<<

Analytical 295, 158–167 (2001) doi:10.1006/abio.2001.5129, available online at http://www.idealibrary.com on

Binding of Bovine to Heparin Determined by Turbidimetric Titration and Frontal Analysis Continuous Capillary Electrophoresis

Toshiaki Hattori,1 Kozue Kimura, Emek Seyrek, and Paul L. Dubin2 Department of Chemistry, Indiana University–Purdue University, Indianapolis, Indiana 46202

Received October 23, 2000; published online July 19, 2001

The binding of to proteoglycans (PGs)3 is an The association of proteins with glycosaminogly- important biological phenomenon. For example, the cans is a subject of growing interest, but few tech- interaction between chondroitin sulfate or keratan sul- niques exist for elucidating this interaction quantita- fate proteoglycans and collagen fibrils in the cartilage tively. Here we demonstrate the application of matrix defines the final shape of the tissue and pro- capillary electrophoresis to the system of serum albu- vides the ability to withstand compressive load (1). min (SA) and heparin (Hp). These two species form Heparan sulfate proteoglycans (HSPGs) interact with soluble complexes, the interaction increasing with re- fibronectin–collagen complexes in plasma membranes duction in pH and/or ionic strength (I). The acid–base to assist cell adhesion (1). HSPGs in the glomerular property of Hp was characterized by potentiometric basement membrane (GBM) also provide selective fil- titration of ion-exchanged Hp. Conditions for complex tration of proteins (1, 2). These interactions appear to formation with SA were qualitatively determined by have a strong electrostatic component, arising from the turbidimetry, which revealed points of incipient bind- interactions between charges on the proteins and the ing (pHc) and phase separation (pH␾), both of which intense negative charge on glycosaminoglycans (GAGs) depend on I. At pH > pH␾, i.e., prior to phase separa- in PGs. tion, frontal analysis continuous capillary electro- GAGs are highly diversified sulfated and carboxy- phoresis was used to measure the concentration of lated linear polysaccharides that constitute the major free and to determine the protein–HP binding components of PGs. The backbone of the extracellular isotherm. The binding isotherms were well fit by the matrix PG aggregate is a high-molecular-weight (ca. McGhee–von Hippel model to yield quantitative bind- 2 ϫ 106) hyaluronic acid, to which are bound a large ing information in the form of intrinsic binding con- number of core proteins; each core protein, in turn, is stants (Kobs) and binding site size (n). The strong in- covalently bound to numerous GAGs (3–5). Many of the crease in Kobs with decrease of pH or I could be roles of PGs in extracellular signaling appear to de- explained on the basis of electrostatic interactions, pend on the interaction between GAGs and other mac- considering the effects of protein charge heterogene- romolecules, in particular proteins (3, 6). Conse- ity. The value of n, independent of pH, was rational- quently, the study of GAG–protein interactions can ized on the basis of size considerations. The implica- facilitate a physical understanding of important biolog- tions of these findings for clinical applications of Hp ical phenomena. and for its physiological behavior are discussed. © 2001 Many techniques have been used to investigate the Academic Press interaction between GAGs and proteins. For example, the formation of complexes has been detected using fluorescence (7, 8) and UV (8) absorption. Diffusion

3 Abbreviations used: PGs, proteoglycans; HSPGs, heparin sulfate 1 Current address: Research Center for Chemometrics, Toyohashi proteoglycans; GBM, glomerular basement membrane; GAGs, gly- University of Technology, Toyohashi, Japan 441-8580. coaminoglycans; CE, capillary electrophoresis; Hp, heparin; FACCE, 2 To whom correspondence should be addressed. Fax: (317) 274- frontal analysis continuous capillary electrophoresis; BSA, bovine 6879. E-mail: [email protected]. ; SEC, size-exclusion chromatography.

158 0003-2697/01 $35.00 Copyright © 2001 by Academic Press All rights of reproduction in any form reserved. ELECTROPHORETIC DETERMINATION OF SERUM ALBUMIN–HEPARIN BINDING 159 coefficients and sizes of complexes have been measured gand and the complex can then be fit to appropriate using light scattering (9, 10). Protein structural binding isotherms to yield binding constants and changes as a result of complexation between GAGs and binding site size, two parameters that are very use- proteins have been observed by circular dichroism ful in elucidation of the binding mechanism. More- measurement (CD) (11, 12) and NMR (13), and the over, sample requirements are smaller, and the mea- binding sites of proteins and GAGs have been identi- surement time is short. FACCE thus appears to be a fied by X-ray crystallography of complexes between simple, rapid, economical, precise, and reproducible proteins and GAG oligomers (14). Complex mobilities method for quantitative characterization of protein– have been measured by capillary electrophoresis (CE) polyelectrolyte binding. (15). Previous studies with FACCE focused on combina- Although numerous methods have been used to ex- tions of proteins and synthetic polyelectrolytes (21). amine protein–GAG complexes, few techniques are ca- Here we are interested in developing FACCE method- pable of quantitatively determining an important vari- ology for combinations of proteins and biological poly- able: the GAG–protein complex formation constant. electrolytes. While the ultimate goal is to apply this These potentiometric, spectroscopic, or chromato- technique to physiological cognates, our purpose here graphic methods all pose respective difficulties. For is to develop the methodology, as opposed to addressing example, the heparin(Hp)–protamine binding constant a specific biochemical question. To do this, we have was determined using a Hp-sensitive electrode after chosen (BSA) as the protein and neutralization of the complex (16), but the mechanism heparin (Hp) as the biological polyelectrolyte. of this electrode response is not well understood, and The interaction of Hp and Hp-like GAGs with pro- other bulky anions e.g., free proteins of negative net teins is clearly of considerable interest. Hp–protein charge, can act as interferents in this assay. Quantita- interactions regulate biological processes as diverse tion of complexation has been carried out using fluo- as complement activation, coagulation, cell adhe- rescence (17–19), but this approach depends on a large sion, angiogenesis, platelet aggregation, lipolysis, change in absorption before and after complexation. and smooth muscle proliferation. Hp regulates an- While affinity chromatography has been used to mea- giogenesis, the growth of new capillary vessels, sure the dissociation constant (20), large amounts of by interacting with peptide growth factors (27). It sample are required and the dissociation constant can has an anticoagulant effect in coagulation, where be perturbed by the immobilization of protein or GAG. binding to a plasma Hp cofactor is required for its Frontal analysis continuous capillary electro- activity (28); consequently, Hp is often used during phoresis (FACCE) (21) is a new technique that has surgical procedures or as an anticoagulant for been demonstrated as a way to obtain precise and reliable binding data for protein–polyelectrolyte thrombosis. Hp also has a direct aggregatory effect complexes. This technique resembles other capillary on platelets and lipoproteins; the interaction with electrophoresis methods, namely the Hummel– the platelets causes platelet proaggregating and po- Dreyer method, conventional frontal analysis tentiating effects (29), while the binding of Hp to method, affinity capillary electrophoresis, and va- lipoprotein leads to aggregation or fusion of the lipo- cancy peak and vacancy affinity capillary electro- protein during lipolysis (30). Hp is also responsible phoresis; however, each of these measurements of for the antiproliferative activity in smooth muscle quantitative binding parameters poses some prob- proliferation (31). While these diverse and complex lems (22). Because the perturbation of the binding biological roles are not fully understood, they clearly equilibria intrinsic to conventional frontal analysis involve the binding of many proteins. method is overcome by continuous sampling, FACCE Serum albumin is not a biochemical cognate of Hp, can be performed even when the mobility of the but it is the most abundant protein in the body and in complex is not equal to the mobility of ligand or particular is present at high concentration for intrave- acceptor. Affinity capillary electrophoresis is useful nously administered Hp. Because the molecular weight for determining the dissociation constant of pro- of serum albumin is large (even in the absence of mul- tein–Hp binding (23–26), but the resulting constant timerization), complex formation of serum albumin is not regarded as a true quantitative binding pa- with polymers is easily observed by many methods rameter when multiple ligands bind to the acceptor based on scattering or electrical mobility (32). The (22). On the other hand, FACCE may be used even in structure of serum albumin is well understood and its multiple complexation equilibria to measure the con- properties, particularly its ionization behavior, are centration of free ligand because its concentration is thoroughly documented. Selection of the Hp/albumin not determined from mobility; instead, the peak pair as a demonstrative model of the quantitation of height directly indicates free ligand concentration. GAG–protein interactions using FACCE is thus based The stoichiometric relationship between bound li- on numerous considerations. 160 HATTORI ET AL.

EXPERIMENTAL 10-min wash with Milli-Q water. This ensured a Materials negative capillary surface charge under the experi- mental conditions. Heparin (sodium salt, porcine intestinal mucosa, Lot Sample solutions were made from freshly prepared B27591), nominal Mr 13,500–15,000 was from Calbio- stock solutions of BSA and Hp dissolved in CE run chem (La Jolla, CA). BSA ( free, Lot 8656224, buffer solution. The run buffer solution of I ϭ 0.01 Mr 68,000) was purchased from Boehringer Mannheim consisted of disodium hydrogen phosphate and sodium Corp. (Indianapolis, IN). Analytical-grade dibasic so- dihydrogen phosphate adjusted to pH’s 6.5, 6.8, and dium phosphate was obtained from J. T. Baker Chem- 7.0. The concentration range of BSA was 0.2–8.0 g/L ical Co. (Phillipsburg, NJ) and monobasic sodium phos- and the concentration of Hp was constant at 0.2 g/L. phate was from Mallinckrodt Inc. (Paris, KY). Other The pH values of the samples were always adjusted to chemicals and reagents were obtained from current same pH as the run buffer. commercial sources at the highest level of purity avail- The FACCE (21) experiment was initialized by able. All buffers and solutions were prepared with equilibrating the capillary with buffer solution for 5 Milli-Q water (Millipore, Milford, MA). min. The inlet end of the capillary was then placed in a vial containing the equilibrated sample solution, Potentiometric Titration of Ion-Exchanged Hp and the outlet end was placed into a vial containing buffer solution. Constant voltage was applied, and Hp was ion-exchanged from the sodium salt to the separation, manifested in continuous plateaus, was hydrogen form using 45 ml of Dowex 50W-X8 (cation observed. The first eluting plateau was the free pro- exchange resin; 5.1 meq/dry gram, 1.8 meq/ml resin tein, and the second was the free protein and the bed) in a 50-ml buret column. A 10-ml solution contain- protein–Hp complex. Although a multiple peak pat- ing 0.05 g of heparin sodium salt was poured into the tern sometimes appeared instead of the second pla- column and washed through with Milli-Q water. After teau, the first plateau was always obtained. After 10 ml of void volume was removed, at least 50 ml of each electrophoretic run, a 5-min wash with 0.1 M solution was recovered. Collection of larger volumes NaOH followed by a 5-min rinse with water was did not increase the amount of measurable acid, prov- performed. The concentration of free protein was ing that all the Hp was collected. The ion-exchanged determined from the height of first plateau, using a Hp solution was titrated with 0.5 M NaOH using a calibration curve constructed by measuring the pla- Corning pH meter 240 with a FUTURA refillable com- teau height of known concentrations of protein ob- bination pH electrode (Beckman, Fullerton, CA). tained under the same experimental conditions as for the protein–Hp mixture. The amount of bound Turbidimetric Titration BSA was obtained by subtracting the amount of mea- Turbidity measurements, reported as 100-%T, were sured free BSA from the total amount used. performed at 420 nm using a Brinkman PC800 probe colorimeter equipped with a 1-cm path-length fiber Computational Methods optics probe. Turbidimetric titrations were carried out by adding 0.5 N HCl to mixed solutions of BSA (1 g/L) Molecular modeling was done using DelPhi v98.0 and Hp (0.1 g/L) containing variable concentrations of (Molecular Simulations Inc.), where the electrostatic NaCl (0.01, 0.02, 0.05, 0.1, or 0.2 M NaCl). All titra- potential in and around the protein is calculated by tions were carried out with gentle magnetic stirring, nonlinear solution of Poisson–Boltzmann equation. and the observed pH and transmittance were deter- The protein was placed in the center of a grid box with mined after the values had been stable for at least 1 its largest dimension occupying 40% of the grid length. min. The measured values were corrected by subtract- The resolution was set at 101 grid points per axis. The ing the turbidity of a Hp-free blank. dielectric constants of the solvent and the protein were set to 80 and 2.5, respectively. Using the fractional charges for each charged amino acid residues, the elec- Frontal Analysis Continuous Capillary trostatic potential is then calculated at every point Electrophoresis inside the grid box. Capillary electrophoresis was performed using a The charges of amino acid residues were determined P/ACE 5500 CE (Beckman, Fullerton, CA) with UV using as a starting point the simple model put forward detector, operating at 8 kV and 25°C. The fused silica by Tanford (33). In this spherical-smeared-charge capillary (Polymicro Technologies Inc., Phoenix, AZ) model, the titration curve of a protein is a superposi- of dimensions 50 ␮m ϫ 27 cm (effective length 20 cm) tion of the curves for each of the seven groups of amino was prepared prior to each set of experiments by acids. The fraction ␣ of dissociated groups of any washing with 0.1 M NaOH for 10 min followed by a groups is given by ELECTROPHORETIC DETERMINATION OF SERUM ALBUMIN–HEPARIN BINDING 161 ៮ TABLE 1 for by 0.868wZ, with w being essentially an empirical Intrinsic Dissociation Constants of Titratable Groups fitting parameter. Previous titration data of BSA could in BSA be well fit by Eq. [3] using ionic-strength-dependent w values (35). Titration curves were thus first con- Number Number structed using these along with some w values inter- Type of group in groupa pK a in groupb pK c int int polated with respect to ionic strength, in conjunction

␤,␥-Carboxyl 99 4.02 98 4.1 with pKint values that represented approximately the ⑀ -Amino 57 9.8 59 10.3 mean of literature data (35). Then, the pKint value for Imidazole 16 6.9 16 6.7 each group was adjusted in the range of previously Phenolic 19 10.35 18 11 found values until the differences between calculated Guanidine 22 Ͼ12 24 12 ␣-Carboxyl 1 3.75 1 3.75 and experimental titration curves were minimized at ␣-Amino 1 7.75 1 7.75 all ionic strengths. Adjusted pKint values (see Table 1) were then used along with Eq. [3] to calculate the a Values from Ref. (37). charges of each ionizable group at the desired pH and I. b Actual number in group, taken from RCSB (crystal structure 1AO6). c Best fit values (see text for explanation). RESULTS AND DISCUSSION Potentiometric Titration of Ion-Exchanged Hp ␣ 1 ѨW Hp, a polydisperse polysaccharide consisting of sev- pH Ϫ logͩ ͪ ϭ pK Ϫ , [1] eral different disaccharide units, cannot be repre- 1 Ϫ ␣ int 2.303kT ѨZ sented by any single structure (36). The contents of sulfate groups and carboxylate groups are thus vari- where Kint is an intrinsic dissociation constant charac- able and also dependent on the source (36). Since we teristic of each group and W is the electrostatic free find the binding between BSA and Hp to be mainly energy of the protein. All of the charges are considered based upon electrostatic interaction, characterization to be smeared evenly over the surface so that positive of the ionic content of Hp is an important parameter and negative charges can cancel and only the average ៮ which can formally be expressed by an equivalent net charge Z remains. The total electrostatic free en- weight (mass per unit charge group). ergy is then the work of charging the sphere from ៮ Figure 1 shows potentiometric titration curves of zero to Z and the work of charging due to the presence ion-exchanged Hp in the absence and presence of so- of the surrounding ions and is given by dium chloride. The first inflection point must be the end point of the strong acid (sulfate and sulfamido) Z៮ 2⑀ 2 1 ␬ groups, and the second inflection that of the weak acids W ϭ ͩ Ϫ ͪ , [2] 2D b 1 ϩ ␬a (carboxylate groups). The pKa of the latter is elevated by the electrostatic influence of the strongly ionized groups. Consequently, the disappearance of the first where b is the radius of the sphere, a is the distance of closest approach between the centers of two ions (i.e., a Ϫ b is the dimensions of a salt ion radius), ␬ is the Debye–Huckel parameter proportional to the square root of the ionic strength, ⑀ is the protonic charge, and D is the dielectric constant of the solvent. With this value of W, Eq. [1] becomes

␣ pH Ϫ logͩ ͪ ϭ pK Ϫ 0.868wZ៮ [3] 1 Ϫ ␣ int

⑀ 2 1 ␬ w ϭ ͩ Ϫ ͪ . 2DkT b 1 ϩ ␬a

Equation [3] has been used extensively and success- fully in the interpretation of potentiometric titrations FIG. 1. Potentiometric titration of ion-exchanged heparin. (ᮀ) studies of globular proteins such as , serum 0.0548 g of heparin in the absence of NaCl; (■) 0.0252 g of Hp in 1 M , ribonuclease, hemoglobin, and ␤-lactoglobu- NaCl. Solid line is the differential of the titration curve without lin (34). Electrostatic interactions are thus accounted NaCl. 162 HATTORI ET AL.

tion of the dissociation of carboxylate by the high neg- ative charge of the sulfate groups persists even at high salt.

Turbidimetric Titration In order to establish appropriate conditions for FACCE, qualitative information about the binding of BSA–Hp is required. For a large protein such as BSA, turbidimetry is sufficiently sensitive to detect the sol- uble polyelectrolyte complex and so can be used to define three pH regions corresponding to no interac- tion, soluble complex, and precipitate or coacervate (38). Figure 3 shows the turbidimetric titration curve of the mixture of 1 g/L of BSA and 0.1 g/L of Hp in 0.03 M NaCl. In region 1, there is no complexation between BSA and Hp because of the strong coulombic repulsion Ϸ FIG. 2. The ␣ dependence of the pKa of the carboxylate groups of between the net negatively charged BSA (pI 4.9) and heparin (ᮀ) in the absence of NaCl and (■)atI ϭ 1.0 M NaCl. the negatively charged Hp. In region 2, with decreasing negative charge, BSA forms a soluble complex with Hp. However, the soluble complex may form “on the wrong end point upon addition of 1 M sodium chloride is a side” of pI (39), i.e., at pH Ͼ pI. The abrupt transition result of the salt-induced depression of this pKa so that from region 1 to region 2, i.e., a well-defined “pHc” for the titration of the strong acid and carboxylate groups soluble complex formation, can be verified by dynamic are no longer separable. The amount of NaOH con- light scattering (39). In region 3, the complex aggre- sumed after the second end point is equivalent to the gates and forms visible precipitate, with an abrupt summation of the strong and weak acidic groups, so the increase in turbidity at “pH␾.” The dependences of pHc equivalent weight was 5.0 meq/g. The relative ratio of and pH␾ on I are shown as BSA–Hp “phase bound- carboxylate groups to strong acid groups was obtained aries” in Fig. 4. Obviously, FACCE can only be done in as 1:2.7. This ratio corresponds to 2.7 sulfate and sul- region 2. The increase of I strongly depresses pHc since foamido groups per disaccharide, in agreement with a the salt screens electrostatic interactions, necessitat- literature value (14). ing a larger protein positive charge. On the other hand, The dissociation constant of polycarboxylates such pH␾ is independent of I, and this constant value is close as poly(acrylic acid) in general depends on ␣ and I (37). to pI (38, 39). Figure 2 shows the apparent dissociation constant, pKa of the carboxylate groups, defined by pKa ϭ pH ϩ FACCE Electropherograms Ϫ ␣ ␣ ␣ log[(1 )/ ], as a function of and I. In pure water, In contrast to conventional frontal analysis, FACCE ␣ pKa strongly depends upon and is larger than the involves continuous sampling, so the electropherogram intrinsic pKa; the value of pKa is 5.1 at ␣ ϭ 0.5, which agrees with the value of 5.1 measured by CD and potentiometric titration (38). This value is consider- ably larger than the pKa of 3 determined by NMR (39). Possible limitations of the NMR method, particularly at concentrations as high as 10% w/v may be indicated by another report that the pKa strongly decreases with

Hp concentration (40), despite the fact that pKa’s of polyacids are generally not reported to show strong concentration dependence. In any event, it is clear that intrinsic properties of polymer chains are best ob- served at high dilution, which can account for the good agreement between our value and the CD result, both measured at about 0.1% w/v. The addition of 1.0 M sodium chloride solution causes, as expected, a de- crease in our measured pKa from 5.1 to 4.3. This value is higher than pKa ϭ 3.28 of D-glycopyranosyluronic ϭ acid (45) or the intrinsic dissociation constant, pK0 FIG. 3. Turbidimetric titration curve at I ϭ 0.03, for 1 g/L of BSA 2.9, of hyaluronic acid (46), indicating that the inhibi- with 0.1 g/L of heparin. ELECTROPHORETIC DETERMINATION OF SERUM ALBUMIN–HEPARIN BINDING 163

FIG. 4. Phase boundary for BSA (1 g/L) and heparin (0.1 g/L): (ᮀ) pHc;(■) pH-. (*) point corresponding to conditions used for compu- tational model in Fig. 8.

FIG. 6. Binding isotherms for BSA and heparin. (Œ) pH 6.5; (F)pH is a step-like separation profile, as shown in Fig. 5. 6.8; (■) pH 7.0; at I ϭ 0.01. v is the number of bound BSA per Since the electrophoretic mobility of the complex is negative heparin group. more negative than that of free BSA, the first plateau in the electropherogram is due to the elution of free BSA, and the second plateau is due to free BSA and Binding Isotherms complex. In conventional frontal analysis, the loss of Binding isotherms of BSA-Hp at I ϭ 0.01 M and at free BSA in the mixture region (corresponding to sec- various pH values are shown in Fig. 6. Here [BSA] is ond plateau) would make the complex decompose into free free BSA and free Hp. In FACCE, however, the loss is the concentration of free BSA in the mixture, and v is made up by supplying free protein in the following the number of protein molecules bound per charged eluent. The concentration of free BSA is equal and group. v is defined in this manner for the following constant in both first and second plateaus and there is reason. In the case of nonspecific binding arising from no perturbation of the binding equilibrium. The con- long-range electrostatic forces, the binding site does centration of free BSA in the mixture of BSA and Hp is not correspond to a well-defined portion of the host then determined by calibrating the peak height of first macromolecule. The corresponding binding site size plateau. could be somewhat arbitrarily defined as a number of An unexplained spike peak appears for Hp-free BSA, sugar rings or disaccharides. However, since the bind- possibly due to limited adsorption of BSA onto the ing is primarily electrostatic, we choose here to discuss capillary. In order to minimize such adsorption FACCE the size of the apparent binding site in terms of the was carried out only at pH Ͼ 6.5. number of charges it encompasses, a procedure similar to the one used in describing the binding of oligolysines to poly(rI) ϩ poly(rU) (43). The relationship of v, the number of BSA bound per Hp charge group, to v*, the number of BSA bound per disaccharide, is simply v* ϭ 3.7v. The amount of BSA bound depends strongly on the pH. The diminution of the binding with pH must be the result of increased repulsion between Hp and BSA due to the increasing negative charge of BSA. The addition of salt also depresses the amount of BSA bound: even at I ϭ 0.03, the concentration of BSA bound in pH 6.7 buffer solution was too low to obtain a significant re- sult. These strong effects of pH and ionic strength indicate that the interaction is mainly due to coulombic forces between Hp and BSA. FIG. 5. Electropherograms obtained for (a) 1.0 g/L BSA ϩ 0.2 g/L heparin, (b) 4.0 g/L BSA ϩ 0.2 g/L heparin in phosphate buffer The binding isotherms were fit to the McGhee–von solution of I ϭ 0.01 at pH 6.5. Hippel (45) equation 164 HATTORI ET AL.

1 Ϫ nv nϪ1 v/L ϭ K ͑1 Ϫ nv͒ͩ ͪ , [4] obs 1 Ϫ ͑n Ϫ 1͒v where v is the number of bound BSA per ionic site of

Hp, L is the concentration of free BSA, Kobs is the observed binding constant, and n is the number of binding sites. The two binding parameters of Kobs and n were obtained by nonlinear curve fitting of v/L vs v, and their values are listed in Table 2. Although the McGhee–von Hippel (MvH) model can be made to en- compass cooperative binding (45), the binding iso- therms in the present study were well fit without an additional cooperativity term. As shown by the solid lines in Fig. 6, the fitted curves conform very well to the experimental points. Similarly good two-parameter binding isotherm fits were reported for the binding between protein and synthetic polymers (46, 47). The intrinsic binding constant decreases strongly with in- FIG. 7. Binding ratio of BSA per disaccharide of Hp. (‚) pH 6.5; (E) ᮀ ϭ ϭ creasing pH, while the binding site size expressed in pH 6.8; ( ) pH 7.0; all at I 0.01. The solid line is at Nlimit 8 and ϭ terms of Hp ionic sites is n ϭ 14 independent of pH. the dashed line is at Nsite 4 (see text). This corresponds to a binding site size of 14/3.7 disac- charide units, i.e., an octasaccharide. The binding isotherms in Fig. 6 may be converted interpreted too rigidly. The fact that the MvH model is into Fig. 7, which shows the number of BSA bound per in principle strictly relevant to chains of infinite length disaccharide increasing with added BSA to a limiting (48) is a similar reason to avoid strict interpretation of value of 0.125. This corresponds to approximately the fitting parameters in terms of the theoretical three BSA bound per Hp molecule, based on an Mr of model. ϭ 14,000. The limiting value of 0.125 indicates that BSA From Nsite 4 and the length per disaccharide of ca. occupies on average 8 disaccharides in Hp, Nlim, which 1 nm (49), we can estimate that the contour length of is larger than the value Nsite ϭ 4 found above; i.e., the protein-binding site on Hp is about 4 nm, which is efficient packing in which every octasaccharide site is smaller than the Stokes diameter of BSA of ca. 7 nm occupied does not occur. This result is consistent with (50). Thus, Hp does not wrap around BSA, but rather the MvH “overlapping binding site” model, in which binds to some site on the protein, presumably a domain complete saturation of BSA is inhibited by the entropic that bears a positive charge (see below). The consis- resistance based on the “parking problem”: the binding tency of n as a function of pH suggests constancy of the possibilities for BSA depend upon the binding site of Hp-binding site of BSA; on the other hand, we find no previously bound BSA. This effect, however, could also evidence for a unique protein-binding site on Hp. arise from the repulsive force between negatively charged neighboring bound proteins (see below) lead- Computational Model of the Binding Site ing to anticooperative binding. Analysis of the binding The binding of a strong polyanion such as Hp to BSA isotherms does not make it possible to distinguish be- at a pH well above pI indicates the presence of a pos- tween these two types of anticooperativity. Since reso- itive “patch” in an otherwise negatively charged pro- lution between these two effects is not possible at the tein (44). Computer graphics allows us to identify one present time, the MvH plots should be recognized as or more domains of positive potential, which must be effective empirical fits which, however, should not be involved in Hp binding by a protein of net negative charge, and also to visualize the implications of some of the results obtained from FACCE. However, if the po- TABLE 2 tential was displayed on the protein surface, positive Intrinsic Binding Constant and Binding Site Size ϭ charges appeared in the negative domains and vice of BSA–Hp at I 0.01 M versa: no continuous and extensive positive potential domain could be observed. But when the potential was pH log Ki n displayed 5 Å away from the van der Waals surface, 6.5 4.12 (Ϯ0.04) 15.5 (Ϯ0.7) positive and negative regions could clearly be distin- 6.8 3.56 (Ϯ0.02) 13.9 (Ϯ0.7) Ϯ Ϯ guished. Figure 8 shows such an image of BSA, from 7.0 3.23 ( 0.02) 13.7 ( 0.9) three different views, for the conditions corresponding ELECTROPHORETIC DETERMINATION OF SERUM ALBUMIN–HEPARIN BINDING 165

for glycosaminoglycan–protein pairs will be the subject of future investigations.

Physiological and Clinical Implications The present results may be relevant to some of the clinical applications of Hp, as well as its physiological functions. Hp has been administered as a treatment for (abnormal urinary excretion of albumin) (51, 52) in diabetics. It has been suggested that albu- minuria, which is a loss of filtration selectivity in the glomerular basement membrane GBM, could arise from a diminution of the heparin sulfate proteoglycan content of the GBM (53), consequently, albumin-free filtration might be restored if Hp were to somehow ameliorate this loss. Figure 4 shows that physiological blood conditions (pH 7.4, I ϭ 0.15 M) correspond to strong repulsive interactions between BSA and Hp, even under acidosis or hyponatremia (the decrease of sodium concentration in blood), and Hp as a component of the GBM should reduce serum albumin filtration. On the other hand, severe reduction of I can lead to attraction (region 2 or 3 in Fig. 4). I values as low as 40 mM have been suggested for cartilage because of the ionic exclusion properties of proteoglycans (54). In os- teoarthritic cartilage, in which Hp has been found to constitute 3–4% of proteoglycans (55), the amount of albumin increases significantly (56). In such tissues, FIG. 8. (a) Electrostatic potential surface of BSA at 5 Å away from attractive serum albumin–Hp interactions may exist. the van der Waals surface, under conditions of complex formation: I ϭ 0.01 and pH 7. (b) 90° rotated view of image in (a). (c) 270° rotated ACKNOWLEDGMENTS view of image in (a). The crystal structure of BSA (PDB ID: 1AO6) is taken from Research Collaboratory for Structure, Bioinformatics This research was supported by NSF Grant CHE-9987891. The Protein Data Bank (http://www.rcsb.org) (57). authors thank Kristopher Grymonpre´ for his help with molecular modeling. to the point represented by (*) of Fig. 4 on the phase REFERENCES boundary in the region of complex formation (pH 7, I ϭ 1. Rhoades, R., and Pflanzer, R. (1996) Human Physiology, 3rd ed., 0.01). The image in Fig. 8b shows a continuation of the pp. 704–735, Saunders College, Orlando. positive potential patch in Fig. 8a. 2. Farquhar, M. G. (1985) The glomerular basement membrane: A As has been discussed earlier, Hp does not wrap selective macromolecular filter. In Cell Biology of the Extracel- around BSA, so the Hp-binding site is some portion of lular Matrix (Hay, E. D., Ed.), pp. 335–378, Plenum, New York. this patch, corresponding—as discussed above—to a 3. Sames, K. (1994) The Role of Proteoglycans and Glycosamino- size of about 4 nm, consistent with the view in Fig. 8b. glycans in Aging, Karger, Basel. The modest curvature of this site is in accord with the 4. Dannell, J., Lodish, H., and Baltimore, D. (1990) Molecular Cell flexibility of the Hp molecules. The side of the protein Biology, Chap. 23, Freeman, New York. opposite to the binding side shown in Fig. 8c is strongly 5. Jackson, R. L., Busch, S. J., and Cardin, A. D. (1991) Glycosami- noglycans: Molecular properties, protein interactions, and role in negative. Thus, repulsive interactions among bound physiological processes. Physiol. Rev. 71, 481–539. BSA could lead to anticooperative binding. 6. Kjellen, L., and Lindahl, U. (1991) Proteoglycans: Structures and The binding of Hp to protein cognates has been dis- interactions. Annu. Rev. Biochem. 60, 443–475. cussed in terms of specific amino acid sequences that 7. Teramoto, A., Takagi, Y., Hachimori, A., and Abe, K. (1999) constitute the Hp-binding site(s) (14). This description Interaction of albumin with polysaccharides containing ionic of specific binding differs fundamentally from loose groups. Polym. Adv. Technol. 10, 681–686. nonspecific binding at the 5-Å surface proposed here 8. Finotti, P., and Pagetta, A. (1997) Heparin-induced structural modifications and oxidative cleavage of in for BSA and Hp. In our model multiple configurations the absence and presence of glucose: Implications for transcap- of the Hp chain which approximate the minimum en- illary leakage of albumin in hyperglycaemia. Eur. J. Biochem. ergy state may exist. Whether such a model is tenable 247, 1000–1008. 166 HATTORI ET AL.

9. Eigner, W., Mitterer, A., Schurz, J., and Gu¨ nther, J. (1982) 27. Folkman, J., and Shing, Y. (1992) Control of angiogenesis by Light-scattering investigations on LDL–heparin complexes. Bio- heparin and other sulfated polysaccharides. Adv. Exp. Med. Biol. sci. Rep. 2, 413–417. 313, 355–364. 10. Mach, H., Volkin, D. B., Burke, C. J., and Middaugh, R. (1993) 28. Hirsh, J., and Fuster, V. (1994) Guide to anticoagulant-thera- Nature of the interaction of heparin with acidic fibroblast growth py.1. Heparin. Circulation 89, 1449–1466. factor. Biochemistry 32, 5480–5489. 29. Burgess, J. K., and Chong, B. H. (1997) The platelet proaggre- 11. Lima, L. T. R., and De Prat-Gay, G. (1997) Conformational gating and potentiating effects of unfractionated heparin, low changes and stabilization induced by ligand binding in the DNA- molecular weight heparin and heparinoid in intensive care pa- binding domain of the E2 protein from human papillomavirus. tients and healthy controls. Eur. J. Haematology 58, 279–285. J. Biol. Chem. 272, 19295–19303. 30. Hakala, J. K., Oorni, K., Ala-Korpela, M., and Kovanen, P. T. 12. Vyas, A. A., Pan, J., Patel, H. V., Vyas, K. A., Chiang, C., Sheu, (1999) Lipolytic modification of LDL by phospholipase A(2) in- Y., Hwang, J., and Wu, W. (1997) Analysis of binding of cobra duces particle aggregation in the absence and fusion in the ␤ cardiotoxins to heparin reveals a new -sheet heparin-binding presence of heparin. Arterioscler Thromb. Vasc. Biol. 19, 1276– structural motif. J. Biol. Chem. 272, 9661–9670. 1283. 13. Pineda-Lucena, A., Jime´nez, M. A., Lozano, R. M., Nieto, J. L., 31. Garg, H. G., Joseph, P. A. M., Thompson, B. T., Hales, C. A., Santoro, J., Rico, M., and Gime´nez-Gallego, G. (1996) Three- Toida, T., Imanari, T., Capila, I., and Linhardt, R. J. (1999) dimensional structure of acidic fibroblast growth factor in solu- Effect of fully sulfated glycosaminoglycans on pulmonary artery tion: Effects of binding to a heparin functional analog. J. Mol. smooth muscle cell proliferation. Arch. Biochem. Biophys. 371, Biol. 264, 162–178. 228–233. 14. Hileman, R. E., Fromm, J. R., Weiler, J. M., and Linhardt, R. J. 32. Gao, J. Y., Dubin, P. L., and Muhoberac, B. B. (1998) Capillary (1998) Glycosaminoglycan–protein interaction: Definition of con- sensus sites in glycosaminoglycan binding proteins. BioEssay electrophoresis and dynamic light scattering studies of structure 20, 156–167. and binding characteristics of protein-polyelectrolyte complexes. J. Phys. Chem. B 102, 5529–5535. 15. Gunnarsson, K., Valtcheva, L., and Hjerte´n, S. (1997) Capillary zone electrophoresis for the study of the binding of 33. Tanford, B. C., and Kirkwood, J. G. (1957) Theory of protein to low-affinity heparin. Glycoconjugate J. 14, 859–862. titration curves. I. General equations for impenetrable spheres. J. Am. Chem. Soc. 79, 5333–5339. 16. Ramamurthy, N., Baliga, N., Wakefield, T. W., Andrews, P. C., Yang, V. C., and Meyerhoff, M. E. (1999) Determination of low- 34. Edsall, J., and Wyman, J. (1958) Biophysical Chemistry, Aca- molecular-weight heparins and their binding to protamine and a demic Press, New York, NY. protamine analog using polyion-sensitive membrane electrodes. 35. Tanford, C., Swanson, S. A., Shore, W. S. (1955) Hydrogen ion Anal. Biochem. 266, 116–124. equilibria of bovine serum albumin. J. Am. Chem. Soc. 77, 6414– 17. Crescenzi, V., Rizzo, R., and Manzini, G. (1981) Model system 6421. studies of polysaccharide–proteins interactions. Period. Biol. 83, 36. Casu, B. (1985) Structure and biological activity of heparin. Adv. 31–38. Carbohydr. Chem. Biochem. 43, 51–134. 18. Patel, H. V., Vyas, A. A., Vyas, K. A., Liu, Y., Chiang, C., Chi, L., 37. Nagasawa, M., Murase, T., and Kondo, K. (1965) Potentiometric and Wu, W. (1997) Heparin and heparan sulfate bind to snake titration of stereoregulator polyelectrolytes. J. Phys. Chem. 69, cardiotoxin. J. Biol. Chem. 272, 1484–1492. 4005–4012. 19. Olson, S. T., Halvorson, H. R., and Bjork, J. (1991) Quantitative 38. Park, J. W., and Chakrabarti, B. (1977) Acid–base and optical characterization of the thrombin–heparin interaction. J. Biol. properties of heparin. Biochem. Biophys. Res. Commun. 78, 604– Chem. 266, 6342–6352. 608. 20. Maccarana, M., and Lindahl, U. (1997) Mode of interaction be- 39. Wang, H., Loganathan, D., and Linhardt, R. J. (1991) Determi- tween platelet factor 4 and heparin. Glycobiology 3, 271–277. nation of the pKa of glucuronic acid and the carboxy groups of 21. Gao, J. Y., Dubin, P. L., and Muhoberac, B. B. (1997) Measure- heparin by 13C-nuclear-magnetic resonance spectroscopy. Bio- ment of the binding of proteins to polyelectrolytes by frontal chem. J. 278, 689–695. analysis continuous capillary electrophoresis. Anal. Chem. 69, 40. Gatti, G., Casu, B., Hamer, G. K., and Perlin, A. S. (1979) 2945–2951. Studies on the conformation of heparin by 1H and 13C NMR 22. Busch, M. H. A., and Kraak, H. P. (1997) Principles and limita- spectroscopy. Macromolecules 12, 1001–1007. tions of methods available for the determination of binding con- 41. Kohn, R., and Kovac, P. (1978) Dissociation constants of D- stants with affinity capillary electrophoresis. J. Chromatogr. A. 777, 329–353. galacturonic and D-glucuronic acid and their O-methyl deriva- tives. Chem. Zvesti 32, 478–485. 23. Gunnarsson, K., Valtcheva, L., and Hjerten, S. (1997) Capillary zone electrophoresis for the study of the binding of antithrombin 42. Cleland, R. L., Wang, J. L., and Detweiler, D. M. (1982) Poly- to low-affinity heparin. Glycoconjugate J. 14, 859–862. electrolyte properties of sodium hyaluronate. 2. Potentiometric titration of hyaluronic acid. Macromolecules 15, 386–395. 24. Heegaard, N. H. H. (1998) A heparin-binding peptide from hu- man serum P component characterized by affinity cap- 43. Mattison, K. W., Brittain, I. J., and Dubin, P. L. (1995) Protein– illary electrophoresis. Electrophoresis 19, 442–447. polyelectrolyte phase boundaries. Biotech. Prog. 11, 632–637. 25. VanderNoot, V. A., Hileman, R. E., Dordick, J. S., and Linhardt, 44. Park, J. M., Muhoberac, B. B., Dubin, P. L., and Xia, J. (1992) R. J. (1998) Affinity capillary electrophoresis employing immo- Effect of protein charge heterogeneity in protein–polyelectrolyte bilized glycosaminoglycan to resolve heparin-binding peptides. complexation. Macromolecules 25, 290–295. Electrophoresis 19, 437–441. 45. McGhee, J. D., and von Hipple, P. H. (1974) Theoretical aspects 26. Wu, X., and Linhardt, R. J. (1998) Capillary affinity chromatog- of DNA–protein interaction: Co-operative and non-co-operative raphy and affinity capillary electrophoresis of heparin binding binding of large ligands to a one-dimensional homogeneous lat- proteins. Electrophoresis 19, 2650–2653. tice. J. Mol. Biol. 86, 469–489. ELECTROPHORETIC DETERMINATION OF SERUM ALBUMIN–HEPARIN BINDING 167

46. Hallberg, R. K., and Dubin, P. L. (1998) Effects of pH on the 52. Myrup, B., Hansen, P. M., Jensen, T., Kofoed-Enevoldsen, A., binding of ␤-lactoglobulin to sodium polystyrene sulfonate. J. Feldt-Rasmussen, B., Gram, J., Kluff, C., Jespersen, J., and Phys. Chem. B 102, 8629–8633. Deckert, T. (1995) Effect of low-dose heparin on urinary albumin 47. Hattori, T., Hallberg, R. K., and Dubin, P. L. (2000) Roles of excretion in insulin-dependent mellitus. Lancet 345, electrostatic interaction and polymer structure in the binding of 421–422. ␤-lactoglobulin to anionic polyelectrolytes: Measurement of bind- 53. Yokoyama, H., Høyer, P. E., Hansen, P. M., van den Born, J., ing constants by frontal analysis continuous capillary electro- Jensen, T., Berden, J. H. M., Deckert, T., and Garbarsch, C. phoresis. Langmuir 16, 9738–9743. (1997) Immunohistochemical quantification of heparan sulfate 48. Epstein, I. R. (1978) Cooperative and noncooperative binding of proteoglycan and collagen IV in skeletal muscle capillary base- large ligands to a finite one-dimensional lattice: A model for ment membrane of patients with . Diabetes ligand–oligonucleotide interactions. Biophys. Chem. 8, 327–339. 46, 1875–1880. 49. Comper, W. D., and Laurent, T. C. (1978) Physiological func- 54. Urban, J., and Hall, A. (1992) Physical modifiers of cartilage tion of connective tissue polysaccharides. Physiol. Rev. 58, metabolism. in Aricular Cartilage and Osteoarthritis (Kuettner, 255–315. K., et al., Eds.), pp. 393–409, Raven Press, New York. 50. Kronman, M. J., and Foster, J. F. (1957) Sedimentation and 55. Yusipova, N. A., and Kriuk, A. S. (1979) Articular cartilage, optical rotation behavior of bovine plasma albumin at low pH in blood serum glycosaminoglycans, and glycoproteins in osteoar- presence of various anions: Effect of charge on molecular expan- thritis deformans. Clin. Chim. Acta 94, 9–21. sion. Arch. Biochem. Biophys. 72, 205–218. 56. Vilamitjana-Ame´de´e, J., and Harmand, M. F. (1990) Biochemical 51. Gambaro, G., Cavazzana, A. O., Luzi, P., Piccooli, A., Borsatti, analysis of normal and osteoarthritic human cartilage. Clin. A., Crepaldi, G., Marchi, E., and Venturini, A. P. (1992) Glycos- Physiol. Biochem. 8, 221–230. aminoglycans prevent morphological renal alterations and albu- 57. Carter, D. C., and Mo, J. X. (1994) Structure of serum albumin, minuria in diabetic rats. Int. 42, 285–291. Adv. Protein Chem. 45, 153–158.