
Changes in the structure and sizes of human and bovine serum albumins as well as polyvinyl alcohol macromolecules in aqueous solutions depending on temperature, concentration, and acid-base balance (pH) of the solutions are discussed. It is taken into consideration that the change in the hydrodynamic radius of a macromolecule is one of the indicators of structural phase transformations of globular proteins. The methods of the macromolecular radii determination from the shear viscosity and the self-diffusion of macromolecules in solutions are discussed. The hydrodynamic radius values of albumin and polyvinyl alcohol macromolecules obtained by the above methods are thoroughly compared. Consideration of these questions provides us with important information on the nature of the binding of water molecules with protein macromolecules.
Citation: Oleksii V. Khorolskyi, Nikolay P. Malomuzh. Macromolecular sizes of serum albumins in its aqueous solutions[J]. AIMS Biophysics, 2020, 7(4): 219-235. doi: 10.3934/biophy.2020017
[1] | Mostean Bahreinipour, Hajar Zarei, Fariba Dashtestani, Jamal Rashidiani, Khadijeh Eskandari, Seyed Ali Moussavi Zarandi, Susan Kabudanian Ardestani, Hiroshi Watabe . Radioprotective effect of nanoceria and magnetic flower-like iron oxide microparticles on gamma radiation-induced damage in BSA protein. AIMS Biophysics, 2021, 8(2): 124-142. doi: 10.3934/biophy.2021010 |
[2] | Yu-Ting Huang, Hui-Fen Liao, Shun-Li Wang, Shan-Yang Lin . Glycation and secondary conformational changes of human serum albumin: study of the FTIR spectroscopic curve-fitting technique. AIMS Biophysics, 2016, 3(2): 247-260. doi: 10.3934/biophy.2016.2.247 |
[3] |
Maria Teresa Caccamo, Giuseppe Mavilia, Letterio Mavilia, Pietro Calandra, Domenico Lombardo, Salvatore Magazù .
Thermal investigation of montmorillonite/BSA by fourier transform infrared spectroscopy measurements |
[4] | Anatoliy I. Fisenko, Oleksii V. Khorolskyi, Nikolay P. Malomuzh, Artur A. Guslisty . Relationship between the major parameters of warm-blooded organisms' life activity and the properties of aqueous salt solutions. AIMS Biophysics, 2023, 10(3): 372-384. doi: 10.3934/biophy.2023022 |
[5] | Nily Dan . Bilayer degradation in reactive environments. AIMS Biophysics, 2017, 4(1): 33-42. doi: 10.3934/biophy.2017.1.33 |
[6] | Thomas Schubert, Gernot Längst . Studying epigenetic interactions using MicroScale Thermophoresis (MST). AIMS Biophysics, 2015, 2(3): 370-380. doi: 10.3934/biophy.2015.3.370 |
[7] | Lajevardipour Alireza, W. M. Chon James, H. A. Clayton Andrew . Determining complex aggregate distributions of macromolecules using photobleaching image correlation microscopy. AIMS Biophysics, 2015, 2(1): 1-7. doi: 10.3934/biophy.2015.1.1 |
[8] | Shan-Yang Lin . Salmon calcitonin: conformational changes and stabilizer effects. AIMS Biophysics, 2015, 2(4): 695-723. doi: 10.3934/biophy.2015.4.695 |
[9] | Klemen Bohinc, Leo Lue . On the electrostatics of DNA in chromatin. AIMS Biophysics, 2016, 3(1): 75-87. doi: 10.3934/biophy.2016.1.75 |
[10] | Shivakumar Keerthikumar . A catalogue of human secreted proteins and its implications. AIMS Biophysics, 2016, 3(4): 563-570. doi: 10.3934/biophy.2016.4.563 |
Changes in the structure and sizes of human and bovine serum albumins as well as polyvinyl alcohol macromolecules in aqueous solutions depending on temperature, concentration, and acid-base balance (pH) of the solutions are discussed. It is taken into consideration that the change in the hydrodynamic radius of a macromolecule is one of the indicators of structural phase transformations of globular proteins. The methods of the macromolecular radii determination from the shear viscosity and the self-diffusion of macromolecules in solutions are discussed. The hydrodynamic radius values of albumin and polyvinyl alcohol macromolecules obtained by the above methods are thoroughly compared. Consideration of these questions provides us with important information on the nature of the binding of water molecules with protein macromolecules.
The structure and properties of proteins continue to be in the focus of attention of physicians, chemists, physicists, and other researchers. Features of changes in the protein structure, being dissolved in water and aqueous solutions, are the main research questions for physicists [1]. Herein we want to dwell on transformations that occur with albumin macromolecules in solutions. In this article we will focus on the macromolecules of human serum albumin (HSA) and bovine serum albumin (BSA).
Human serum albumin (HSA) consists of 585 amino acid residues combined into a single macromolecular chain with a molecular weight of 66.5 kDa [2]. In the crystalline state, the HSA macromolecule is folded into a compact conformation of a heart-shaped regular triangular prism with sizes of ∼80 Å and ∼30 Å [3]. Bovine serum albumin (BSA) consists of 583 amino acid residues combined into a macromolecule with a molecular weight of 66.5 kDa and a spatial structure similar to that of HSA [2],[4]. The amino acid sequences of HSA and BSA are 75.52% identical [5].
The domain structure of HSA and BSA is generally accepted. At physiological pH values, the secondary structure of HSA and BSA consists of alpha helices (50–68%) and beta folds (16–18%) stabilized by hydrogen bonds, as well as an unordered part of the macromolecular chain [2],[4],[6]. Due to 17 disulfide bonds between the cysteine residues of the alpha helices, the tertiary structure of these albumins is formed: three domains are formed, each of which is formed by subdomains of three alpha helices, and hydrophobic interactions between the domains determine the globular structure of proteins [2],[4].
Differences in the nucleotide sequences of albumins lead to some change in the hydrophobicity and conformational mobility of the HSA macromolecule compared to the BSA macromolecule [5]. One of the differences in the HSA structure is the presence of two tryptophan residues (Trp134 and Trp213) in the BSA macromolecular sequence, in contrast to one tryptophan residue in the HSA macromolecule (Trp214) [2].
When dissolved in water, as well as in aqueous and biological solutions due to interaction with water, the rigid conformation of albumin macromolecules should violate. To the greatest extent, this applies to those structural features of albumin that are due to Coulomb forces. Upon dissolution, a slight redistribution of locally charged sections of the albumin macromolecule occurs, which changes its hydrophilic and hydrophobic properties. The latter can be the cause of compaction of albumin macromolecules, i.e. reduction in the size of the macromolecule, while the predominance of hydrophilic interactions can contribute to the transformation of a compact domain structure into a quasi-linear structure [2],[5]. Therefore, both hydrophobic and hydrophilic interactions can contribute to the destruction of the domain structure of albumin in solution, which is inherent in the crystalline state of a protein.
The dissolution of albumin macromolecules in water can be considered as an internal structural phase transition. The structural transformations of the albumin macromolecule will also change the structure of the water environment [7],[8]. One of the features of such a phase transition is a change in the size of the macromolecule.
One of the indicators of the structural phase transformations is the hydrodynamic radius change. The latter is studied by various physicochemical methods: dynamic light scattering [9]–[11], small-angle neutron scattering [12]–[14], atomic force microscopy [15], small-angle X-ray scattering [16], pulsed-field gradients NMR spectroscopy [17],[18], gel permeation chromatography [19], capillary viscometry [20].
We focused on methods that allow us to determine the radius of albumin macromolecules as an indicator of conformational transitions using the shear viscosity of macromolecular solutions and the self-diffusion coefficients of macromolecules in them. These methods allow 1) to study the change in the hydrodynamic radius of the macromolecule depending on temperature, concentration, acid-base balance (pH), the presence of salts and the nature of the buffer solutions; 2) to analyze the dynamics of solvent molecules associated by intermolecular interactions with a macromolecule.
To differentiate the temperature, concentration, and acid-base balance (pH) effects on the internal structure of a macromolecule, a comparison is made of the temperature-concentration dependences of the hydrodynamic radius of albumin macromolecules and polyvinyl alcohol (PVA) macromolecules. It is important to note that the PVA macromolecule has an internally disordered, statistically determined structure of the macromolecular coil in solution, while the internal structure of albumin has a strictly determined domain structure of the macromolecular globule, determined by the amino acids sequence.
The macromolecular radius dependence on the conformation changes of albumin macromolecules forms a separate problem and provides a basis for the next step—the study of changes of the macromolecular radius versus solution acid-base balance, which is one of the factors of conformation changes in proteins. At present, the concentration and temperature dependencies of the macromolecular radius of albumin at constant pH values have been obtained, which correlate well with the data obtained by other experimental methods.
Here, the ratio between weight concentration cm and volume concentration φ in the solution of macromolecules:
With certain reservations macromolecules in dilute solutions can be approximated as spherical particles. In this assumption, the average viscosity of the dilute solution can be estimated from Einstein's theory
It follows from
Hydrodynamic interactions between macromolecules become essential as the solution concentration increases. Their accounting was first performed by Batchelor
Batchelor formula generalizes Einstein formula and turns out to be applicable up to volume concentrations φ ≤ 0.2, which corresponds to the viscosity ratio: η/η0 < 1.5. Unfortunately, a further generalization of the Batchelor method faces considerable difficulties in finding the higher-order members of the perturbation theory, finding the sum of the series, and even in solving the question of its convergence.
The progress in determining the effective radii of macromolecules is associated with the use of cellular approaches
We use
The effective radius of macromolecules is found directly by the formula (2). An independent method for determining the effective radius of macromolecules is related to their self-diffusion coefficient.
The dependence of the self-diffusion coefficient on the volume concentration φ is usually presented in the form:
To define δR value, let us present the left part of
As before, we're looking for his solution in the form: δR/R0 = b1φ + b2φ2 + …. Equating the coefficients at the same degrees of φ, we find:
In case of the linear nature of the function f(φ): f(φ) = aφ + O(φ2), the effective radius of the macromolecule is changed in the simplest way:
Recall that the above solutions for the effective radius of the macromolecule are fair for volume concentrations of φ ≤ 0.5.
If the dependence of the self-diffusion coefficient on the volume concentration φ is presented in the form of equation (15), where φ turns out to be significantly different from zero, then the application of perturbation theory is unacceptable.
In this case we represent RD as RD = R0 + δR ≡ R0(1 + u). Then
In this part, we will consider the successively effective hydrodynamic radii of polyvinyl alcohol (PVA), bovine serum albumin (BSA), and human serum albumin (HSA) macromolecules. In comparison with the latter two, PVA macromolecules have the simplest internal structure, which boils down to a sufficiently dense spherical core and a relatively rarefied periphery. It is assumed that the dense PVA core has a dominant influence on the shear viscosity behavior of its solutions. In the case of BSA and HSA, the character of temperature and temperature dependence of the effective hydrodynamic radius is complicated by the influence of their more complex internal structure. Our results on the size of albumin macromolecules refer to the temperature range up to 318 K, which corresponds to the stage of pre-denaturation of protein macromolecules.
This section presents the results of the effective hydrodynamic sizes investigation of polyvinyl alcohol (PVA) macromolecules in aqueous solutions [25]. The PVA of Mowiol 6–98 (Kuraray) brand with the average molecular weight Mw = 47 kDa that corresponds to the average number of monomers in a macromolecule p = 1000 and the concentration of OH-groups h = (98.4 ± 0.4) mol.% in a macromolecule was used.
From the small-angle X-ray scattering
Temperature and concentration dependencies of the effective radius of PVA macromolecules are restored by experimental values of shear viscosity of aqueous solutions from [25] according to formulas (14), (15), and (2) corresponding to the cell approach [22]–[24]. The results obtained in this way are presented in Figure 1.
As we can see, with increasing temperature and concentration, the effective hydrodynamic radius of a core Rη decreases. Our neglect of the periphery influence on the solution viscosity is confirmed by the analysis performed in [40] by the molecular dynamics method. In this work it is shown that the shape of PVA macromolecule in aqueous solution is close to a spherical one with weakly blurred borders.
The concentration dependencies on temperature, at which the PVA macromolecule radius remains unchanged, were studied in [25]. These lines are characteristic curves that allow tracking the role of the effects of the interaction of macromolecules with solvent and among themselves at the overlap of their edges, as well as the disorder that increases with increasing temperature. Two straight lines intersecting at temperature (315 ± 2) K approximate the characteristic curves for aqueous PVA solutions [25]. As shown in [41],[42] there is a dynamic phase transition in pure water at this temperature, when a significant change in the character of thermal motion of water molecules occurs. Thus, PVA macromolecules in aqueous solutions are indicators of changes in water properties [43],[44].
The results obtained by us specify those given in other publications. In
In work
Deviation from our calculated value of 65 Å for diluted PVA solutions may be caused by ignoring the additives specified by us (see equations (21)–(24)). Moreover, viscosity values lead to the size of the dense core, which is slightly smaller than the size of the PVA macromolecule with consideration of the rarefied periphery.
The results of work [47] testify to the growth of the hydrodynamic radius of macromolecules with temperature growth, which is also predicted in [39] using the method of molecular dynamics. Presented on Figure 1 surface of effective radius of PVA macromolecules in aqueous solutions shows constancy of effective radius of macromolecules up to concentrations φ ∼ 0.2 and temperatures T∼330 K. Growth of fluctuations in the peripheral part of the macromolecule with temperature growth should lead to the increase in the periphery size. At the same time, the core size decreases or remains unchanged (see Figure 1).
To obtain effective radii of BSA macromolecules, we use experimental values of shear viscosity of BSA aqueous solutions obtained in [48]. Using equations (13), (14), and (2) of the cell approach we obtained temperature and concentration dependences of an effective radius of BSA macromolecules. The surface of effective radius of BSA macromolecules in the concentration range (2.0÷20.0) wt.% (φ = (0.05÷0.49)) and temperatures (278÷318) K at constant pH = 5.2, corresponding to the vicinity of isoelectric point of BSA is presented in Figure 2.
As we can see, at all temperatures studied, the concentration dependencies of the effective radius of BSA macromolecules
The observed concentration maximum of effective radii
Let us compare the sizes of BSA macromolecules obtained from the analysis of shear viscosity of its aqueous solutions with the available literature data. In
The method of dynamic light scattering to determine the values of self-diffusion coefficient of albumin macromolecules is also used in
In
In
A comparison of the values of effective radii of BSA macromolecules calculated from viscosity data according to equation (13) and self-diffusion coefficient according to equation (17) is presented in Table 1. The properties of the solution – temperature T, acid-base balance pH, and concentration of ions in the solution I–are given in brackets.
cm, wt.% | φ | Rη(T, pH, I), Å | RD(T, pH, I), Å |
1.00 | 0.024 | 35.6 (296 K; 5.2; 0 M) | 37.9 (296 K; 5.0; 0,023 M) |
2.14 | 0.06 | 39.90 (298 K; 5.2; 0 M) | 39.92 (298 K; 5.0; 0,15 M) |
2.79 | 0.08 | 41.5 (298 K; 5.2; 0 M) | 36.9 (298 K; 4.7; 0,1 M) |
40.3 (298 K; 4.7; 0,1 M)* |
* calculation result using corrections due to
Thus, the results obtained by us correlate well with data from other sources. Also, correction of the hydrodynamic radius of albumin macromolecules by equations (20)–(21) allows us to achieve almost complete consistency of data obtained from shear viscosity and self-diffusion. Some ambiguity of macromolecule radii comparison is determined by a mismatch between the pH of a solution and the presence of ions of salts, to which albumin structure is sensitive.
To obtain temperature and concentration dependencies of effective radii of HSA macromolecules we use cell approach equations (13), (14), and (2) and experimental values of shear viscosity in [57]. The surface of effective radii of HSA macromolecules as a function of temperature in the interval (278÷318) K and concentration in the interval (0.82÷18.2) wt.% (φ = (0.03÷0.48)) for the value pH = 7.0 is presented in Figure 4.
Work [57] shows that three concentration intervals can be identified over the entire temperature range 1) (0.82÷3.65) wt.% (φ = (0.03÷0.14)), where the effective radii of human serum albumin remain unchanged; 2) (4.67÷9.45) wt.% (φ = (0.18÷0.31)), where the effective radii of albumin in aqueous solution decrease and 3) (10.2÷18.2) wt.% (φ = (0.33÷0.48)), where the effective radii of albumin macromolecules with increasing concentration linearly decrease, and the slope is poorly dependent on temperature (Figure 5).
In
Let us compare the sizes of HSA macromolecules obtained from the analysis of the shear viscosity of its aqueous solutions with the available literature data. In [13] using the small-angle neutron scattering some dilute HSA aqueous solutions (I = 0.15 M) were investigated: the results indicate the conformation probability of HSA macromolecule as an asymmetric prolate ellipsoid with a radius of 42.5 Å , which correlates well with our results.
In
In
The cellular approach (1)–(14) turns out to be a rather simple, efficient method for the determination of the macromolecular radii from the shear viscosity data for the macromolecular solutions. In other hand, the determination of the macromolecular size in the framework of the dynamic light scattering method is a multistage process, each step of which requires the application of certain physical models with their own approximations and errors. An experimental accessibility of the capillary viscometry method and an ability to quickly change the solution composition during the study make it possible to obtain the dependence of the macromolecular size on the temperature, concentration, pH, and ionic composition of the solution, which is important for understanding the complicated dynamics of biomacromolecules in the aqueous solutions.
Thus, the results obtained by us are comparable with the literature sources data. As above, the comparison of macromolecule radii is complicated by the mismatch between pH solution and the presence of salt ions, to which albumin structure is sensitive.
The article is devoted to the discussion of changes in the structure and size of PVA, BSA, and HSA macromolecules in aqueous solutions depending on temperature, concentration, and acid-base balance (pH) of the solution. Special attention is paid to the comparison of hydrodynamic radius values obtained from data on self-diffusion of macromolecules and shear viscosity of aqueous macromolecular solutions. The importance of these issues is determined by the necessity of careful consideration of water influence on protein structure.
The determination of macromolecule radii from the shear viscosity of its solutions using cellular approaches proves to be one of the simplest and most effective methods. An important feature of the cell approach used by us [23],[24] is that it leads to the results that exactly coincide with those obtained by the Einstein equation [21] and the Batchelor equation [22]. It complements and clarifies the features of macromolecule dynamics in aqueous solutions obtained by other methods.
It should be emphasized that the analysis of macromolecule sizes obtained by different methods is extremely important for investigating changes in the macromolecular coil structure due to their interaction with water at the dissolution in it, as well as considering the interaction between the peripheries of different coils, especially at their overlap, and interaction with ions of acids and alkalis dissolved in water.
[1] | Fenimore PW, Frauenfelder H, Magazù S, et al. (2013) Concepts and problems in protein dynamics. Chem Phys 424: 2-6. |
[2] | Peters T (1996) All about Albumin: Biochemistry, Genetics, and Medical Applications New York: Academic Press, 1-432. |
[3] | He XM, Carter DC (1992) Atomic structure and chemistry of human serum albumin. Nature 358: 209-215. |
[4] | Majorek KA, Porebski PJ, Dayal A, et al. (2012) Structural and immunologic characterization of bovine, horse, and rabbit serum albumins. Mol Immunol 52: 174-182. |
[5] | Akdogan Y, Reichenwallner J, Hinderberger D (2012) Evidence for water-tuned structural differences in proteins: an approach emphasizing variations in local hydrophilicity. PLoS ONE 7: e45681. |
[6] | Carter DC, Ho JX (1994) Structure of serum albumin. Adv Protein Chem 45: 153-203. |
[7] | Pagnotta SE, Ricci MA, Bruni F, et al. (2008) Water structure around trehalose. Chem Phys 345: 159-163. |
[8] | Magazù S, Migliardo F, Benedetto A, et al. (2011) Thermal behavior of hydrated lysozyme in the presence of sucrose and trehalose by EINS. J Non-Cryst Solids 357: 664-670. |
[9] | Luik AI, Naboka YuN, Mogilevich SE, et al. (1998) Study of human serum albumin structure by dynamic light scattering: two types of reactions under different pH and interaction with physiologically active compounds. Spectrochim Acta A Mol Biomol Spectrosc 54: 1503-1507. |
[10] | Maslova MN, Zaritskiy AR, Chaykov LL (2014) The blood plasma particles sizes oscillations observed by dynamic light scattering. Biophys J 106: 457a-458a. |
[11] | Magazù S, Maisano G, Migliardo P, et al. (1999) Experimental simulation of macromolecules in trehalose aqueous solutions: a photon correlation spectroscopy study. J Chem Phys 111: 9086-9092. |
[12] | Sjöberg B, Mortensen K (1994) Interparticle interactions and structure in nonideal solutions of human serum albumin studied by small-angle neutron scattering and Monte Carlo simulation. Biophys Chemist 52: 131-138. |
[13] | Kiselev MA, Gryzunov YA, Dobretsov GE, et al. (2001) The size of human serum albumin molecules in solution. Biophysics 46: 423-427. |
[14] | Varga B, Migliardo F, Takacs E, et al. (2008) Neutron scattering studies on dUTPase complex in the presence of bioprotectant systems. Chem Phys 345: 250-258. |
[15] | Chicea D, Chicea R, Chicea LM (2013) HSA particle size characterization by AFM. Rom Reports Phys 65: 178-185. |
[16] | Olivieril JR, Craievich AF (1995) The subdomain structure of human serum albumin in solution under different pH conditions studied by small angle X-ray scattering. Eur Biophys J 24: 77-84. |
[17] | Wilkins DK, Grimshaw SB, Receveur V, et al. (1999) Hydrodynamic radii of native and denatured proteins measured by pulse field gradient NMR techniques. Biochem 38: 16424-16431. |
[18] | Magazù S (2000) NMR, static and dynamic light and neutron scattering investigations on polymeric aqueous solutions. J Mol Struct 523: 47-59. |
[19] | Qian J, Tang Q, Cronin B, et al. (2008) Development of a high performance size exclusion chromatography method to determine the stability of human serum albumin in a lyophilized formulation of Interferon alfa-2b. J Chromatogr A 1194: 48-56. |
[20] | Monkos K (2004) On the hydrodynamics and temperature dependence of the solution conformation of human serum albumin from viscometry approach. BBA-Proteins Proteom 1700: 27-34. |
[21] | Einstein A (1906) Eine neue bestimmung der moleküldimensionen. Ann Phys 19: 289-305. |
[22] | Batchelor GK (1967) An Introduction to Fluid Dynamics Cambridge: Cambridge university press, 1-615. |
[23] | Malomuzh NP, Orlov EV (2005) Static shear viscosity of a bimodal suspension. Ukr J Phys 50: 618-622. |
[24] | Orlov EV (2010) Shear viscosity of dispersions of particles with liquid shells. Colloid J 72: 820-824. |
[25] | Khorolskyi OV (2018) Effective radii of macromolecules in dilute polyvinyl alcohol solutions. Ukr J Phys 63: 144-149. |
[26] | Mooney RCL (1941) An X-ray study of the structure of polyvinyl alcohol. J Am Chem Soc 63: 2828-2832. |
[27] | Bunn CW (1948) Crystal structure of polyvinyl alcohol. Nature 161: 929-930. |
[28] | Colvin BG (1974) Crystal structure of polyvinyl alcohol. Nature 248: 756-759. |
[29] | Gowariker VR, Viswanathan NV, Sreedhar J (1986) Polymer science New York: Wiley, 1-505. |
[30] | Rumyantsev M, Gushchin A, Zelentsov S (2012) Effect of the type of hydrogen bonding on the reactivity of hydroxyl groups in the acetalization of poly(viny1 alcohol) with butanal. Polym Sci Series B 54: 464-471. |
[31] | Karunakaran DJSA, Ganesh T, Sylvester MM, et al. (2014) Dielectric properties and analysis of H-bonded interaction study in complex systems of binary and ternary mixtures of polyvinyl alcohol with water and DMSO. Fluid Phase Equil 382: 300-306. |
[32] | Tamai Y, Tanaka H, Nakanishi K (1996) Molecular dynamics study of polymer-water interaction in hydrogels. 1. Hydrogen-bond structure. Macromolecules 29: 6750-6760. |
[33] | Budhlall BM, Landfester K, Sudol ED, et al. (2003) Characterization of partially hydrolyzed poly(vinyl alcohol). Effect of poly(vinyl alcohol) molecular architecture on aqueous phase conformation. Macromolecules 36: 9477-9484. |
[34] | Fileti EE, Chaudhuri P, Canuto S (2004) Relative strength of hydrogen bond interaction in alcohol-water complexes. Chem Phys Lett 400: 494-499. |
[35] | Pal S, Kundu TK (2012) Theoretical study of hydrogen bond formation in trimethylene glycol-water complex. Int Scholarly Res Notices 2012: 570394. |
[36] | Rumyantsev M (2013) Influences of co-solvent on hydrogen bond reorganization in ternary poly(vinyl alcohol) solutions. Eur Polym J 49: 2257-2266. |
[37] | Rumyantsev M, Zelentsov SV, Gushchin AV (2013) Retardation effect in acetalization of poly(vinyl alcohol) with butyraldehyde. Eur Polym J 49: 1698-1706. |
[38] | Malomuzh NP, Bulavin LA, Gotsulskyi VYa, et al. (2020) Characteristic changes in the density and shear viscosity of human blood plasma with varying protein concentration. Ukr J Phys 65: 151-156. |
[39] | Rumyantsev M, Rumyantsev S, Kazantsev OA, et al. (2020) Managing of hydrogen bonding in aqueous poly(vinyl alcohol) solutions: new perspectives towards preparation of more homogeneous poly(vinyl butyral). J Polym Res 27: 52. |
[40] | Groda YG, Vikhrenko VS, Poghosyan AH, et al. (2015) Conformation and diffusion properties of polyvinyl alcohol and polyvinylpyrrolidone molecules. Elect J Nat Sci 24: 60-67. |
[41] | Fisenko AI, Malomuzh NP (2009) To what extent is water responsible for the maintenance of the life for warm-blooded organisms? Int J Mol Sci 10: 2383-2411. |
[42] | Bulavin LA, Malomuzh NP (2010) Dynamic phase transition in water as the most important factor of provision of denaturation of proteins in heat blood organisms. Phys Alive 18: 16-22. |
[43] | Migliardo F, Magazù S, Caccamo MT (2013) Infrared, raman and INS studies of poly-ethylene oxide oligomers. J Mol Struct 1048: 261-266. |
[44] | Caccamo MT, Magazù S (2016) Tagging the oligomer-to-polymer crossover on EG and PEGs by infrared and raman spectroscopies and by wavelet cross-correlation spectral analysis. Vibr Spectr 85: 222-227. |
[45] | Garvey M, Tadros T, Vincent B (1974) A comparison of the volume occupied by macromolecules in the adsorbed state and in bulk solution: Adsorption of narrow molecular weight fractions of poly(vinyl alcohol) at the polystyrene/water interface. J Colloid Interf Sci 49: 57-68. |
[46] | Wang B, Mukataka S, Kokufuta E, et al. (2000) Viscometric, light scattering, and size exclusion chromatography studies on the structural changes of aqueous poly(vinyl alcohol) induced by γ-ray irradiation. J Polym Sci Pol Phys 38: 214-221. |
[47] | Wiśniewska M (2012) Temperature effects on the adsorption of polyvinyl alcohol on silica. Cent Eur J Chem 10: 1236-1244. |
[48] | Khorolskyi OV, Moskalenko YuD (2020) Calculation of the macromolecular size of bovine serum albumin from the viscosity of its aqueous solutions. Ukr J Phys 65: 41-49. |
[49] | Magazù S, Calabrò E (2011) Studying the electromagnetic-induced changes of the secondary structure of bovine serum albumin and the bioprotective effectiveness of trehalose by Fourier transform infrared spectroscopy. J Phys Chem B 115: 6818-6826. |
[50] | Magazù S, Calabrò E, Campo S, et al. (2012) New insights into bioprotective effectiveness of disaccharides: an FTIR study of human haemoglobin aqueous solutions exposed to static magnetic fields. J Bio Phys 38: 61-74. |
[51] | Minutoli L, Altavilla D, Bitto A, et al. (2008) Trehalose: a biophysics approach to modulate the inflammatory response during endotoxic shock. Eur J Pharm 589: 272-280. |
[52] | Yadav S, Shire SJ, Kalonia DS (2011) Viscosity analysis of high concentration bovine serum albumin aqueous solutions. Pharm Res 28: 1973-1983. |
[53] | Gaigalas AK, Hubbard JB, McCurley M, et al. (1992) Diffusion of bovine serum albumin in aqueous solutions. J Phys Chem 96: 2355-2359. |
[54] | Keller KH, Canales ER, Yum SI (1971) Tracer and mutual diffusion coefficients of proteins. J Phys Chem 75: 379-387. |
[55] | Jachimska B, Pajor A (2017) Physico-chemical characterization of bovine serum albumin in solution and as deposited on surfaces. Bioelectrochem 87: 138-146. |
[56] | Brenner H (1974) Rheology of a dilute suspension of axisymmetric brownian particles. J Multiphase Flow 1: 195-341. |
[57] | Khorolskyi OV (2019) Calculation of the effective macromolecular radii of human serum albumin from the shear viscosity data for its aqueous solutions. Ukr J Phys 64: 287-292. |
[58] | Bardik V, Gotsulskii V, Pavlov E, et al. (2012) Light scattering study of human serum albumin in pre-denaturation: relation to dynamic transition in water at 42 °C. J Mol Liq 176: 60-64. |
[59] | Atamas N, Bardik V, Bannikova A, et al. (2017) The effect of water dynamics on conformation changes of albumin in pre-denaturation state: photon correlation spectroscopy and simulation. J Mol Liq 235: 17-23. |
[60] | Atamas N, Bardik VY, Komisarenko S, et al. (2019) Water dynamics and stability of major blood proteins at pre-denaturation stage. Atti Accad Pelorit Pericol Cl Sci Fis Mat Nat 97: A16. |
[61] | Hushcha TO, Luik AI, Naboka YN (2000) Conformation changes of albumin in its interaction with physiologically active compounds as studied by quasi-elastic light scattering spectroscopy and ultrasonic method. Talanta 53: 29-34. |
1. | Nikolay Mukhin, Georgii Konoplev, Aleksandr Oseev, Marc-Peter Schmidt, Oksana Stepanova, Andrey Kozyrev, Alexander Dmitriev, Soeren Hirsch, Label-Free Protein Detection by Micro-Acoustic Biosensor Coupled with Electrical Field Sorting. Theoretical Study in Urine Models, 2021, 21, 1424-8220, 2555, 10.3390/s21072555 | |
2. | Anchali Kalidason, Kaori Saito, Yuki Nanbu, Hideki Sasaki, Rina Ohsumi, Akihiko Kanazawa, Takashi Kuroiwa, Biodegradable Crosslinked Chitosan Gel Microbeads with Controlled Size, Prepared by Membrane Emulsification-External Gelation and Their Application as Reusable Adsorption Materials, 2022, 55, 0021-9592, 61, 10.1252/jcej.21we061 | |
3. | Oleh Kuzyk, Ihor Stolyarchuk, Olesya Dan’kiv, Roman Peleshchak, Baric properties of quantum dots of the type of core (CdSe)-multilayer shell (ZnS/CdS/ZnS) for biomedical applications, 2022, 2190-5509, 10.1007/s13204-022-02604-5 | |
4. | Xiaojun Wei, Qian Wang, Chang Liu, Nanopore sensing of γ‐cyclodextrin induced host‐guest interaction to reverse the binding of perfluorooctanoic acid to human serum albumin, 2022, 22, 1615-9853, 2100058, 10.1002/pmic.202100058 | |
5. | Srdjan Pusara, Peyman Yamin, Wolfgang Wenzel, Marjan Krstić, Mariana Kozlowska, A coarse-grained xDLVO model for colloidal protein–protein interactions, 2021, 23, 1463-9076, 12780, 10.1039/D1CP01573G | |
6. | L.A. Bulavin, N.P. Malomuzh, O.V. Khorolskyi, Температурні та концентраційні залежності показника кислотно-лужного балансу водних розчинів хлориду натрію при розчиненні у них атмосферного вуглекислого газу, 2023, 67, 2071-0194, 833, 10.15407/ujpe67.12.833 | |
7. | Hanhee Cho, Seong Ik Jeon, Cheol-Hee Ahn, Man Kyu Shim, Kwangmeyung Kim, Emerging Albumin-Binding Anticancer Drugs for Tumor-Targeted Drug Delivery: Current Understandings and Clinical Translation, 2022, 14, 1999-4923, 728, 10.3390/pharmaceutics14040728 | |
8. | Pavel Yanev, Geralda A.F. van Tilborg, Kristel W. M. Boere, Ann M. Stowe, Annette van der Toorn, Max A. Viergever, Wim E. Hennink, Tina Vermonden, Rick M. Dijkhuizen, Thermosensitive Biodegradable Hydrogels for Local and Controlled Cerebral Delivery of Proteins: MRI-Based Monitoring of In Vitro and In Vivo Protein Release, 2023, 9, 2373-9878, 760, 10.1021/acsbiomaterials.2c01224 |
cm, wt.% | φ | Rη(T, pH, I), Å | RD(T, pH, I), Å |
1.00 | 0.024 | 35.6 (296 K; 5.2; 0 M) | 37.9 (296 K; 5.0; 0,023 M) |
2.14 | 0.06 | 39.90 (298 K; 5.2; 0 M) | 39.92 (298 K; 5.0; 0,15 M) |
2.79 | 0.08 | 41.5 (298 K; 5.2; 0 M) | 36.9 (298 K; 4.7; 0,1 M) |
40.3 (298 K; 4.7; 0,1 M)* |
* calculation result using corrections due to