11.06.2013 Views

Self-Assembly of Synthetic and Biological Polymeric Systems of ...

Self-Assembly of Synthetic and Biological Polymeric Systems of ...

Self-Assembly of Synthetic and Biological Polymeric Systems of ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

properties induce solvent-adapted structural “responses” <strong>of</strong> the<br />

aggregating protein, which leads to a structural diversity <strong>of</strong> the<br />

resulting aggregates/fibers. 21-23 In particular, alcohols are<br />

included among the most commonly studied cosolvents. Several<br />

studies have shown that the effects <strong>of</strong> alcohol on very different<br />

systems <strong>and</strong> processes such as the thermal denaturation <strong>of</strong> nucleic<br />

acids <strong>and</strong> proteins, protein folding, micellization <strong>of</strong> surfactant<br />

molecules, or the solubility <strong>of</strong> nonelectrolytes are strikingly<br />

similar. 24-26 Also, water/alcohol mixtures at moderately low pH<br />

values have been suggested as model systems for studying the<br />

joint action <strong>of</strong> the local decrease in both pH <strong>and</strong> dielectric<br />

constant on the protein structure near the membrane surface; 27-32<br />

similarly, new ways <strong>of</strong> protein administration <strong>and</strong> delivery based<br />

on inhalation systems with non-aqueous solvents as suspension<br />

media also reinforce the necessity <strong>of</strong> clarifying how the presence<br />

<strong>of</strong> the solvent affects the protein conformation <strong>and</strong> biological<br />

24, 25<br />

activity.<br />

From a thermodynamic st<strong>and</strong>point, the protein accessible<br />

surface area (ASA), the solvent exposure <strong>of</strong> hydrophobic<br />

residues, <strong>and</strong> a solvophobic backbone 33 can be taken as barriers<br />

hampering protein-solvent interactions. 34 Therefore, the<br />

hierarchical assembly <strong>of</strong> amyloids can perfectly be understood as<br />

an alternative to the native packing conformational struggle <strong>of</strong> a<br />

polypeptide chain, reducing ASA <strong>and</strong> saturating hydrogen<br />

bonding, while a simultaneous decrease in the configurational<br />

entropy <strong>of</strong> the protein is <strong>of</strong>fset by gains in solvent entropy. The<br />

role <strong>of</strong> hydrational forces in protein aggreation in vivo, while still<br />

poorly understood, may be instrumental in elucidating the<br />

molecular basis <strong>of</strong> amyloidosis <strong>and</strong>, as such, attracts considerable<br />

interest. 35 In this way, the propensity for amyloid formation in<br />

mixed solvents <strong>of</strong> different proteins as, for example, human<br />

serum albumin (HSA), hen egg white lysozyme, bovine βlactoglobulin,<br />

histone H2A, insulin, <strong>and</strong> bovine trypsinogen<br />

amongst many others, has been characterized. 36-39 Amongst them,<br />

HSA has been proposed as a good model for protein aggregation<br />

studies 40-42 as a consequence <strong>of</strong> both its physiological<br />

implications as a carrier protein <strong>and</strong> blood pressure regulator,<br />

together with its propensity to easily aggregate in vitro. Previous<br />

reports have addressed the HSA fibrillation kinetics, fibrillation<br />

pathway, <strong>and</strong> intermediate <strong>and</strong> final protein aggregated structures<br />

under varying pH <strong>and</strong> ionic strength solution conditions. 40-42<br />

However, a detailed characterization <strong>of</strong> the aggregation/fibril<br />

process <strong>and</strong> the structure <strong>of</strong> the resulting aggregates in mixed<br />

solvent solutions under varying external conditions is still lacked.<br />

In the present work, we present a detailed study on the<br />

aggregation/fibrillation pathway <strong>of</strong> the protein human serum<br />

albumin (HSA) in water/ethanol mixed solvent solutions at<br />

varying pH <strong>and</strong> temperature conditions. This enables us to make<br />

a clear comparison between the fibrillation pathway <strong>and</strong><br />

intermediate/final protein aggregates in the mixed solvent with<br />

those obtained in pure aqueous solution in order to define the role<br />

<strong>of</strong> hydrational forces on the protein fibrillation mechanism. In<br />

fact, changes on the fibrillation mechanism from a downhill to a<br />

nucleated-growth polymerization mechanism at acidic pH <strong>and</strong><br />

high temperature solution conditions are confirmed. Overall, as a<br />

result <strong>of</strong> differences in packing depending on the solution<br />

condition, different structures for the resulting fibrils are<br />

observed, from classically thin <strong>and</strong> very long straight to shorter<br />

5<br />

10<br />

15<br />

20<br />

25<br />

30<br />

35<br />

40<br />

45<br />

50<br />

55<br />

<strong>and</strong> more curly fibers. In this regard, it comes as no surprise that<br />

factors favoring or disfavoring exposure <strong>of</strong> solvophobic groups<br />

must affect thermodynamic preferences for particular<br />

conformations.<br />

2 | Journal Name, [year], [vol], 00–00 This journal is © The Royal Society <strong>of</strong> Chemistry [year]<br />

186<br />

60<br />

65<br />

Materials <strong>and</strong> Methods<br />

Materials.<br />

Human serum albumin (70024-90-7), <strong>and</strong> Thi<strong>of</strong>lavin T (ThT)<br />

were obtained from Sigma Chemical Co. ThT was used as<br />

70 received. All other chemicals were <strong>of</strong> the highest purity available.<br />

75<br />

80<br />

85<br />

90<br />

95<br />

100<br />

105<br />

110<br />

115<br />

Preparation <strong>of</strong> HSA solutions.<br />

HSA was purified by liquid chromatography using a<br />

Superdex 75 column equilibrated with 0.01 M phosphate buffer<br />

before use. To prepare the protein solutions, increasing volumes<br />

<strong>of</strong> ethanol were added to 1.0 mL <strong>of</strong> HSA stock solution either at<br />

pH 7.4 (sodium monophosphate-sodium diphosphate buffer) or<br />

pH 2.0 (glycine + HCl buffer), to get the desired ethanol<br />

concentration in the mixed solvent. The final protein<br />

concentration in all cases was 10 mg/mL with ionic strength 10<br />

mM. pH was kept constant by adding HCl or NaOH when<br />

needed. The protein concentration was determined<br />

spectrophotometrically, using a molar absorption coefficient <strong>of</strong><br />

35219 M -1 cm -1 at 280 nm. 43 Aliquots were added through an<br />

electronic disperser unit Dosimat Metrohm 765. Experiments<br />

were carried out using double distilled, deionized <strong>and</strong> degassed<br />

water. Before incubation, solutions were filtered through a 0.2<br />

µm filter into sterile test tubes. Samples were incubated at a<br />

specified temperature in a refluxed reactor for 15-20 days as<br />

required.<br />

Light Scattering.<br />

DLS <strong>and</strong> SLS intensities were measured at 25 °C by means <strong>of</strong><br />

an ALV-5000F (ALV-GmbH) instrument working with a<br />

vertically polarized incident light <strong>of</strong> wavelength λ = 488 nm<br />

supplied by a CW diode-pumped Nd; YAG solid-state laser<br />

supplied by Coherent, Inc. <strong>and</strong> operated at 400 mW. The intensity<br />

scale was calibrated against scattering from toluene.<br />

Measurements were carried out at a scattering angle θ = 90° to<br />

the incident beam. Solutions were equilibrated for 30 min before<br />

measurements to reach thermal stabilization. Experiment duration<br />

was in the range <strong>of</strong> 3-5 min, <strong>and</strong> each experiment was repeated<br />

two or more times. The correlation functions from DLS were<br />

analyzed using the CONTIN method to obtain intensity<br />

distributions <strong>of</strong> decay rates (Γ). 44 The decay rate distribution<br />

functions gave distributions <strong>of</strong> apparent diffusion coefficient<br />

(D app = Γ /q 2 ,where the scattering vector q = (4πns/λ)sin(θ /2),<br />

<strong>and</strong> ns is the refractive index <strong>of</strong> solvent) <strong>and</strong> integrating over the<br />

intensity distribution gave the intensity-weighted average <strong>of</strong> D . app<br />

Values <strong>of</strong> the apparent hydrodynamic radius (r , radius <strong>of</strong><br />

h,app<br />

hydrodynamically equivalent hard sphere corresponding to D ) app<br />

were calculated from the Stokes-Einstein equation:

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!