) was tested by surface plasmon
detected because no newly formed virus particles had been resonance (SPR). The sensorgrams are divided into three
released yet (Supplementary Figure 6a). Furthermore, in a distinct phases; (i) phase A, interaction of positive-NGs with
multicycle infection (MOI 0.01) over the course of 48 h, both poly-L-lysine (PLY)-coated gold surface; (ii) phase B,
NGs reduced infection to 40−60% as at least two or three interaction of sulfated NGs with positive-NGs, and (iii)
additional rounds of replication had been inhibited (Supple- phase C, dissociation of sulfated NGs (Supplementary Figure
mentary Figure 6b). Preincubation of NGs with cells for 24 h 8). Positive NGs were first injected, and the corresponding
and subsequent infection with HSV-1 (MOI of 1) for 6 h change in SPR response units (RU; 500−700 RU) indicates
confirmed that NGs are active for a long time as they still the immobilization and saturation of the SPR chip with
inhibited the infection to the same extent (Supplementary positive NGs. Sulfated and control NGs were then injected,
Figure 6c). and significant interaction with positive NGs was seen for R-
Similarly, the inhibitory potential of the NGs was confirmed NGs (Req = 7000 ± 1000 RU) and F-NGs (Req = 2000 ±
by plaque reduction assay. Here, NGs at 200 μg/mL were 500 RU). As expected, C-NGs did not bind significantly to
either mixed with virus at 100 PFU (plaque-forming units) positive NGs (Req = 775 RU). The increased signal obtained
incubated with the cells for 1 h, or the cells were infected first for R-NGs compared to that for F-NGs is in line with our
before the NGs were added. To restrict virus cell-to-cell steric shielding and flexibility hypothesis. For the same
spread, the cell monolayer was overlaid with DMEM concentration of NGs, the results suggest that less F-NGs
containing 0.5% methylcellulose. In accordance with the flow are required to cover the chip surface and to reach saturation.
cytometry experiments, F-NGs and R-NGs were able to reduce This is consistent with our AFM results, which showed that F-
plaque numbers by 50% in the preinfection setting, whereas in NGs spread out more (with a larger contact area) once they
the postinfection setup, no significant difference in plaque are bound. Furthermore, due to the lower aspect ratio h/w of
numbers was observed for the NGs preparations (Figure 3d) the adhered F-NGs, they form a thinner layer onto the surface.
because we used an overlay medium, which abolishes a This explains the lower signal obtained with F-NGs compared
subsequent second round of infection by newly made virus to R-NGs that form a thicker layer. The sensorgrams further
particles. show the dissociation of C-NGs and, to lesser extent, of R-NGs
To confirm the broad-range antiviral activity of our NGs, we from the positive NGs; however, this does not stand for F-NGs
conducted preinfection experiments with the unrelated RNA where the height of the signal remains constant. In conclusion,
virus, PRRSV, which uses HS as an attachment receptor.28,51 F- sulfated NGs with a similar high multivalent capacity bind
NGs were able to reduce infectivity of the input virus to significantly to positive surfaces; however, those with enhanced
around 10%. The titer of the virus stock used was determined flexibility resulted in stronger and stable binding.
to be 3.5 × 104 TCID50/mL, and mixing with 200 μg/mL F- While this work was in progress to be submitted, Cagno and
NGs reduced the infectivity to around 4 × 103 TCID50/mL colleagues published results describing the use of gold
(Supplementary Figure 7b). Thus, our sulfated NGs were able nanoparticles (NPs; core size 2−5 nm) decorated with long
not only to inhibit entry of HSV-1 into cells but also to flexible linkers mimicking HS as potent inhibitors of virus
strongly reduce infection with another unrelated, HS-depend- infection at the entry level.52 They have shown that the
ent virus. This finding strengthens the concept of sulfated NGs introduction of flexible linkers in the hard gold nanoparticle
being mimicking and competing with cellular HS in virus core leads to better entry inhibition compared to the use of a
binding. rigid linker, whereas in this report, our aim was to develop
The contact area (CA) of inhibitors is a very important anionic nanoparticles with inherent flexibility and degrad-
parameter while studying the interaction of soft multivalent ability. Furthermore, we have correlated how deformation of
architectures with pathogens. A study on the dependence of nanogels can lead to better steric shielding of virus particles. It
multivalency and steric shielding effects with size for hard is worth noting that the use of hard-core nanoparticles may
multivalent globular inhibitors has been reported recently from lead to a long time accumulation in the body. In line with our
our group.38 Hereby, a robust first approximation to estimate a results, the synthesized gold NPs, mainly, inhibited infection
value for the NGs contact area CA can be made by considering when first incubated with the virus before infection. The claim
NGs’ attachment to be the area of a circle with diameter w, that the sulfonated gold NPs can inhibit viruses after infection
thus CA = π(w/2)2, which leads to 22686 and 38687 nm2 for needs further investigation as the effect was not assessed in the
R-NGs and F-NGs, respectively (Supplementary Table 1). real-time assay. The infected materials (infected cells/super-
Now if we consider the size of HSV-1 particles in solution to natants and nanoparticles) were collected, and virus titration
be around 200 nm and using the calculated values for the was further conducted in another assay.52 It is expected that
average contact area CA of NGs, an estimation of the number the collected nanoparticles in the infected materials can still
(NB) of NGs required to sterically shield the virus can be inhibit the initially added viruses in the solution (not
made. A spherical virus with a diameter of around 200 nm necessarily inside the cells).
would have a surface area of about 125 600 nm2. This Nanogels Interact with HSV-1 and Inhibit Virus
estimated contact areas of nanogels (F-NGs and R-NGs) Attachment to Cells. To better understand the mechanism
would imply that to sterically shield a virus particle, around 5− of inhibition as well as the fate of free NGs, we have
6 R-NGs are needed, whereas only 3 F-NGs would be required introduced a fluorophore into the nanogels by conjugating
(Figure 6 and Supplementary Table 1). This theoretical cyanine dye in the dPGS azide and dPG azide (Supplementary
observation, based on mathematical calculations, matches with Figure 9). Similarly, three nanogels were prepared containing
our experimental findings as discussed above. cyanine dye by inverse nanoprecipitation as discussed above.
To further determine the affinity of sulfated NGs to interact Experiments using confocal microscopy clearly show the
with the virus, binding of sulfated nanogels (R-NGs and F- uptake of the functionalized (sulfated) NGs after 30 min
NGs) with virus-mimicking positively charged nanogels incubation at 37 °C into Vero E6 cells (Figure 4e,f). This
6434 DOI: 10.1021/acsnano.8b01616
ACS Nano 2018, 12, 6429−6442
ACS Nano Article

Figure 4. Nanogels block virus attachment to cell membranes. Representative images of C-NGs as well F-NGs and R-NGs (green) (each at
10 μg/mL concentration). NGs were incubated with Vero E6 cells at 4 °C (0 min; a) or for 30 min at 37 °C (30 min; e) alone (top row) or
with HSV-1 (red) as well as virus alone (bottom row). Scale bars indicate 5 μm. (b−d) Median number of (b) nanogels or (c) virus particles
per cell in total, at the plasma membrane or intracellular, or (d) virus particles colocalizing with NGs for time point 0 min and (e−h) time
point 30 min. Virus was used at a multiplicity of infection (MOI) of 5. NGs and virus signals in 52 cells from three independent experiments
were counted; *p < 0.05, **p < 0.01, ***p < 0.001 (Kruskal−Wallis with Dunns post-test).

uptake can be completely blocked in cells kept continuously at complexes in the supernatant. These effects are highly
4 °C (Figure 4a,b and Supplementary Figure 10). Addition of significant for F-NGs, whereas tendencies with limited
virus particles leads to significantly reduced numbers of both statistical significance were detected for R-NGs. Interestingly,
virus and NGs at the plasma membrane (Figure 4a,c) and NGs were rarely taken up in the presence of virus but
inside the cells (Figure 4e,g). For F-NGs, the effect was more remained at the plasma membrane, and once viral particles had
prominent than for R-NGs, whereas C-NGs were not entered a cell even in the presence of the nanogels, they did
detectable on or inside the cells. The number of virus particles not colocalize with NGs anymore. Both observations can be
and NGs in the supernatant was significantly reduced when explained with the interaction of NGs with viral glycoproteins
added to cells alone, whereas co-presence of both virus and that the virus leaves behind at the plasma membrane when
NGs recovers concentrations in the supernatant to almost entering Vero E6 cells via direct fusion.53
preinfection levels (Supplementary Figure 11). The low Nanogels Enter Cells through Clathrin-Mediated
number of virus particles present at the membrane colocalizes Endocytosis and Route through the Classical Endo-
with almost 90% of F-NGs and with around 40% of R-NGs somal Pathway. To tackle the question of intracellular fate of
(Figure 4d,h). Because HSV-1 uses its positively charged the NGs, subcellular structures were investigated for
glycoproteins to adhere to negatively charged heparan sulfate colocalization with NGs. Internalization of extracellular cargos
on cell surfaces, shielding these glycoproteins by binding to into the cells occurs mainly through the endocytic pathway,
polyglycerol nanogels, providing high densities of negatively which is dependent on scaffolding proteins like clathrin and
charged sulfate groups, could neutralize the ability of the virus caveolin.54 Macropinocytosis is also adopted by the cells to
and NGs to bind to cells, explaining the significantly reduced uptake larger volumes of extracellular fluids.55 No colocaliza-
amount of both on and inside the cells. Consequently, the tion was detected with caveolin-coated pits or macro-
added input material (virus−nanogel complex) was found in pinosomes (Supplementary Figure 12); however, 20−30% of
the supernatant, whereas viruses or NGs alone adsorb to and F-NGs colocalized with clathrin after 10 min incubation at 37
enter the cells, which decreases the amount of virus−nanogel °C with gradually decreasing numbers over the course of 2 h
6435 DOI: 10.1021/acsnano.8b01616
ACS Nano 2018, 12, 6429−6442
ACS Nano Article

Figure 5. Subcellular localization of F-NGs and R-NGs. Representative images of immunostained cells to detect colocalization of F-NGs
(green; 10 μg/mL) after 10 min (top row), 30 min (middle row), and 120 min (bottom row) at 37 °C with (a) clathrin, (b) EEA-1, (c)
RAB7−7, and (d) LAMP-1 (red). Scale bars indicate 5 μm and are 1 μm in the zoomed details. Median portion of F-NGs colocalizing and
noncolocalizing with (e) clathrin, (f) EEA-1, (g) RAB-7, and (h) LAMP-1. Median portion of R-NGs colocalizing and noncolocalizing with
(i) clathrin, (j) EEA-1, (k) RAB-7, and (l) LAMP-1 normalized to the total amount after 10, 30, and 120 min. NGs signals were counted in
52 cells from three independent experiments; ***p < 0.001 (Kruskal−Wallis with Dunns post-test).

(Figure 5a,e). Around the same percent of colocalization was 2 h, 70% of internalized F-NGs colocalized with markers
detected with an early endosome marker (early endosome defining the endosomal−lysosomal pathway, although clearly
antigen 1; EEA-1) (Figure 5b,f) and to a lesser extent with a other, not yet identified, means of uptake are involved. It is
late endosomal marker (Ras-related protein; RAB-7) (Figure worth mentioning that not all sulfated NGs are uptaken by the
5c,g). Detection of lysosomes with the lysosome-associated cells, and they remain active for a long time (24 h) outside the
membrane protein (LAMP)-1 revealed increasing percent of cells (Supplementary Figure 6c). The proposed mechanism of
colocalization with up to 50% after 2 h (Figure 5d,h). R-NGs action is summarized in Figure 6.
behaved much the same way as F-NGs albeit with lower levels To further elucidate the mechanism of uptake, we used
of colocalization (Figure 5i−l). We showed that sulfated NGs pharmacological inhibitors to study various cellular functions.
attach to the cell surface and are internalized at least in part by None of the inhibitors showed significant cytotoxicity at the
clathrin-mediated endocytosis. They follow the classical concentrations used (Supplementary Figure 13a). Vero E6
endosomal pathway accumulating in lysosomal structures as cells were preincubated with chlorpromazine, which inhibits
seen for doxorubicin-loaded polyglycerol-based nanogels.56 assembly of the clathrin adaptor protein 2, thereby preventing
The nonsulfated control C-NGs, however, failed to bind to formation of clathrin-coated pits,57,58 dynasore, an inhibitor of
cells and were only found in the supernatant. Over a course of a cellular GTPase (dynamin) involved in membrane fission
6436 DOI: 10.1021/acsnano.8b01616
ACS Nano 2018, 12, 6429−6442
ACS Nano Article

Figure 6. Mechanism of inhibition. (i) HSV-1 initially attaches to heparan sulfate (HS) proteoglycan on the cell surface through
glycoproteins B (gB) and gC. Specific binding is next mediated by gD, and subsequent fusion releases the nucleocapsid into the host cell. (ii)
Sulfated nanogels (mimicking HS) shield viral particles and inhibit virus binding to cells. (iii) Only around three flexible as opposed to five
rigid nanogels are needed to completely cover HSV-1 particles. (iv) In the absence of virus particles, NGs are taken up by the cell via
clathrin-mediated endocytosis. Passed on to early endosomes, they are shuttled through late endosomes to lysosomes where they
accumulate. In the presence of virus particles, however, NG−virus complexes remain mainly in the supernatant and cannot internalize into
the cell.

and release of vesicles, as well as ethyl isopropyl amiloride allows for an efficient gelation. The different flexibilities of the
(EIPA), which inhibits macropinocytosis by targeting a nanogels were successfully proven by comparing the spreading
sodium-proton exchanger.59 Chlorpromazine was the most of nanogels to a positively charged mica surface. As expected,
potent inhibitor by reducing NG uptake to around 50%, the flexible particles have attained a smaller height and
whereas 25% reduction was seen in cells treated with dynasore increased width compared to those of the rigid ones.
or EIPA. Drugs (filipin or mβCD) targeting cholesterol- Furthermore, the deformation data confirm that the
dependent pathways, such as caveolin-mediated endocytosis,60 application of similar force (500 pN) to the NGs leads to
as well as the tyrosine kinase inhibitor (genistein) had no deformation in flexible NGs higher than that of the rigid ones.
significant effect on NG uptake (Supplementary Figure 13b). Our experimental results showed that globular multivalent
The fact that NG uptake is sensitive to chlorpromazine and inhibitors with different flexibility have a considerable impact
dynasore treatment supports the role of clathrin-dependent on pathogen inhibition through their increase in contact area.
endocytosis as the primary means of internalization. It remains Higher flexibility of these nanogels increases their antiviral
unclear whether the low reduction of uptake caused by EIPA is activity by better shielding the interactions of virus particles
due to the inhibition of macropinocytosis, which would be with cellular surfaces, thereby inhibiting virus attachment to
contradicted by the fact that no colocalization with the cells and subsequent uptake of viral particles. Free NGs,
macropinosome marker dextran was detected. However, however, are endocytosed and routed through the classical
inhibition of the sodium-proton exchanger 3 (NHE3), which endosomal−lysosomal pathway. As many human viruses attach
is the primary target of EIPA, has been shown to block apical to the cells through HS interactions, these flexible multivalent
receptor-mediated endocytosis in kidney cells and to prevent dPGS nanogels have a great potential as robust inhibitors for
cytoskeleton rearrangement;61,62 both can affect clathrin- many virus species.
mediated uptake of NGs.
METHODS
CONCLUSION Chemical Synthesis. All chemicals were reagent grade obtained
from Acros Organics (Geel, Belgium) or Sigma-Aldrich (Munich,
In summary, polyglycerol-based sulfated nanogels with differ- Germany) and used as received without further purification. The
ent flexibility were developed with defined sizes (100−200 SO3−pyridine complex was used as received from Fluka Production
nm) using a bio-orthogonal cross-linking chemistry, which GmbH (Buchs, Switzerland). Reactions sensitive to air or moisture

6437 DOI: 10.1021/acsnano.8b01616


ACS Nano 2018, 12, 6429−6442
ACS Nano Article

were performed under anhydrous conditions in flame-dried glassware. Approximations Applied to Correct Width and Height in
Detailed procedures for the synthesis of all compounds and their AFM. Determination of nanogel size using AFM imaging usually
characterization are provided in the Supporting Information (for 1H considers the height of the measured particle as a more reliable value
spectra of all compounds, Supplementary Methods and Supplemen- than its width because the lateral resolution of the imaged object is
tary Figures 14−20). affected by the AFM tip size and induces an overestimation of the
Preparation of dPGS Nanogels. For R-NGs, dPGS-N3 (10 mg, particles’ “true” width w (Supplementary Figure 21). Geometrical
0.65 μmol) and dPG-cyclooctyne (15 mg, 2.0 μmol) were dissolved models to account for the induced convolution Δ requires knowledge
separately in Milli-Q water (2.5 mL). The solutions were cooled to 4 of the AFM tip radius, opening angle θ, and the size and shape of the
°C, mixed, and added quickly to magnetically stirred acetone at 900 particle being scanned. In the case of R-NGs, an approximation to
rpm (200 mL). Precipitated polyglycerol sulfate nanoparticles were obtain Δ and subsequently to estimate the “true” width w can be
obtained as blue shining dispersions, and the particle size was made considering R-NGs as spherical particles of height h and h >
determined by DLS (Table 1). The reaction was quenched by the radius of tip (rtip). Here, we used the approach reported by Canet-
addition of excess water after keeping the mixture overnight. Acetone Ferrer and co-workers64 and obtained a value of Δ = 2.64 nm for the
was evaporated to obtain blue shining nanogel dispersions in water. overestimation of the width and a true value of w = 170.14 ± 56.29
The nanogels were dialyzed in water for 2 days and concentrated by nm (Supplementary Table 1). A schematic representation of this
evaporation of water. Concentration was determined by lyophilizing approach and additional details to obtain Δ are provided in
the solution. Supplementary Figure 21. For the case of F-NGs, the AFM tip
The other nanogels, F-NGs and C-NGs, were prepared using the could delineate a gradual increase in the NGs’ surface topography
same procedure, and the macromers used are described in Table 1. from the very distal polymer extensions toward the center of the
Similarly, dye-functionalized nanogels were prepared using the dye- particle, as shown in Figure 2d. Although small, the edge of these
functionalized dPGS azide and dPG azide following the procedure distal extensions showed vertical deflections in the range of 1−2 nm,
described above. which is still distinguishable from the substrate background. In this
Nanosight Tracking Analysis and Dynamic Light Scattering. particular situation, Δ was calculated using the approximation applied
The particle size distribution and ζ-potential of the nanogels were by Winzer and co-workers for the case of h > rtip when the AFM tip
measured at a concentration of 1 mg/mL in Milli-Q water or contacts the imaged object at its maximum height.65 Here, the vertical
phosphate-buffered saline (PBS) at 25 °C using a Zetasizer (Malvern height of the edge of F-NGs is about 1.5 nm. With this approximation,
Zetasizer-Nano ZS, Malvern Instruments Limited, Worcestershire, we obtain a value of Δ = 5.52 nm for the tip convolution and of w =
UK) and temperature equilibration for 60 s. The intensity by size 222.04 ± 47.18 nm for a more accurate value of the F-NGs’ width.
distribution and the PDI values were recorded. In addition, the It is worth mentioning that the geometrical approximations we
hydrodynamic diameter of the nanogels was measured by NTA with a used to estimate the tip convolution will hold more accurate for rigid
Nanosight NS500. The concentration used for NTA was 1 μg/mL. objects and could have serious deviations for the case of very soft
Cryo-Transmission Electron Microscopy. Five microliters of particles as in the present case. The complex surface details of the
nanogel solution (0.5 mg/mL) was drop-casted on hydrophilized NGs’ shape or their deformation at the edges as a result from the
carbon film grids (Quantifoil R1/4) at room temperature. Then the scanning process will obviously introduce uncertainties. Furthermore,
excess fluid was discarded using filter paper to generate an ultrathin determination of “true” contact area for particles with a height larger
layer of the solution covering the holes in the carbon film. The grids than the AFM tip could be well overestimated as determination of the
were vitrified in liquid ethane using a standard plunging device and particle shape below its equator is not possible, and the NGs’ contact
stored in liquid nitrogen prior to the measurement. The vitrified angle could strongly vary depending on the surface interactions and
samples were transferred under liquid nitrogen into a Tecnai F20 NGs’ elasticity. Moreover, our robust approximation of the contact
transmission electron microscope (FEI Company, Oregon, USA) area CA takes into account an idealized circular region with diameter
using a Gatan (Gatan Inc., California, USA) cryo-holder and stage w, which again is a simplification of the observed NG attachment.
(model 626). Micrographs were recorded at a sample temperature of Regardless of this, we consider that our approximation of the CA
around 100 K using the microscope’s low-dose protocol at a primary provides a step forward in the characterization of the spreading of
magnification of 29000× and an acceleration voltage of 160 kV. Image multivalent inhibitors with different flexibility.
recording was done using an EAGLE 2k-CCD device (FEI Company, Surface Plasmon Resonance. Positively charged dPG nanogels
Oregon, USA) at full 2048 × 2048 pixel size, and the defocus was (positive-NGs; size determined by DLS was 214 nm, and the ζ-
chosen to be ca. −10 μm in all cases to create enough phase contrast. potential was +24 mV)66 were immobilized on a PLY-coated SPR
Graphical analysis of the images was performed using ImageJ v1.50b. gold sensor chip (XanTec bioanalytics GmbH). Prior to use, the
Atomic Force Microscopy. PeakForce mode was used for surface of the PLY-SPR chip was cleaned by injection of sodium
imaging at high resolution with SNL-10 tips from Bruker (tip radius hydroxide (NaOH; 20 mM, 200 μL) and hydrochloric acid (HCl; 80
2−12 nm and spring constant 0.3 N/m). Before each experiment, mM, 200 μL). Positive NGs (2.5 mg/mL) were first injected at low
cantilevers were first calibrated on cleaved mica surfaces to obtain the flow rate (16 μL/min) to allow their immobilization on the PLY-SPR
cantilever sensitivity, and then the cantilever spring constant was chip. Association with the chip surface and saturation equilibrium
determined using the thermal noise method.63 Briefly, the negatively levels (Req) were monitored for 10 min. The corresponding change
charged nanogels were electrostatically attached to the surface of mica in resonance units was monitored and recorded in the sensorgrams.
coated with the cationic polymer poly-L-lysine. Round discs of The dissociation phase was monitored for 5 min. C-NGs, R-NGs, and
Muscovite mica of about 1 cm in diameter were cleaved with regular F-NGs were injected at a concentration of 1 mg/mL at low flow rate
adhesive tape, and 5 μL of poly-L-lysine (0.01 wt %, MW 70000 (16 μL/min). Regeneration of the surface of the SPR chip was
Sigma-Aldrich) was deposited and left to dry. Then the surface was performed by short injection of 100 mM HCl then 50 mM NaOH.
rinsed several times with Millipore water to remove any free unbound SPR measurements were performed at 25 °C in 0.5× PBS (pH 7.5) in
polymer, and the discs were glued to round metal pucks. Ten a Biacore J SPR analyzer (Biacore AB, Uppsala, Sweden). SPR signals
microliters of a 20-fold diluted solution (concentration ∼3 mg/mL) were presented as a set of sensorgrams, which is a plot of resonance
was deposited on the mica-coated substrates and incubated for only units versus time. All solutions were freshly prepared, degassed, and
10 min. The samples were finally mounted on the AFM head, and a filtered through 0.22 μm pore filters.
fluid cell was assembled. Millipore water was introduced into the Cells and Viruses. African green monkey kidney cells (Vero E6,
sample chamber, and consequently, samples were never allowed to not listed as misidentified by ICLAC) were grown in MEM medium
dry. Scan Assyst was used during imaging, but maximal loading forces (PAN Biotech P04-08050) supplemented with 10% fetal calf serum
were manually set to 500 pN in order to minimize damage on the (FCS, PAN Biotech P30-1402) and 100 U/mL penicillin (Roth
sample surface but still obtain reliable images. Imaging was taken at CN28.3) and 100 μg/mL streptomycin (Alfa Aesar J61299) in a 37
512 points per line and a scan rate of 0.7−0.8 Hz. °C incubator with 5% CO2 atmosphere. Cells were grown to

6438 DOI: 10.1021/acsnano.8b01616


ACS Nano 2018, 12, 6429−6442
ACS Nano Article

confluency, washed with PBS, and detached from the culture vessel acid (Roth 7624.1), 135 mM NaCl (VWR 27868.297), and 10 mM
surface by the addition of 0.25% trypsin supplemented with 2.5 μM KCl (Roth HNO2.3)] for 30 s, neutralized with medium, and washed
EDTA. Cells were counted in a Neubauer counting chamber and again with PBS before they were overlaid with MEM containing 0.5%
seeded at appropriate densities. MARC-145 (simian kidney epithelial methylcellulose (Sigma M0262).69 At 2 days postinfection,
cells derived from MA-104, ATCC CRL-6489) were maintained as fluorescent plaques were counted using an inverted fluorescence
adherent culture in Dulbecco’s modified Eagle’s medium (DMEM, microscope (Zeiss Axiovert 100). As controls, mock-infected cells and
PAN Biotech P04-01548S1) supplemented with 10% FCS, 100 U/mL virus-infected cells in the absence of nanogels were included. All
penicillin, and 100 μg/mL streptomycin at 37 °C in an atmosphere plaques were counted, and the mean of three independent
with 5% CO2. experiments was calculated. Plaque numbers induced by parental
HSV-1 BAC (pYEbac102)67 kindly provided by Dr. Yasushi viruses were set to 100.
Kawaguchi, Institute of Medical Science, The University of Tokyo, TCID50 of PRRSV. A 10-fold serial dilution of PRRSV stock was
was reconstituted after transfecting BAC DNA into Vero E6 cells prepared, mixed with NGs at 200 μg/mL, and added to MARC-145
using Lipofectamine 2000 (Invitrogen). HSV-1 encodes the gene for cells in eight replicates. After 1.5 h, cells were washed with PBS, and
enhanced green fluorescent protein to facilitate detection of infected medium (DMEM supplemented with 2% FCS, 100 U/mL penicillin,
cells. PRRSV strain Lelystad virus (low pathogenic PRRSV-I and 100 μg/mL streptomycin) with NGs at 200 μg/mL was supplied.
prototype strain,68 (GenBank accession number: M96262.2) was After 5 days, individual wells were analyzed for cytopathic effect. Viral
maintained on MARC-145 cells. titers (TCID50/mL) were calculated by the Spearman-Kärber
Cell Viability Assay. Viability of Vero E6 and MARC-145 cells in algorithm.70,71
the presence of nanogels (1 mg/mL and 200 μg/mL, respectively) as Confocal Microscopy. For binding assay, cells were seeded in an
well as pharmacological inhibitors (see below) was determined using 8-well μ-slide (IBIDI) at low densities (1 × 104) and allowed to settle
the WST-1 assay after 20−24 h.48 Briefly, cells were seeded in 96-well for 2 h at 37 °C and 5% CO2. On ice, 1.5 μg (10 μg/mL) of
plates and incubated with NGs for 1 h on ice. Temperature was fluorescent NGs (excited at 560 nm) with or without HSV-1 (1 × 105
shifted to 37 °C, and cells were incubated for 30 min or 20 h with PFU) or virus alone was added to cells. Cells were rested at 4 °C for 1
inhibitors or NGs, respectively. The WST-1 mixture was added h, washed with cold or prewarmed medium, and either kept at 4 °C (0
(Cayman Chemicals 10008883), and after 2 h, absorbance of each min time point) or shifted to 37 °C for 30 min (30 min time point).
sample was read using a microplate reader at 450 nm wavelength. Cells were washed twice with PBS and fixed with 4% paraformalde-
Controls used were untreated cells (mock) and cells treated with 0.1% hyde (PFA) in PBS for 15 min at room temperature. Cells were
H2O2 (Sigma H1009) for 30 s. blocked with blocking buffer [BB; 4% FCS and 1% bovine serum
Flow Cytometry. A mixture of HSV-1 at an MOI of 1 and albumin (VWR 0332) in PBS]. BB containing 0.25% saponin
nanogels (ranging from 50 to 200 μg/mL for determination of EC50, (AppliChem A2542) was used to permeabilize cell membranes for
and at 200 μg/mL for inhibition experiments) were diluted in culture 30 min at 37 °C. Virus particles were detected using anti-gD (1:200 in
medium and added to a confluent monolayer of Vero E6 cells and
BB; Genetex GTX18160) or anti-VP5 (1:200 in BB; Abcam ab6508)
incubated at 37 °C for 20 h (preinfection). Alternatively, cells were
antibodies for the 0 and 30 min time points, respectively, and
infected with HSV-1 at an MOI of 1 for 1 h, and then cells were
incubated for 30 min at 37 °C. Cells were washed twice with PBS and
washed with PBS. NGs (200 μg/mL) were added and incubated for 6
incubated with secondary goat anti-mouse IgG-Alexa Fluor 647
or 20 h at 37 °C. For multicycle experiments, cells were infected with
antibody (Abcam ab150115) or goat anti-rabbit IgG-Alexa Fluor 647
HSV-1 at an MOI of 0.01 for 1 h, washed, and incubated with NGs at
(Abcam ab150079) in a 1:1000 dilution in BB for 30 min at 37 °C.
200 μg/mL for 48 h. For preincubation experiments, cells were
incubated with NGs at 200 μg/mL for 24 h before virus at an MOI of For colocalization experiments, cells were stained with different
1 was added for 6 h. Subsequently, cells were trypsinized and washed antibodies for 1 h at 37 °C to detect colocalization of NGs with
twice with cold PBS. After centrifugation, cells were resuspended in subcellular structures as described above. Antibodies used were
PBS, and 10 000 cells were analyzed with a CytoFLEX flow cytometer anticlathrin (Abcam ab21679), anticaveolin (Abcam ab2910), anti-
(Beckman Coulter) to determine the percentage of infected cells by EEA-1 (Abcam ab2900), anti-RAB-7 (Abcam ab126712), and anti-
fluorescence emission. Infection induced by parental viruses was set to LAMP-1 (Abcam ab24170) in 1:200 dilutions. Dextran-FITC
100%. Flow cytometry data were analyzed with FlowJo software with conjugates (1 μg/mL; Sigma 46945) were used as markers of
equal gate settings according to the controls (Supplementary Figure macropinosomes. Finally, cells were washed three times with PBS, and
22). the nucleus was stained with 2 μg/mL 4′,6-diamidino-2-phenylindole
Pharmacological Inhibitors. Cells were pretreated with different (DAPI, Life Technologies 62247).
drugs for 30 min at 37 °C before addition of fluorescent NGs (50 μg/ To evaluate NGs and virus presence in the supernatants, samples
mL) in the presence of inhibitors. After 30 min incubation on ice, were prepared as described above. Twenty microliters of all samples
cells were washed with prewarmed medium and reincubated at 37 °C before addition to the cells (preinfection) and after 1 h incubation on
for 30 min in the presence of inhibitors. Cells were then trypsinized ice with the cells (postinfection) was dropped on prewet coverslips.
and washed twice with PBS. After centrifugation, cells were The samples were allowed to dry for 2 h at 37 °C and fixed with 4%
resuspended in PBS, and 10 000 cells were analyzed with a CytoFLEX PFA for 15 min. Subsequently, cells were blocked and stained with
flow cytometer (Beckman Coulter) to determine the uptake of NGs. anti-gD antibody and secondary antibody as described above.
The drug concentrations used were 50 μg/mL genistein (Sigma Coverslips were put on standard glass slides using Vectashield with
G6649) dissolved in DMSO, 10 μg/mL chlorpromazine (Sigma DAPI mounting medium (Vector Laboratories).
C8138) in PBS, 5 μg/mL filipin (Sigma 11078−21−0) in DMSO, 20 All samples were analyzed with a VisiScope spinning disc confocal
mM methyl-β-cyclodextrin (MβCD; Sigma 332615) in PBS, 50 μM microscope (Visitron Systems GmbH). Acquisition settings (exposure
5-(N-ethyl-N-isopropyl)amiloride (EIPA; Sigma 1154−25−2) in times, laser voltages) in each channel were kept constant within
ethanol, and 150 μM dynasore (Sigma 1202867-00-2) in DMSO. experiments with small variations between experiments due to
Mock samples were treated with 1% DMSO. Toxicity panels were technical reasons. Images were taken with the VisiView software
performed to ensure that the inhibitors did not cause an adverse effect (Visitron Systems GmbH) and analyzed with ImageJ (ImageJ).
when used at the given concentrations. Brightness and contrast settings were adjusted according to the
Plaque Reduction Assay. Cells were grown in 24-well plates. controls and kept constant within experiments.
Virus suspensions (100 PFU per well) were mixed with the nanogels Statistical Analysis. All viability and flow-cytometry-based
(200 μg/mL), added to the cells, and incubated at 37 °C for 1 h. experiments were blinded before analysis. All experiments involving
Alternatively, cells were infected first with 100 PFU for 1 h, washed, confocal microscopy were carried out in a blinded fashion.
and treated with 200 μg/mL of the NGs for 1 h. Subsequently, cells Normalization was done in Microsoft Excel (Microsoft), and
were washed with PBS, treated with citrate buffer pH 3 [40 mM citric statistical evaluation was performed using GraphPad Prism 5

6439 DOI: 10.1021/acsnano.8b01616


ACS Nano 2018, 12, 6429−6442
ACS Nano Article

(GraphPad software). Values of p less than 0.05 are considered (9) Nadelman, C. M.; Newcomer, V. D. Herpes Simplex Virus
significant. Infections. New Treatment Approaches Make Early Diagnosis even
more Important. Postgrad. Med. 2000, 107, 189−200.
ASSOCIATED CONTENT (10) Laquerre, S.; Argnani, R.; Anderson, D. B.; Zucchini, S.;
Manservigi, R.; Glorioso, J. C. Heparan Sulfate Proteoglycan Binding
*
S Supporting Information
by Herpes Simplex Virus Type 1 Glycoproteins B and C, which Differ
The Supporting Information is available free of charge on the in their Contributions to Virus Attachment, Penetration, and Cell-to-
ACS Publications website at DOI: 10.1021/acsnano.8b01616. Cell Spread. J. Virol. 1998, 72, 6119−6130.
(11) Azab, W.; Osterrieder, K. Initial Contact: The First Steps in
Synthesis procedure and characterization data such as Herpesvirus Entry. Adv. Anat., Embryol. Cell Biol. 2017, 223, 1−27.
1
H NMR, IR, and UV−vis spectra, additional results and (12) Gruenheid, S.; Gatzke, L.; Meadows, H.; Tufaro, F. Herpes
Simplex Virus Infection and Propagation in a Mouse L Cell Mutant
tables (PDF)
Lacking Heparan Sulfate Proteoglycans. J. Virol. 1993, 67, 93−100.
(13) Herold, B. C.; Visalli, R. J.; Susmarski, N.; Brandt, C. R.; Spear,
AUTHOR INFORMATION P. G. Glycoprotein C-Independent Binding of Herpes Simplex Virus
to Cells Requires Cell Surface Heparan Sulphate and Glycoprotein B.
Corresponding Authors
J. Gen. Virol. 1994, 75, 1211−1222.
*E-mail: haag@zedat.fu-berlin.de. (14) Mettenleiter, T. C.; Zsak, L.; Zuckermann, F.; Sugg, N.; Kern,
*E-mail: walid.azab@fu-berlin.de. H.; Ben-Porat, T. Interaction of Glycoprotein gIII with a Cellular
ORCID Heparinlike Substance Mediates Adsorption of Pseudorabies Virus. J.
Pradip Dey: 0000-0002-1302-0874 Virol. 1990, 64, 278−286.
(15) Okazaki, K.; Matsuzaki, T.; Sugahara, Y.; Okada, J.; Hasebe, M.;
Walid Azab: 0000-0003-2880-7255 Iwamura, Y.; Ohnishi, M.; Kanno, T.; Shimizu, M.; Honda, E.; Kong,
Author Contributions Y. BHV-1 Adsorption is Mediated by the Interaction of Glycoprotein
⊥ gIII with Heparinlike Moiety on the Cell Surface. Virology 1991, 181,
P.D. and T.B. contributed equally.
666−670.
Notes (16) Osterrieder, N. Construction and Characterization of an
The authors declare no competing financial interest. Equine Herpesvirus 1 Glycoprotein C Negative Mutant. Virus Res.
1999, 59, 165−177.
ACKNOWLEDGMENTS (17) Shieh, M. T.; WuDunn, D.; Montgomery, R. I.; Esko, J. D.;
Spear, P. G. Cell Surface Receptors for Herpes Simplex Virus are
We thank Florian Paulus for PG synthesis, Cathleen Schlesener
Heparan Sulfate Proteoglycans. J. Cell Biol. 1992, 116, 1273−1281.
and Anja Stöshel for measuring gel permeation chromatog- (18) WuDunn, D.; Spear, P. G. Initial Interaction of Herpes Simplex
raphy and protected linear polyglycerol synthesis, Dr. Mathias Virus with Cells is Binding to Heparan Sulfate. J. Virol. 1989, 63, 52−
Dimde for providing positively charged dPG nanogels, and Dr. 58.
Pamela Winchester for proofreading the manuscript. We (19) Azab, W.; Tsujimura, K.; Maeda, K.; Kobayashi, K.; Mohamed,
especially acknowledge Prof. Klaus Osterrieder, Dr. Sumati Y. M.; Kato, K.; Matsumura, T.; Akashi, H. Glycoprotein C of Equine
Bhatia, and Dr. Dirk Steinhilber for the useful discussions. This Herpesvirus 4 Plays a Role in Viral Binding to Cell Surface Heparan
study was supported by the SFB 765 and the core-facility Sulfate. Virus Res. 2010, 151, 1−9.
biosupramol (www.biosupramol.de) to R.H. This study was (20) Herold, B. C.; Spear, P. G. Neomycin Inhibits Glycoprotein C
partly supported by grant from the DFG to W.A. (AZ 97/3-2). (gC)-Dependent Binding of Herpes Simplex Virus Type 1 to Cells
P.D. thanks DRS-Honors Post-Doctoral Fellowship (2016-1), and also Inhibits Postbinding Events in Entry. Virology 1994, 203,
166−171.
Freie Universität Berlin (https://www.fu-berlin.de/en/sites/
(21) Herold, B. C.; WuDunn, D.; Soltys, N.; Spear, P. G.
drs/postdocs/drs_fellowships/index.html). Glycoprotein C of Herpes Simplex Virus type 1 Plays a Principal
Role in the Adsorption of Virus to Cells and in Infectivity. J. Virol.
REFERENCES 1991, 65, 1090−1098.
(1) De Clercq, E. Antiviral Drugs in Current Clinical Use. J. Clin. (22) Lycke, E.; Johansson, M.; Svennerholm, B.; Lindahl, U. Binding
Virol. 2004, 30, 115−133. of Herpes Simplex Virus to Cellular Heparan Sulphate, an Initial Step
(2) Fields, B.; Knipe, D.; Howley, P.; Griffin, D. Fields Virology, 5th in the Adsorption Process. J. Gen. Virol. 1991, 72, 1131−1137.
ed.; Wolters Kluwer Health/Lippincott Williams & Wilkins: (23) Liang, X. P.; Babiuk, L. A.; van Drunen Littel-van den Hurk, S.;
Philadelphia, 2007. Fitzpatrick, D. R.; Zamb, T. J. Bovine Herpesvirus 1 Attachment to
(3) Bhatia, S.; Camacho, L. C.; Haag, R. Pathogen Inhibition by Permissive Cells is Mediated by its Major Glycoproteins gI, gIII, and
Multivalent Ligand Architectures. J. Am. Chem. Soc. 2016, 138, 8654− gIV. J. Virol. 1991, 65, 1124−1132.
8666. (24) Schreurs, C.; Mettenleiter, T. C.; Zuckermann, F.; Sugg, N.;
(4) Fasting, C.; Schalley, C. A.; Weber, M.; Seitz, O.; Hecht, S.; Ben-Porat, T. Glycoprotein gIII of Pseudorabies Virus is Multifunc-
Koksch, B.; Dernedde, J.; Graf, C.; Knapp, E.-W.; Haag, R. tional. J. Virol. 1988, 62, 2251−2257.
Multivalency as a Chemical Organization and Action Principle. (25) Luscher-Mattli, M. Polyanions–a Lost Chance in the Fight
Angew. Chem., Int. Ed. 2012, 51, 10472−10498. against HIV and Other Virus Diseases? Antiviral Chem. Chemother.
(5) Bartlett, A. H.; Park, P. W. Proteoglycans in Host-Pathogen 2000, 11, 249−259.
Interactions: Molecular Mechanisms and Therapeutic Implications. (26) O’Donnell, C. D.; Shukla, D. The Importance of Heparan
Expert Rev. Mol. Med. 2010, 12, e5. Sulfate in Herpesvirus Infection. Virol. Sin. 2008, 23, 383−393.
(6) Liu, J.; Thorp, S. C. Cell Surface Heparan Sulfate and its Roles in (27) Cavanagh, D. Nidovirales: A New Order Comprising
Assisting Viral Infectious. Med. Res. Rev. 2002, 22, 1−25. Coronaviridae and Arteriviridae. Arch. Virol. 1997, 142, 629−633.
(7) Fatahzadeh, M.; Schwartz, R. A. Human Herpes Simplex Virus (28) Delputte, P. L.; Vanderheijden, N.; Nauwynck, H. J.; Pensaert,
Infections: Epidemiology, Pathogenesis, Symptomatology, Diagnosis, M. B. Involvement of the Matrix Protein in Attachment of Porcine
and Management. J. Am. Acad. Dermatol. 2007, 57, 737−763. Reproductive and Respiratory Syndrome Virus to a Heparinlike
(8) Brady, R. C.; Bernstein, D. I. Treatment of Herpes Simplex Virus Receptor on Porcine Alveolar Macrophages. J. Virol. 2002, 76, 4312−
Infections. Antiviral Res. 2004, 61, 73−81. 4320.

6440 DOI: 10.1021/acsnano.8b01616


ACS Nano 2018, 12, 6429−6442
ACS Nano Article

(29) Baram-Pinto, D.; Shukla, S.; Perkas, N.; Gedanken, A.; Sarid, R. (49) Sodeik, B.; Ebersold, M. W.; Helenius, A. Microtubule-
Inhibition of Herpes Simplex Virus Type 1 Infection by Silver Mediated Transport of Incoming Herpes Simplex Virus 1 Capsids to
Nanoparticles Capped with Mercaptoethane Sulfonate. Bioconjugate the Nucleus. J. Cell Biol. 1997, 136, 1007−1021.
Chem. 2009, 20, 1497−1502. (50) Sandri-Goldin, R. M. Replication of the Herpes Simplex Virus
(30) Hu, R. L.; Li, S. R.; Kong, F. J.; Hou, R. J.; Guan, X. L.; Guo, F. Genome: does it Really Go Around in Circles? Proc. Natl. Acad. Sci. U.
Inhibition Effect of Silver Nanoparticles on Herpes Simplex Virus 2. S. A. 2003, 100, 7428−7429.
GMR, Genet. Mol. Res. 2014, 13, 7022−7028. (51) Delputte, P. L.; Nauwynck, H. J. Porcine Arterivirus Entry in
(31) Baram-Pinto, D.; Shukla, S.; Gedanken, A.; Sarid, R. Inhibition Macrophages: Heparan Sulfate-Mediated Attachment, Sialoadhesin-
of HSV-1 Attachment, Entry, and Cell-to-Cell Spread by Function- Mediated Internalization, and a Cell-Specific Factor Mediating Virus
alized Multivalent Gold Nanoparticles. Small 2010, 6, 1044−1050. Disassembly and Genome Release; Springer: Boston, MA, 2006; pp
(32) Deokar, A. R.; Nagvenkar, A. P.; Kalt, I.; Shani, L.; Yeshurun, 247−252.
Y.; Gedanken, A.; Sarid, R. Graphene-Based “Hot Plate” for the (52) Cagno, V.; Andreozzi, P.; D’Alicarnasso, M.; Jacob Silva, P.;
Capture and Destruction of the Herpes Simplex Virus Type 1. Mueller, M.; Galloux, M.; Le Goffic, R.; Jones, S. T.; Vallino, M.;
Bioconjugate Chem. 2017, 28, 1115−1122. Hodek, J.; Weber, J.; Sen, S.; Janecek, E. R.; Bekdemir, A.; Sanavio, B.;
(33) Ziem, B.; Azab, W.; Gholami, M. F.; Rabe, J. P.; Osterrieder, Martinelli, C.; Donalisio, M.; Rameix Welti, M. A.; Eleouet, J. F.; Han,
N.; Haag, R. Size-Dependent Inhibition of Herpesvirus Cellular Entry Y.; et al. Broad-Spectrum Non-Toxic Antiviral Nanoparticles with a
by Polyvalent Nanoarchitectures. Nanoscale 2017, 9, 3774−3783. Virucidal Inhibition Mechanism. Nat. Mater. 2017, 17, 195−203.
(34) Barras, A.; Pagneux, Q.; Sane, F.; Wang, Q.; Boukherroub, R.; (53) Nicola, A. V.; McEvoy, A. M.; Straus, S. E. Roles for
Hober, D.; Szunerits, S. High Efficiency of Functional Carbon Endocytosis and Low pH in Herpes Simplex Virus Entry into HeLa
Nanodots as Entry Inhibitors of Herpes Simplex Virus Type 1. ACS and Chinese Hamster Ovary Cells. J. Virol. 2003, 77, 5324−5332.
Appl. Mater. Interfaces 2016, 8, 9004−9013. (54) Doherty, G. J.; McMahon, H. T. Mechanisms of Endocytosis.
(35) Sametband, M.; Kalt, I.; Gedanken, A.; Sarid, R. Herpes Annu. Rev. Biochem. 2009, 78, 857−902.
Simplex Virus Type-1 Attachment Inhibition by Functionalized (55) Al Soraj, M.; He, L.; Peynshaert, K.; Cousaert, J.; Vercauteren,
Graphene Oxide. ACS Appl. Mater. Interfaces 2014, 6, 1228−1235. D.; Braeckmans, K.; De Smedt, S. C.; Jones, A. T. siRNA and
(36) Ji, H.; Yang, Z.; Jiang, W.; Geng, C.; Gong, M.; Xiao, H.; Wang, Pharmacological Inhibition of Endocytic Pathways to Characterize the
Z.; Cheng, L. Antiviral Activity of Nano Carbon Fullerene Differential Role of Macropinocytosis and the Actin Cytoskeleton on
Lipidosome against Influenza Virus in vitro. J. Huazhong Univ. Sci. Cellular Uptake of Dextran and Cationic Cell Penetrating Peptides
Technol., Med. Sci. 2008, 28, 243−246. Octaarginine (R8) and HIV-Tat. J. Controlled Release 2012, 161,
(37) Pachota, M.; Klysik, K.; Synowiec, A.; Ciejka, J.; Szczubialka, 132−141.
K.; Pyrc, K.; Nowakowska, M. Inhibition of Herpes Simplex Viruses (56) Zhang, X.; Achazi, K.; Steinhilber, D.; Kratz, F.; Dernedde, J.;
Haag, R. A Facile Approach for Dual-Responsive Prodrug Nanogels
by Cationic Dextran Derivatives. J. Med. Chem. 2017, 60, 8620−8630.
(38) Vonnemann, J.; Liese, S.; Kuehne, C.; Ludwig, K.; Dernedde, J.; Based on Dendritic Polyglycerols with Minimal Leaching. J. Controlled
Release 2014, 174, 209−216.
Böttcher, C.; Netz, R. R.; Haag, R. Size Dependence of Steric
(57) Azab, W.; Lehmann, M. J.; Osterrieder, N. Glycoprotein H and
Shielding and Multivalency Effects for Globular Binding Inhibitors. J.
alpha4beta1 Integrins Determine the Entry Pathway of Alphaherpes-
Am. Chem. Soc. 2015, 137, 2572−2579.
viruses. J. Virol. 2013, 87, 5937−5948.
(39) Jastrzębska, A. M.; Kurtycz, P.; Olszyna, A. R. Recent Advances
(58) Wang, L. H.; Rothberg, K. G.; Anderson, R. G. Mis-Assembly
in Graphene Family Materials Toxicity Investigations. J. Nanopart.
of Clathrin Lattices on Endosomes Reveals a Regulatory Switch for
Res. 2012, 14, 1320−1341.
Coated Pit Formation. J. Cell Biol. 1993, 123, 1107−1117.
(40) Oberdörster, G.; Oberdörster, E.; Oberdörster, J. Nano- (59) Harris, C.; Fliegel, L. Amiloride and the Na(+)/H(+)
toxicology: An Emerging Discipline Evolving from Studies of Exchanger Protein: Mechanism and Significance of Inhibition of the
Ultrafine Particles. Environ. Health Perspect. 2005, 113, 823−839. Na(+)/H(+) Exchanger. Int. J. Mol. Med. 1999, 3, 315−321.
(41) Papp, I.; Sieben, C.; Sisson, A. L.; Kostka, J.; Böttcher, C.; (60) Orlandi, P. A.; Fishman, P. H. Filipin-Dependent Inhibition of
Ludwig, K.; Herrmann, A.; Haag, R. Inhibition of Influenza Virus Cholera Toxin: Evidence for Toxin Internalization and Activation
Activity by Multivalent Glycoarchitectures with Matched Sizes. through Caveolae-like Domains. J. Cell Biol. 1998, 141, 905−915.
ChemBioChem 2011, 12, 887−895. (61) Gekle, M.; Freudinger, R.; Mildenberger, S. Inhibition of Na
(42) Steinhilber, D.; Sisson, A. L.; Mangoldt, D.; Welker, P.; Licha, +-H+ Exchanger-3 Interferes with Apical Receptor-Mediated
K.; Haag, R. Synthesis, Reductive Cleavage, and Cellular Interaction Endocytosis via Vesicle Fusion. J. Physiol. 2001, 531, 619−629.
Studies of Biodegradable, Polyglycerol Nanogels. Adv. Funct. Mater. (62) Turner, J. R.; Black, E. D.; Ward, J.; Tse, C. M.; Uchwat, F. A.;
2010, 20, 4133−4138. Alli, H. A.; Donowitz, M.; Madara, J. L.; Angle, J. M. Transepithelial
(43) Steinhilber, S.; Rodriguez, J. M.; Estrada, C. A.; Kraemer, R. R. Resistance can be Regulated by the Intestinal Brush-Border Na+/H+
A 76-year-old Woman with Diaphoresis and Anxiety. J. Gen. Int. Med. Exchanger NHE3. Am. J. Physiol.: Cell Physiol. 2000, 279, C1918−
2013, 28, 1376−1380. C1924.
(44) Zhang, X.; Malhotra, S.; Molina, M.; Haag, R. Micro- and (63) Sader, J. E.; Chon, J. W. M.; Mulvaney, P. Calibration of
Nanogels with Labile Crosslinks - from Synthesis to Biomedical Rectangular Atomic Force Microscope Cantilevers. Rev. Sci. Instrum.
Applications. Chem. Soc. Rev. 2015, 44, 1948−1973. 1999, 70, 3967−3969.
(45) Cuvelier, D.; Thery, M.; Chu, Y. S.; Dufour, S.; Thiery, J. P.; (64) Canet-Ferrer, J.; Coronado, E.; Forment-Aliaga, A.; Pinilla-
Bornens, M.; Nassoy, P.; Mahadevan, L. The Universal Dynamics of Cienfuegos, E. Correction of the Tip Convolution Effects in the
Cell Spreading. Curr. Biol. 2007, 17, 694−699. Imaging of Nanostructures Studied through Scanning Force
(46) Odenthal, T.; Smeets, B.; Van Liedekerke, P.; Tijskens, E.; Van Microscopy. Nanotechnology 2014, 25, 395703−395711.
Oosterwyck, H.; Ramon, H. Analysis of Initial Cell Spreading Using (65) Winzer, A. T.; Kraft, C.; Bhushan, S.; Stepanenko, V.; Tessmer,
Mechanistic Contact Formulations for a Deformable Cell Model. I. Correcting for AFM Tip Induced Topography Convolutions in
PLoS Comput. Biol. 2013, 9, e1003267. Protein−DNA Samples. Ultramicroscopy 2012, 121, 8−15.
(47) Zhao, B.; Song, Y.; Wang, S.; Dai, B.; Zhang, L.; Dong, Y.; Lu, (66) Dimde, M.; Neumann, F.; Reisbeck, F.; Ehrmann, S.; Cuellar-
J.; Hu, J. Mechanical Mapping of Nanobubbles by PeakForce Atomic Camacho, J. L.; Steinhilber, D.; Ma, N.; Haag, R. Defined pH-
Force Microscopy. Soft Matter 2013, 9, 8837−8843. sensitive Nanogels as Gene Delivery Platform for siRNA Mediated in
(48) Segawa, K.; Suzuki, J.; Nagata, S. Constitutive Exposure of vitro Gene Silencing. Biomater. Sci. 2017, 5, 2328−2336.
Phosphatidylserine on Viable Cells. Proc. Natl. Acad. Sci. U. S. A. (67) Tanaka, M.; Kagawa, H.; Yamanashi, Y.; Sata, T.; Kawaguchi, Y.
2011, 108, 19246−19251. Construction of an Excisable Bacterial Artificial Chromosome

6441 DOI: 10.1021/acsnano.8b01616


ACS Nano 2018, 12, 6429−6442
ACS Nano Article

Containing a Full-Length Infectious Clone of Herpes Simplex Virus


Type 1: Viruses Reconstituted from the Clone Exhibit Wild-Type
Properties in vitro and in vivo. J. Virol. 2003, 77, 1382−1391.
(68) Meulenberg, J. J.; Hulst, M. M.; de Meijer, E. J.; Moonen, P. L.;
den Besten, A.; de Kluyver, E. P.; Wensvoort, G.; Moormann, R. J.
Lelystad Virus, the Causative Agent of Porcine Epidemic Abortion
and Respiratory Syndrome (PEARS), is Related to LDV and EAV.
Virology 1993, 192, 62−72.
(69) Azab, W.; Osterrieder, N. Glycoproteins D of Equine
Herpesvirus type 1 (EHV-1) and EHV-4 Determine Cellular Tropism
Independently of Integrins. J. Virol. 2012, 86, 2031−2044.
(70) Kärber, G. Beitrag zur Kollektiven Behandlung Pharmakolo-
gischer Reihenversuche. Naunyn-Schmiedeberg's Arch. Pharmacol.
1931, 162, 480−483.
(71) Spearman, C. The Method of ‘Right and Wrong Cases’
(‘Constant Stimuli’) Without Gauss’s Formulae. Br. J. Med. Psychol.
1908, 2, 227−242.

6442 DOI: 10.1021/acsnano.8b01616


ACS Nano 2018, 12, 6429−6442

You might also like