You are on page 1of 8

Modifications in the nanoparticle-protein

interactions for tuning the protein


adsorption and controlling the stability of
complexes
Cite as: Appl. Phys. Lett. 118, 153701 (2021); https://doi.org/10.1063/5.0046745
Submitted: 06 February 2021 . Accepted: 21 March 2021 . Published Online: 13 April 2021

Sugam Kumar, Debasish Saha, Shin-ichi Takata, Vinod K. Aswal, and Hideki Seto

ARTICLES YOU MAY BE INTERESTED IN

Electric-field effect on magnetic moments in Co ultra-thin films deposited on Pt


Applied Physics Letters 118, 152405 (2021); https://doi.org/10.1063/5.0049052

Axis-dependent carrier polarity in polycrystalline NaSn2As2


Applied Physics Letters 118, 153903 (2021); https://doi.org/10.1063/5.0047469

Substrate-controlled dynamics of spin qubits in low dimensional van der Waals materials
Applied Physics Letters 118, 154003 (2021); https://doi.org/10.1063/5.0048399

Appl. Phys. Lett. 118, 153701 (2021); https://doi.org/10.1063/5.0046745 118, 153701

© 2021 Author(s).
Applied Physics Letters ARTICLE scitation.org/journal/apl

Modifications in the nanoparticle-protein


interactions for tuning the protein adsorption
and controlling the stability of complexes
Cite as: Appl. Phys. Lett. 118, 153701 (2021); doi: 10.1063/5.0046745
Submitted: 6 February 2021 . Accepted: 21 March 2021 .
Published Online: 13 April 2021

Sugam Kumar,1 Debasish Saha,1 Shin-ichi Takata,2 Vinod K. Aswal,1,3,a) and Hideki Seto4

AFFILIATIONS
1
Solid State Physics Division, Bhabha Atomic Research Centre, Mumbai 400 085, India
2
J-PARC Center, Japan Atomic Energy Agency, Tokai, Ibaraki 319-1195, Japan
3
Homi Bhabha National Institute, Mumbai 400 094, India
4
J-PARC Center, High Energy Accelerator Research Organization, Tokai, Ibaraki 319-1106, Japan

a)
Author to whom correspondence should be addressed: vkaswal@barc.gov.in

ABSTRACT
We report the pathways to suppress or enhance the protein adsorption on nanoparticles and thereby control the stability of the
nanoparticle-protein complexes with the help of selective additives. This has been achieved by tuning the electrostatic interaction between
the nanoparticles and proteins, in the presence of surfactant and multivalent counterions. The preferential binding of the proteins with the
surfactant and multivalent ions induced charge reversibility of nanoparticles can lead to adsorption of an otherwise non-adsorbing protein
and vice versa. The findings are demonstrated for anionic silica nanoparticles and two globular proteins [lysozyme (cationic) and bovine
serum albumin (BSA) (anionic)] as model systems, in the presence of two ionic surfactants [anionic sodium dodecyl sulfate (SDS) and cat-
ionic dodecyltrimethylammonium bromide (DTAB)], and ZrCl4 as multivalent salt. Small-angle neutron scattering with the unique advan-
tage of contrast variation has been used to probe the role of individual components in the multi-component system. It is shown that the
non-adsorbing behavior of BSA with silica nanoparticles changes into adsorbing in the presence of oppositely charged DTAB surfactant,
whereas the strong adsorbing behavior of lysozyme on nanoparticles modifies to be non-adsorbing in the presence of oppositely charged SDS
surfactant. The presence of multivalent counterions (ZrCl4) leads the charge reversal of the nanoparticles, transforming the lysozyme from
adsorbing to non-adsorbing, and no significant change in the behavior of BSA. The results presented can find potential applications in the
field of nanobiotechnology.
Published under license by AIP Publishing. https://doi.org/10.1063/5.0046745

The integration of nanoparticles with proteins has a prime inter- the conjugate and its response toward the biological target is governed
est in the field of nanobiotechnology, where these complexes are aimed by the adsorbed layer of the proteins.1,10,11 The adsorption of human
to be utilized in different applications such as targeted drug delivery serum protein on iron oxide or calcium carbonate nanoparticles is
and biosensing.1–5 The proteins have the tendency to immediately shown to be improving the uptake of the complex by cancerous cells
cover the nanoparticles in the biological milieu to form the protein in comparison to bare nanoparticles.12,13 It has also been shown that
corona (a tightly packed layer of protein molecules) around the nano- the interaction of gold nanoparticles with different cells is decided by
particles. However, the uncontrolled formation of a protein corona the type of the protein corona.11,14
around nanoparticles can limit applications due to prompt clearance However, despite the tendency of the protein to adsorb on nano-
from the systemic circulation by macrophages, invoking the need of particles, in some cases the repulsive interactions between nanopar-
controlled and selective protein adsorption.1,6,7 ticles and proteins (e.g., electrostatic repulsion between similarly
The protein adsorption on nanoparticles is governed by several charged nanoparticles and proteins) are so strong that they do not
interactions, such as hydrogen bonding, electrostatic complexion, allow the protein adsorption.15 For example, the adsorption of the
covalent bonding, and hydrophobic attraction, depending on the phys- serum protein on the silver nanoparticles is suppressed by introducing
ical nature of the nanoparticles and protein.8–10 The functionality of additional steric repulsion between nanoparticles and serum proteins

Appl. Phys. Lett. 118, 153701 (2021); doi: 10.1063/5.0046745 118, 153701-1
Published under license by AIP Publishing
Applied Physics Letters ARTICLE scitation.org/journal/apl

for reducing the cellular uptake of the nanoparticles and its cytotoxic- component systems (like in the present study) by suppressing the scat-
ity.16 The tunability between adsorption and non-adsorption with tering contribution of others.26,27
respect to some physicochemical parameters can enable the unique The ability of contrast variation in SANS originates due to the
advantages that can be exploited for various purposes.1,10,11,17,18 For presence of the contrast term (qp – qs)2 in the equation for the intensity
example, the idea of utilizing these complexes as carriers for targeted of the scattered neutrons, which can be expressed in terms of differen-
drug delivery requires both the adsorption as well as release of the pro- tial scattering cross section as27,28 dX
dR
ðQÞ ¼ UVðqp  qs Þ2 PðQÞSðQÞ
tein (biomolecules, in general) at different instances.17,18 Similarly, in þB, where Q [¼ (4p sinh)/k, 2h is the scattering angle, and k is the
an interesting study, adsorption of fusion protein on silica nanopar- mean wavelength of neutrons] is the scattering vector, U is the volume
ticles has been utilized as a shield to suppress the adsorption of serum fraction of scatterers, and V is the volume of the individual scatterer. qp
proteins. The MD simulations carried out in this study suggest that and qs are scattering length densities of scatterer and solvent, respec-
the electrostatic interaction between the shield protein coated nano- tively. If qs is equal to the qp for any component (e.g., nanoparticle) in
particles and serum proteins is an important factor in minimizing the the multi-component system (e.g., nanoparticle-protein system), the
serum proteins’ adsorption.7 scattering from that component will have no contribution, and hence
The interaction between the nanoparticles and protein also dic- the SANS data will only be governed by the scattering from the other
tates the stability of the system, which is another important parameter component (e.g., protein).26,27 In the present study, the scattering
for the application of the nanoparticle-protein complexes.15,19 In the length density of the solvent is varied by mixing the suitable amount of
case of oppositely charged nanoparticles and protein, the strong elec- H2O (qH2O ¼ 0.56  1010/cm2) and D2O (qD2O ¼ 6.38  1010/cm2)
trostatic attraction leads to the protein-mediated aggregation of the to match them with the scattering length density of silica nanoparticles
nanoparticles.15 The non-adsorption of the protein on the nanopar- (qsilica ¼ 3.80  1010/cm2).26 P(Q) is intraparticle structure factor and
ticles may also give rise to nanoparticle aggregation via protein- S(Q) is interparticle structure factor. B is a constant term representing
induced depletion attraction between nanoparticles.20 The aggregation incoherent background. P(Q) depends on shape and size of the scat-
in the nanoparticle-protein systems may be useful in some instances terer, whereas S(Q) provides the information on the interparticle inter-
such as to generate the multi-scale equilibrium clusters,21 however, actions. It may be mentioned here that the use of D2O instead of the
most of the time it is undesired as it hinders the interaction of the H2O is known to give rise some minor effects on protein-solvent inter-
complex with the biological environment. Nevertheless, the interac- actions. However, in the present study, no significant effect of the H2O
tions between nanoparticle and protein may be tuned to gain control or D2O or H2O/D2O mixed solvent was observed on the nanoparticle-
over the protein adsorption on the nanoparticles as well as colloidal protein interactions, as confirmed by DLS.29,30
stability of the resulting complexes.1,10,22,23 SANS measurements were performed at the small and wide-
In this Letter, we have tuned the interaction between nanopar- angle neutron scattering instrument (TAIKAN) at J-PARC, Japan.31
ticles and protein to control the adsorption/desorption of proteins on The details of the experiments and data analysis can be found in the
nanoparticles and colloidal stability of the system using oppositely supplementary material. Figure 1 shows the SANS data of 1 wt. % sil-
charged ionic surfactants and multivalent counterions. Anionic silica ica nanoparticles (polydisperse sphere, mean radius ¼ 8.0 nm, supple-
(SiO2) nanoparticles [Ludox HS40 (diameter 16 nm and zeta poten- mentary Fig. S3 and supplementary Table S1)26 with 1 wt. % lysozyme
tial 60 mV), 40 wt. % dispersion in water] and two globular pro- protein (ellipsoidal shape, 2.4  1.4  1.4 nm3, supplementary Fig. S3
teins [anionic bovine serum albumin (BSA) and cationic lysozyme], and supplementary Table S1)32 in the absence [Fig. 1(a)] and presence
which are oppositely charged at pH 7.0 were used as model systems [Figs. 1(b) and 1(c)] of SDS surfactant, in the two contrast conditions,
(Fig. S1 in the supplementary material). The silica nanoparticles are where (i) all the components are visible (solvent is 100% D2O)
one of the most widely studied model nanoparticles in the field of [Fig. 1(b)] and (ii) nanoparticles are contrast-matched (solvent is 62%
nanobiotechnology due to their several relevant properties such as low D2O in H2O/D2O mixture) [Fig. 1(c)]. The interaction of the anionic
toxicity, high stability, good compatibilities with other materials, and silica nanoparticles with the cationic lysozyme protein (in the absence
ability to be functionalized with a range of macromolecules.24,25 The of the SDS) leads to the protein-mediated aggregation of the nanopar-
lysozyme protein adsorbs on the silica nanoparticles due to strong ticles. This is evident as the sample becomes turbid (supplementary
electrostatic attraction, resulting in protein-mediated nanoparticle Fig. S4), and the SANS data [Fig. 1(a)] show a linear rise in the
aggregation. On the other hand, BSA protein experiences electrostatic scattering intensity in the low Q region due to the formation of the
repulsion with nanoparticles and hence remains non-adsorbing. The fractal-like aggregates.32,33 The data are analyzed by employing P(Q)
non-adsorbing nature of the BSA gives rise to depletion attraction of core-shell for the building block (nanoparticle core with adsorbed
between the nanoparticles, which again causes particle aggregation but protein shell) and S(Q) for the mass fractal-like aggregation (see the
at significantly higher concentration, compared to lysozyme.15 Two supplementary material). On the other hand, SDS surfactant does not
surfactants (Fig. S2 in the supplementary material), namely, anionic have any physical interaction with the nanoparticles due to similar
sodium dodecyl sulfate [SDS: CH3(CH2)11SO4Na] and cationic dode- charge of the two components (supplementary Fig. S5).26 The addition
cyltrimethylammonium bromide [DTAB: CH3(CH2)11N(CH3)4Br] of 5 mM SDS, in 1 wt. % HS40 þ 1 wt. % lysozyme system, leads to the
were used as both the surfactants have strong interactions with pro- increase in scattering intensity compared to that in the absence of
teins, and the results are supported by dynamic light scattering (DLS) SDS, without changing the scattering features [Fig. 1(b)]. The physical
measurements. Small-angle neutron scattering (SANS) has been state of the sample also remains turbid (supplementary Fig. S4). The
employed to investigate the nanoparticle-protein systems with suitable data hence in this case (1 wt. % HS40 þ 1 wt. % lysozyme þ 5 mM
contrast conditions. The unique advantage of contrast variation ena- SDS) are analyzed using the same model, which provides much larger
bles to probe the role of individual component in such multi- shell thickness than that obtained for the 1 wt. % HS40 þ 1 wt. %

Appl. Phys. Lett. 118, 153701 (2021); doi: 10.1063/5.0046745 118, 153701-2
Published under license by AIP Publishing
Applied Physics Letters ARTICLE scitation.org/journal/apl

FIG. 1. (a) SANS data of 1 wt. % HS40 þ 1 wt. % lysozyme along with pure components, (b) SANS data of 1 wt. % HS40 þ 1 wt. % lysozyme with varying concentration of
SDS in 100% D2O (all components are visible), (c) in 62% D2O in D2O/H2O (nanoparticles are contrast-matched), and (d) SANS data of 1 wt. % lysozyme with varying SDS
concentrations.

lysozyme system (supplementary Table S2). This may happen if the the disappearance of the linearity in the scattering profile. In fact, the
nanoparticle aggregation is driven by the lysozyme-SDS complex data have typical features of the two-form factors governed scattering
instead of the lysozyme alone. The lysozyme protein also interacts contributions from two different sizes [Fig. 1(b)]. To understand the
strongly with the SDS, since lysozyme and SDS are oppositely charged. system behavior, we have contrast-matched silica nanoparticles with
The lysozyme-SDS complex forms hexagonally closed packed (HCP) the solvent [Fig. 1(c)]. In the absence of the SDS, the scattering inten-
fiber structures at the mentioned SDS to lysozyme molar ratio sity is quite weak of aggregated core-shell structure, as the contrast-
[Fig. 1(d) and supplementary Table S2].34 The lysozyme protein and match points of the silica nanoparticles and the lysozyme protein are
SDS are known to form different structures at the different molar ratio nearly same.38 On addition of 5 mM SDS, the data show strong linear
of the lysozyme and SDS.35 When SDS is added in the scattering with a small oscillation at Q 0.035 Å1, which can be
HS40 þ lysozyme system, the lysozyme preferably binds with the SDS, understood from a fractal-like structure formed by the shell-like build-
which may be due to the combined dominating effect of the hydro- ing blocks (adsorbed protein shell with nanoparticle core contrast
phobic and electrostatic interactions in the protein-surfactant system matched). The shell thickness in this case is consistent with that
in comparison to electrostatic attraction alone in the nanoparticle- obtained in the no contrast-match condition [Fig. 1(b)]. However, at
protein system.36 The complex is then attached to the nanoparticle 15 mM SDS concentration the scattering profile shows typical features
surface and leads to the complex-mediated nanoparticle aggregation. of the pure SDS micelles, suggesting the coexistence of SDS micelles
It should be noted that about 6 mM SDS is required to neutralize all with nanoparticles. To understand whether lysozyme interacts with
the molecules of the lysozyme (charge þ 8e.u.) in 1 wt. % solution, SDS or nanoparticles or both, SANS data of 1% lysozyme with 15 mM
which implies that at lower SDS concentration (5 mM) the lysozyme- SDS are examined [Fig. 1(d)]. The data of 1 wt. % lysozyme þ 15 mM
SDS complex still carry positive charge, and therefore is attracted SDS show typical features of the SDS micelles, suggesting the forma-
toward the anionic silica nanoparticles.37 On the contrary, the physical tion of transparent and soluble complexes of SDS micelles with
state of the solution becomes clear, with no signature of the nanoparti- unfolded lysozyme at this concentration.35 These complexes carry an
cle aggregation, on raising the SDS concentration to 15 mM. In this overall negative charge and comprises a decorated core-shell structure,
case, the scattering intensity in the low Q region decreases along with where the core is made-up of dodecyl chains while the shell is formed

Appl. Phys. Lett. 118, 153701 (2021); doi: 10.1063/5.0046745 118, 153701-3
Published under license by AIP Publishing
Applied Physics Letters ARTICLE scitation.org/journal/apl

FIG. 2. (a) SANS data of 1 wt. % HS40 þ 1 wt. % BSA system with varying concentration of the DTAB in 100% D2O (all components are visible), (b) in 62% D2O in D2O/H2O
(nanoparticles are contrast-matched), and (c) SANS data of 1 wt. % BSA with varying concentration of DTAB.

by SDS head groups along with protein in molten globule state.35 The 1 wt. % nanoparticles (supplementary Fig. S7) and with 1 wt. % BSA
charge-charge repulsion between the nanoparticles and SDS-lysozyme [Fig. 2(c)], separately are measured. The physical appearance of the
complex (15 mM) does not allow adsorption of the complex on nano- nanoparticles-DTAB system was turbid while that of the DTAB-BSA
particles and the complex coexists with individual nanoparticles with- system was transparent (supplementary Fig. S4). It is found that the
out any significant interaction. In accordance with this, the data of the DTAB micelles strongly adsorb on the nanoparticles and aggregate
1 wt. % HS40 þ 1 wt. % Lysozyme þ 15 mM SDS system could be ana- them, due to the micelles are oppositely charged to the nanoparticles.26
lyzed by combining the scattering contributions from the nanopar- On the other hand, DTAB micelles unfold the protein and form a
ticles and SDS micelles (supplementary Fig. S6). The fitted parameters bead-necklace kind of complex, where the surfactant micelles are
are shown in supplementary Tables S1 and S2. arranged along the unfolded protein chain.39 It is therefore clear that
So far, our results show that the addition of the SDS in higher in the three-component (1 wt. % HS40 þ 1 wt. % BSA þ 5 or 15 mM
concentration could suppress the strong adsorption of the lysozyme DTAB) system, the nanoparticles either interact through DTAB
on the nanoparticles and hence the protein-mediated nanoparticle micelles alone or DTAB-BSA complex.
aggregation. At this SDS concentration, the protein preferably interacts In order to confirm the actual condition, SANS measurements
with the surfactant via hydrophobic interaction instead of charge were performed at a condition where nanoparticles were contrast-
interaction with the nanoparticles, as overall charge of lysozyme is matched. The SANS data again show the signature of fractal-like
expected to be neutralized by anionic SDS (section S.3 in the supple- structure with a core-shell kind of building block, where the core
mentary material).37 Now, we illustrate that the suitable surfactant (nanoparticles) is contrast-matched to the solvent.26 The fitted shell
(cationic DTAB) can also be used to adsorb (or to enhance the protein thickness suggests that the aggregation is mediated by the BSA-DTAB
adsorption) a protein (BSA) on nanoparticles (anionic silica), which is complex, similar to the case of lysozyme and SDS. Therefore, the addi-
otherwise non-adsorbing, as both protein and nanoparticles are simi- tion of the suitable surfactant leads to the adsorption of the protein
larly charged at pH 7. which is otherwise non-adsorbing.
SANS data of 1 wt. % HS40 þ 1 wt. % BSA without and with In the above cases of nanoparticle protein interactions in the
DTAB surfactant, in the two contrast conditions, where all compo- presence of surfactants, we could achieve the tuning of the protein
nents are visible [Fig. 2(a)] and nanoparticles are contrast-matched to adsorption from strongly adsorbing to non-adsorbing for lysozyme
the solvent [Fig. 2(b)] are shown in Fig. 2. The scattering data of the and from non-adsorbing to adsorbing for BSA, via complex formation
1 wt. % HS40 þ 1 wt. % BSA system show a slight increase in the low with the surfactant, but the proteins lose their native structure. To
Q region compared to that of pristine 1 wt. % HS40. It is known that avoid this, the multivalent counterions, which are known to anoma-
the BSA gives rise to depletion attraction between the nanoparticles at lously alter the double layer around the nanoparticles as well as pro-
higher concentrations due to its (the BSA’s) non-adsorbing nature teins, to tune the protein adsorption on the nanoparticles, can also be
(arising from the strong electrostatic repulsion between nanoparticle employed.40,41 Recently, a multivalent counterions driven re-entrant
and protein), high number density and smaller size (SizeHS40: SizeBSA phase behavior has been observed in the both nanoparticles and pro-
2.8) compared to nanoparticles.15 Therefore, the low Q scattering teins (individually), where these systems undergo a transformation
build-up can be attributed to the changes in the interparticle potentials from one-phase (individual stabilized) to two-phase (aggregation or
hence is modeled by a single Yukawa potential taking account of liquid-liquid phase separation) and back to one-phase (individual)
depletion attraction (supplementary Table S3). On addition of the with monotonic increase in the multivalent ions concentration.42–46
DTAB, the scattering intensity increases in the low Q region in both Such re-entrant phase behavior arises due to excessive condensation of
cases and the samples become turbid (supplementary Fig. S4), sugges- the multivalent counterions on the nanoparticles/proteins, which in
ting the nanoparticle aggregation. Unlike SDS, the cationic DTAB turn reverse the charge of the nanoparticles/proteins.47 Such excessive
interacts with both anionic nanoparticles as well as anionic BSA pro- condensation is believed to be arising from the strong ion-ion correla-
tein (supplementary Figs. S1 and S2). Hence, to understand the forma- tions, prevailing in the multivalent counterions and is usually not
tion of the complexes in this system, the SANS data of the DTAB with observed for the monovalent ions.48 The emergence of the re-entrant

Appl. Phys. Lett. 118, 153701 (2021); doi: 10.1063/5.0046745 118, 153701-4
Published under license by AIP Publishing
Applied Physics Letters ARTICLE scitation.org/journal/apl

behavior is beyond the standard theories of the colloidal stability, e.g., The nanoparticle protein interaction has been tuned by adding
DerjaguinLandauVerweyOverbeek (DLVO) or Debye H€ uckel 20 mM concentration of the ZrCl4, where the negative charge of both
(DH) theories (section S.4 in the supplementary material) and is find- nanoparticles and BSA changes to positive (as both of these reverse
ing profound applications in physical as well as natural sciences.48–51 their charges, supplementary Fig. S9), while that of lysozyme remains
We have utilized this ability of the multivalent ions to reverse the unaltered (in this case, Cl- are the counterions, which unlike multiva-
charge of nanoparticles/protein facilitating the tuning of the electro- lent counterions do not adsorb on the lyzozyme due to lower charge
static interaction between nanoparticles and the two proteins. density). As a result, the interaction between nanoparticle and
Figure 3(a) shows the charge reversal in the nanoparticles in the lysozyme is changed from attractive to repulsive. The SANS data
presence of ZrCl4, where the measured zeta potentials for 1 wt. % [Fig. 3(c)] of the 1 wt. % HS40 þ 1 wt. % lysozyme system in the pres-
HS40 silica nanoparticle solution with varying ZrCl4 concentration are ence of 20 mM ZrCl4, exhibit no signature of the aggregation, as
plotted. At low ZrCl4 concentrations (CZr < 1 mM), the nanoparticles observed in the absence of ZrCl4 [Figs. 1(a) and 3(c)]. In fact, the data
are stable and having a negative zeta-potential. On increasing salt con- almost match with that of the pure nanoparticles in the low Q region,
centration (1 mM < CZr < 2 mM), the nanoparticles aggregate and suggesting the non-adsorption of the protein on the nanoparticles
zeta-potential could not be measured in this range. Further addition of (also verified by DLS, supplementary Figs. S8 and S10). A small scat-
the ZrCl4 (CZr > 2 mM) leads to a re-entrant stable phase with a posi- tering build-up in the high Q region arises from the free protein, which
tive zeta-potential, suggesting the charge reversal in the system.42 The again demonstrating that the protein mostly remains free in the solu-
SANS data of the 1 wt. % HS40 nanoparticles in the presence of tion.32 It should be noted that the presence of the 200 mM NaCl does
20 mM ZrCl4 and 200 mM NaCl (same ionic strengths) are shown in not show any significant changes in the scattering data, implying that
Fig. 3(b). As can be observed, the data in the presence and absence of the monovalent ions are not able to alter the nanoparticle-protein
the salt remain more or less same, and without any signature of the interactions.44,47–50 On the other hand, SANS data of the 1 wt. %
aggregation, suggesting that the nanoparticles are all stable (supple- HS40 þ 1 wt. % BSA system show similar features as observed without
mentary Fig. S8).42 addition of 20 mM ZrCl4. An increase in the low Q scattering can be

FIG. 3. (a) Zeta-potential of 1 wt. % HS40 nanoparticles as a function of the ZrCl4 salt concentration and the physical appearance of the nanoparticle dispersion with increasing
salt concentration. Insets show the physical appearance of the nanoparticle dispersion with increasing salt concentration, (b) SANS data of 1 wt. % HS40 nanoparticles in
the absence and presence of salts (200 mM NaCl and 20 mM ZrCl4), (c) SANS data of the 1 wt. % HS40 þ 1 wt. % lysozyme without and with salts, and (d) SANS data of the
1 wt. % HS40 þ 1 wt. % BSA system without and with salts. SANS data, presented in this figure, are collected in contrast condition where all the components are visible
(solvent is 100% D2O).

Appl. Phys. Lett. 118, 153701 (2021); doi: 10.1063/5.0046745 118, 153701-5
Published under license by AIP Publishing
Applied Physics Letters ARTICLE scitation.org/journal/apl

FIG. 4. Schematic showing the nanoparticle-protein interactions in the presence of surfactants (SDS/DTAB) and multivalent ions (Zrþ4) and resultant structures formed at each
step.

observed, suggesting the evolution of the attractive depletion interac- nanobiotechnology for creating tunable nano-biohybrids with
tion. In this case, the protein adsorption on nanoparticles does not directed response, for predicting and controlling the phase stability
enhance as both are expected to be charge reversed. This shows that and for producing multiscale equilibrium clusters.
the suitable multivalent counterions in appropriate concentrations can
be employed to tune the protein adsorption.41 See the supplementary material for details of experiments, data
To conclude, the present study shows pathways to tune the analysis (procedure and fitted parameters), and additional data.
interaction between nanoparticles and proteins and thereby protein
adsorption on nanoparticles as well as colloidal stability of the com- The neutron experiments at the Materials and Life Science
plexes formed. Figure 4 shows schematics of nanoparticle-protein Experimental Facility at J-PARC were approved by the Neutron
complexes formed in the presence of surfactants (SDS/DTAB) and Science Program Review Committee (Proposal No. 2019B0033).
multivalent ions (Zrþ4). It has been demonstrated that the preferen- This work was supported by a Grant-in-Aid for Scientific Research
tial binding of the protein with the surfactant and the multivalent on Innovative Areas (No. JP19H05717: Aquatic Functional
ion driven charge inversion of the nanoparticles/proteins can be uti- Materials).
lized to create switching between the protein adsorption and non- The authors declare no conflict of interest.
adsorption. Moreover, these parameters can also enable the control
over the undesired protein adsorption and nanoparticle aggregation
in the nanoparticle-protein systems. Since the methodology utilizes DATA AVAILABILITY
tuning of the electrostatic interaction between the silica nanopar- The data that support the findings of this study are available
ticles and proteins, we believe that the findings are transferable to from the corresponding author upon reasonable request.
the other charged nanoparticles (e.g., metal nanoparticles). In this
regard, other additives such as ionic biomolecules (DNA, lipids),
REFERENCES
ionic liquids, and charged polymers, which are able to tune the elec- 1
trostatic interaction between the nanoparticles and proteins, can M. Hadjidemetriou and K. Kostarelos, Nat. Nanotechnol. 12, 288 (2017).
2
M. Kopp, S. Kollenda, and M. Epple, Acc. Chem. Res. 50, 1383 (2017).
also be utilized. However, different strategies are to be designed for 3
I. Lynch and K. A. Dawson, Nanotoday 3, 40 (2008).
the systems, where the protein adsorption is facilitated by other 4
M. Papi, V. Palmieri, S. Palchetti, D. Pozzi, L. Digiacomo, E. Guadagno, M. d.
interactions such as covalent or hydrogen bonding. Overall, our B. D. Caro, M. D. Domenico, S. Ricci, R. Pani, M. Mahmoudi, A. Di Carlo, and
studies can find excellent potential applications in the field of G. Caracciolo, Appl. Phys. Lett. 114, 163702 (2019).

Appl. Phys. Lett. 118, 153701 (2021); doi: 10.1063/5.0046745 118, 153701-6
Published under license by AIP Publishing
Applied Physics Letters ARTICLE scitation.org/journal/apl

5 28
R. D. Santo, E. Quagliarini, S. Palchetti, D. Pozzi, V. Palmieri, G. Perini, M. L. Marichal, G. G. Colas, F. Cousin, A. Thill, J. Labarre, Y. Boulard, J.-C. Aude,
Papi, A. L. Capriotti, A. Lagana, and G. Caracciolo, Appl. Phys. Lett. 114, S. Pin, and J. P. Renault, Langmuir 35, 10831 (2019).
29
233701 (2019). A. S. Goryunov, Gen. Physiol. Biophys. 25, 303 (2006).
6 30
A. K. Barui, J. Y. Oh, B. Jana, C. Kim, and J.-H. Ryu, Adv. Therap. 3, 1900124 Y. M. Efimova, S. Haemers, B. Wierczinski, W. Norde, and A. A. van Well,
(2020). Biopolymers 85, 264 (2007).
7 31
J. Y. Oh, H. S. Kim, L. Palanikumar, E. M. Go, B. Jana, S. A. Park, H. Y. Kim, S. Takata, J. Suzuki, T. Shinohara, T. Oku, T. Tominaga, K. Ohishi, H. Iwase, T.
K. Kim, J. K. Seo, S. K. Kwak, C. Kim, S. Kang, and J.-H. Ryu, Nat. Commun. Nakatani, Y. Inamura, T. Ito, K. Suzuya, K. Aizawa, M. Arai, T. Otomo, and M.
9, 4548 (2018). Sugiyama, JPS Conf. Proc. 8, 036020 (2015).
8 32
E. Sanfins, J. Dairou, F. R. Lima, and J. M. Dupret, J. Phys.: Conf. Ser. 304, S. Kumar, V. K. Aswal, and P. Callow, Langmuir 30, 1588 (2014).
33
012039 (2011). E. G. Iashina, M. V. Filatov, R. A. Pantina, E. Y. Varfolomeeva, W. G.
9
M. Wang, C. Fu, X. Liu, Z. Lin, N. Yang, and S. Yu, Nanoscale 7, 15191 (2015). Bouwman, C. P. Duif, D. Honecker, V. Pipich, and S. V. Grigoriev, J. Appl.
10
F. Scaletti, J. Hardie, Y.-W. Lee, D. C. Luther, M. Ray, and V. M. Rotello, Cryst. 52, 844 (2019).
34
Chem. Soc. Rev. 47, 3421 (2018). A. Stenstam, A. Khan, and H. Wennerstr€ om, Langmuir 17, 7513 (2001).
11 35
Y. Li and J.-S. Lee, Materials 13, 3093 (2020). Y. Sun, P. L. O. Filho, J. C. Bozelli, Jr., J. Carvalho, S. Schreier, and C. L. P.
12
M. M. Yallapu, N. Chauhan, S. F. Othman, V. K. Sharghi, M. C. Ebeling, S. Oliveira, Soft Matter 11, 7769 (2015).
36
Khan, M. Jaggi, and S. C. Chauhan, Biomaterials 46, 1–12 (2015). D. Winogradoff, S. John, and A. Aksimentiev, Nanoscale 12, 5422 (2020).
13 37
V. Vergaro, I. Pisano, R. Grisorio, F. Baldassarre, R. Mallamaci, A. Santoro, G. P. A. Stenstam, G. Montalvo, I. Grillo, and M. Gradzielski, J. Phys. Chem. B 107,
Suranna, P. Papadia, F. P. Fanizzi, and G. Ciccarella, Materials 12, 1481 (2019). 12331 (2003).
14 38
Y. Li and N. A. Monteiro-Riviere, Nanomedicine (London) 11, 3185 (2016). J. Gummel, F. Boue, B. Deme, and F. Cousin, J. Phys. Chem. B 110, 24837
15
S. Kumar, I. Yadav, V. K. Aswal, and J. Kohlbrecher, Langmuir 34, 5679 (2018). (2006).
16 39
M. Barbalinardo, F. Caicci, M. Cavallini, and D. Gentili, Small 14, 1801219 D. Saha, D. Ray, J. Kohlbrecher, and V. K. Aswal, ACS Omega 3, 8260 (2018).
40
(2018). C. Pasquier, M. Vazdar, J. Forsman, P. Jungwirth, and M. Lund, J. Phys. Chem.
17
T. T. Chen, J. T. Yi, Y. Y. Zhao, and X. Chu, J. Am. Chem. Soc. 140, 9912 B 121, 3000 (2017).
41
(2018). M. R. Fries, D. Stopper, M. K. Braun, A. Hinderhofer, F. Zhang, R. M. J. Jacobs,
18
D. Shao, M. Li, Z. Wang, X. Zheng, Y. H. Lao, Z. Chang, F. Zhang, M. Lu, J. M. W. A. Skoda, H. H. Goos, R. Roth, and F. Schreiber, Phys. Rev. Lett. 119,
Yue, H. Hu, H. Yan, L. Chen, W. F. Dong, and K. W. Leong, Adv. Mater. 30, 228001 (2017).
42
1801198 (2018). S. Kumar, I. Yadav, S. Abbas, V. K. Aswal, and J. Kohlbrecher, Phys. Rev. E 96,
19
S. D. Medina, L. Kisley, L. J. Tauzin, A. Hoggard, B. Shuang, A. S. D. S. 060602(R) (2017).
43
Indrasekara, S. Chen, L. Y. Wang, P. J. Derry, A. Liopo, E. R. Zubarev, C. F. S. Kumar, D. Ray, S. Abbas, D. Saha, V. K. Aswal, and J. Kohlbrecher, Curr.
Landes, and S. Link, ACS Nano 10, 2103 (2016). Opin. Colloid Interface Sci. 42, 17 (2019).
20 44
I. Yadav, S. Kumar, V. K. Aswal, and J. Kohlbrecher, Phys. Rev. E 89, 032304 F. Zhang, M. W. A. Skoda, R. M. J. Jacobs, S. Zorn, R. A. Martin, C. M. Martin,
(2014). G. F. Clark, S. Weggler, A. Hildebrandt, O. Kohlbacher, and F. Schreiber, Phys.
21
S. T. Moerz, A. Kraegeloh, M. Chanana, and T. Kraus, ACS Nano 9, 6696 Rev. Lett. 101, 148101 (2008).
45
(2015). F. Zhang, R. Roth, M. Wolf, F. R. Runge, M. W. A. Skoda, R. M. J. Jacobs, M.
22
I. Yadav, S. Kumar, V. K. Aswal, and J. Kohlbrecher, Langmuir 33, 1227 Stzuckie, and F. Schreiber, Soft Matter 8, 1313 (2012).
46
(2017). C. Molitor, A. Bijelic, and A. Rompel, IUCrJ 4, 734 (2017).
23 47
R. Podila, R. Chen, P. C. Ke, J. M. Brown, and A. M. Rao, Appl. Phys. Lett. 101, T. T. Nguyen, A. Y. Grosberg, and B. I. Shklovskii, Phys. Rev. Lett. 85, 1568 (2000).
48
263701 (2012). A. Y. Grosberg, T. T. Nguyen, and B. I. Shklovskii, Rev. Mod. Phys. 74, 329
24
M. Yu, Z. Gu, T. Ottewell, and C. Yu, J. Mater. Chem. B 5, 3241 (2017). (2002).
25 49
M. E. Shetehy, A. Moradi, M. Maceroni, D. Reinhardt, A. P. Fink, B. R. O. Matsarskaia, F. R. Runge, and F. Schreiber, ChemPhysChem 21, 1742 (2020).
50
Rutishauser, F. Mauch, and F. Schwab, Nat. Nanotechnol. 16, 344 (2021). S. Kumar, I. Yadav, D. Ray, S. Abbas, D. Saha, V. K. Aswal, and J. Kohlbrecher,
26
S. Kumar, V. K. Aswal, and J. Kohlbrecher, Langmuir 28, 9288 (2012). Biomacromolecules 20, 2123 (2019).
27 51
D. I. Svergun and M. H. J. Koch, Rep. Prog. Phys. 66, 1735 (2003). X. Chen, L. Liu, and C. Jiang, Acta Pharm. Sin. B 6, 261 (2016).

Appl. Phys. Lett. 118, 153701 (2021); doi: 10.1063/5.0046745 118, 153701-7
Published under license by AIP Publishing

You might also like