throbber
OF BIOLOGICAL CHEMISTRY
`THE JOURNAL
`0 1987 by The American Society for Biochemistry
`
`and Molecular Biology, Inc.
`
`Vol. 262, No. 23, Issue of August 15, PP. 10971-10979,1987
`Printed in U.S.A.
`
`Carbon 13 NMR Studies of Saturated Fatty Acids Bound to Bovine
`Serum Albumin
`I. THE FILLING OF 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
`
`13C NMR chemical shift and intensity results for a
`series of carboxyl “C-enriched saturated fatty acids
`(8-18 carbons) bound to bovine serum albumin (BSA)
`are presented as a function of increasing fatty acid
`(FA)/BSA mole ratio. Spectra for long-chain (212 car-
`bons) FA-BSA complexes exhibited up to five FA car-
`boxyl resonances, designated a, b, b’, c, and d. Only
`three resonances (peaks b, b’, and d ) were observed
`below 3:l FA*BSA mole ratio, and at r3:l mole ratio,
`two additional resonances were observed (peaks c and
`a). In a spectrum of 5:l stearic acid-BSA complexes,
`peaks b, b’, and d each represented approximately one-
`fifth, and peak c approximately two-fifths, of the total
`FA carboxyl intensity. Plots of total carboxyl/carbonyl
`intensity ratio as a function of FA-BSA mole ratio were
`linear up to 7-9 mole ratio. Deviation from linearity
`at mole ratios 27 was accompanied by the detection of
`crystalline unbound FA (as 1:l acidlsoap) by x-ray
`diffraction. In contrast to
`long-chain FA*BSA com-
`plexes, 13C NMR spectra of octanoic acid*BSA com-
`plexes yielded only one FA carboxyl resonance (peak
`c ) at FA-BSA mole ratios between 1 and 20. We con-
`clude: (i) peaks b, b’, and d represent FA bound to three
`individual high affinity (primary) long-chain FA bind-
`ing sites on BSA; (ii) peak c represents FA bound to
`several secondary long-chain (or primary short-chain)
`FA binding sites on BSA; (iii) peak a represents long-
`chain FA bound to an additional lower affinity binding
`site. We present a model that correlates the observed
`13C NMR resonances with individual binding site lo-
`cations predicted by a recent three-dimensional model
`of BSA.
`
`Utilization of circulating FFA’ by tissues is influenced not
`only by the avidity of FFA for and the blood flow through
`
`*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, 1985), and preliminary accounts of
`portions of this work have been published in abstract form (Cistola
`et al., 1983). 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.
`The abbreviations used are: FFA, free (nonesterified) fatty
`acid(s); FA, fatty acid(s); BSA, bovine serum albumin; NOE, nuclear
`Overhauser enhancement, C8.0, octanoic acid; C12.0, dodecanoic (lauric)
`acid; Clr.o, tetradecanoic (myristic) acid; Cla.o, hexadecanoic (palmitic)
`acid; C18.0, octadecanoic (stearic) acid; C18.1, oleic acid. 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 FA carboxyl group.
`
`each tissue, but also by the mole ratio of FFA to albumin in
`the circulation (Scow and Chernick, 1970; Spector and
`Fletcher, 1978). In normal human subjects, this ratio is vari-
`able and is elevated under certain metabolic or environmental
`conditions such as fasting (Frederickson and Gordon, 1958)
`and/or prolonged exercise (Have1 et al., 1967). Under certain
`pathological conditions, FFA/albumin ratios may be
`tran-
`siently or consistently elevated secondary to increased FFA
`mobilization (diabetic ketoacidosis, myocardial infarction,
`acute anxiety) or decreased circulating albumin (nephrotic
`syndrome, liver disease, familial hypoalbuminemia). It is con-
`ceivable that increased FFA production and/or decreased
`circulating albumin could result in abnormal partitioning of
`FFA into other components of the circulation (lipoproteins,
`blood cell membranes, endothelial cell membranes; Spector
`and Fletcher, 1978). This abnormal FFA partitioning might
`result in detrimental structural and/or functional alterations
`such as decreased neutrophil phagocytic and bacteriocidal
`activity (Hawley and Gordon, 1976), platelet aggregation
`(Hoak et al., 1970), and endothelial cell damage (Zilversmit,
`1976).
`As one approach to predicting FFA/albumin interactions
`at different mole ratios in uiuo, the FA binding properties of
`albumin have been extensively examined in vitro using several
`approaches (for a review, see Spector, 1975). Binding data
`obtained from equilibrium partitioning methods have been
`analyzed by the Scatchard model (Scatchard, 1949) or the
`stepwise association model (Klotz et al., 1946; Spector et al.,
`1971). The Scatchard analyses for long-chain FA bound to
`human (Goodman, 1958) and bovine (Spector et al., 1969)
`albumin have yielded the concept of three classes of FA
`binding sites with respect to relative affinities. In contrast,
`the stepwise association analyses assumed no grouping of
`binding constants into classes and suggested that this group-
`ing is somewhat arbitrary (Spector et al., 1971; Ashbrook et
`al., 1975). Second, mapping studies using peptide fragments
`et al., 1975),
`(King and Spencer, 1970; King, 1973; Reed
`chemical modifications (Koh and Means, 1979), or affinity
`labeling (Lee and McMenamy, 1980) have aided in the local-
`ization of ligand binding sites to general regions on
`the
`polypeptide sequence. However, pitfalls include possible dis-
`ruptive changes in protein conformation and binding proper-
`ties following fragmentation, inconsistencies in affinity label-
`ing, and a
`lack of specificity with chemical modification
`(Brown and Schockley, 1982). Third, spectroscopic studies
`using fluorescence (Sklar et al., 1977; Berde et al., 1979), ESR
`(Kuznetsor et al., 1975; Morrisett et al., 1975; Rehfield et al.,
`1978; Perkins et al., 1982), and NMR (Muller and Mead, 1973;
`Inoue et al., 1979) spectroscopy have yielded information
`concerning the physicochemical interactions of ligands and
`
`10971
`
`MPI EXHIBIT 1008 PAGE 1
`
`MPI EXHIBIT 1008 PAGE 1
`
`

`

`13C NMR of Fatty Acid-Albumin Interactions
`albumin. However, pitfalls include the need for fatty acids
`phoresis. I3C NMR spectra for C14.0.BSA complexes using monomeric
`or unfractionated BSA (obtained under identical conditions) were
`containing structure-perturbing spin-label probes (ESR) or
`indistinguishable. This result is consistent with our previous obser-
`conjugated double bonds (fluorescence). Also, 13C NMR at
`vations for C,,.,.BSA complexes (Parks et al., 1983).' Therefore,
`natural abundance has been hampered by a lack of sensitivity
`unfractionated BSA was used throughout this study.
`(Kragh-Hansen and Riisom, 1976).
`13C carboxyl-enriched (90%) fatty acids were purchased from KOR
`As an alternative approach, we have utilized 13C NMR
`Isotopes (Cambridge, MA) ( G o , Clz.o, c16.0, C18.,) and Merck Sharp
`spectroscopy with 13C-enriched fatty acids to investigate the
`and Dohme Isotopes (St. Louis, MO) (C14.0). Sample purity as deter-
`mined from thin layer chromatography (hexanediethyl ether:acetic
`interactions of biologically important FA with bovine albu-
`acid, 90:9:1) was >98%. In addition, no impurities were visualized by
`min. Carbon 13 enrichment greatly enhances spectral sensi-
`'H NMR for I3C-enriched FA samples dissolved in deuterated chlo-
`tivity and permits investigation of FA/albumin interactions
`roform.
`in the range of physiologically relevant FA/albumin mole
`Sample Preparation-BSA solutions (7%, w/v) were prepared using
`ratios. Using this approach, we have shown that the carboxyl
`doubly distilled deionized water. After adjusting the pH to 7.4, the
`chemical shift of oleic acid bound to BSA is highly sensitive
`solutions were centrifuged at 10,000 rpm for 30 min to remove trace
`amounts of particulate matter. The protein concentration was deter-
`to the FA binding environment on albumin; spectra revealed
`mined from the absorbance at 279 nm of 1:lOO dilutions using an
`multiple FA carboxyl resonances corresponding to multiple
`extinction coefficient of 6.67 for a 1% sample (Janatova et al., 1968).
`FA binding environments (Parks et al., 1983). In addition, we
`Crystals of 13C-enriched FA were dissolved in 2:l chloroform/meth-
`have utilized fatty acids with 13C enrichment in hydrocarbon
`anol, and the concentrations were determined by measuring dry
`chain carbons (C-3, (2-14) to probe the interactions of myristic
`weights on an electrobalance (Cahn model 25, Cerritos, CA). Stoichi-
`acid with BSA (Hamilton et al., 1984).
`ometric amounts of fatty acids in solvent were added to 10-mm NMR
`tubes, and the solvent was evaporated under N,. D20 (200 pl) and 1.2
`This paper presents 13C NMR results for a series of carboxyl
`eq of 1 N KOH were added, and samples were thoroughly mixed until
`I3C-enriched saturated fatty acids (8-18 carbons) bound to
`all FA crystals dissolved to form an optically clear micellar soap
`BSA as a function of increasing FA.BSA mole ratio. The
`solution. Hydrated BSA samples (1.8 ml, pH 7.4) were added to the
`results provide direct physicochemical information regarding
`soap solutions (0.2 ml) with continuous vortexing for several minutes,
`the order of filling and saturation of individual FA binding
`and samples were equilibrated and intermittently vortexed for 30
`min. Potassium stearate solutions formed a gel phase at room tem-
`sites with increasing FA.BSA ratio as well as the relative
`perature and had to be gently and briefly heated before BSA solutions
`occupation of individual sites at a given mole ratio. Second,
`were added. Sample pH was adjusted from pH 7.5-7.6 (following
`the results delineate differences between shorter chain and
`mixing) to 7.4 and samples were equilibrated at room temperature
`longer chain FA with regard to the number and type of FA
`(25 "C) for 8-12 h prior to I3C NMR experiments. The NMR results
`binding sites. Third, accompanying powder x-ray diffraction
`were independent of equilibration time.
`results provide information about the physical state of un-
`All pH measurements were made directly in the NMR tube using
`a pH meter (Beckman 3560, Fullerton, CA) equipped with a 29 cm x
`bound FA at high mole
`ratios. Finally, the results permit
`4 mm glass combination electrode (Markson MiraMark, Phoenix,
`direct correlation of observed NMR resonances with the FA
`AZ). Values measured before and after obtaining NMR spectra agreed
`binding sites predicted by recent models of BSA structure
`within 0.1 pH unit.
`(Brown and Shockley, 1982) and provide insights into the
`The final FA.BSA samples used for NMR contained from 0.5 to
`binding locations of FA at different FA/albumin ratios i n
`20 mol of FA/mol of BSA and 7% w/v protein. As reported elsewhere
`uiuo.
`(Cistola, 1985), I3C NMR spectra (FA as well as protein resonances)
`for C16.0.BSA complexes at 3.8, 7.5, and 11.4% w/v BSA (all at 5:l
`mol ratio, pH 7.4, 35 "C) were essentially identical. Hence, it is
`unlikely that noncovalent protein aggregation occurred over this
`concentration range. The salt concentrations (as determined by flame
`photometry) of 7% hydrated samples with no added salt were 7 mM
`(sodium) and 6 mM (potassium). No salt was added to FA.BSA
`samples in this study. Addition of KC1 to FA.BSA samples up to a
`final concentration of 0.1 M does not change 13C NMR results pro-
`vided that the sample temperature is kept below 38 "C (Cistola, 1985).
`Carbon 13 NMR Spectro~copy-~~C NMR spectra were obtained
`on a Bruker WP-200 NMR spectrometer at 50.3 MHz as described
`elsewhere (Hamilton and Small, 1981; Cistola et al., 1982). Internal
`D,O was used as a lock and shim signal. Chemical shift values were
`measured digitally with an estimated uncertainty of f O . l ppm. The
`chemical shift (6 = 39.57 ppm) of the narrow resonance from protein
`c-Lys/P-Leu carbons (Gurd and Keim, 1973) was used as an internal
`reference after calibrating this resonance against external tetrameth-
`ylsilane. To enhance spectral resolution in selected cases, the convo-
`lution difference method was used (Campbell et al., 1973). FA car-
`boxyl/BSA carbonyl intensity ratios were measured using the inte-
`gration routine provided in the Bruker DISNMR program. NMR
`sample temperatures were controlled to 34 "C and measured as de-
`scribed previously (Cistola et al., 1982). Spin-lattice relaxation times
`( Tl) were measured using a fast inversion recovery technique (Canet
`et al., 1975) and calculated using a three-parameter fitting routine
`(Sass and Ziessow, 1977). Nuclear Overhauser enhancements (NOE)
`were determined from comparisons of peak heights from spectra
`accumulated with broad-band and inverse-gated decoupling (Opella
`et al., 1976). For all spectral accumulations, pulse intervals were equal
`to the TI value of the largest FA carboxyl peak in the spectrum, and
`90 "C pulses (15 p s ) were used.
`
`EXPERIMENTAL PROCEDURES
`Materials-Essentially FA-free crystallized lyophilized BSA was
`purchased from Sigma (A-7511, lot 22F-9340). The supplier used
`method IV of Cohn et al. (1947) to recrystallize fraction V albumin
`and the charcoal defatting procedure of Chen (1967) to remove bound
`fatty acid. The content of bound fatty acid, as determined by gas-
`liquid chromatography, was C0.02 mol of FA/mol of BSA prior to the
`addition of FA. The protein content of the BSA sample was analyzed
`by sodium dodecyl sulfate-polyacrylamide gel electrophoresis; the gels
`were overloaded with sample in order to search for minor impurities.
`In addition to the major albumin band at 66,000 daltons, minor bands
`were observed at -120,000 (-5%), 55,000 (<1%), and 160,000 ( 4 % ) .
`These minor bands most likely corresponded to BSA dimers (Fried1
`and Kistler, 1970; Foster, 1977), a,-antitrypsin (Laurel1 and Jeppsson,
`1975), and immunoglobulins, respectively (Putnam, 1975; Peters,
`1975). Although apoprotein A-I is often present as a contaminant in
`commercial albumin preparations (Fainaru and Deckelbaum, 1979),
`no bands were observed at the appropriate molecular weight (28,000).
`The dimer/polymer content of Sigma A-7511 BSA (with added
`FA), as determined by column chromatography (Parks et al., 1983),
`was 25%. To determine whether NMR results were affected by the
`presence of disulfide-linked dimers and polymers, Sigma BSA was
`further fractionated by gel filtration chromatography. A 2-ml aliquot
`of hydrated BSA (100 mg/ml) was applied to a column of Sephadex
`G-150 (90 X 2.6 cm) equilibrated with 20 mM KC1, 0.1% NaN3 at
`4 "C. Fractions of 6 ml were collected at a flow rate of 7 ml/h. The
`protein concentration of each fraction was determined by its absorb-
`ance at 279 nm. The elution profile was essentially identical to one
`ly published (Morrisett et al., 1975). Fractions containing
`prcr .'I
`monomeric BSA were pooled
`and concentrated by ultrafiltration
`using Amicon (Danvers, MA) PM-10 filters. The final monomeric
`BSA sample contained 1.8 ml of 60 mg/ml BSA. No reformation of
`dimers/oligomers occurred in the concentrated monomeric fraction,
`as determined by sodium dodecyl sulfate-polyacrylamide gel electro-
`
`10972
`
`Note that in our previous study, a 90 X 2.5-cm column was used
`for preparative fractionation of monomeric BSA, rather than a 90 x
`1.5-cm column (incorrectly noted in Fig. 1 caption of Parks et al.,
`1983).
`
`MPI EXHIBIT 1008 PAGE 2
`
`MPI EXHIBIT 1008 PAGE 2
`
`

`

`I 3 C NMR of Fatty Acid-Albumin Interactions
`containing suspended crystalline
`X-ray Diffraction-For samples
`material (>7:1 FA. BSA), the material was pelleted by centrifugation
`at 10,000 rpm for 30 min at 30 "C. The pellet was transferred to 1-
`mm quartz capillary tubes (Charles Supper Co., Natick, MA), and
`the capillaries were placed in
`a sample holder kept
`at constant
`temperature (30 "C) by a circulating antifreeze/water bath. Nickel-
`(X = 1.5418 A) from a microfocus x-ray
`filtered CuKa x-radiation
`generator (Jarrell-Ash, Waltham, MA) was focused by a single nickel-
`coated mirror and further collimated by a Luzzati-Bar0 camera with
`slit optics. Low-angle powder x-ray diffraction patterns were recorded
`with a position-sensitive detector (Tennelec PSD-1100, Oak Ridge,
`TN) and a computer-based analysis system (Tracor Northern TN-
`1710, Middleton, WI).
`
`h
`
`10973
`
`711 - 511
`
`d
`
`,
`
`D'
`
`311 - 111
`
`111 - 011
`
`RESULTS
`NMR spectra at various C14.0-BSA mole ratios (at fixed
`pH, BSA concentration, ionic strength, and temperature) are
`shown in Fig. 1, A-E. The broad envelope centered at -176
`ppm represents carbonyl carbons of glutamine, asparagine,
`and the peptide backbone (Gurd and Keim, 1973) as well as
`aspartate carboxyl carbons (Shindo and Cohen, 1976) of BSA.
`The narrower resonances falling between 179 and 184 ppm
`primarily represent carboxyl carbons of 13C-carboxyl-enriched
`FA bound to BSA (Parks et al., 1983; Cistola et al., 1983;
`Hamilton et al., 1984), except for the resonance at
`180.9-
`181.1 ppm, which represents protein glutamate carboxyl car-
`bons (Shinodo and Cohen, 1976). Protein-free saturated FA
`(>lo carbons) exist as crystalline 1:1 acid/soap compounds a t
`pH 7.4 and 35 "C (Cistola et al., 1986), and crystalline phases
`do not give rise to high resolution NMR resonances. Further-
`more, none of the observed carboxyl resonances had chemical
`shifts coincident with those of soluble short-chain FA in water
`et al.,
`(without protein) under the same conditions (Cistola
`1982; Cistola, 1985). Hence, the observed carboxyl resonances
`do not represent unbound
`FA. In addition, none
`of these
`resonances represented BSA peaks that shifted into the car-
`boxyl region upon FA binding, since spectra
`of FA.BSA
`complexes using C14.0 with no 13C enrichment were identical
`to FA-free BSA samples (Fig. lA). In order to compare FA
`carboxyl peaks in different FA. BSA spectra, we have named
`analogous FA carboxyl peaks, based on their chemical shifts
`at pH 7.4 (and their ionization behavior; Cistola et al., 1987),
`as follows: peak a (183.7-184.1 ppm), peak b (182.5 ppm),
`peak b' (182.2-182.4 ppm), peak c (181.8-182.1 ppm), and
`peak d (180.4-180.7 ppm). In addition, we have named the
`glutamate carboxyl resonance as peak p r (180.9-181.1 ppm).
`b' was not observed for Clal .BSA
`It is notable that peak
`complexes (Parks et al., 1983) either because peak b' was not
`present or because it was not resolved from peaks b and c.
`For most of the FA.BSA spectra presented in this study,
`the intensitivies of individual FA carboxyl peaks could not be
`
`quantitatively measured as peak areas or peak heights because
`of closely overlapping FA or protein carboxyl resonances.
`Therefore, increases in FA carboxyl peak intensity with in-
`creasing mole ratio are represented qualitatively in the form
`of difference spectra. Digital subtraction of the upper spectra
`from the lower spectra yielded the difference spectra shown
`in the right column of Fig. 1 (F-Z). The difference spectra
`contained only those FA carboxyl peaks which increased or
`decreased in intensity or changed chemical shifts between the
`two different mole ratios. The intensities resulting from un-
`perturbed carboxyl or carbonyl resonances of BSA were sub-
`tracted out by this procedure.
`At 1:l mole ratio (Fig. l B ) , peaks b and b' were clearly
`visible, but peak d was difficult
`to distinguish because of
`closely overlapping protein glutamate resonance(s). However,
`subtraction of FA-free BSA (Fig. lA) from 1:1 C14.0. BSA (Fig.
`1B) revealed that peak d was present a t 1:1 mole ratio (Fig.
`
`188
`
`I82
`
`178
`
`I74
`
`chemical shlft
`(ppm)
`FIG. 1. Carboxyl/carbonyl region of ''C NMR spectra (A-
`E ) and difference spectra (F-I) for C14.,,.BSA complexes with
`different C14.0.BSA mole ratios at pH 7.4 and 34 "C. Difference
`spectra were obtained by digitally subtracting a spectrum at a given
`mole ratio from one at a higher mole ratio. This method removes
`BSA resonances from
`spectra and shows which FA carboxyl reso-
`nances increased between two corresponding FA.BSA mole ratios.
`The pairs of spectra which were
`subtracted are indicated by the
`dashed lines in the middle of the figure. The lower case letters above
`each peak indicate specific FA carboxyl resonances with characteristic
`chemical shifts (see "Results"). For all samples, the BSA concentra-
`tion was 7% (w/v). All spectra were recorded after 6,000 accumula-
`tions with a pulse interval of 2.0 s, 16,384 time domain points, and a
`spectral width of 10,000 Hz. Line broadening (3 Hz) was used in all
`spectrum; B, 1:l C14.n.BSA
`spectral processing. A, FA-free BSA
`spectrum; C, 3:l C14.0.BSA spectrum; D, 5:l CI4.,.BSA spectrum; E,
`7:l C1,n.BSA spectrum; F, difference spectrum, 1:l C14.n.BSA minus
`FA-free BSA; G, difference spectrum, 3:l C14.,.BSA minus 1:1 Clro.
`BSA; H, difference spectrum, 5:l CI4.,.BSA minus 3:l Cl4,,.BSA; I,
`difference spectrum, 7:l C14.0'BSA minus 5:l CI4.,.BSA.
`
`1F). Between 1:l and 3:l mole ratio (Fig. 1, B, C, and G),
`peaks b, b', and d increased, and peaks c and a appeared above
`2:l mole ratio (2:l spectrum not shown). Between 3:l and 5:l
`mole ratio (Fig. 1, C, D, and H), intensity increases occurred
`in all five FA carboxyl peaks. Between 5:l and 7:l mole ratio
`(and up to 13:l mole ratio), only peak c increased in intensity
`(Fig. 1, D, E, and I). Peak d decreased in intensity between
`5:l and 7:l mole ratio (Fig. 1Z).
`A plot of total carboxyl/carbonyl area ratio as a function of
`C14.0 .BSA mole ratio is shown in Fig. 2 (circles). The plot is
`linear up to 8-9:l mole ratio, above which point the sample
`downward
`became increasingly
`turbid (suspended crystals;
`pointing arrow in Fig. 2). The samples were centrifuged and
`yielded a transparent supernatant and a crystalline pellet.
`
`MPI EXHIBIT 1008 PAGE 3
`
`MPI EXHIBIT 1008 PAGE 3
`
`

`

`13C NMR of Fatty Acid-Albumin Interactions
`
`i
`
`i
`
`,""
`
`Mole Ratio FAIBSA
`FIG. 2. Plots of total "C NMR carboxyl intensity relative
`to total carbonyl intensity as a function of FA.BSA mole ratio
`for Cl,.o.BSA (o"-o) and Cla.o.BSA (A-A).
`The arrows
`indicate the mole ratio at and above which turbidity and crystalline
`precipitate were visualized in Clao.BSA (arrow pointing downward)
`and Clao.BSA (arrow pointing upward) samples. The dashed line
`indicates the expected (extrapolated) intensity ratio if essentially all
`FA were protein associated. FA not associated with protein under
`these conditions was crystalline and would not contribute to the total
`NMR carboxyl intensity. Note that the absolute values of the inten-
`sity ratios are arbitrary and have no direct stoichiometric meaning.
`
`Examination of the pellet by powder x-ray diffraction showed
`first order long spacings (41.0 f 0.9A) characteristic for
`crystalline 1:l potassium (or sodium) hydrogen dimyristate, a
`1:l acid/soap compound (Piper, 1929; Cistola et al., 1986).
`Therefore C14.0'BSA samples at high mole ratio contained
`unbound FA in the form of crystalline 1:l acid/soap. In
`general, long-chain fatty acids in water (>micromolar concen-
`trations) form 1:1 acid/soap crystals or fatty acid/soap lamel-
`lar liquid crystals between pH 7 and 10 (Small, 1986)?
`13C NMR spectra and difference spectra at various Cla.o.
`BSA mole ratios are shown in Fig. 3, A-I. At 1:l mole ratio,
`peaks b, b', and d were present (Fig. 3, B and F ) , although
`peak d was barely detectable. Peak d was much more clearly
`seen at 2:1 (spectrum not shown) and 31 (Fig. 3, C and G).
`Between 1:l and 3:l mole ratio, peaks b, b', and d increased
`in intensity (Fig. 3, B, C, and G ) . Between 3:l and 5:l mole
`ratio, peaks c and a became visible and all five FA carboxyl
`peaks increased in intensity (Fig. 3, C, D, and H), and between
`5:l and 7:l (spectrum not shown), peaks c, b/b', and a in-
`creased in intensity (Fig. 3, D, E, and I ) .
`A plot of total carboxyl/carbonyl area ratio as a function of
`Clao-BSA mole ratio (not shown) was linear up to about 7:l
`mole ratio. Deviation from linearity was accompanied by the
`visual appearance of sample turbidity (suspended crystals) at
`and above 7:l mole ratio. As with C14.0/BSA (see above), these
`crystals most likely represent unbound Ciao in the form of
`crystalline 1:l acid/soap.
`13C NMR spectra (but not difference spectra) for Clz,o. BSA
`complexes at four mole ratios are shown in Fig. 4. These
`spectra were very similar to corresponding spectra for c14.0'
`BSA complexes. At 1:l mole ratio, peaks b/b' and pr/d were
`visualized; at 3:1, three FA peaks were visualized (b, b', d).
`At 5:l and 7:l mole ratios, all five FA carboxyl peaks were
`distinguishable. Convolution difference spectra (not shown)
`permitted greater resolution of peaks b and b'.
`
`D. P. Cistola, J. A. Hamilton, D. Jackson, and D. M. Small (1987)
`Biochemistry, submitted for publication.
`
`",
`
`186
`
`I
`
`I
`170
`
`188
`
`I \
`
`h
`
`3/1 - 111
`
`170
`
`1
`I78
`I78
`174
`1 8 2
`182
`114
`chemical shift
`chemical shift (pprn)
`(pprn)
`FIG. 3. Carboxyl/carbonyl region of "C NMR spectra (A-
`E ) and difference spectra (F-I) for Cle.o.BSA complexes with
`increasing Cls.o.BSA mole ratio at pH 7.4, 34 OC. The lower
`case letters above each peak indicate specific FA carboxyl resonances
`with characteristic chemical shifts (see "Results"). Spectra were
`recorded after 6000 accumulations with a pulse interval of 2.7 s. All
`other spectral conditions and explanations are as described in the
`legend to Fig. 1.
`A plot of total carboxyl/carbonyl area ratio as a function of
`C12.0-BSA mole ratio (not shown) was essentially identical to
`that for C14.0+BSA (Fig. 2, circles). Deviation from linearity
`and the appearance of sample turbidity (crystal formation)
`occurred above 8-9:l mole ratio. Centrifugation and exami-
`nation of crystals by powder x-ray diffraction revealed first
`order long spacings (35.4 f 0.7 A) characteristic for crystalline
`1:l potassium (or sodium) hydrogen laurate, an acid/soap
`compound (Cistola et al., 1986).
`13C NMR spectra for C16.0. BSA complexes as a function of
`mole ratio (Cistola, 1985) are not shown here and were essen-
`tially identical to those for Cl8,, BSA complexes (Parks et al.,
`1983). A plot of total carboxyl/carbonyl area ratio as a func-
`tion of C16.0-BSA mole ratio is shown in Fig. 2 (triangles).
`Deviation from linearity and sample turbidity (crystals) ap-
`peared at and above 7:l mole ratio. The crystals most likely
`represented crystalline 1:l acid/soap, as was demonstrated by
`x-ray diffraction, for C12.,.BSA and C14,o'BSA samples.
`Although peaks b, b', and c closely overlapped in many FA.
`BSA spectra, peaks d and a were nearly completely resolved
`from the b/b'/c envelope, and their relative intensities (areas)
`could be quantitatively estimated (Table I). To determine the
`relative area of peak d, the area from the glutamate carboxyl
`resonance (peak p r ) had to be subtracted out (see legend to
`Table I). The results indicate that the relative intensities of
`peak d increased and peak a decreased, with increasing FA
`chain length. However, the relative intensity of the b/b'/c
`
`MPI EXHIBIT 1008 PAGE 4
`
`MPI EXHIBIT 1008 PAGE 4
`
`

`

`13C NMR of Fatty Acid-Albumin Interactions
`
`10975
`
`TABLE I
`Relative I 3 C NMR intensities of peaks a and d
`The areas of all resonances were measured by integration (see
`“Experimental Procedures”). The areas of peak d and the total FA
`carboxyl region were determined by subtracting out the contribution
`of BSA glutamate carboxyl resonances. This glutamate contribution
`(glutamate/carbonyl ratio) was determined from spectra of FA-free
`BSA accumulated under the same conditions as FA/BSA spectra. All
`three samples reported in this table contained 5:l mole ratio of FA/
`BSA. Ciao. BSA and C1s.O.BSA results were derived from spectra
`shown in Figs. 1D and 30, respectively, and C16.0.BSA results were
`derived from a spectrum not shown here (Cistola, 1985). The esti-
`mated uncertaintv is +lo%.
`
`dla ratio
`
`Ciao. BSA
`Cxo.BSA
`C1s.o.BSA
`
`0.09
`0.7
`2.3
`
`:I
`
`Total FA carboxyl intensity
`b + b ‘ + c
`a
`
`d
`
`2
`12
`19
`
`%
`
`20
`19
`8
`
`78
`69
`73
`
`4 IC
`
`4 / 1
`
`b
`
`‘
`I
`I
`186 1 8 2 1 7 8 1 7 4 1 7 0
`
`{
`
`‘
`
`I
`
`
`
`chemical shift (ppm)
`FIG. 5. Carboxyl/carbonyl region of “C NMR spectra for
`
`l-”C CB.O.BSA with increasing CS.~.BSA qple ratio at pH 7.4
`and 34 “C. The numbers at the right of each spectrum indicate the
`n\ole ratio of Cs.o’BSA. The narrow peaks labeled c are FA carboxyl
`resonances, and the very broad peaks centered at -176 ppm corre-
`spond to BSA carbonyl resonances. Spectra were recorded after 4000
`accumulations with a pulse interval of 2.0 s All other spectral
`conditions are as described in the legend to Fig. 1. The BSA concen-
`tration was 7.5% (w/v).
`
`182.1 ppm; this value suggested that this FA resonance was
`analogous to peak c in the other FA.BSA spectra. This
`correlation was also supported by NMR titration results
`which demonstrated that peak c was the only m e of the five
`observed FA carboxyl peaks to exhibit a titration shift be-
`tween pH 7.4 and 3.0 (Cistola et d., 1987). The chemical shift
`of peak c in C,,.BSA spectra increased with increasing mole
`ratio from 182.0 ppm (1:l and 3:l mole ratio) to 182.1 ppm
`(5:1), 182.2 ppm (7:1), 182.3 ppm (9:1), and 182.9 ppm (20:1).4
`The line widths of this resonance remained narrow ( 4 0 Hz)
`at all mole ratios studied.
`‘ Unbound Ca.0 under these conditions would have existed as mon-
`omers in solution, rather than crystalline or liquid-crystalline aggre-
`gates or micelles. For C8., BSA samples at >3:1 mole ratio, a progres-
`sive increase in the chemical shift of peak c (see “Results”) toward
`the chemical shift of monomeric aqueous C S . ~ (184.2 ppm) provided
`evidence that monomeric unbound Cs.0 was present at concentrations
`>>1 ~ L M and in rapid exchange (>>lo0 exchanges/s) with protein-
`bound C8.0.
`
`I
`
`I
`
`I
`
`I
`
`I
`
`I
`
`I
`
`I
`
`
`
`I
`
`
`
`1 8 2 1 7 8 1 7 4 1 7 0
`chemical shift (ppm)
`FIG. 4. Carboxyl/carbonyl region of “C NMR spectra for
`l-”C C12.0.BSA with increasing C12.,.BSA mole ratio at pH
`7.4 and 34 “C. The numbers at the right of each spectrum indicate
`the mole ratio of Clz.o.BSA. Spectra were recorded after 4000 accu-
`mulations with a pulse interval of 2.8 s. All other spectral conditons
`are as described in the legend to Fig. 1.
`
`FA
`envelope exhibited little or no change with increasing
`chain length. (Results
`for C,,,.BSA were not included be-
`cause peak d was not well resolved from the b-b’-c complex;
`Fig. 4.)
`13C NMR spectra for Ca.o.BSA complexes (Fig. 5 ) , unlike
`those for all other FA.BSA complexes studied, showed only
`one FA carboxyl peak at all mole ratios (up to 20:l). At 5:l
`mole ratio (pH 7.4), the chemical shift of this resonance was
`
`MPI EXHIBIT 1008 PAGE 5
`
`MPI EXHIBIT 1008 PAGE 5
`
`

`

`TABLE I1
`FA carboxyl Tl and NOE values for various FAIBSA complexes
`All measurements were made at pH 7.4 and 35 'C. The Tl value
`for the BSA c-Lys/p-Leu resonance (39.57 ppm) was 0.3 s for all
`samples shown in this table.
`BSA
`Sample,
`mole ratio
`concentration
`%, w/u
`7.5
`1.5
`1.5
`3.8
`1.5
`11.4
`7.5
`
`S
`
`C8.o. BSA, 7:l
`C12.0.BSAI 5:l
`C,d.o.BSA, 5:l
`Cl'.o.BSA, 5:l
`Cl'.o.BSA, 5:l
`C16.o.BSA, 5:l
`Cls.o.BSA, 5:l
`
`C
`b/b'/c
`b'lc
`b'lc
`b'lc
`b'lc
`b'lc
`
`2.2 f 0.1 1.6 f 0.2
`3.1 f 0.1
`2.5 f 0.1 1.7 f 0.2
`2.7 f 0.1 1.7 f 0.2
`2.8 f 0.1 1.7 f 0.2
`2.2 f 0.1 1.9 f 0.2
`2.9 k 0.1
`
`13C NMR of Fatty Acid-Albumin Interactions
`ratio, significant amounts o

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