throbber
THE JOURNAL
`OF BIOLOGICAL CHEMISTRY
`0 1987 by The American Society for Biochemistry and Molecular Biology, Inc.
`
`Vol. 262, No. 23, Issue of August 15, pp. 10980-10985,1987
`Printed in W.S.A.
`
`Carbon 13 NMR Studies of Saturated Fatty Acids Bound to Bovine
`Serum Albumin
`11. ELECTROSTATIC INTERACTIONS IN
`
`INDIVIDUAL FATTY ACID BINDING SITES*
`
`(Received for publication, September 10, 1986)
`
`David P. Cistola, Donald M. Small, and James A. Hamilton
`From the Biophysics Institute, Housman Medical Research Center, Departments of Medicine and Biochemistry,
`Boston University School of Medicine, Boston, Massachusetts 02118-2394
`
`“C NMR chemical shift results as a function of pH
`for a series of carboxyl “C-enriched saturated fatty
`acids (8-18 carbons) bound to bovine serum albumin
`(BSA) are presented. For octanoic acid bound to BSA
`(6:1, mol/mol), the chemical shift of the only FA car-
`boxyl resonance (designated as peak c), plotted as a
`function of pH, exhibited a complete sigmoidal titration
`curve that deviated in shape from a corresponding
`theoretical Henderson-Hasselbach curve. However, FA
`carboxyl chemical shift plotted as a function of added
`HC1 yielded a linear titration curve analogous to those
`obtained for protein-free monomeric fatty acid (FA) in
`water. The apparent pK of BSA-bound octanoic acid
`was 4.3 f 0.2. However, the intrinsic pK (corrected
`for electrostatic effects resulting from the net positive
`charge on BSA) was approximately 4.8, a value iden-
`tical to that obtained for monomeric octanoic acid in
`water in the absence of protein. For long-chain FA
`(212 carbons) bound to BSA (6:1, mol/mol), chemical
`shift titration curves for peak c were similar to those
`obtained for octanoic acid/BSA. However, the four
`additional FA carboxyl resonances observed (desig-
`nated as peaks a, b, b‘, and d ) exhibited no change in
`chemical shift between pH 8 and 3. For C14.,,.BSA
`complexes (3:l and 6:1, mol/mol) peaks b’ and a exhib-
`ited chemical shift changes between pH 8.8 and 11.5
`concomitant with chemical shift changes in the t-car-
`bon (lysine) resonance. In contrast, peaks c and d ex-
`hibited no change and peak b only a slight change in
`chemical shift over the same pH range. We conclude:
`(i) the carboxyl groups of bound FA represented by
`peaks a, b, b’, and d were involved in ion pair electro-
`static interactions with positively charged amino acyl
`residues on BSA; (ii) the carboxyl groups of bound FA
`represented by peak c were not involved in electro-
`static interactions with BSA; (iii) the similarity of the
`titration curves of peak c for BSA-bound octanoic acid
`and long-chain FA suggested that short-chain and
`long-chain FA represented by peak c were bound to
`the same binding site(s) on BSA; (iv) bound FA repre-
`sented by peaks b’ and a (but not d or b) were directly
`adjacent to BSA lysine residues. We present a model
`which correlates NMR peaks b, b’, and d with the
`
`*This research was supported by United States Public Health
`Service Grants HL-26335 and HL-07291. This work was originally
`submitted in partial fulfillment of the degree of Doctor of Philosophy
`at Boston University (Cistola, 1985a), and preliminary accounts of
`portions of this work have been published in abstract form (Cistola,
`1985b). The costs of publication of this article were defrayed in part
`by the payment of page charges. This article must therefore be hereby
`marked “advertisement” in accordance with 18 U.S.C. Section 1734
`solely to indicate this fact.
`
`putative locations of three individual high-affinity
`binding sites in a three-dimensional model of BSA.
`
`In biological systems, free (nonesterified) fatty acids often
`associate with macromolecules or macromolecular assemblies.
`For example, FFA’ in the circulation readily associate with
`albumin and, under certain physiological or pathological con-
`ditions, may associate with lipoproteins, platelets, red blood
`cells, and neutrophils (Spector and Fletcher, 1978). In order
`to predict the physical states formed by FFA in these complex
`systems, the ionization state of the FA carboxyl group must
`be known (Small, 1968, 1986; Cistola, 1985a, 1985b; 1986).
`Knowledge of the ionization behavior of bound FFA would
`aid in understanding not only the mechanism of binding of
`FFA to a given component of the circulation but also the
`mechanism of partitioning of FFA between these components
`and the uptake of FFA by tissue parenchymal cells. However,
`determination of the ionization behavior of FFA in systems
`containing multiple ionizable groups and complex macromo-
`lecular aggregates is difficult using conventional potentiomet-
`ric methods and simply assumptions about pK, values, and
`more specific methods, such as NMR titration, must be used.
`The sensitivity of NMR chemical shifts to the ionization
`state of chemical groups is well established (Jardetzky and
`Jardetzky, 1958) and has been extensively utilized to follow
`the ionization behavior of specific residues in amino acids and
`proteins (for a review, see Jardetzky and Roberts, 1981). Of
`the various NMR resonances which exhibit titration shifts,
`I3C carboxyl resonances are among the most useful because
`of their large titration shift range and resolution from ali-
`phatic, aromatic, and even carbonyl carbons.
`This paper presents 13C NMR titration results for a series
`of carboxyl I3C-enriched saturated fatty acids (8-18 carbons)
`bound to bovine serum albumin. NMR chemical shift titration
`curves at low pH (pH 8-3) indicated differences between
`high-affinity and low-affinity binding sites with respect to the
`ionization behavior of bound FA and the involvement of FA/
`BSA electrostatic interactions in individual FA binding sites.
`NMR chemical shift titration curves at high pH (pH 7-12)
`monitored the ionization behavior of lysine €-ammonium
`groups and the effects of lysine ionization on the carboxyl
`
`The abbreviations used are: FFA, free nonesterified fatty acid($;
`FA, fatty acid(s); BSA, bovine serum albumin; Ce.0, octanoic acid;
`C1z.o, dodecanoic (lauric) acid; C14.0, tetradecanoic (myristic) acid;
`C,,,, hexadecanoic (palmitic) acid; C1e.o, octadecanoic (stearic) acid;
`6, chemical shift; A& total change in chemical shift. The use of the
`abbreviation “FA” or the numerical abbreviations for individual FA
`compounds is not meant to imply anything about the ionization state
`of the carboxyl group.
`
`10980
`
`MPI EXHIBIT 1009 PAGE 1
`
`MPI EXHIBIT 1009 PAGE 1
`
`

`

`DH
`
`"I IN ncI ADDED
`FIG. 1. "C NMR chemical shift titration curves for 1.6 mM
`[ l-'sC]Ca.o in water at 40 "C. A, FA carboxyl chemical shift plotted
`as a function of pH. The circles represent experimentally measured
`values and the triangles theoretically calculated (Henderson-Hassel-
`bach) values.' B, FA carboxyl chemical shift plotted as a function of
`added 1 N HC1. The lines represent a least squares fit of the data
`points ( r = 0.99).
`
`chemical shifts of bound FA in individual binding sites. These
`results permitted further
`correlation of FA
`carboxyl reso-
`nances with the putative locations of FA binding sites in a
`three-dimensional model of BSA (Brown and Schockley, 1982;
`Cistola et al., 1987).
`
`EXPERIMENTAL PROCEDURES
`All materials and sample preparation procedures were described in
`detail in the accompanying paper (Cistola et al., 1987). 13C NMR
`spectra of FA. BSA complexes (3:l and 6:l mole ratio) were recorded
`as a function of decreasing or increasing pH using instrumentation
`and techniques described elsewhere (Cistola et al., 1982a, 1987). All
`FA used in this study were 90% 13C enriched at the carboxyl carbon
`position.
`
`13C NMR Ionization Behavior of Fatty Acid-Albumin Complexes
`10981
`obtained for shorter-chain carboxylic acids (Cistola et al.,
`1982a). Fig. 1B presents the same chemical shift data plotted
`as a function of added HCl rather than pH. A linear decrease
`in carboxyl chemical shift was observed between 0.4 and -3.2
`p1 of added HC1 with break points at or near the ionization
`end points. This linear titration curve is analogous to those
`obtained for water-miscible carboxylic acids in the absence of
`protein (Cistola et al., 1982a, 1982b).
`13C NMR titration curves for C8.,.BSA complexes ( 6 1 mol
`ratio) as a function of pH and added HCl are shown in Fig.
`2, A and B, respectively. The chemical shifts of the only FA
`carboxyl peak observed (peak c; Cistola et al., 1987) exhibited
`a complete sigmoidal titration curve (Fig. 2 A , circles) but
`deviated somewhat from the corresponding theoretical Hen-
`derson-Hasselbach curve2 (Fig. 2 A , triangles). The apparent
`pK, derived from the experimental curve, was 4.3 k 0.2. The
`chemical shift versus microliter HC1 (Fig. 2B) plot exhibited
`a linear decrease in chemical shift from 35 to 147 p1 of added
`HCl, with break points at the FA ionization end points, and
`suggested an ionization behavior analogous to protein-free FA
`monomers in water (Fig. 1B and Cistola et al., 1982a).
`13C NMR spectra for C14.0.BSA complexes (6:l mole ratio)
`at selected pH values are shown in Fig. 3. At pH 7.1 (Fig. 3A),
`four partially resolved FA resonances (peaks a, b, b', and c;
`Cistola et al., 1987) were observed. Below pH 7.1, only peak c
`exhibited chemical shift changes; it shifted upfield with de-
`creasing pH from 181.94 ppm (pH 7.1) to 177.53 ppm (pH
`2.9). In contrast, peaks a, b, and b' exhibited no chemical
`shift changes with decreasing pH. (However, peak a decreased
`in intensity below pH 5 and peaks b and b' decreased in
`intensity below pH 4.)
`Fig. 4 displays chemical shift titration curves for CI4.,, . BSA
`complexes. For peak c, a plot of FA carboxyl chemical shift
`as a function of pH exhibited a sigmoidal titration curve (Fig.
`4A, solid circles) which deviated from the calculated theoret-
`ical Henderson-Hasselbach curve' (Fig. 44, triangles). The
`apparent pK, estimated from the experimental curve, was 4.1
`f 0.2. In contrast, peaks a, b, and b' exhibited little or no
`chemical shift changes with decreasing pH (Fig. 44, open
`circles). Fig. 4B shows a linear dependence of the chemical
`shift of peak c on the quantity of added HC1 (solid circles),
`with break points at the FA ionization end points, analogous
`to data for Cao bound to BSA (Fig. 2B) and protein-free C,,
`(Fig. 1B). In contrast, FA carboxyl peaks a, b, and b' showed
`little or no chemical shift changes with added HC1 (Fig. 4B,
`open circles). Thus, with respect to ionization behavior (as
`followed by 13C NMR), C14.,. BSA spectra exhibited two types
`of FA carboxyl peaks: one type (peak c ) exhibited complete
`titration curves similar to those for protein-free FA in water,
`and the second type (peaks b, b', and a ) exhibited little or no
`chemical shift change with decreasing pH or added HC1.
`
`RESULTS
`As a basis for understanding the more complex NMR
`titration curves for FA.BSA complexes, NMR titration curves
`for protein-free aqueous [l-'3C]C8.0 were determined (Fig. 1).
`The concentration of CS., used (1.7 mM) was far below the
`critical micelle concentration of potassium octanoate (400
`mM; Mukerjee and Mysels, 1970) and slightly below the
`solubility limit of fully protonated octanoic acid (2.2 mM; Bell,
`1973). The samples exhibited with no visible
`turbidity or
`phase separation at any pH value studied. Therefore, Cao was
`apparently in monomeric solution throughout the titration.
`Two types of plots are shown in Fig. 1. In Fig. lA, the
`carboxyl chemical shift of Cao was plotted as a function of
`pH. The experimental points (circles) exhibited a sigmoidal
`curve and were essentially consistent with the calculated
`points (triangles) for a theoretical Henderson-Hasselbach ti-
`tration curve.2 The pK derived from the experimental curve
`was 4.8 f 0.1; this value is essentially the same as those
`' The theoretical Henderson-Hasselbach curve was calculated from
`an NMR version of the Henderson-Hasselbach equation
`
`where amax represents the carboxyl chemical shift for fully ionized FA;
`the chemical shift for fully protonated FA, and 8, the chemical
`shift at a given pH value. Equation 1 was derived from the more
`commonly used form of the Henderson-Hasselbach equation
`pH = pK + log[A-]/[HA]
`(2)
`by equating the following expressions for the fractions of ionized FA
`present (Cistola et al., 1982a)
`[A-I/([HAl + [AI) = (8 - 8min)/(8max - 8 m d
`(3)
`and by solving for [A-]/[HA]. LA-] and [HA] represent the concen-
`trations (or activities) of ionized and protonated FA, respectively. Bv
`substituting an experimentally derived pK value into Equation i, pH
`values can be calculated for a given 6 value.
`
`$
`
`t
`
`
`
`PI tN HCI ADDED
`BH
`FIG. 2. "C NMR chemical shift titration curves for [l-''C]
`Cs.o.BSA, 6:l mol ratio, at 35 OC. A, FA carboxyl chemical shift
`plotted as a function of pH. 0--0, experimentally determined curve;
`A- - -A, theoretically calculated Henderson-Hasselbach curve.' B.
`FA carboxyl chemicalshift plotted
`as a function of added HCl.
`
`MPI EXHIBIT 1009 PAGE 2
`
`MPI EXHIBIT 1009 PAGE 2
`
`

`

`10982
`
`C
`
`bl "1\
`
`
`
`I3C NMR Ionization Behavior
`
`of Fatty Acid-Albumin
`
`Complexes
`
`C
`
`b b' I\
`
`FIG. 5. "C NMR chemical shift titration curves for [1-13C]
`C12.0.BSA, 6:l mol ratio, at 35 "C. Symbols are defined as de-
`scribed in the legend to Fig. 4. A, FA carboxyl chemical shifts as a
`function of pH. B, FA carboxyl chemical shifts plotted as a function
`of added HCI.
`
`A
`
`A
`
`d4 3.3
`
`b-t
`
`.n 2.9
`170
`
`188 n 182 T
`
`-
`
`178
`
`-
`7
`
`,
`
`.
`.
`174 7"-
`~
`
`
`7
`
`T
`
`.
`
`
`
`. . 170 \
`
`"
`
`;
`
`r
`
`I88
`
`182
`178
`174
` (PPM)
`CHEMICAL SHIFT IPPM)
`CHEMICAL W T
`spectra for
`FIG. 3. Carboxyl/carbonyl region of "C NMR
`[1-'3C]C14.0~BSA, 6:l mol ratio, at different pH values at
`34 "C. Spectra were recorded after 6,000 accumulations with a pulse
`interval of 2.5 s, 16,384 time domain points, and a spectral width of
`10,000 Hz. Identical scaling and processing factors (including 3.0-Hz
`line broadening) were used for each spectrum. The lower case letters
`above certain peaks indicate specific FA carboxyl resonances with
`characteristic chemical shifts at pH 7.4 (Cistola et al., 1987).
`
`179.0-
`
`178.0
`3.0
`
`'
`
` '
`
`I
`4.0
`
`
`
`'
`
`1
`6.0
`
` 1
`
`1
`7.0
`
`8
`
`
`80
`
`
`
`120
`
`4
`6.0
`40
`pl 1N HCl
`PH
`FIG. 6. "C NMR chemical shift titration curves for [1-13C]
`Cle.o.BSA, 6:l mol ratio, at 35 "C. A, FA carboxylehemical shift
`plotted as a function of pH. B, FA carboxyl chemical shift plotted as
`a function of added HCI. The FA carboxyl peaks became undetectable
`below pH 4.0 (above 100 pl of HCI).
`
`I
`
`0
`
`.--.,
`
`plot of chemical shift as a function of pH for peak c (Fig. 5A,
`solid circles) was sigmoidal with an apparqt pK, of 4.1 k 0.2
`but deviated from the calculated Henderson-Hasselbach curve
`(Fig. 5A, triangles). A plot of chemical shift as a function of
`added HC1 for peak c (Fig. 5B, solid circles) was linear ( r =
`-0.99) with break points at the FA ionization end points,
`analogous to data for C14.0'BSA (peak c; Fig. 3), Cao. BSA
`(peak c, Fig. 2), and protein-free C8., (Fig. 1).
`13C NMR spectra for Cl6.,-BSA and C18,,.BSA complexes
`(not shown) were similar to those for C14.0. BSA and C,,,.
`BSA except for one major difference: for C16.,.BSA and C18,,.
`BSA, peak c became undetectable below pH 3.9. Therefore,
`complete titration curves for peak c could not be generated.
`However, a partial titration curve for c16.0. BSA (peak c) is
`shown in Fig. 6. As with C14.O'BSA and C12.0.BSA, only peak
`c exhibited chemical shift changes with decreasing pH or
`added HCl for C16.,-BSA (Fig. 6) and Clao.BSA (not shown).
`Peaks a, b, and d yielded no chemical shift changes with
`decreasing pH or added HC1.
`Two C14.,.BSA samples (3:l and 6:l mole ratio) were ti-
`trated with 1.0 N KOH from pH 7.4 to 11.9. At 3:l mole ratio,
`essentially all of the c14.0 was bound to the three high-affinity
`binding sites on BSA (represented by peaks b, b', and d;
`Cistola et oL, 1987). The chemical shift of the e-carbon of
`lysine increased with increasing pH beginning at pH 8.8
`concomitant with a decrease in chemical shift of FA carboxyl
`peak b' above 8.8 (Fig. 7). FA carboxyl peak b exhibited a
`
`PH
`p l l N HCl
`FIG. 4. 13C NMR chemical shift titration curves for [l-"C]
`CI4.,.BSA, 6:l mol ratio, at 35 "C. 0--0, experimentally de-
`termined curves for peaks a, b, and b' (see Fig. 3);
`experi-
`mentally determined curves for peak c (Pig. 3); A- - -A, theoretically
`calculated Henderson-Hasselbach curves for peak c.' The experimen-
`tal data points were derived from the NMR spectra shown in Fig. 3
`as well as other spectra not shown. A, FA carboxyl chemical shifts
`plotted as a function of pH. E, FA carboxyl chemical shifts plotted
`as a function of added HCI.
`
`13C NMR spectra (not shown) of Clz.o.BSPv complexes (6:l
`mole ratio) obtained at pH values between 8.0 and 2.8 were
`very similar to those for C14,o.BSA complexes. The corre-
`sponding NMR chemical shift titration curves for C12.0 e BSA
`complexes are shown in Fig. 5. One FA carboxyl peak I(c)
`exhibited complete titration curves with decreasing pH (Fig.
`5A) or added HCl (Fig. 5B). In contrast, peaks b and b'
`exhibited no change in chemical shift over this pH range. A
`
`MPI EXHIBIT 1009 PAGE 3
`
`MPI EXHIBIT 1009 PAGE 3
`
`

`

`13C NMR Ionization
` Behavior
`
`-
`
`i ----.-I
`
`t
`
`181.0
`
`180.5
`
`5
`
`1
`
`
`1
`7.0 8.6
`
`1
`7.8
`
`
`
`1
`
`1
`
`1
`
`1
`
`
`
`
`1
`
`1
`
`1
`
`1
`1
`zT4
`10.2 11.0
`V H
`FIG. 7. "C NMR chemical shift titration curves at high pH
`for [l-"C]C,,.,.BSA, 3:l mol ratio, at 34 "C. The NMR sample
`was titrated with 1 N KOH from pH 7.4 to 11.9. External tetrameth-
`ylsilane in a capillary was used as a chemical shift reference. Vertical
`error bars represent estimated uncertainties in chemical shift deter-
`minations. A , chemical shift of €-carbon (lysine) resonance as a
`function of pH. B, carboxyl chemical shift of FA carboxyl peaks b, b',
`and d (cf. Fig. 3) and BSA glutamate carboxyl peak p r (Cistola et al.,
`1987) as a function of pH.
`
`Complexes
`of Fatty Acid-Albumin
`ing protein residues in individual FA binding sites.
`Based on studies of spin-label FA analogues or detergents
`bound to albumin (Morrisett et al., 1975), it has been sug-
`gested that hydrophobic interactions are the dominant mech-
`anism by which FA bind to albumin (Spector, 1975). For alkyl
`sulfate detergents, binding affinities for albumin increase as
`the alkyl chain length increases (Karush and Sonnenberg,
`1949). Aminoazobenzene, an uncharged derivative of methyl
`orange, binds nearly as well as methyl orange itself (Klotz
`and Ayers, 1952). For organic anions, there is a large positive
`entropy change for the first mole bound but only a small
`enthalpy change (Klotz and Urquhart, 1949), suggesting that
`binding is primarily an entropy effect resulting from a release
`of water when the ligand protein complex forms (Spector,
`1975).
`However, there is also evidence that electrostatics play at
`least a minor role in ligand/albumin interactions. Based on
`indirect evidence, it has long been thought that FA bound to
`albumin are anionic at pH 7.4 (Ballou et ab, 1945). This
`conclusion was directly confirmed by the 13C NMR titration
`curves presented previously for oleic acid (Parks et al., 1983)
`and myristic acid (Hamilton et al., 1984) along with curves
`presented in this study, at least for bound FA represented by
`relative absorption of an
`NMR peak c. Furthermore, the
`anionic dye (methyl orange) bound to BSA was altered over
`a pH range in which the €-ammonium group of lysine residues
`would be expected to ionize (Klotz and Walker, 1947). In
`addition, studies using spin-label analogues of fatty acids have
`shown that methyl ester analogues bind with lower affinity
`than carboxylate ion analogues (Morrisett et al., 1975). How-
`ever, few studies have utilized native FA, and little informa-
`tion is available about the relative importance of hydrophobic
`and electrostatic interactions in individual binding sites. The
`NMR titration results presented in this study address both of
`these concerns.
`In titration from pH 8 to pH 3 (Figs. 2-6), FA bound to
`BSA exhibited two types of ionization behavior. The first
`type (peak c only) was characterized by NMR titration curves
`analogous to those obtained for protein-free monomeric FA
`in water. In contrast, the second type of ionization behavior
`(peaks a, b, b', and d ) was characterized by no changes in FA
`carboxyl chemical shift with decreasing pH.
`The ionization behavior of peak c was analogous to protein-
`free FA monomers in water according to three criteria: A6
`values, apparent pK values, and the shape of NMR titration
`curves. First, the A6 value for FA.BSA complexes (except
`c8.0.BSA) was 4.7 ppm, identical to that for aqueous protein-
`free C8.0 (Fig. 1) and aqueous short-chain carboxylic acids
`(Cistola et al., 1982a). Second, the experimentally determined
`apparent pK values of FA bound to BSA (pK = 4.1-4.3) were
`somewhat lower than values obtained for protein-free mono-
`meric FA in water (pK = 4.7-4.9; Fig. 1 and Cistola et al.,
`1982a). However, after correction of these pK, values for
`electrostatic effects3 resulting from the net positive charge on
`
`10983
`
`1
`
`
`
`~
`11.8
`
`1
`
`
`
`small but significant increase with increasing pH, and peaks
`d andpr (glutamate carboxyl) demonstrated no chemical shift
`change with increasing pH. In spectra of 6:l mole ratio
`complexes, in which peaks a and c were observed (in addition
`to peaks b, b', and d), the chemical shift of peak a decreased
`above pH 8.8 (data not shown) in a manner similar to peak
`b' (Fig. 7).
`
`DISCUSSION
`As shown in the accompanying paper (Cistola et al., 1987),
`13C NMR spectra of 13C-enriched saturated long-chain FA
`bound to BSA at pH 7.4 contained multiple FA carboxyl
`resonances corresponding to multiple FA binding sites. Three
`of these resonances (b, b', and d ) were correlated with the
`three high-affinity long-chain FA binding sites deduced from
`Scatchard analyses of binding Qata (Spector et al., 1969) and
`from Brown and Schockley's (1982) three-dimensional model
`of BSA. In the latter model, NMR peaks b, b', and d appar-
`ently represented long-chain FA bound to binding sites 1-C,
`3-C, and 2-C, respectively (Cistola et al., 1987). In contrast,
`the FA carboxyl resonance labeled c was correlated with the
`nonspecific sites such as these located in subdomains 1-AB,
`2-AB, and/or 3-AB of Brown and Schockley's model. Peak a
`represented a low-affinity (secondary) long-chain FA binding
`site in subdomain 2-AB. In this study, the ionization behavior
`of ['3C]carboxyl-enriched FA bound to BSA was determined
`by 13C NMR spectroscopy in order to assess whether ion pair
`interactions occur between FA carboxyl groups and neighbor-
`
`Intrinsic pK values (pKi,,) for FA (peak c) bound to BSA were
`estimated using the equation
`= pH + 0.868wZ
`where the term 0.868wZ takes into account the electrostatic interac-
`tions between the net charge ( Z ) of BSA at a given pH value and
`dissociating hydrogen ions (Tanford et al., 1955). Values for the
`electrostatic interaction factor (w) can be determined empirically or
`calculated from an equation derived from Debye-Huckel theory (Tan-
`ford et al., 1955). We used values of 0.052 and 0.065, respectively, in
`these calculations. The net charge Z was determined from the equa-
`tion
`
`MPI EXHIBIT 1009 PAGE 4
`
`MPI EXHIBIT 1009 PAGE 4
`
`

`

`10984
`
`b
`
`- 1-c
`
`Site
`
`- 3-c
`b '
`
`Site
`
`13C NMR Ionization Behavior of Fatty Acid-Albumin Complexes
`BSA at pH 4.1, the resulting intrinsic pK values (4.5-4.8)
`resonances. Since conformational changes have not been de-
`were essentially the same as values for protein-free mono-
`tected over this pH range (Foster, 1977), this explanation is
`unlikely. (ii) FA carboxyl ionization may have given rise to
`meric FA in water. Third, the shape of NMR chemical shift
`titration curves for peak c were very similar to those obtained
`the observed chemical shift changes in peak 6' and also
`for aqueous protein-free FA (Fig. 1). Although plots of chem-
`affected nearby lysine residues. However, the pK values of
`ical shifts (peak c) as a function of pH yielded titration curves
`carboxyl groups are generally much lower than the pH range
`which deviated from idealized Henderson-Hasselbach ioni-
`being considered here. In addition, the chemical shift of peak
`b' exhibited a much smaller A6 value than, and changed in
`zation behavior, chemical shift versus added HCI exhibited a
`the opposite direction of, that expected for FA carboxyl ioni-
`linear decrease analogous to that obtained for protein-free
`monomeric FA in water (Fig. 1 and Cistola et al., 1982a).
`zation. Therefore, this second explanation is highly unlikely.
`Thus, changes in FA chemical shifts were sensitive only to
`The most plausible explanation for changes in the molec-
`changes in FA ionization state, rather than the
`carboxyl
`ular environment of the FA carboxyl (between pH 8.8 and
`group's molecular environment (e.g. alterations caused by
`11.3) is ionization of a basic amino acid residue such as lysine,
`BSA conformational changes). The latter would have been
`tyrosine, or arginine. Arginine guanidinium groups have ex-
`characterized by a nonlinearity, a change in slope of chemical
`tremely high intrinsic pK values (>13) in small molecules
`shift titration curves, or a A6 unequal to the expected value.
`such as guanidine and substituted guanidines (Cohn and
`The linear plots also suggested that the deviation from Hen-
`Edsall, 1943; Neivelt et al., 1951) and pK values >12 in
`derson-Hasselbach ionization behavior shown in Figs. 2 A , 4A,
`proteins such as lysozyme and BSA (Tanford et al., 1955).
`and 5A arose in the pH term, not the chemical shift term, of
`For tyrosine phenolic hydroxyl groups in BSA, the experi-
`Equation I.* The most likely explanation for anomalous pH
`mentally observed pH value at the ionization midpoint (un-
`values in the carboxyl ionization region is the unmasking of
`corrected for electrostatic effects resulting from the net charge
`BSA carboxylate groups during the N-F conformational tran-
`of the protein) was 11.5 (Tanford and Roberts, 1952; Decker
`sition below pH 5 (Foster, 1977). Apparently, the N-F tran-
`and Foster, 1967). Since the observed effect on FA occurs
`sition does not give rise to anomalous FA carboxyl chemical
`between pH 8.8 and 11.5, these amino acids probably would
`shift values for the bound FA represented by peak c. However,
`not be responsible. In contrast, €-ammonium groups in BSA
`this transition does result in a loss of intensity of FA carboxyl
`have a pH midpoint for lysine ionization of approximately
`peaks (Fig. 3).4
`10.7 (Tanford et al., 1955). Chemical shift changes in the t-
`The similarities between the ionization behavior of FA
`carbon of lysine occurred over this pH range (Fig. 7A). There-
`bound to BSA (peak c) and the ionization behavior of protein-
`fore, we conclude that lysine ionization occurs between pH
`free FA monomers in water suggested that the FA carboxyl
`8.8 and 11.5 and affects the chemical environment of adjacent
`groups of bound FA represented by peak c were freely acces-
`FA carboxyl groups represented by peak b' (Fig. 7).
`sible to the aqueous solvent and, therefore, free of specific
`Taken together, the titration results for different FA allow
`electrostatic interactions with side chain residues on BSA. In
`contrast, the lack of NMR titration shifts
`for bound FA
`represented by peaks a, b, b', and d suggested that FA carboxyl
`groups in these binding sites were solvent-inaccessible and
`were probably involved in ion pair electrostatic interactions
`with basic amino acid residues that line the mouths of several
`of the putative FA binding sites on BSA (Brown and Schock-
`ley, 1982).
`In an attempt to determine which positively charged BSA
`side chain residues interact with the carboxylate groups of
`bound FA (represented by peaks a, b, b', and d ) , CM.~.BSA
`complexes (3:l and 6:l mole ratio) were titrated with KOH
`from pH 7.4 to pH 12.2. For 3:l mole ratio complexes (Fig.
`7), peak b' and the €-carbon resonance (lysine) exhibited
`substantial chemical shift changes between pH 8.8 and 11.3.
`In contrast, peak b exhibited only slight chemical shift
`changes and peaks d andpr no chemical shift changes between
`pH 8.8 and 11.3. There are three possible explanations for
`these results. (i) C,,,o. BSA may have undergone gross confor-
`mational change(s) over this pH
`range which could have
`altered the local molecular environment around
`lysine c-
`carbons and FA carboxyl carbons. Spectral changes charac-
`teristic of protein unfolding and/or peptide cleavage did occur
`at pH values >11.5. However, these changes did not occur
`between pH 8.8 and 11.3. Alternatively, a localized change in
`BSA structure might have affected only certain portions of
`the molecule and, hence, only certain of the observed NMR
`Z = 96 - r -
`where r was taken to be 80 and +cl (the moles of bound chloride ions/
`mol of BSA) 6 (Carr, 1953; Steinhardt and Reynolds, 1969). Final
`estimatedp& values were 4.6-4.7 for long-chain FA.BSA complexes
`and 4.8-4.9 for C8.,,-BSA complexes.
`The effect of BSA conformational changes on 13C NMR results
`will be addressed in detail in a future paper.
`
`d-1
`Hslir 3C.X
`0 8 - Hiss,i Thr,=- GIus3;-/
`Helix 3C.Y
`FIG. 8. Schematic diagram depicting putative locations of
`bound FA molecules in three high-affinity long-chain FA
`binding sites on BSA. The nomenclature used for a-helices and
`the amino acid sequences in the putative binding sites are
`from
`Brown and Schockley (1982). We propose that the FA represented
`by NMR peak b' is bound to site 3-C, the FA represented by peak b
`at site I-C, and the FA represented by peak d at site 2-C (see also
`Cistola et al., 1987).
`
`- d
`
`Site
`2-c
`
`@
`Qly,,~ Hi8=3- Cvs,,- C y s , d / H.lia 2A.v
`(Asp - Louzc Lou,,;- Q W , d /
`e =
` Hdix 2AB.Z
`93-1
`@&&&,= Swwj/ HJlx 2C.X
`
`BoundFA
`
` pro,,^ Q l u , ~ Tyr,-
`
`Ahm-/
`
`W i x 2C.Y
`
`Qlu,-
`
`pm,,
`
`(D
`L w . ~ l h r a d /
`Matw+/
`
` Hdir 3A.V
`Halix 3AB.Z
`
`(b.rn; Q&--F&
`9 3 C -1
`
`Bound FA
`
`Pro -&,&,g&,,"J/
`
`Qlu,, Aspw- AhU- Valu-/
`Ahc Oly,s"----J/
`cSsrc,,
`
`OY e
`
`la&,n&,U**au+/
`roln- Tyr,,~ Pho,~ TW,~+/
`
`Hslix 1 A.V
`H d i r 1 AB.2
`
`FA
`
`
`
`H*lix 1c.x
`Hdir 1 C.Y
`
`e?C -- Bound
`
`MPI EXHIBIT 1009 PAGE 5
`
`MPI EXHIBIT 1009 PAGE 5
`
`

`

`10985
`
`13C NMR Ionization Behavior of Fatty Acid-Albumin Complexes
`vidual binding sites on other proteins (e.g. intracellular fatty
`speculative correlation of BSA binding sites and NMR FA
`acid binding proteins, a-fetoprotein, etc.)
`carboxyl peaks. Inspection of the amino acid sequence at the
`mouths of the putative high affinity FA binding sites in Brown
`REFERENCES
`and Schockley's model (1982) reveals clusters of three basic
`Ballou, G. A., Boyer, P. D., and Luck, J. M. (1945) J. Biol. Chem.
`residues (Fig. 8). Binding sites 1-C and 2-C contain the triplets
`159, 111
`His-Arg-Arg (amino acids 143-145 and 334-336, respectively),
`Bell, G. H. (1973) Chem. Phys. Lipids 10,l-10
`and site 3-C contains the triplet Lys532-His533-Lys534. Of the
`Brown, J. R., and Schockley, P. (1982) in Lipid-Protein Interactions
`three NMR peaks which represent high affinity binding
`(Jost, P. C., and Griffith, 0. H., eds) Vol. 1, pp. 26-68, John Wiley
`(peaks b, b', and d; Cistola et al., 1987), only one (peak b')
`and Sons, New York
`Carr, C. W. (1953) Arch. Biochem. Biophys. 46, 417-423
`exhibits chemical shift changes which correspond with lysine
`Cistola, D. P. (1985a) Ph.D. thesis, Boston University School of
`ionization. Therefore, we propose that peak b' represents FA
`Medicine
`carboxyl carbons adjacent to the Lys-His-Lys cluster in sub-
`Cistola, D. P. (198513) Biophys. J. 47, 252a
`domain 3-C. The remaining high-affinity FA peaks ( b and d ) ,
`Cistola, D. P., Small, D. M., and Hamilton, J. A. (1982a) J. Lipid
`Res. 23, 795-799
`therefore, represent FA carboxyl carbons adjacent to His-Arg-
`Cistola, D. P., Small, D. M., and Hamilton, J. A. (1982b) Biophys. J.
`Arg clusters in sites 1-C and
`2-C (not necessarily in that
`37,204a
`order). Further inspection of the amino acid sequences in
`Cistola, D. P., Atkinson, D.,

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket