Soft and Hard Interactions between Polystyrene Nanoplastics and

Mar 1, 2019 - School of Chemical Sciences, The University of Auckland , Auckland ... The MacDiarmid Institute for Advanced Materials and Nanotechnolog...
0 downloads 0 Views 1MB Size
Subscriber access provided by WEBSTER UNIV

Article

Soft and hard interactions between polystyrene nanoplastics and human serum albumin protein corona Shinji Kihara, Nadine J van der Heijden, Chris Kinglsey Seal, Jitendra Mata, Andrew Whitten, Ingo Köper, and Duncan J. McGillivray Bioconjugate Chem., Just Accepted Manuscript • DOI: 10.1021/acs.bioconjchem.9b00015 • Publication Date (Web): 01 Mar 2019 Downloaded from http://pubs.acs.org on March 3, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Bioconjugate Chemistry

Soft and Hard Interactions between Polystyrene Nanoplastics and Human Serum Albumin Protein Corona Shinji Kihara1, 2, Nadine J. van der Heijden1, 2, Chris K. Seal1, 2, Jitendra P. Mata3, Andrew E. Whitten3, Ingo Köper4, and Duncan J. McGillivray*1, 2

1

School of Chemical Sciences, The University of Auckland, Auckland 1010, New Zealand

2 The

3

MacDiarmid Institute for Advanced Materials and Nanotechnology, Wellington 6140, New Zealand

Australian Centre for Neutron Scattering, Australian Nuclear Science and Technology Organisation, Lucas

Heights, NSW 2234, Australia 4

Institute for Nanoscale Science and Technology, College for Science and Engineering, Flinders

University, Adelaide, SA 5042, Australia KEYWORDS: nanoplastics, protein corona, soft corona, hard corona, human serum albumin, nanotoxicology

ACS Paragon Plus Environment

1

Bioconjugate Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 30

ABSTRACT

Upon contact with biological fluids, the surface of nanoparticles is surrounded by many types of proteins, forming a so-called “protein corona”. The physicochemical properties of the nanoparticle/corona complex depend predominantly on the nature of the protein corona. An understanding of the structure of the corona and the resulting complex provides insight into the structure-activity relationship. Here, we structurally evaluate the soft and hard components of protein corona, formed from polystyrene (PS) nanoplastics and human serum albumin (HSA). Using circular dichroism (CD) spectroscopy to elucidate the structure of HSA within the complex, we establish the effect of nanoparticle size and pH on the nature of the protein corona formed- whether hard, or soft. Despite the weak interaction between PS and the HSA corona, small angle neutron scattering (SANS) revealed the formation of complex structure that enhance the intermolecular interactions between HSA proteins, PS particles and the HS/PSA complexes. Fractal formation occurred under conditions where the interaction between PS and HSA was strong; and, increasing HSA concentrations suppressed the degree of aggregation. The size of the nanoparticles directly influenced the nature of the protein corona, with larger particles favouring the formation of a soft corona, due to the decreased PS-HSA attraction.

ACS Paragon Plus Environment

2

Page 3 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Bioconjugate Chemistry

INTRODUCTION With the significant amount of plastic waste in the environment, there is concern that the environmental concentration of polymeric nanoparticles (nanoplastics) will increase considerably in the future, due to the natural weathering1-2 and biodegradation3 of bulk plastic. A recent study4 also demonstrated that microplastics can fragment into nanoplastics when ingested by Antarctic krill (Euphausia superba). The toxicity of nanoplastics is known to be highly dependent on their physical and chemical properties. Nanoparticles that are intended to be taken up into biological systems (particularly nanoparticles used for biomedical applications) are carefully designed in order to ensure they are not toxic. Unlike these engineered nanoparticles, nanoplastics present in the environment have a mixture of sizes, elemental compositions, geometries, and surface functional groups because of uncontrolled manner in which they form. This uncontrolled nature of environmental nanoplastics contrasts with the highly controlled nature of engineered nanoparticles5-6. Engineered nanoparticles are designed to minimise the risk to health for intentional uses in humans. The random nature of environmental nanoplastics lack this design of protection. Toxicological studies conducted on marine organisms have demonstrated the potential adverse effects resulting from nanoplastics2, 5-8. Nanoplastics are also known to propagate up to higher ranks in food chains15, which may cause successively hazardous effects to

ACS Paragon Plus Environment

3

Bioconjugate Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 30

humans. Despite these established physiological effects, the mechanisms behind their toxic actions has been studied infrequently. Much attention has been paid to interactions between nanoparticles (in general) and cellular membranes9-12 in an attempt to elucidate the membrane disruption and penetration mechanisms. However, the research is still far away from identifying the origin of these adverse effects, on a molecular level. It is well established that the surfaces of nanoparticles are covered by various kinds of proteins and other biological molecules when in contact with a biological system13-17. As a result, the nanoparticles now exist as a new entity, a complex of nanoparticles with layers of proteins (so called protein corona13-14, 18), whose surface characteristics alter depending on its composition. The nature of protein coronae is also dependent on the intrinsic properties of the nanoparticles, their chemical compositions14, 19, shape20, and surface chemistry14, 19, 21. Proteins within coronae are divided into two groups: “hard” and “soft” coronae. Those that are adsorbed tightly on the surface are referred to as “hard” and are thought to have a more significant impact on the surface properties of the nanoparticles. Loosely bound nanoparticle/proteins complexes are called “soft” coronae. Although the biological relevance of the protein coronae is not fully understood, the proteins within hard coronae are known to partially lose their biochemical activities, resulting from modification to the secondary structure of proteins22. Intriguingly, proteins that interact softly with nanoparticle surfaces have also been reported to show decreasing enzymatic activities, despite the lack of conformational changes23. The

ACS Paragon Plus Environment

4

Page 5 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Bioconjugate Chemistry

biological relevance of the hard interaction has been frequently investigated22, 24-25 due to the their distinctive stability on nanoparticle surfaces18, 26. While the modified protein functionality in hard coronae can be explained by the disruption of protein structure, the same association cannot be made for soft coronae. Further assessment of protein interactions with nanoparticles is essential to elucidate the origin of modified protein functionality in soft coronae, and importantly the role of corona interactions with biological molecules. The present work investigates the structure of soft and hard protein coronae that form around nanoplastics which could shed light on different roles that each plays. This scientific question has been laid out in the past27-28, although to the best of our knowledge is yet to be studied in detail. Particularly, we investigate soft and hard protein coronae, which form from human serum albumin (HSA) on polystyrene (PS) colloids with a particular focus on the geometry of each corona, PS/HSA complex, and interaction between complexes. HSA was chosen as it has previously been found to be a component in both soft and hard corona19. This is, in part, due to its relative abundance in blood plasma19. PS colloids were chosen as the model nanoplastic, as the particle size can be changed without modifying the surface functional groups. Using PS allows to develop a controlled nanoparticle (NP), varying both surface chemistry and size. PS nanoparticles of two different sizes were used in order to assess the influence of size on the formation of protein coronae. In addition, the relative abundance of PS in the environment and its reported adverse

ACS Paragon Plus Environment

5

Bioconjugate Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 30

physiological effects make the results acquired from this model system biologically relevant29-30. CD spectroscopy was used to probe the secondary structure of HSA before and after the interaction with PS nanoplastics and to distinguish the formation soft corona from hard corona. The protein adsorption behaviour and resulting geometries of the PS/HSA complexes were characterised by DLS and SANS. In SANS the H2O/D2O ratio of solvent was adjusted to match the neutron scattering length density (SLDn) of the solvent with the PS particles (Table S1) to highlight the geometry of HSA component of the PS/HSA complexes for both the soft and the hard coronae.

RESULTS AND DISCUSSION Circular Dichroism (CD) Spectroscopy The secondary structure of HSA before and after the introduction of HSA was evaluated by CD spectroscopy. A distinction between soft and hard coronae is made by looking at the preservation of the secondary structure of the protein after introducing nanoparticles. If the secondary structure is preserved, then we define the corona as soft; if it is lost then we define the corona as hard. Irrespective of the PS nanoparticle size, the secondary structure of HSA was unaffected at pH 7.4 (Figure 1 left), making the interaction between PS particles and HSA soft. However, the CD spectrum observed for PS-s/HSA complex at pH 5.0 showed a partial unfolding of 𝛼 helices, 51% ( ± 2%) to 46% ( ± 2%) (Figure 1

ACS Paragon Plus Environment

6

Page 7 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Bioconjugate Chemistry

right), leading to a hard interaction. This conformational change is attributed to the direct adsorption of HSA onto the PS-s surface and the free energy minimisation within this newly formed complex31. The absence of the change in the CD spectra, before and after the introduction of PS-L at pH 5.0, confirmed the presence of soft interaction (Figure 1 right). The summary of secondary structure content of all the samples is found in Table S2 and S3.

Figure 1. CD spectra of HSA, before and after the introduction of PS-s and PS-L at pH 7.4 (left) and pH 5.0 (right).

Dynamic Light Scattering (DLS) and zeta potential measurements

ACS Paragon Plus Environment

7

Bioconjugate Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 30

Figure 2. Z-average distributions obtained by DLS for HSA with PS-s in pH 7.4 (a), in situ Z-average study of PS-s/HSA at pH 7.4 (b), PS-s/HSA at a higher concentration of HSA (c), PS-s/HSA complex at pH 5.0 (d), PS-L/HSA complex at pH 7.4 and 5.0 (e-f). The final concentration of PS and HSA used was 1.0 mg mL-1, unless otherwise, stated.

DLS is a frequently used technique to characterise the adsorbed layer thickness of colloidal suspension and resulting aggregates23, 32-33. Steady-state DLS experiments were conducted for the HSA (1.0 mg mL-1) and small PS-s (1.0 mg mL-1) prior to mixing, and the hydrodynamic diameter (Z) of the HSA/PS-s was measured after 1.0 h when the system was equilibrated (Figure 2a). A small increase of the average Z (~4.0 nm) was observed when the PS-s particles were mixed with HSA. This was followed by an in situ

ACS Paragon Plus Environment

8

Page 9 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Bioconjugate Chemistry

DLS experiment (Figure 2b) under the same conditions which revealed that the incorporation process took approximately 30 min to reach an increase in approximately 4.0 nm. The size distribution of PS-s/HSA complex at pH 5.0, acquired by multimodal analyses, is presented in Figure 2d. An increasing Z-average from 33 nm to 120 nm was confirmed as well as the formation of large aggregates (< 5 𝜇m). Consistent with the subtle increase of Z-average for PS-s/HSA soft corona complex at pH 7.4 (Figure 2), the PS-L/HSA soft corona complexes showed a small increase in the Z-average regardless of the pH, without larger aggregate formation.

Additionally, zeta potentials of these single components and complexes were measured to elucidate the nature of interactions that PS and HSA undertake. PS-s and PS-L both showed negative surface charges at each pH (Table S4), whereas, the zeta potential of HSA significantly became neutral as the pH was changed from 7.4 to 5.0 (-12.2 ± 3.2 mV to -4.4 ± 4.5 mV). The neutralisation of the HSA is caused by two factors: protonation of aspartic acid and glutamic acid, and protonation of histidine. Based on the calculation using Equation S1-2, 5% of aspartic acid (~2 molecules) and 17% of glutamic acid (~10 molecules) are protonated at pH 5.0, resulting in the suppression of net negative charge. Additionally, 90% of histidine undergo protonation (~14 molecules), and so the net charge becomes more positive. The zeta potentials of PS-s/HSA complex were -15.5 mV (

ACS Paragon Plus Environment

9

Bioconjugate Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 30

± 3.0 mV) and -15.8 mV ( ± 9.8 mV) at pH 7.4 and 5.0, respectively, both of which showed a similar extent of charge suppression in comparison with the free PS-s particles. Intriguingly, the zeta potential of PS-L/HSA at pH 7.4 was -43.4 mV ( ± 7.1 mV), showing a smaller degree of interaction between PS-L and HSA at pH 7.4, while at pH 5.0, the zeta potential value of -9.0 mV ( ± 6.8 mV) was observed. SANS characterisation of single components The diameters of the PS-s and PS-L were found to be 230 Å (with a polydispersity of 17%) and 2374 Å (with a polydispersity of 2%) based on the sphere model (Figure S1 and S2, and Table S5 and S6). The fitted diameters were smaller than the average size determined in the DLS study, which was expected due to the fact that the size resolved by DLS also takes into account of the movement of the particles34. The neutron scattering data for PSL

(9.0 mg mL-1) was collected in 25% d-buffer to ensure that the PS particles are contrast

matched, and no significant scattering feature was observed (Figure S3). The dimensions of HSA estimated by the fit of an ellipsoid to the neutron scattering data were 15.6 × 43.7 Å, which is in a reasonable agreement with previously reported values from the literature35 (Figure S4 and Table S7). SANS characterisation of PS/HSA complex system

ACS Paragon Plus Environment

10

Page 11 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Bioconjugate Chemistry

Figure 3. SANS profiles of PS-s/HSA complexes at pH 7.4 (left) and pH 5.0 (right, andd their corresponding individual components). The inset in both plots show the residual between the scattering intensities from the PS-s/HSA complex and the sum of individual components. The concentration of the PS-s used in the complex was fixed at 9.0 mg mL-1.

Figure 4. SANS profiles of PS-s/HSA complex at pH 7.4 in 100% d-buffer (left) and in 25% d-buffer with HSA (3.0 mg mL-1) collected in 100% d-buffer as a comparison (right). The SANS profile was fitted with sphere model with a structure factor (sticky hard sphere).

ACS Paragon Plus Environment

11

Bioconjugate Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 30

Table 1. Fitting parameters of sphere model with structure factor (sticky hard sphere) for PS-s/HSA complex in pH 7.4. The radius and the SLD contrast fitted here takes into account of both PS and HSA.

100% d-buffer at

Radius / Å

perturb

Stickiness

∆SLDn / 10-6 Å-2

135a ± 0.4

0.100 ±

0.0856 ±

4.45

0.003

0.0008

pH 7.4 PS-s/HSA

aThe

optimised polydispersity was 17%.

SANS profiles of HSA/PS complexes were divided into two distinctive q ranges for the convenience of describing two distinctive structural features at different length scales (Figure 3). The q range I encodes information relating to the structural features from individual PS-s and HSA, while q-range II encodes information relating to the geometry of PS-s/HSA complexes. The insets of Figure 3 show the residual scattering intensity after subtracting the sum of scattering from PS-s and HSA from PS-s/HSA complex. The small residual at q-range I at pH 7.4 reflects the absence or small portion of PS-s and HSA experiencing a direct interaction. The resulting scattering at q range I is best described by a sum of a sphere model and an ellipsoid model for q range I (Figure S5 and Table S8). Nevertheless, the q range II could not be described, suggesting that these models do not capture the geometrical arrangement of the complexes and/or its interparticle interaction. Hence, a combination of a sphere model and a structure factor (hard sticky sphere) was

ACS Paragon Plus Environment

12

Page 13 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Bioconjugate Chemistry

used to describe the q range II (Figure 4) with “stickiness” and “perturb” as a measure of the strength of interparticle interaction (PS-s/HSA – PS-s/HSA). To provide further insight, a SANS experiment was carried out in 25% d-buffer, which contrast matches the PS-s particles (Figure 4 right and Figure 5). Rg of the HSA/PS-s complex in 25% d-buffer resolved by Guinier analysis was approximately 134 Å which compares with 116 Å and 400 Å for the PS-s and PS-s/HSA complexes in 100% d-buffer, respectively (Figure S6). The resulting neutron scattering reflects that HSA participated in the formation of complexes, and likely that the HSA is decorated on the surface of PS-s spheres.

Figure 5. Proposed structural model of PS-s/HSA complexes in pH 7.4 (top) and pH 5.0 (bottom). The background of the figures on the right are coloured to portray the neutron scattering data collected using 25% d-buffer.

ACS Paragon Plus Environment

13

Bioconjugate Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 30

Figure 6. SANS profiles of PS-s/HSA complex in 100% and 0% d-buffer at pH 5.0, fitted with a fractal model. Table 2. Fitting parameters of fractal moodel for PS-s/HSA complex in pH 5.0.

PS-s/HSA at pH

Radius / Å

5.0

Fractal

Correlation length

dimension



∆SLDn / 10-6 Å-2

100% d-buffer

135 ± 0.23

2.3 ± 0.041

1556 ± 0.2

5.0

0% d-buffer

135 ± 0.35

2.3 ± 0.064

1556 ± 0.23

1.8

aOptimised

value of PD was 2%.

ACS Paragon Plus Environment

14

Page 15 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Bioconjugate Chemistry

Figure 7. SANS profiles of PS-s/HSA complex in 100% d-buffer at pH 5.0 with various HSA concentrations (left) and PS-s/HSA complex in 25% d-buffer at pH 5.0 with various HSA concentrations.

The SANS profiles of PS/HSA complex at pH 5.0 (Figure 3 inset right) shows that the scattering from PS-s/HSA complex can no longer be explained by the summation of scattering from individual components. This change in the q range I can be attributed to the change in the PS-HSA interface via direct adsorption. Fitting the q-range I at pH 5.0 with a core shell model supports the direct adsorption model with a partial coverage (Figure S7 and Table S9). Unlike pH 7.4, the upturn in q range II cannot be accommodated by a sphere model with structure factor (sticky hard sphere) with a given volume fraction. The aggregation of the colloidal system is better described by a fractal model, where both primary particles (PS-s/HSA complex) and aggregates (geometric arrangements of PSs/HSA

complexes) can be considered in q range I and q range II, respectively. Global

fitting was used to fit the scattering curves obtained in 100% d-buffer and 0% d-buffer in

ACS Paragon Plus Environment

15

Bioconjugate Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 30

Figure 6 (fitted parameters in Table 2). In particular, the SLDn of the solvent was fixed according to the D2O content in each, and the fractal dimension (Dm) and correlation length were fitted with appropriate constraints. The radius obtained for the model was 135 Å ( ± 0.4 Å) which took into account of the radius of primary particles (PS-s and HSA). For mass fractals, the power law I(q) ∝ q-Dm relationship holds, as opposed to the power law relationship for surface fractal36, described as I(q) ∝ q-6+Dm. Here, the power law fit for PS-s/HSA (with HSA concentration of 3.0 mg mL-1, presented in ) complex shows a relationship I(q) ∝ q-2.3. The observation that this characteristic upturn in the low q region lasts more than one decade in q and given the fractal dimension of 2.5 ( ± 0.04) for the power law fit, supports a mass fractal formation. SANS curves with various concentrations of HSA (0 – 15 mg mL-1) and a fixed concentration of PS-s (9.0 mg mL-1) are presented in Figure 7. The increasing scattering intensity in q region I is attributed to the scattering from individual HSA. The change in the steepness of the upturn in q region II, however, shows a modification to the geometry of the PS-s/HSA complex, which depends on the HSA concentration. Power law fitting in q region II showed a decreasing coefficient from q-2.3 to q-1.8, as the HSA concentration increased from 3.0 mg mL-1 to 15.0 mg mL-1. If the assumption that the power law relationship I(q) ∝ q-Dm holds for the PS-s/HSA complexes at all the HSA concentrations, the power law coefficients, q-P, are interchangeable with mass fractal coefficients. The mass fractal dimension describes the “compactness” of the fractal aggregates36-38, with

ACS Paragon Plus Environment

16

Page 17 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Bioconjugate Chemistry

more compact aggregates, with a higher local concentration, having a higher dimension, Dm (cartoon shown in Figure S8). The decrease of fractal dimension with an increasing HSA concentration can be explained by charge screening of the PS-s by HSA bound to the surface. With an increasing concentration of HSA on the PS-s surface, additional HSA is more likely to experience electrostatic repulsion from HSA bound to other PS/HSA complexes, leaving a spatially limited adsorption site on PS. This results in a less branched fractal aggregate (low Dm), suggesting that minimal aggregation would be observed with an excess HSA concentration. Further evidence of this structural model can be found in SANS curves for PS-s/HSA with two HSA concentrations (3.0 and 9.0 mg mL-1) contrast matched in 25% d-buffer. The increase from q-1.6 to q-2.1 in these SANS curves (Figure 7 right) reflects a local aggregation of HSA on PS-s surfaces. The SANS profile of PS-L/HSA complexes at pH 7.4 shows that the interaction between the PS-L and HSA is minimal (Figure 8 left and that a combination of spherical and ellipsoid model could accommodate the scattering profiles over the entire q range (Figure 8, Table S10). For pH 5.0, structure factor (sticky hard sphere) was added to the model to describe the upturn at a low q range (Fitting parameters shown in table S11).

ACS Paragon Plus Environment

17

Bioconjugate Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 30

Figure 8. SANS profiles of PS-L/HSA complexes at pH 7.4 (left) and pH 5.0 (right). The scattering profiles of PS-L is fitted with sphere model. The scattering profile of PS-L/HSA (3.0 mg mL-1) complex at pH 5.0 is fitted with a combined model of sphere, ellipsoid, and structure factor (sticky hard sphere), else the scattering profiles are fitted with a combination of sphere and ellipsoid models.

Figure 9. Crystal structure (1AO6)39 of HSA from two different angles with colour codes (red for positively chargeable amino acids and blue for negatively chargeable amino acids), portraying the potential sites for PS particles to experience an electrostatic interaction.

ACS Paragon Plus Environment

18

Page 19 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Bioconjugate Chemistry

Soft and hard HSA corona around PS nanoplastics The origin of the interaction between the HSA soft corona with PS surface is maintained by a subtle balance of the electrostatic interaction. Despite having an overall negative charge, the HSA surface is a mixture of negatively and positively charged amino acids as shown in Figure 9. This weak interaction observed in HSA soft corona is attributed to an attractive interaction between the positively charged residues and the PS surface, while negatively charged residues, particularly those in the vicinity of the positively charged residues, inhibit a strong interaction. When the pH was lowered down to 5.0, protonated histidine (providing positive charge), aspartic acid and glutamic acid (losing negative charges) enhanced the electrostatic attraction. As a result, HSA experienced a strong attraction to PS particles, forming larger aggregates. However, greater HSA concentrations led to a lesser degree of aggregation. The finding suggested that an excessive HSA concentration would result in an extensive coverage of HSA, and that protein molecules on nanoparticle surface mitigate interactions with other PS particles. We expect that a real biological system follows such case. When compared with the smaller analogue, PS-L exhibited a weaker attraction to HSA. In the literature, the binding affinities of proteins increase as the binding site becomes less curved due to the reduced free energy22-23, 40. However, the opposite effect was reported

ACS Paragon Plus Environment

19

Bioconjugate Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 30

for TiO2 nanoparticles with pepsin, where smaller size facilitated more proteins on the surface than its larger analogue41. Our finding for PS/HSA complexes follows such an example. Many of the protein-protein interactions are known to be transient, yet significant in their functions42-44. For soft corona, despite the absence of conformational change, the interparticle interaction between the PS/HSA – PS/HSA and HSA/HSA, could alter the way native proteins interact with other proteins. However, to correlate the observed interparticle interaction between soft coronae to their biological relevance, an assessment of protein functionality, such as enzymatic activities, is required. Nevertheless, the presence of interparticle interactions would be an important parameter to add to the current understanding of soft coronae.

CONCLUSION We studied the interaction between negatively charged PS nanoparticles and HSA, with a particular focus on the structure of soft and hard protein coronae. CD spectroscopy was first used as a diagnostic tool to identify the nature of protein corona at each condition. A hard interaction was observed for PS-s/HSA complex at pH 5.0, though only soft interaction was observed with PS-L/HSA at pH 5.0. DLS revealed differences in this

ACS Paragon Plus Environment

20

Page 21 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Bioconjugate Chemistry

interaction, based on size distributions and kinetic studies, highlighting a subtle increase in the Z-average for soft interactions and aggregate formation for hard interactions. Specific geometries of HSA soft and hard corona within PS/HSA complexes were evaluated by SANS. The SANS profiles of the complex system strongly supported the presence of interparticle interaction between PS/HSA-PS/HSA complexes. The proposed structural model of a soft corona provides additional information to the existent understanding of the soft corona. At pH 5.0, the formation of larger aggregates for PSs/HSA

complexes was confirmed both by DLS and SANS. The resulting aggregates were

described by fractal model. Furthermore, we found that an increasing HSA concentration lessens the degree of aggregation by charge screening. The finding suggested that with excess HSA, the aggregation caused by hard corona could be minimised. Although the soft and hard components of has corona around PS nanoplastics were characterised in this work, it should be noted that these structures may change when soft and hard components co-exist, and the types of nanoplastics and proteins in the corona should also play an important role. In parallel to structural analyses, protein functionality and their effect on other biological molecules would be crucial to illustrate the biological relevance of protein coronae. METHODS AND MATERIALS

ACS Paragon Plus Environment

21

Bioconjugate Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 30

Carboxylic-terminated polystyrene nanoplastics (PS-s) were purchased from SigmaAldrich (CML Latex beads, 0.02 𝜇m). Larger polystyrene nanoplastics functionalised with carboxylic acid (PS-L) were synthesised via an emulsion-free polymerisation method described elsewhere45. The synthesised PS-L was dialysed extensively against ultrapure water (Milli-Q, resistivity = 18.2 M𝛺 cm) prior to further characterisation. NaCl (ECP, 99.5%), KCl (99.5%, ECP), Na2HPO4 (99%, Sigma Aldrich), KH2PO4 (99.5%), and citric acid (99.5%, Sigma Aldrich) were used to prepare buffer solutions. For neutron experiments, both PS-s and PS-L were dialysed twice, over a period of 24 h, against D2O (provided by the Australian Nuclear Science and Technology Organisation (ANSTO)). Unless stated, all the reagents were used without further purification.

Circular Dichroism (CD) spectroscopy Circular dichroism (CD) experiments were performed on a Chirascan spectrometer (Applied Photophysics, UK) over the range of 180 nm to 360 nm, at 0.5 nm intervals, using a quartz cuvette of 1.0 mm path length (Hellma cells Pacific, Singapore). The spectra were recorded for the samples (400 µL) at a scan rate of 0.5 s per point and 1.0 nm bandwidth. A minimum of five scans were averaged prior to the data conversion to absolute CD values. HSA and PS nanoplastics were dispersed in 10 mM sodium phosphate buffer

ACS Paragon Plus Environment

22

Page 23 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Bioconjugate Chemistry

without saline at pH 7.4. The raw CD signal (𝜃) was, then, converted to mean residual ellipticity ([𝜃]MRE) using the equation below:

[𝜃]𝑀𝑅𝐸 =

𝜃 𝐶𝑟𝑙

where, Cr is protein concentration and 𝑙 is the cuvette path length (1.0 mm). Spectral deconvolution was performed using CONTIN algorithm46 with a reference set of SP17547 available on DICHROWEB46.

Dynamic Light Scattering (DLS) and zeta potential measurements Particle size distribution and zeta potential analyses were carried out on a Malvern Instruments Nanozetasizer ZS equipped with a 633 nm laser. The size distribution of PS nanoplastics and HSA in PBS buffer (pH 7.4) was measured in non-invasive back scatter (NIBS) mode, with the photomultiplier detector placed at 173° relative to the incident beam, a measurement path length of 3.00 mm and an attenuation factor of 9. The temperature of the samples was maintained at 23 °C with 120 s of temperature equilibration time. The measurements comprised of 14 runs and with three repeats combined to provide average results. For the time resolved experiments, each run comprised of 10 s data acquisition with six repeats, making the time interval at each point 60 s. For zeta potential determination, the Smoluchowski approximation was used.

ACS Paragon Plus Environment

23

Bioconjugate Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 30

Small-angle Neutron Scattering (SANS) measurements and data analysis SANS measurements were carried out on BILBY, the time-of-flight small-angle neutron scattering (SANS) instrument, at ANSTO, Lucas Heights, Australia48. Samples were suspended in three different deuterated buffers (d-buffer) at the following H2O:D2O ratios- 0%, 25%, and 100%. The 25% d-buffer was used to mask the neutron signals from PS particles. The samples were placed in a 1.0 mm path-length quartz Hellma cell. The buffer solutions at pH 7.4 contained Na2HPO4 (10 mM), KH2PO4 (3 mM), NaCl (132 mM) and KCl (2.7 mM), and the pH was adjusted with NaOH dissolved in D2O (50 𝜇M). For the acidic buffers, citric acid (10 mM) and NaCl (132 mM) was used with NaOH dissolved in D2O to make pH 5.0 solution. Scattering data were radially averaged (under the assumption of isotropic scattering) and placed on an absolute scale using the direct beam intensity. The value of q was defined as

𝑞=

4𝜋𝑠𝑖𝑛𝜃 𝜆

where 𝜆 is the wavelength of the incident neutron beam and 2𝜃 is the angle of scattering. For the PS-s and HSA, neutrons with wavelengths between 2–20 Å were collected on 2dimensional detector panels positioned at 7.000 (rear), 4.500 (horizontal curtains) and 3.500 m (vertical curtains) from the sample (q-range = ~0.002 – 0.340 Å-1, 𝛥𝜆/𝜆 = 12.4-

ACS Paragon Plus Environment

24

Page 25 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Bioconjugate Chemistry

14.3%). For PS-L and PS/HSA complexes, the rear detector was moved to 17.000 m from the sample (q-range = ~0.001 – 0.340 Å-1, 𝛥𝜆/𝜆 = 9.0 - 14.3%). Neutron

scattering

curves

were

fitted

using

the

SasView

software

(http://www.sasview.org/). The scattering intensity from bare PS particles were fitted with a sphere model 49, and that of HSA was fitted with a ellipsoid model50. For a PS/HSA complex system, a sphere model with a structure factor (sticky hard sphere)51, and a fractal model37 were used. The radius of gyration (Rg) was resolved using the AutoRg function in the ATSAS package (2.8.2)52.

ACS Paragon Plus Environment

25

Bioconjugate Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 30

ACKNOWLEDGEMENT S. Kihara thanks The University of Auckland for the doctoral scholarship and would also like to thank AINSE Limited for providing financial assistance (Award - PGRA) to enable work on the research. We would also like to thank ANSTO for the provision of neutron beamtime (P6925 & P6766). Prof. John W. White (Australian National University) is also acknowledged for valuable discussions.

SUPPORTING INFORMATION Supporting information is available. CORRESPONDING AUTHOR Duncan J. McGillivray Email: [email protected]

REFERENCES 1. Lambert, S.; Wagner, M., Characterisation of nanoplastics during the degradation of polystyrene. Chemosphere 2016, 145, 265-268. 2. Hüffer, T.; Hofmann, T., Sorption of non-polar organic compounds by microsized plastic particles in aqueous solution. Environ. Pollut. 2016, 214, 194-201. 3. Harshvardhan, K.; Jha, B., Biodegradation of low-density polyethylene by marine bacteria from pelagic waters, Arabian Sea, India. Mar. Pollut. Bull. 2013, 77, 100-106. 4. Dawson, A. L.; Kawaguchi, S.; King, C. K.; Townsend, K. A.; King, R.; Huston, W. M.; Nash, S. M. B., Turning microplastics into nanoplastics through digestive fragmentation by Antarctic krill. Nat. Comm. 2018, 9, 1001. 5. Nolte, T. M.; Hartmann, N. B.; Kleijn, J. M.; Garnaes, J.; van de Meent, D.; Hendriks, A. J.; Baun, A., The toxicity of plastic nanoparticles to green algae as

ACS Paragon Plus Environment

26

Page 27 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Bioconjugate Chemistry

influenced by surface modification, medium hardness and cellular adsorption. Aquat. Toxicol. 2017, 183, 11-20. 6. Pitt, J. A.; Kozal, J. S.; Jayasundara, N.; Massarsky, A.; Trevisan, R.; Geitner, N.; Wiesner, M.; Levin, E. D.; Di Giulio, R. T., Uptake, tissue distribution, and toxicity of polystyrene nanoparticles in developing zebrafish (Danio rerio). Aquat. Toxicol. 2018, 194, 185-194. 7. Besseling, E.; Wang, B.; Lürling, M.; Koelmans, A. A., Nanoplastic affects growth of S. obliquus and reproduction of D. magna. Environ. Sci. Technol. 2014, 48, 1233612343. 8. Chen, Q.; Gundlach, M.; Yang, S.; Jiang, J.; Velki, M.; Yin, D.; Hollert, H., Quantitative investigation of the mechanisms of microplastics and nanoplastics toward zebrafish larvae locomotor activity. Sci. Total Environ. 2017, 584, 1022-1031. 9. Verma, A.; Stellacci, F., Effect of surface properties on nanoparticle–cell interactions. Small 2010, 6, 12-21. 10. Bulpett, J. M.; Snow, T.; Quignon, B.; Beddoes, C. M.; Tang, T. Y. D.; Mann, S.; Shebanova, O.; Pizzey, C. L.; Terrill, N. J.; Davis, S. A., et al., Hydrophobic nanoparticles promote lamellar to inverted hexagonal transition in phospholipid mesophases. Soft Matter 2015, 11, 8789-8800. 11. Karlsson, H. L.; Cronholm, P.; Hedberg, Y.; Tornberg, M.; De Battice, L.; Svedhem, S.; Wallinder, I. O., Cell membrane damage and protein interaction induced by copper containing nanoparticles—Importance of the metal release process. Toxicology 2013, 313, 59-69. 12. Pogodin, S.; Slater, N. K.; Baulin, V. A., Surface patterning of carbon nanotubes can enhance their penetration through a phospholipid bilayer. ACS Nano 2011, 5, 11411146. 13. Cedervall, T.; Lynch, I.; Lindman, S.; Berggård, T.; Thulin, E.; Nilsson, H.; Dawson, K. A.; Linse, S., Understanding the nanoparticle–protein corona using methods to quantify exchange rates and affinities of proteins for nanoparticles. Proceedings of the National Academy of Sciences 2007, 104, 2050-2055. 14. Casals, E.; Pfaller, T.; Duschl, A.; Oostingh, G. J.; Puntes, V., Time evolution of the nanoparticle protein corona. ACS Nano 2010, 4, 3623-3632. 15. Jayaram, D. T.; Runa, S.; Kemp, M. L.; Payne, C. K., Nanoparticle-induced oxidation of corona proteins initiates an oxidative stress response in cells. Nanoscale 2017, 9, 7595-7601. 16. Winzen, S.; Schoettler, S.; Baier, G.; Rosenauer, C.; Mailaender, V.; Landfester, K.; Mohr, K., Complementary analysis of the hard and soft protein corona: sample preparation critically effects corona composition. Nanoscale 2015, 7, 2992-3001. 17. Berts, I.; Fragneto, G.; Porcar, L.; Hellsing, M. S.; Rennie, A. R., Controlling adsorption of albumin with hyaluronan on silica surfaces and sulfonated latex particles. J. Colloid Interface Sci. 2017, 504, 315-324.

ACS Paragon Plus Environment

27

Bioconjugate Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 30

18. Vilanova, O.; Mittag, J. J.; Kelly, P. M.; Milani, S.; Dawson, K. A.; Rädler, J. O.; Franzese, G., Understanding the kinetics of protein–nanoparticle corona formation. ACS Nano 2016, 10, 10842-10850. 19. Tenzer, S.; Docter, D.; Kuharev, J.; Musyanovych, A.; Fetz, V.; Hecht, R.; Schlenk, F.; Fischer, D.; Kiouptsi, K.; Reinhardt, C., Rapid formation of plasma protein corona critically affects nanoparticle pathophysiology. Nat. Nanotechnol. 2013, 8, 772. 20. Carnovale, C.; Bryant, G.; Shukla, R.; Bansal, V., Impact of nanogold morphology on interactions with human serum. Phys. Chem. Chem. Phys. 2018. 21. Nattich-Rak, M.; Sadowska, M.; Adamczyk, Z.; Ciesla, M.; Kakol, M., Formation mechanism of human serum albumin monolayers on positively charged polymer microparticles. Colloid Surf. B-Biointerfaces 2017, 159, 929-936. 22. Vertegel, A. A.; Siegel, R. W.; Dordick, J. S., Silica nanoparticle aize influences the structure and enzymatic activity of adsorbed lysozyme. Langmuir 2004, 20, 6800-6807. 23. Wang, J.; Jensen, U. B.; Jensen, G. V.; Shipovskov, S.; Balakrishnan, V. S.; Otzen, D.; Pedersen, J. S.; Besenbacher, F.; Sutherland, D. S., Soft interactions at nanoparticles alter protein function and conformation in a size dependent manner. Nano Lett. 2011, 11, 4985-4991. 24. Deng, Z. J.; Liang, M.; Monteiro, M.; Toth, I.; Minchin, R. F., Nanoparticleinduced unfolding of fibrinogen promotes Mac-1 receptor activation and inflammation. Nat. Nanotechnol. 2011, 6, 39. 25. Kharazian, B.; Lohse, S. E.; Ghasemi, F.; Raoufi, M.; Saei, A. A.; Hashemi, F.; Farvadi, F.; Alimohamadi, R.; Jalali, S. A.; Shokrgozar, M. A., Bare surface of gold nanoparticle induces inflammation through unfolding of plasma fibrinogen. Sci. Rep. 2018, 8, 12557. 26. Walczyk, D.; Bombelli, F. B.; Monopoli, M. P.; Lynch, I.; Dawson, K. A., What the cell “sees” in bionanoscience. J. Am. Chem. Soc. 2010, 132, 5761-5768. 27. Monopoli, M. P.; Walczyk, D.; Campbell, A.; Elia, G.; Lynch, I.; Baldelli Bombelli, F.; Dawson, K. A., Physical− chemical aspects of protein corona: relevance to in vitro and in vivo biological impacts of nanoparticles. J. Am. Chem. Soc. 2011, 133, 2525-2534. 28. Monopoli, M. P.; Åberg, C.; Salvati, A.; Dawson, K. A., Biomolecular coronas provide the biological identity of nanosized materials. Nat. Nanotechnol. 2012, 7, 779. 29. Andrady, A. L., Persistence of plastic litter in the oceans. In Marine anthropogenic litter, Springer: Cham, 2015; pp 57-72. 30. Koelmans, A. A.; Besseling, E.; Shim, W. J., Nanoplastics in the Aquatic Environment. Critical Review. In Marine Anthropogenic Litter, Bergmann, M.; Gutow, L.; Klages, M., Eds. Springer: Cham, 2015; pp 325-340. 31. Norde, W. In Driving forces for protein adsorption at solid surfaces, Macromolecular Symposia, Wiley Online Library: 1996; pp 5-18.

ACS Paragon Plus Environment

28

Page 29 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Bioconjugate Chemistry

32. Yadav, I.; Aswal, V.; Kohlbrecher, J., Electrolyte effect on the phase behavior of silica nanoparticles with lysozyme and bovine-serum-albumin proteins. Phys. Rev. E 2015, 91, 052306. 33. Tripathi, B. P.; Dubey, N. C.; Subair, R.; Choudhury, S.; Stamm, M., Enhanced hydrophilic and antifouling polyacrylonitrile membrane with polydopamine modified silica nanoparticles. RSC Advances 2016, 6, 4448-4457. 34. Follens, L.; Aerts, A.; Haouas, M.; Caremans, T.; Loppinet, B.; Goderis, B.; Vermant, J.; Taulelle, F.; Martens, J.; Kirschhock, C. E., Characterization of nanoparticles in diluted clear solutions for Silicalite-1 zeolite synthesis using liquid 29Si NMR, SAXS and DLS. Phys. Chem. Chem. Phys. 2008, 10, 5574-5583. 35. Leis, D.; Barbosa, S.; Attwood, D.; Taboada, P.; Mosquera, V., Influence of the pH on the complexation of an amphiphilic antidepressant drug and human serum albumin. J. Phys. Chem. B 2002, 106, 9143-9150. 36. Besselink, R.; Stawski, T.; Van Driessche, A.; Benning, L., Not just fractal surfaces, but surface fractal aggregates: Derivation of the expression for the structure factor and its applications. J. Chem. Phys 2016, 145, 211908. 37. Teixeira, J., Small‐angle scattering by fractal systems. J. Appl. Crystallogr. 1988, 21, 781-785. 38. Thouy, R.; Jullien, R., Structure factors for fractal aggregates built off-lattice with tunable fractal dimension. Journal de Physique I 1996, 6, 1365-1376. 39. Sugio, S.; Kashima, A.; Mochizuki, S.; Noda, M.; Kobayashi, K., Crystal structure of human serum albumin at 2.5 Å resolution. Protein Eng. 1999, 12, 439-446. 40. Zhao, Y.; Sun, X.; Zhang, G.; Trewyn, B. G.; Slowing, I. I.; Lin, V. S.-Y., Interaction of mesoporous silica nanoparticles with human red blood cell membranes: size and surface effects. ACS Nano 2011, 5, 1366-1375. 41. Allouni, Z. E.; Gjerdet, N. R.; Cimpan, M. R.; Høl, P. J., The effect of blood protein adsorption on cellular uptake of anatase TiO2 nanoparticles. Int. J. Nanomed. 2015, 10, 687. 42. Nooren, I. M.; Thornton, J. M., Structural characterisation and functional significance of transient protein–protein interactions. J. Mol. Biol. 2003, 325, 991-1018. 43. Perkins, J. R.; Diboun, I.; Dessailly, B. H.; Lees, J. G.; Orengo, C., Transient protein-protein interactions: structural, functional, and network properties. Structure 2010, 18, 1233-1243. 44. Sato, T.; Fukasawa, T.; Shimozawa, T.; Komatsu, T.; Sakai, H.; Ishiwata, S., Protein-protein interactions in solution and their interplay with protein specific functions. J. Phys. Soc. Jpn. 2012, 81, 11. 45. Wu, Y. n.; Li, F.; Zhu, W.; Cui, J.; Tao, C. a.; Lin, C.; Hannam, P. M.; Li, G., Metal– organic frameworks with a three‐dimensional ordered macroporous structure: dynamic photonic materials. Angewandte Chemie International Edition 2011, 50, 12518-12522.

ACS Paragon Plus Environment

29

Bioconjugate Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 30

46. Whitmore, L.; Wallace, B., DICHROWEB, an online server for protein secondary structure analyses from circular dichroism spectroscopic data. Nucleic Acids Res. 2004, 32, W668-W673. 47. Lees, J. G.; Miles, A. J.; Wien, F.; Wallace, B., A reference database for circular dichroism spectroscopy covering fold and secondary structure space. Bioinformatics 2006, 22, 1955-1962. 48. Sokolova, A.; Christoforidis, J.; Eltobaji, A.; Barnes, J.; Darmann, F.; Whitten, A. E.; de Campo, L., BILBY: time-of-flight small angle scattering instrument. Neutron News 2016, 27, 9-13. 49. Guiner, A.; Fournet, G.; Walker, C., Small angle scattering of X-rays. J. Wiley & Sons, New York 1955. 50. Feigin, L.; Svergun, D. I.; Taylor, G. W., General principles of small-angle diffraction. In Structure analysis by small-angle X-ray and neutron scattering, Springer: New York, 1987; pp 25-55. 51. Menon, S.; Manohar, C.; Rao, K. S., A new interpretation of the sticky hard sphere model. J. Chem. Phys 1991, 95, 9186-9190. 52. Petoukhov, M. V.; Franke, D.; Shkumatov, A. V.; Tria, G.; Kikhney, A. G.; Gajda, M.; Gorba, C.; Mertens, H. D.; Konarev, P. V.; Svergun, D. I., New developments in the ATSAS program package for small-angle scattering data analysis. J. Appl. Crystallogr. 2012, 45, 342-350.

For Table of Content

ACS Paragon Plus Environment

30