You are on page 1of 8

Journal of Molecular Liquids 282 (2019) 169–176

Contents lists available at ScienceDirect

Journal of Molecular Liquids

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

Biopolymer nanoparticle surface chemistry dictates the nature and


extent of protein hard corona
Aalok Basu a,b, Sonia Kundu a, Chitra Basu c, Sumanta Kumar Ghosh a, Runa Sur c, Arup Mukherjee a,d,⁎
a
Department of Chemical Technology, University of Calcutta, 92 A.P.C. Road, Kolkata 700009, West Bengal, India
b
Dr. B.C. Roy College of Pharmacy and Allied Health Sciences, Bidhannagar, Durgapur 713206, West Bengal, India
c
Department of Biophysics, University of Calcutta, 92 A.P.C. Road, Kolkata 700009, West Bengal, India
d
Department of Biotechnology, Maulana Abul Kalam Azad University of Technology, BF 142, Sector 1, Salt Lake City, Kolkata 700064, West Bengal, India

a r t i c l e i n f o a b s t r a c t

Article history: Biopolymer nanoparticles functionalized with targeting ligands serve as carriers for a wide range of therapeutic
Received 27 August 2018 payloads to desired sites of the body. Upon introduction into the blood stream, the nanoparticles tend to bind
Received in revised form 5 February 2019 with the plasma proteins in its immediate vicinity, forming a protein corona. In this work, we have investigated
Accepted 4 March 2019
the effects of differently functionalized PLGA nanoparticles on the formation of protein corona. It was observed
Available online 5 March 2019
that nanoparticles of uniform size range adsorbed protein in varying amounts and the final hard corona differed
Keywords:
structurally with respect to nanoparticle surface functionalization. A combination of fluorescence spectroscopy
PLGA nanoparticles and FTIR studies revealed that the association of plasma proteins with PLGA nanoparticles during corona forma-
Protein corona tion was intricately guided by the nanoparticle surface chemistry. Densitometric analysis through one dimen-
Nanoparticle-plasma protein interaction sional gel electrophoresis further showed that the plasma proteins present in the hard corona were different
for each case of surface functionalization. The aggregation of plasma protein during corona formation as studied
through Thioflavin T fluorescence and circular dichorism spectroscopy was also found to significantly differ with
each nanoparticle type. It is evident that surface chemistry of biopolymeric nanoparticles defines the final corona
form in the physiological environment, and this study on detailed understanding of nanoparticle-protein corona
would help to develop nanomedicines for clinical applications.
© 2019 Elsevier B.V. All rights reserved.

1. Introduction intended target site [5]. Whenever a nanoparticle presents itself in a bi-
ological milieu, its surface is known to get quickly embedded in plasma
Advancements in the field of biopolymeric nanoparticle engineering proteins, thus altering the biological identity of the nanoparticle. This
have rapidly promoted their utility in medical diagnostics and thera- form, termed by K. A. Dawson as “protein corona”, affects the cellular
peutic interventions, such as pin pointed targeting of drug, enhanced uptake and subsequent fate of the nanoparticle in vivo [6,7]. Nanoparti-
cellular uptake, reduced toxicity, and programmed release of their pay- cle protein corona is currently conceived as dynamic physiological ac-
load [1]. Poly (lactic-co-glycolic acid) (PLGA) is one widely used bio- tivity which ultimately equilibrates in a plasma protein adsorption [8].
polymer in designing of nano-scale materials, and can be differentially Current concepts indicates that the corona comprise of a tightly
tailored to serve specific needs. Functionalization of PLGA nanoparticles bound, “hard” near-monolayer of proteins which is again shielded by
with site-specific ligands to hit cellular targets, polyelectrolyte deposi- a “soft” layer of loosely associated proteins [9]. The proteins adsorbed
tions to improve cellular uptake or PEGlyation to reduce elimination in the hard corona have longer residence time making it both physiolog-
by phagocytosis are of the common strategies to develop smarter drug ically relevant [10] and easier to isolate [11]. Proteins upon adsorption
delivery devices [2–4]. on the nanosurface experience imbalance in entropy and change in
Despite of advances in preclinical development of nanoparticle drug structural conformations [12]. Such changes can be either reversible or
delivery systems, there has always been a therapeutic shortfall upon ex- irreversible [13] and is a function of nanoparticle surface curvature
posure of nanoparticles in the physiological environment. A survey con- [14–16]. A number of reports in past few years laid down the fact that
ducted by Warren C.W. Chan and co-workers revealed that only 0.7% of nanoparticles have a profound effect on protein fibrillogenesis, which
administered dose of nanoparticles are effectively delivered to its is characterized by formation of cross β sheets at the expense of native
α helical structure [17].
⁎ Corresponding author at: Department of Chemical Technology, University of Calcutta,
Most of the initial works are concentrated upon formation and pro-
92 A.P.C. Road, Kolkata 700009, West Bengal, India. teomics of the corona around different metal nanoparticles [18–20].
E-mail address: arupm1234@gmail.com (A. Mukherjee). More recent studies depicted that surface charge and size of

https://doi.org/10.1016/j.molliq.2019.03.016
0167-7322/© 2019 Elsevier B.V. All rights reserved.
170 A. Basu et al. / Journal of Molecular Liquids 282 (2019) 169–176

nanoparticles play a significant role in corona formation and cellular in- 2.3. Nanoparticle characterization
teractions thereafter [21,22]. Several workers have attempted to regu-
late the protein corona formations around the nanoparticle by tuning Particle analysis in all cases were carried out in a Zetasizer (Malvern
their surface-bound ligands [23], introduction of zwitterionic function- Nano ZS, Malvern, UK) and measurements were completed in dispos-
alities [24], or through synthesis of PEGylated copolymers [25]. How- able polystyrene cuvettes against 4 mW He-Ne laser using a back scat-
ever, works on nano-bio interface of biopolymeric nanoparticles and tering angle of 173°. Electrophoretic mobility was recorded in PBS
their functionalized counterparts are fewer in number. Kreuter (2013) (50 mM, pH 7.4) and zeta potentials were recorded for comparative
and Langer (2015) have investigated the proteins eluted from the co- analysis.
rona around PLGA nanoparticles through SDS PAGE, followed by diges- TEM micrographs of nanoparticles were obtained using JEOL JEM
tion, purification and peptide identification with mass spectroscopy 2100 HR transmission electron microscope capable of point to point res-
[19,26]. Coronated nanoparticles are though the final form in physiolog- olution. Highly diluted samples were deposited on carbon coated cop-
ical environment, a proper insight in that domain is rather sparse. PLGA per grid (Ted Pella Inc., US) and micrographs were collected at a low
is a Food and Drug Administration (FDA) approved biopolymer and is accelerating voltage of 120 kV.
garnered lot of attention for its biocompatibility and degradability Polyelectrolytes on nanoparticles were quantified using colorimetric
[27]. Our work has been intended to investigate the effect of surface assays. Chitosan interaction with alizarin red dye (0.1% w/v) at 571 nm
functionalization of PLGA nanoparticles on the formation of their hard was used to develop a concentration dependent standard graph, y =
corona structures. Initially, bare pluronics stabilized and polyelectro- 0.002x + 0.084, R2 = 0.998. The chitosan deposit on PlnC was estimated
lytes such as heparin and chitosan coated PLGA nanoparticles of uni- from the concentration difference between the initial solution and the
form size range were synthesized. The structure of the hard corona post harvested supernatants. Three separate batches of experiments
around these nanoparticles and their underlying chemical interactions were run and the data were averaged for comparative purposes. Hepa-
formed the basis of our study. Affinity constants of different nanoparti- rin was quantified using toluidine blue color reactions [30]. Different
cle types derived from fluorescence spectroscopy were collectively used concentrations of heparin standard solutions were first reacted with to-
with FTIR data to explore the molecular interactions occurring in luidine blue (0.025% w/v) and the absorbance at 562 nm was used to
nanoparticle-protein interfaces. Further investigations to identify the generate a standard graph y = 0.038x + 0.127, R2 = 0.992. Heparin
effects of these molecular interactions on protein secondary structures content on nanoparticles was derived from the difference of polyelec-
were also carried out through Thioflavin T fluorescence and circular di- trolyte initial solution concentration and the post harvesting accumu-
chroism spectroscopy. It is expected that this work would shed some in- lated supernatant content. An average of three batch experiment was
sight into mechanism of protein corona on functionalized biopolymeric recorded and the data were presented in percentage.
nanoparticles commonly used in clinical settings.

2. Materials and methods 2.4. Nanoparticle protein corona

2.1. Materials Lyophilized human plasma was initially dissolved in 1 mL of PBS


(50 mM, pH 7.4). The reconstituted solution was then diluted to 1% v/
PLGA (50:50), 85% deacetylated chitosan (MW medium), heparin v for use in subsequent experiments [31]. Nanoparticles dispersed in
sodium, pluronic F -127, Bradford reagent for protein assay, lyophilized PBS (500 μg mL−1) were incubated with the plasma solutions for 1 h
human plasma and human serum albumin (purified) were all procured at 37 ± 0.5 °C. The solutions were then centrifuged at 14,500 rpm in
from Sigma Aldrich (St. Louis, US). All solvents, including water, were of order to obtain hard corona nanoparticles. Individual pellet deposits
HPLC grade and obtained from Spectrochem Pvt. Ltd. (Mumbai, India). were dispersed in 1 mL PBS and washed thrice by centrifugation in
Potassium bromide for IR analysis was purchased from Merck (Darm- order to ensure complete removal of any loosely bound proteins. The
stadt, Germany). Thioflavin T dye for fibrillation experiments were pur- final hard corona nanoparticles were re-suspended in PBS for DLS mea-
chased from TCI Chemicals Pvt. Ltd. (Chennai, India). surements prior to in-depth analysis.

2.2. Nanoparticle synthesis 2.5. Protein quantification

PLGA nanoparticles (Pln) were prepared in solvent diffusion tech- Bradford assay (Sigma Aldrich, US) was performed to estimate the
niques perfected earlier [28]. Briefly, 50 mg of the polymer was dis- mass of protein on nanoparticles surface before and after the formation
solved in 5 mL of acetone, and dropped into 10 mL aqueous pluronic of hard corona [32]. Briefly, 1 mL of 1% plasma solution was incubated
F-127 (1% w/v) solution under magnetic stirring. Stirring was continued with or without 500 μg of nanoparticles for 1 h. 0.1 mL of the superna-
for 3 more hours and the nanoparticles formed were harvested by cen- tant acquired after centrifugation was mixed with 3 mL of Bradford re-
trifugation at 14,500 rpm. Pln were re-dispersed in 10 mL of water and agent, and analyzed at 595 nm using a UV–Visible spectrophotometer
collected after centrifugation for subsequent polyelectrolyte coating. (UV 1800 Shimadzu, Japan). An absorbance (y) vs. concentration
The chitosan or heparin polyelectrolyte (0.05% w/v) was dissolved in (x) graph, y = 0.690x − 0.071, R2 = 0.999, was initially developed
water and the solution pH was maintained at 6. Pln nanosuspension in from standard HSA solutions in PBS and used for protein quantification
water (3% w/v, 2 mL) was then added drop-wise into a 10 mL polyelec- throughout.
trolyte solution under magnetic stirring. Stirring was continued for
5 min more and the resultant chitosan or heparin surfaced nanoparti-
cles (PlnC or PlnH) were harvested by centrifugation at 14,500 rpm. 2.6. Steady-state fluorescence quenching studies
A solute free technique was perfected to obtain nanoparticles in nar-
row size window [29]. Gravity driven sedimentation cycle was carried Intrinsic fluorescence from 1% v/v human plasma in PBS was first re-
out to separate nanoparticles from solutions. Briefly, nanoparticle sus- corded in a spectrofluorimeter (Perkin-Elmer LS-5, PerkinElmer Inc.,
pension was first subjected to a mild centrifugation at 6000 rpm to sed- US). Different concentrations (0 to 40 μg mL−1) of each nanoparticle
iment any large particles. The supernatant was collected and types were incubated with plasma at 37 °C for 1 h [33]. The quenching
centrifuged again at 14,500 rpm. Sediment particles were re-dispersed of emission spectra were recorded over a range of 290–400 nm with ex-
in particle free water and the stocks were maintained at 4 ± 0.5 °C for citation wavelength set at 280 nm, and the data were plotted in a log-
further studies. log graph to derive plasma protein affinity constant in each case.
A. Basu et al. / Journal of Molecular Liquids 282 (2019) 169–176 171

2.7. FT IR studies thus eliminated the possibility of affecting the corona structure due
morphological differences [15].
FT IR spectral analysis was performed using Jasco-670 Plus FTIR in- A thorough characterization of the nanomaterials was performed
strument (Jasco, Japan). Spectral analysis was carried out at a resolution through independent experimental procedures to assert polyelectrolyte
of 4 cm−1 and an average of 128 scans were recorded over the mid IR deposition on the particle surface. Polyelectrolyte mass deposition on
ranges of 4000–400 cm−1. Individual sample of nanoparticles were ly- the nanoparticles was quantified using alizarin red and toluidine blue
ophilized and then pressed in IR grade KBr to develop transparent pel- based colorimetric reactions for chitosan and heparin, respectively
lets [34]. Gathered data were stacked using Bio-Rad software (Bio- [36]. In case of chitosan coating, the entrapment efficiency was calcu-
Rad, KnowItAll, USA) for analysis of bare nanoparticles and protein co- lated to be 74.3%, while for heparin the value was 51.7%.
rona interactions. FT IR spectra also confirmed the deposition of chitosan and heparin
on the PLGA nanoparticle as in their individual cases. The pluronics sta-
2.8. Gel electrophoresis bilized Pln spectrum exhibited a sharp characteristics peak at
1759 cm−1 due to presence of carbonyl group, while vibrations for
Plasma protein binding onto nanoparticle surface was determined C\\O and C\\H were recorded at 1456 cm−1 and 1169 cm−1. Coating
by 1 D sodium dodecyl sulfate polyacrylamide gel electrophoresis of chitosan upon Pln could be easily identified from the presence of
(SDS PAGE) following a method explained elsewhere [21]. \\OH peak at 3438 cm−1, and amide band stretch at wavenumbers
Nanoparticle-protein hard coronas obtained from the procedure as de- 1636 cm−1, and 1533 cm−1. Heparin functionalized Pln exhibited char-
scribed in Section 2.4 was re-suspended in 20 μL of 1× loading buffer. acteristics stretching of CH2SO3− at 959 cm−1 and vibrations of COO−
Samples were incubated at 90 °C for 2 min, cooled to room temperature. groups at 1632 cm−1. A wide band at around 3426 cm−1 was prominent
Nanoparticle surface bound proteins were separated on 10% polyacryl- due to presence overly expressed hydroxyl moieties in heparin mole-
amide SDS PAGE in an electric field run under constant voltage on 100 cules. The peak at 1759 cm−1 from Pln was preserved in both cases of
V for 45 min. The resulting gels were fixed by Coomassie blue stain PlnC and PlnH.
and recorded using a gel documentation system (BioRad, Hercules), Further, the zeta potential for Pln, PlnC and PlnH were found to be
and gel densitometry was performed using Bio-Rad Gel Doc using −19.7, +22.3 and −38.4 mV, respectively. These recordings revealed
Quantity One software. Plasma and purified human serum albumin the changes in nano-surface chemistry in each type of particles and
were also run along-side for comparison. well correlated with the FT IR observations. The negative charge of Pln
was prevalent due to the exposed carboxylic polymer chain groups of
2.9. Thioflavin T (ThT) dye interactions studies PLGA. Chitosan coating with exposed amide groups provided a positive
charge to PlnC, whereas the presence of sulphonate and carboxylate
Nanoparticles (Pln, PlnC or PlnH) were dispersed in 1% v/v human groups of heparin were responsible for a high charge density on PlnH
plasma in PBS and incubated at physiological temperature (37 ± 0.5 surface.
°C) for 1 h. Samples were then re-incubated with ThT (10 μM) and the
fluorescence emissions were measured at physiological temperature 3.2. Protein corona formation
(37 ± 0.5 °C) in each case [35]. The excitation and emission wavelength
were set at 450 nm and 482 nm, respectively. Amyloid-like plasma fi- Protein corona formation was studied by incubating an equal
brillation was further induced in 1% v/v human plasma in PBS at 60 °C amount (500 μg) of differently functionalized nanoparticles for 1 h in
for 1 h and that effect was assumed as the positive control. PBS containing 1% human plasma. Evolution of protein corona around
the nanoparticle is a dynamic phenomenon involving reversible and ir-
2.10. Circular dichroism reversible attachment of protein entities on the surface [24]. Hard co-
ronas of the differently functionalized nanoparticles obtained through
Far-UV CD measurements of the samples (from Section 2.9) were centrifugation, were of sole interest for our work. Dynamic light scatter-
performed on a JASCO J-815 CD spectro-polarimeter with a thermostat- ing (DLS) analysis was performed under dilute conditions provided the
ically controlled cell holder at 25 °C, in a cuvette of 0.1 cm path length. initial evidences of protein interactions on nanoparticle surface. The re-
The scan speed was set as 100 nm min−1 and ellipticity was recorded sults showed an increase in the hydrodynamic diameter (Fig. 1A) of in-
between 190 and 250 nm from an average of three scans. dividual particles ranging from 52.1% (for Pln) to 171.2% (for PlnC), with
the adsorption of varying amounts of plasma proteins (Fig. 1B). DLS
2.11. Statistical analysis measurements also exhibited a divergence from the original polydisper-
sity index (data not shown), suggesting particle aggregation and
All the experiments were performed in triplicates and their results changes in overall dispersion pattern (Fig. 1C).
are expressed as their mean ± S.D. Data from comparative studies Similar observations recorded from the zeta potential analysis
were interpreted using student t-test and the difference was considered (Fig. S2) were used to interpret the stability of colloids in plasma envi-
significant as p b 0.05. ronment. Exposure of nanoparticles to plasma results in lowering of
charge density with instantaneous binding of oppositely charged pro-
3. Results & discussion teins [37]. However, the prevalence of a slight negative charge on the
nanomaterial surface at pH 7.4 was observed due to the subsequent ad-
3.1. Nanoparticle characterization sorption of serum albumin [38]. The corona of PlnC faced a maximum
electrostatic aggregation effect in biological media, therefore tends to
Our studies were carried out on surface functionalized PLGA nano- settle down more. The results further showed that minimum deviation
particles, which present great potential in various therapeutic applica- of zeta potential values occurred in case of the Pln- protein corona,
tions [2]. Pluronics stabilized PLGA nanoparticles obtained through whose stability is mediated by steric repulsion offered by the amphi-
solvent diffusion techniques were subsequently coated with heparin philic nature of the pluronics [39].
or chitosan. Highly mono-dispersed nanoparticles (Fig. S1), either in
pristine form or coupled with chitosan or heparin, were obtained 3.3. Nanoparticle protein interaction studies
through selective centrifugation and were used to explore the effect of
nanosurface functionalization on protein corona formation. Electron Elucidating the interaction of protein on nanoparticle surface is one
microscopy confirmed that the particles were spherical in shape and important step to gain detailed information on protein corona
172 A. Basu et al. / Journal of Molecular Liquids 282 (2019) 169–176

Fig. 1. (A) Particle size distribution for (i) Pln (ii) PlnC (iii) PlnH in absence and presence of human plasma (1%) (B) Protein adsorption quantification (C) Transmission electron
micrographs of (i) Pln (ii) PlnC (iii) PlnH in presence of human plasma (1%).

formation. Preliminary information about nanoparticle-protein interac- the amide stretch for PlnC were not detected upon co-incubation with
tion was obtained from tryptophan fluorescence spectroscopy, as tryp- 1% human plasma. However, amide I band of plasma exhibited a signif-
tophan emission is highly specific to changes in local environment [40]. icant shift from 1650 cm−1 to 1644 cm−1, along with a shift of amide II
A gradual quenching of fluorescence intensity between 300 and 450 nm peak from 1538 cm−1 to 1551 cm−1. These interactions were expressed
at λex = 295 nm was observed for all three nanoparticle types (Fig. 2), due to intermolecular hydrogen bond formation between carbohydrate
which can be attributed to the conformational changes of human moieties and plasma protein during surface adsorption [43]. Similarly,
serum proteins around their tryptophan residues upon interaction when PlnH were co-incubation with plasma, there has been a marked
with the nanoparticle surface. up-shift (from 1539 cm−1 to 1543 cm−1) of amide II band of plasma
Nanoparticle-protein binding constants in 1% human plasma were protein due to their adsorption on the nanosurface. While peaks
derived from log [(F0-F)/(F-Fs)] vs. log [nanoparticles] plots for all owing to different groups from pluronics and PLGA (C_O, C\\O,
three types of nanoparticles. It was observed that among the other C\\H) remained intact, a disappearance of peak at 959 cm−1 was ob-
types of nanoparticles, PlnC had the maximum affinity towards plasma served. The FTIR results suggested that certain degrees of electrostatic
proteins (Kb = 2.96), due to a typical guest-host complex formation interactions were pronounced between the amino groups of plasma
[41].The binding constant value was found to be at a lesser extent (Kb protein and negatively charged sulphonates and carboxylates of hepa-
= 2.37) in case of PlnH and least (Kb = 1.12) for the pluronics stabilized rin. Such protein-glycosaminoglycans interactions often require expo-
Pln. The results clearly suggested that surface engineered biopolymeric sure of the binding site in protein through re-alignment of basic
nanoparticles associated differently with plasma proteins under similar amino acid residues [44] and may contribute to the alternations in pro-
conditions, despite of identical morphological characteristics. tein tertiary structures.
The underlying interactions leading to the nanoparticle-protein
binding was further investigated through FT IR spectroscopy. Amide I 3.4. Nanoparticle-bound protein profile analysis
band at 1630 cm−1- 1658 cm−1 (due to C_O stretch) and amide II
band at 1525 cm−1- 1547 cm−1 (associated with C\\N stretch and Nanoparticle-protein binding affinities and chemical interactions
N\\H in-plane bending) for plasma protein were recorded and moni- are the key factors in determining the number and types of proteins
tored for subsequent nanoparticle-protein interaction studies. The in- which constitutes the hard corona structure [18]. Proteins present in
tensity of amide I band is related to secondary structure of protein the hard corona of functionalized nanoparticles were therefore sepa-
whereas amide II signal is attributed to protein adsorption on particle rated, using one dimensional gel electrophoresis (Fig. 4a), to perceive
surface [42]. Detailed analysis of protein interactions was done through the nature of protein corona surrounding the nanoparticles. Coomassise
IR experiments, where 1650 cm−1- 1658 cm−1 region of amide I peak is stained gel for 1% v/v human plasma exhibited typical bands corre-
a representation of α-helical structures and 1620 cm−1- 1645 cm−1 re- sponding to most abundant proteins such as transferrin (80 kDa) [45],
gion signifies presence of the β-sheets [13]. Therefore, alterations in serum albumin (67 kDa), IgG (heavy chain, 50 kDa), apolipoprotein E-
these bands characteristics would help explore how nanoparticle sur- 1 (34 kDa) [38,46]. The protein patterns arising from the hard coronas
face chemistry affects the stability of native proteins which participate of the nanoparticles revealed the adsorption of all the major plasma pro-
in corona formation. Pln when exposed to 1% human plasma, exhibited teins with varying intensity for each nanoparticle type. The observations
the dual protein peaks (for amide I and amide II) at their exact positions, were found to be very similar to other works on biopolymeric nanopar-
while retaining the characteristic peaks of pluronics stabilized nanopar- ticles [47,48].
ticles (Fig. 3). This indicated the absence of any chemical interactions The data were further verified from the band intensity profiles
among the protein residues and Pln, and that the corona formation (Fig. 4b), obtained from QuantityOne software, of gel image. Albumin
around the Pln was merely due to surface deposition of plasma proteins. is the most abundant plasma protein (~44 g L−1), characterized by var-
In case of chitosan coated nanoparticles, the bands originally assigned to ious binding sites, specific structural conformation and was found to be
A. Basu et al. / Journal of Molecular Liquids 282 (2019) 169–176 173

Fig. 2. (A) Steady state fluorescence quenching of Trp214 signal upon interaction with increasing concentration (0 to 40 μg mL−1) of different nanoparticles at 37 °C (i) Pln (ii) PlnC (iii)
PlnH. (B) Binding constant derived from spectral data.

adsorbed on all the nanoparticles. Albumin attachment to PlnH was The relative density of these three proteins were slightly higher on the
maximum probably due to its high surface charge (−38.4 mV) at surface of chitosan coated PLGA nanoparticles. Chitosan is positively
pH 7.4. Apart from albumin, other proteins such as transferrin, IgG and charged polymer, and can induce innate immunological response in bi-
apolipoprotein E-1 were strongly adsorbed on the nanoparticle surface. ological system [49]. Thus, PlnC was found to be more associated with

Fig. 3. FTIR spectra of different nanoparticles and their hard corona.


174 A. Basu et al. / Journal of Molecular Liquids 282 (2019) 169–176

Fig. 4. Analysis of human plasma protein on nanoparticles (a) Resolved SDS-PAGE of adsorbed plasma proteins on nanoparticles (b) Histograms of band intensities of major adsorbed
proteins.

immunoglobulins (IgG) than other nanoparticles. This was significantly aggregation using a combination of Thioflavin T fluorescence and circu-
lower in case of PlnH because heparin coating, unlike chitosan, ensures lar dichroism spectroscopy (CD). ThT is cross beta sheet specific dye
reduced activation of complement system [50]. Protein adsorption upon used to elucidate whether a certain testing agent modulate the assem-
Pln surface was significantly lower in comparison to other cases, as pre- bly or disassembly of protein aggregates or amyloids [53]. Freshly
viously quantified through the Bradford assay (Fig. 1b). These observa- reconstituted human plasma (1%) exhibited a low degree of ThT fluo-
tions perfectly rationalize the outcomes obtained from the detailed rescence intensity due to presence of proteins (Fig. 5A), such as
protein-nanoparticle interaction studies through IR and fluorescence microglobulins and glycoproteins which feature native beta sheet con-
spectroscopy. The corona surrounding the nanoparticles was found to figurations [54,55]. Plasma upon incubation at 60 °C showed a large in-
composed of plasma proteins of different kinds and amount, the binding crease in ThT fluorescence response at 480 nm. Albumin, the major
of which rested upon the nature of the nanoparticle surface chemistry. protein plasma, undergoes conformation change to amyloid-like struc-
tures upon incubation at elevated temperatures [56], and was greatly
3.5. Effect of nano-surface chemistry on protein aggregation responsible for upsurge for ThT fluorescence intensity. This observation
was regarded as the positive control. Plasma samples co-incubated with
Nanoparticles are potential candidates known to cause major con- Plns beared a low degree of ThT intensity. This is due to the fact that
formational changes in protein structure [37,51] and the hard corona fewer number of proteins were adsorbed on the Pln surface and faced
is one site where nanoparticles come in close vicinity of different pro- structural alteration [57]. Among the different nanoparticles co-
teins. FTIR studies had primarily revealed that adsorption of plasma pro- incubated with plasma, ThT response was maximum in case of PlnC.
teins on the surface engineered nanoparticles was due to chemical Similar responses of chitosan were recorded during its interaction
interactions and could lead to the loss of protein tertiary structures. De- with serum albumin [41]. Chitosan functionalization disrupted alpha
stabilization of native protein structure through misfolding and expo- helical conformations of adsorbed proteins, and led to acceleration of
sure of hydrophobic residues marks the initiation of protein protein aggregation on the nanoparticle surface. Therefore, it can be en-
aggregation and amyloid formation [52]. We, therefore, investigated visaged that the corona surrounding the PlnC comprised mostly of ag-
the effects of nanoparticle surface chemistry on plasma protein gregated protein. A low degree of aggregation was also observed

Fig. 5. (A) ThT fluorescence response of 1% (v/v) blood plasma co-incubated with and without nanoparticles. (B) Far UV CD spectra of 1% (v/v) blood plasma co-incubated with and without
nanoparticles.
A. Basu et al. / Journal of Molecular Liquids 282 (2019) 169–176 175

during association of PlnH with plasma protein, possibly due to exis- [5] S. Wilhelm, A.J. Tavares, Q. Dai, S. Ohta, J. Audet, H.F. Dvorak, W.C.W. Chan, Analysis
of nanoparticle delivery to tumours, Nat. Rev. Mater. 1 (2016), 16014. https://doi.
tence of weak electrostatic interactions. org/10.1038/natrevmats.2016.14.
CD was also used to measure the conformational changes occurring [6] V. Mirshafiee, M. Mahmoudi, K. Lou, J. Cheng, M.L. Kraft, Protein corona significantly
in plasma protein upon nanoparticle interactions. Negative peaks at 208 reduces active targeting yield, Chem. Commun. 49 (2013) 2557–2559, https://doi.
org/10.1039/c3cc37307j.
and 222 nm arise from π → π* and n → π* transitions of carbonyl groups [7] T. Cedervall, I. Lynch, S. Lindman, T. Berggard, E. Thulin, H. Nilsson, K.A. Dawson, S.
present in the polypeptide chains of serum albumin. Plasma incubated Linse, Understanding the nanoparticle-protein corona using methods to quantify
at 60 °C showed a disappearance of the negativity at 208 and 222 nm exchange rates and affinities of proteins for nanoparticles, Proc. Natl. Acad. Sci.
104 (2007) 2050–2055, https://doi.org/10.1073/pnas.0608582104.
due to breakdown of α-helical forms (Fig. 5B). Appearance of single [8] D. Dell'Orco, M. Lundqvist, C. Oslakovic, T. Cedervall, S. Linse, Modeling the time evo-
negative peak at 215 nm indicated an increase of β-sheet content lution of the nanoparticle-protein Corona in a body fluid, PLoS One 5 (2010),
[58,59]. Co-incubation with Pln showed minimum change in the CD e10949. https://doi.org/10.1371/journal.pone.0010949.
[9] M.P. Monopoli, C. Åberg, A. Salvati, K.A. Dawson, Biomolecular coronas provide the
spectra as compared to that of fresh plasma. However, more deviations
biological identity of nanosized materials, Nat. Nanotechnol. 7 (2012) 779–786,
in terms of mdeg values observed in case of PlnC and PlnH revealed that https://doi.org/10.1038/nnano.2012.207.
PlnC and PlnH affected the native conformations of protein in the vicin- [10] C.D. Walkey, W.C.W. Chan, Understanding and controlling the interaction of
ity of the nanoparticles. ThT fluorescence and CD spectroscopy observa- nanomaterials with proteins in a physiological environment, Chem. Soc. Rev. 41
(2012) 2780–2799, https://doi.org/10.1039/c1cs15233e.
tions thus suggested that different nanosurface functionalization affects [11] M. Lundqvist, J. Stigler, G. Elia, I. Lynch, T. Cedervall, K.A. Dawson, Nanoparticle size
the native protein conformations in the corona microenvironment. Fur- and surface properties determine the protein corona with possible implications for
ther studies under in vivo conditions would help understand how pro- biological impacts, Proc. Natl. Acad. Sci. 105 (2008) 14265–14270, https://doi.org/
10.1073/pnas.0805135105.
tein aggregation upon nanoparticle surface would interfere the [12] C. Cabaleiro-Lago, O. Szczepankiewicz, S. Linse, The effect of nanoparticles on amy-
biodistribution of the nanoparticles and affect the efficiency of site- loid aggregation depends on the protein stability and intrinsic aggregation rate,
specific delivery of payload. Langmuir. 28 (2012) 1852–1857, https://doi.org/10.1021/la203078w.
[13] L. Shang, Y. Wang, J. Jiang, S. Dong, PH-dependent protein conformational changes
in albumin:gold nanoparticle bioconjugates: a spectroscopic study, Langmuir. 23
4. Conclusion (2007) 2714–2721, https://doi.org/10.1021/la062064e.
[14] M. Kurylowicz, H. Paulin, J. Mogyoros, M. Giuliani, J.R. Dutcher, The effect of nano-
scale surface curvature on the oligomerization of surface-bound proteins, J. R. Soc.
In this work, we had intended to systematically explore and com- Interface 11 (2014)https://doi.org/10.1098/rsif.2013.0818.
pare the physico-chemical interactions occurring between some typi- [15] A. Albanese, P.S. Tang, W.C.W. Chan, The effect of nanoparticle size, shape, and sur-
cally used PLGA nanoparticles and human plasma proteins, and face chemistry on biological systems, Annu. Rev. Biomed. Eng. 14 (2012) 1–16,
https://doi.org/10.1146/annurev-bioeng-071811-150124.
understand the effect of such interactions on protein corona formations. [16] S. Goy-López, J. Juárez, M. Alatorre-Meda, E. Casals, V.F. Puntes, P. Taboada, V.
It is now a known fact that nanoparticles exposed to biological media Mosquera, Physicochemical characteristics of protein–NP bioconjugates: the role
triggers the formation of a protein rich layer surrounding the nanopar- of particle curvature and solution conditions on human serum albumin conforma-
tion and fibrillogenesis inhibition, Langmuir. 28 (2012) 9113–9126, https://doi.
ticle surface, which re-defines the identity of the particles in terms of its org/10.1021/la300402w.
therapeutic capacities. Our findings suggested that composition and [17] C. Cabaleiro-Lago, F. Quinlan-Pluck, I. Lynch, S. Lindman, A.M. Minogue, E. Thulin,
thickness of the protein corona vary with surface functionalities of bio- D.M. Walsh, K. a Dawson, S. Linse, Inhibition of amyloid β protein fibrillation by
polymeric nanoparticles, J. Am. Chem. Soc. 130 (2008) 15437–15443. doi:https://
polymer nanoparticles. It was intriguing to observe that adsorption of doi.org/10.1021/ja8041806.
plasma protein altered the native protein conformations, and led to for- [18] M. Lundqvist, J. Stigler, T. Cedervall, T. Berggård, M.B. Flanagan, I. Lynch, G. Elia, K.
mation of protein aggregates. These nanosurface-plasma protein inter- Dawson, The evolution of the protein corona around nanoparticles: a test study,
ACS Nano 5 (2011) 7503–7509, https://doi.org/10.1021/nn202458g.
actions will prove key players in determining the immunological
[19] R. Gossmann, E. Fahrländer, M. Hummel, D. Mulac, J. Brockmeyer, K. Langer, Com-
identity, systemic uptake and residence time, and toxicological aspects parative examination of adsorption of serum proteins on HSA- and PLGA-based
of surface engineered nanoparticles. This kind of work is thus expected nanoparticles using SDS – PAGE and LC – MS, Eur. J. Pharm. Biopharm. 93 (2015)
80–87.
to set the stage for further studies by highlighting the behaviour of
[20] J.H. Shannahan, X. Lai, P.C. Ke, R. Podila, J.M. Brown, F.A. Witzmann, Silver nanopar-
biopolymeric nanoparticles in biological environment, and address the ticle protein Corona composition in cell culture media, PLoS One 8 (2013)https://
issues related to translational development of nanomedicines. doi.org/10.1371/journal.pone.0074001.
Supplementary data to this article can be found online at https://doi. [21] G.J. Pillai, M.M. Greeshma, D. Menon, Impact of poly(lactic-co-glycolic acid) nano-
particle surface charge on protein, cellular and haematological interactions, Colloids
org/10.1016/j.molliq.2019.03.016. Surfaces B Biointerfaces. 136 (2015) 1058–1066, https://doi.org/10.1016/j.colsurfb.
2015.10.047.
[22] S. Tenzer, D. Docter, J. Kuharev, A. Musyanovych, V. Fetz, R. Hecht, F. Schlenk, D.
Acknowledgement Fischer, K. Kiouptsi, C. Reinhardt, K. Landfester, H. Schild, M. Maskos, S.K. Knauer,
R.H. Stauber, Rapid formation of plasma protein corona critically affects nanoparti-
The authors would like to thank TEQIP project in the University of cle pathophysiology, Nat. Nanotechnol. 8 (2013) 772–781, https://doi.org/10.
1038/nnano.2013.181.
Calcutta for providing instrumental facilities on no cost basis. A research
[23] S. Khan, A. Gupta, C.K. Nandi, Controlling the fate of protein corona by tuning surface
fellowship sponsored to author CB from University Grant Commission, properties of nanoparticles, J. Phys. Chem. Lett. 4 (2013) 3747–3752, https://doi.org/
India is gratefully acknowledged. The authors would also like to declare 10.1021/jz401874u.
[24] D.F. Moyano, K. Saha, G. Prakash, B. Yan, H. Kong, M. Yazdani, V.M. Rotello, Fabrica-
that there is no conflict of interest in publication of this article.
tion of corona-free nanoparticles with tunable hydrophobicity, ACS Nano 8 (2014)
6748–6755, https://doi.org/10.1021/nn5006478.
References [25] V. Torrisi, A. Graillot, L. Vitorazi, Q. Crouzet, G. Marletta, C. Loubat, J.F. Berret,
Preventing corona effects: Multiphosphonic acid poly(ethylene glycol) copolymers
[1] D. Peer, J.M. Karp, S. Hong, O.C. Farokhzad, R. Margalit, R. Langer, Nanocarriers as an for stable stealth iron oxide nanoparticles, Biomacromolecules. 15 (2014)
emerging platform for cancer therapy, Nat. Nanotechnol. 2 (2007) 751–760, https:// 3171–3179, https://doi.org/10.1021/bm500832q.
doi.org/10.1038/nnano.2007.387. [26] K. Sempf, T. Arrey, S. Gelperina, T. Schorge, B. Meyer, M. Karas, J. Kreuter, Adsorption
[2] Y. Il Chung, J.C. Kim, Y.H. Kim, G. Tae, S.Y. Lee, K. Kim, I.C. Kwon, The effect of surface of plasma proteins on uncoated PLGA nanoparticles, Eur. J. Pharm. Biopharm. 85
functionalization of PLGA nanoparticles by heparin- or chitosan-conjugated Pluronic (2013) 53–60, https://doi.org/10.1016/j.ejpb.2012.11.030.
on tumor targeting, J. Control. Release 143 (2010) 374–382, https://doi.org/10.1016/ [27] X. Song, Y. Zhao, W. Wu, Y. Bi, Z. Cai, Q. Chen, Y. Li, S. Hou, PLGA nanoparticles simul-
j.jconrel.2010.01.017. taneously loaded with vincristine sulfate and verapamil hydrochloride: systematic
[3] S. Alqahtani, L. Simon, C.E. Astete, A. Alayoubi, P.W. Sylvester, S. Nazzal, Y. Shen, Z. study of particle size and drug entrapment efficiency, Int. J. Pharm. 350 (2008)
Xu, A. Kaddoumi, C.M. Sabliov, Cellular uptake, antioxidant and antiproliferative ac- 320–329, https://doi.org/10.1016/j.ijpharm.2007.08.034.
tivity of entrapped α-tocopherol and γ-tocotrienol in poly (lactic-co-glycolic) acid [28] S. Das, P. Roy, R.G. Auddy, A. Mukherjee, Silymarin nanoparticle prevents
(PLGA) and chitosan covered PLGA nanoparticles (PLGA-Chi), J. Colloid Interface paracetamol-induced hepatotoxicity, Int. J. Nanomedicine 6 (2011) 1291–1301,
Sci. 445 (2015) 243–251, https://doi.org/10.1016/j.jcis.2014.12.083. https://doi.org/10.2147/IJN.S15160.
[4] X. Xu, W. Ho, X. Zhang, N. Bertrand, O. Farokhzad, Cancer nanomedicine: from [29] F. Bonaccorso, M. Zerbetto, A.C. Ferrari, V. Amendola, Sorting nanoparticles by cen-
targeted delivery to combination therapy, Trends Mol. Med. 21 (2015) 223–232, trifugal fields in clean media, J. Phys. Chem. C 117 (2013) 13217–13229, https://
https://doi.org/10.1016/j.molmed.2015.01.001. doi.org/10.1021/jp400599g.
176 A. Basu et al. / Journal of Molecular Liquids 282 (2019) 169–176

[30] G.M. Xiong, S. Yuan, J.K. Wang, A.T. Do, N.S. Tan, K.S. Yeo, C. Choong, Imparting [45] S. Milani, F. Baldelli Bombelli, A.S. Pitek, K.A. Dawson, J. Rädler, Reversible versus ir-
electroactivity to polycaprolactone fibers with heparin-doped polypyrrole: modula- reversible binding of transferrin to polystyrene nanoparticles: soft and hard corona,
tion of hemocompatibility and inflammatory responses, Acta Biomater. 23 (2015) ACS Nano 6 (2012) 2532–2541, https://doi.org/10.1021/nn204951s.
240–249, https://doi.org/10.1016/j.actbio.2015.05.003. [46] S. Winzen, S. Schoettler, G. Baier, C. Rosenauer, V. Mailaender, K. Landfester, K. Mohr,
[31] M.J. Hajipour, S. Laurent, A. Aghaie, F. Rezaee, M. Mahmoudi, Biomaterials Science 2 Complementary analysis of the hard and soft protein corona: sample preparation
(2014) 1210–1221, https://doi.org/10.1039/c4bm00131a Biomater. Sci. critically effects corona composition, Nanoscale. 7 (2015) 2992–3001, https://doi.
[32] M.M. Bradford, A rapid and sensitive method for the quantitation of microgram org/10.1039/C4NR05982D.
quantities of protein utilizing the principle of protein-dye binding, Anal. Biochem. [47] S. Dimchevska, N. Geskovski, R. Koliqi, N. Matevska-Geskovska, V. Gomez Vallejo, B.
72 (1976) 248–254, https://doi.org/10.1016/0003-2697(76)90527-3. Szczupak, E.S. Sebastian, J. Llop, D.R. Hristov, M.P. Monopoli, G. Petruševski, S.
[33] L. Fu, Y. Sun, L. Ding, Y. Wang, Z. Gao, Z. Wu, S. Wang, W. Li, Y. Bi, Mechanism eval- Ugarkovic, A. Dimovski, K. Goracinova, Efficacy assessment of self-assembled
uation of the interactions between flavonoids and bovine serum albumin based on PLGA-PEG-PLGA nanoparticles: correlation of nano-bio interface interactions,
multi-spectroscopy, molecular docking and Q-TOF HR-MS analyses, Food Chem. 203 biodistribution, internalization and gene expression studies, Int. J. Pharm. 533
(2016) 150–157, https://doi.org/10.1016/j.foodchem.2016.01.105. (2017) 389–401, https://doi.org/10.1016/j.ijpharm.2017.05.054.
[34] S. Chakraborti, S. Sarwar, P. Chakrabarti, The effect of the binding of zno nanoparticle [48] Y.M. Thasneem, M.R. Rekha, S. Sajeesh, C.P. Sharma, Biomimetic mucin modified
on the structure and stability of α-lactalbumin: a comparative study, J. Phys. Chem. PLGA nanoparticles for enhanced blood compatibility, J. Colloid Interface Sci. 409
B 117 (2013) 13397–13408, https://doi.org/10.1021/jp404411b. (2013) 237–244, https://doi.org/10.1016/j.jcis.2013.07.004.
[35] J. Juárez, S.G. López, A. Cambón, P. Taboada, V. Mosquera, Influence of electrostatic [49] C.D. Hoemann, D. Fong, Immunological Responses to Chitosan for Biomedical Appli-
interactions on the fibrillation process of human serum albumin, J. Phys. Chem. B cations, Elsevier, 2016https://doi.org/10.1016/B978-0-08-100230-8.00003-0.
113 (2009) 10521–10529, https://doi.org/10.1021/jp902224d. [50] N. Weber, H.P. Wendel, G. Ziemer, Hemocompatibility of heparin-coated surfaces
[36] P. Roy, S. Das, a Mondal, U. Chatterji, A Mukherjee, nanoparticle engineering en- and the role of selective plasma protein adsorption, Biomaterials. 23 (2002)
hances anticancer efficacy of andrographolide in MCF-7 cells and mice bearing 429–439, https://doi.org/10.1016/S0142-9612(01)00122-3.
EAC, Curr. Pharm. Biotechnol. 13 (2012) 2669–2681. http://www.ncbi.nlm.nih. [51] M. Mahmoudi, H.R. Kalhor, S. Laurent, I. Lynch, Protein fibrillation and nanoparticle
gov/pubmed/23072387. interactions: opportunities and challenges, Nanoscale 5 (2013) 2570, https://doi.
[37] C. Gunawan, M. Lim, C.P. Marquis, R. Amal, Nanoparticle-protein corona complexes org/10.1039/c3nr33193h.
govern the biological fates and functions of nanoparticles, J. Mater. Chem. B 2 (2014) [52] S.K. Chaturvedi, M.K. Siddiqi, P. Alam, R.H. Khan, Protein misfolding and aggrega-
2060–2083, https://doi.org/10.1039/c3tb21526a. tion: mechanism, factors and detection, Process Biochem. 51 (2016) 1183–1192,
[38] M.M. Yallapu, N. Chauhan, S.F. Othman, V. Khalilzad-sharghi, M.C. Ebeling, S. Khan, https://doi.org/10.1016/j.procbio.2016.05.015.
M. Jaggi, S.C. Chauhan, Implications of protein corona on physico-chemical and bio- [53] H. Levine, Thioflavine T interaction with synthetic Alzheimer's disease β-amyloid
logical properties of magnetic nanoparticles, Biomaterials. 46 (2015) 1–12, https:// peptides: detection of amyloid aggregation in solution, Protein Sci. 2 (1993)
doi.org/10.1016/j.biomaterials.2014.12.045. 404–410, https://doi.org/10.1002/pro.5560020312.
[39] D.H. Napper, A. Netschey, Studies of the steric stabilization of colloidal particles, J. [54] P.D. Gorevic, T.T. Casey, W.J. Stone, C.R. DiRaimondo, F.C. Prelli, B. Frangione, Beta-2
Colloid Interface Sci. 37 (1971) 528–535, https://doi.org/10.1016/0021-9797(71) microglobulin is an amyloidogenic protein in man, J. Clin. Invest. 76 (1985)
90330-4. 2425–2429, https://doi.org/10.1172/JCI112257.
[40] T. Sen, S. Mandal, S. Haldar, K. Chattopadhyay, A. Patra, Interaction of gold nanopar- [55] J. Lozier, N. Takahashi, F.W. Putnam, Complete amino acid sequence of human
ticle with human serum albumin (HSA) protein using surface energy transfer, J. plasma beta 2-glycoprotein I, Proc. Natl. Acad. Sci. U. S. A. 81 (1984) 3640–3644,
Phys. Chem. C 115 (2011) 24037–24044, https://doi.org/10.1021/jp207374g. https://doi.org/10.1073/pnas.81.12.3640.
[41] L. Bekale, D. Agudelo, H.a. Tajmir-Riahi, Effect of polymer molecular weight on [56] P. Taboada, S. Barbosa, E. Castro, V. Mosquera, Amyloid fibril formation and other ag-
chitosan-protein interaction, Colloids Surfaces B Biointerfaces. 125 (2015) gregate species formed by human serum albumin association, J. Phys. Chem. B 110
309–317, https://doi.org/10.1016/j.colsurfb.2014.11.037. (2006) 20733–20736, https://doi.org/10.1021/jp064861r.
[42] W.K. Surewicz, H.H. Mantsch, D. Chapman, Determination of protein secondary [57] J.S. Gebauer, M. Malissek, S. Simon, S.K. Knauer, M. Maskos, R.H. Stauber, W. Peukert,
structure by Fourier transform infrared spectroscopy: a critical assessment, Bio- L. Treuel, Impact of the nanoparticle − protein Corona on colloidal stability and pro-
chemistry. 32 (1993) 389–394, https://doi.org/10.1021/bi00053a001. tein structure, Langmuir 28 (2012) 9673–9679.
[43] J.F. Carpenter, J.H. Crowe, An infrared spectroscopic study of the interactions of car- [58] J. Juárez, P. Taboada, V. Mosquera, Existence of different structural intermediates on
bohydrates with dried proteins, Biochemistry. 28 (1989) 3916–3922, https://doi. the fibrillation pathway of human serum albumin, Biophys. J. 96 (2009) 2353–2370,
org/10.1021/bi00435a044. https://doi.org/10.1016/j.bpj.2008.12.3901.
[44] D.M. Mann, E. Romm, M. Migliorini, Delineation of the glycosaminoglycan-binding [59] M.K. Siddiqi, P. Alam, S.K. Chaturvedi, R.H. Khan, Anti-amyloidogenic behavior and
site in the human inflammatory response protein lactoferrin, J. Biol. Chem. 269 interaction of Diallylsulfide with human serum albumin, Int. J. Biol. Macromol. 92
(1994) 23661–23667. (2016) 1220–1228, https://doi.org/10.1016/j.ijbiomac.2016.08.035.

You might also like