throbber
0~
`
`ri
`~
`.,
`}
`:;;
`'"" &.
`~~ 2..c
`~ ~
`~ ~
`<'I·:::
`co ·5o
`<'I.!!
`~ g
`: .. 6
`,8 ~
`] l
`E :;
`
`~ ~
`<'l ~
`~ {
`~]
`-~ ~
`~]
`2 i
`; q
`8 ~
`1
`"
`e-.E
`8 en
`
`Biochemistry
`
`0 Copyright 1996 by the American Chemical Society
`
`Volume 35, Number 1
`
`January 9, 1996
`
`Accelerated Publications
`
`Direct Measurement of the Aspartic Acid 26 pKa for Reduced Escherichia coli
`Thioredoxin by 13C NMR t
`Mei-Fen Jengt and H. Jane Dyson*
`The Scripps Research Ins titute, 10666 North Torrey Pines Road, La Jolla, California 92037
`
`Received October 6, 1995; Revised Manuscript Received November 8, 1995®
`
`ABSTRACT: Because of interference from the pH-dependent behavior of nearby groups in the active site
`of Escherichia coli thioredoxin, the pKa of the buried carboxyl group of the aspartic acid at position 26
`has been difficult to quantitate. We repo1t a direct 1neasure1nent of this pK3 using an NMR 1nethod
`utilizing the conelation between the cPH proton resonances and the 13CO of the titrating carboxyl group.
`The experiments show unequivocally that the pKa is 7.3-7.5, rather than the value of9 or greater recently
`proposed by Wilson, N. A., et al. [(1995) Biochemistry 34, 8931-8939]. The assignment of the titrating
`resonances to Asp 26 is unambiguous: the values of the cPH chemical shifts conespond exactly to those
`of Asp 26, and their titration in the pH range 5. 7-10.0 is the same as that obse1ved previously for the
`proton resonances alone. In addition, the chemical shift of the carboxyl 13C resonance at pH 5.7 is upfield
`of those of the other carboxyl and carboxamide resonances, diagnostic for a protonated carboxyl group.
`The resonances assigned to Asp 26 are the only ones that titrate in the pH range 5.7-10.5. None of the
`other aspai1ate and glutamate residues in the molecule ai·e titrated in this pH range, consistent with their
`positions on the smface of the molecule. The pK. measmed for Asp 26 in reduced thioredoxin is identical
`within experimental e1rnr to that measm·ed in the oxidized fonn of the protein. This is significant for the
`reductive mechanism of thioredoxin: the bmied salt b1idged/hydrogen-bonded side chains of Asp 26 and
`Lys 57 are likely to contribute to the facility of the reaction by providing a convenient source ai1d sink
`for protons in the hydrophobic enviromnent of the complex between thioredoxin and its substrates.
`
`A number of studies have recently adch-essed the issue of
`the pK. of tl1e buried aspartic acid in Escherichia coli
`thioredoxin. This residue is highly conse1v ed among all
`th.ioredoxins (Eklund et al., 1991) and appears to be
`completely buried in both proka1yotic (Katti et al., 1990;
`Jeng et al., 1994) and eukaiyotic (Qin et al., 1994) tlJ.io(cid:173)
`redoxins. The E-amino group of a lysine residue at position
`57 is also buried in close proximity to Asp 26 and takes
`pai1 in a loose (Jeng et al. , 1994) or water-mediated (Katti
`
`t This work was supported by Grant GM43238 from the National
`Institutes of Health.
`• To whom correspondence should be addressed.
`! Present address: Graduate Institute of Cell and Molecular Biology,
`Taipei Medical College, Taipei, Taiwan.
`® AbstractpublishedinAdvanceACS AbstracL<, December 15, 1995.
`
`et al., 1991) hydrogen-bonding interaction. TlJ.is residue is
`highly conse1ved only among prokaryotes, but its function
`is apparently duplicated in eukaryotic th.ioredoxins by the
`presence of a lysine residue at position 39 (Eklund et al.,
`1991). The pK. of the buried aspai1ate residue has been
`detennined by a number of methods for the oxidized form
`of thioredoxin (Trx-S2) (Dyson et al., 1991; Langsetmo et
`al., 1991a) to be considerably shifted from "normal"
`values: to 7.5 instead of ~4.0. The pK. detemJ.ination for
`the reduced f01m of the protein [Trx-(SH)2] is complicated
`by tl1e effect of the titration of the nearby tlJ.iol groups of
`the active site cysteine residues, Cys 32 and Cys 35, but
`was infetTed from NMR studies to be also in the vicinity of
`7 (Dyson et al., 1991 ). TI1e relationship of these three groups
`is shown in Figure 1.
`
`0006-2960/96/0435-1$12.00/0
`
`© 1996 American Chemical Society
`
`Bausch Health Ireland Exhibit 2043, Page 1 of 6
`Mylan v. Bausch Health Ireland - IPR2022-00722
`
`

`

`2 Biochemistry, Vol. 35, No. 1, 1996
`
`Accelerated Publications
`
`FIGURE 1: Portion of one of the solution structures of Trx-(SH)2 (Jeng et al., 1994) showing the spatial relationship between the two
`cysteines and the buried Asp 26 and Lys 57 side chains.
`
`The unusual pH dependence of the two active site thiols
`in reduced thioredoxin is crucial for the reductive mechanism.
`Reduction of a substrate disulfide by thioredoxin involves a
`two-electron, two-proton reduction. pH control at the active
`site is therefore important to the control of the rate of
`reduction by thioredoxin. We have recently site-specifically
`determined the pKas of the two active site thiols in Trx-(SH)2
`using direct NMR methods to be 7.5 and 9.5 (Jeng et al.,
`1995). The 13C resonance of the atom attached to the titrating
`group should primarily show the influence only of the
`directly bonded titrating group. Surprisingly, the C(cid:226) reso-
`nances of the two cysteine residues show a double titration.
`We have interpreted this as evidence of a shared proton
`between the sulfur atoms of the two cysteines at neutral pH,
`consistent with the otherwise unexplained shifted pKa of the
`solvent-exposed Cys 32 thiol.
`At the same time, another group was working on the same
`system (Wilson et al., 1995). Their 13C-1H data for the
`cysteine thiols of Trx-(SH)2 appear identical to our own, but
`a very different interpretation was placed on them. The
`higher of the two pKas observed in the 13C(cid:226)-1H(cid:226) titration,
`at 9.5, was attributed to the buried Asp 26 carboxyl group
`on the basis of a comparison of the behavior of a mutant in
`which Asp 26 had been replaced by alanine (D26A). In this
`paper we present a direct measurement of the Asp 26 pKa,
`using a modified two-dimensional H(CA)CO NMR experi-
`ment to estimate directly the 13C chemical shift of the
`carboxyl carbon of Asp 26. These measurements establish
`unequivocally that the pKa of Asp 26 is between 7.3 and
`7.5, rather than 9, as inferred by Wilson et al. (1995). This
`has profound implications for the mechanism of pH control
`in the reductive and oxidative reactions of thioredoxin.
`
`MATERIALS AND METHODS
`
`Reduced thioredoxin uniformly labeled with 15N and 13C
`was prepared as previously described (Chandrasekhar et al.,
`1991, 1994), utilizing an algal homogenate (Martek Co.) as
`a basis for an enriched medium. Purification of thioredoxin
`and preparation of D2O samples of the reduced protein using
`dithiothreitol were performed as previously described (Dyson
`et al., 1989). The pH of the sample was varied between 5.7
`
`and 10.6 by the addition of small aliquots of 0.1 M NaOD
`or DCl in D2O. pH values quoted are meter readings
`uncorrected for the deuterium isotope effect.
`NMR experiments were carried out at 308 K on a Bruker
`spectrometer operating at 500 MHz for protons. The
`behavior of the 13CO of the carboxyl and carboxamide groups
`in the protein as a function of pH was monitored using a
`two-dimensional H(CA)CO experiment (Kay et al., 1990),
`modified to optimize the detection of the coupling between
`the 13CO of a carboxyl or carboxamide and the adjacent
`13C(cid:226)H or 13CγH (Yamazaki et al., 1993; Oda et al., 1994).
`Spectra were referenced in ω2 to the pH-independent back-
`bone 13CO-1H cross peaks of Gly 84 at 4.38 ppm and Ala
`108 at 4.09 ppm (Dyson et al., 1989) and indirectly in ω1
`(Wishart et al., 1995). Spectral widths were 6250 Hz with
`2048 complex points in ω2 and 6250 Hz with 128 complex
`points in ω1. Quadrature detection was achieved in ω1 by a
`combined States-TPPI method. Spectra were Fourier
`transformed using Gaussian and exponential window func-
`tions on a Sun workstation using the FTNMR software of
`Dennis Hare.
`Chemical shift values as a function of pH for Asp 26 were
`analyzed using the program Templegraph (Mihalisin As-
`sociates) in terms of a single titration curve of the form
`(Dyson et al., 1991):
`
`δ ) δHA
`
`- ((δHA
`
`- δA)/[1 + 10n(pKa-pH)])
`
`where δ is the observed chemical shift at a given pH, δHA
`and δA are the chemical shifts for the various protonated
`forms of the protein, n is the number of protons transferred,
`and Ka is the acid ionization constant.
`
`RESULTS AND DISCUSSION
`
`A modified H(CA)CO spectrum of Trx-(SH)2 at pH 8.5
`is shown in Figure 2. All of the cross peaks can be identified
`either with the backbone 13CO-HR correlations of glycine
`and other residues whose 13CR frequency is sufficiently low
`to be excited in the experiment or with the side-chain
`13CO-H(cid:226) correlations of
`the aspartate and asparagine
`
`Bausch Health Ireland Exhibit 2043, Page 2 of 6
`Mylan v. Bausch Health Ireland - IPR2022-00722
`
`

`

`Accelerated Publications
`
`Biochemistry, Vol. 35, No. 1, 1996 3
`
`G71 ., G97
`
`N83
`
`@)
`
`6 q
`
`0
`
`D26
`
`~
`
`~
`
`~ (cid:173)
`
`© L
`N63 :c::;
`-
`
`V::::::e
`N106
`
`D15
`~
`
`172
`
`174
`
`176
`
`17813cc
`
`180 ppm
`
`182
`
`184
`
`186
`
`4.5
`
`4.0
`
`2.5
`
`2.0
`
`3.0
`3.5
`1H ppm
`FIGURE 2: A 500 MHz modified two-dimensional H(CA)CO
`spectrum of reduced thioredoxin at pH 8.52, showing the cross
`peaks for the glycine CRH-CO (backbone) and the side-chain
`C(cid:226)H-CO and CγH-CO connectivities for the Asp, Glu, Asn, and
`Gln residues. Cross peaks are also visible for other residues for
`which (like glycine) the CR resonance is at a sufficiently low
`frequency to be excited during the pulse sequence, which has
`otherwise been optimized for the C(cid:226) and Cγ of the side chains.
`
`Table 1:
`
`13CO Assignments for Trx-(SH)2 at pH 5.92, 308 Ka
`CRH
`13CO
`Gly
`Gly 21
`3.99, 3.99
`174.7
`Gly 33
`3.98, 4.31
`176 1
`Gly 51
`4.40, 3.74
`176 2
`Gly 65
`3.98, 3.67
`178.7
`Gly 71
`3.82, 3.67
`176 3
`Gly 74
`4.20, 3.74
`172.4
`Gly 84
`4.38, 3.75
`175 1
`Gly 92
`3.59, 4.38
`173.6
`Gly 97
`3.76, 3.92
`179 1
`C(cid:226)H
`2.27, 2.18
`2.89, 2.23
`2.65, 2.21
`2.72, 2.78
`2.85, 3.11
`2.23, 1.88
`CγH
`13CO
`Glu/Gln
`185.2
`2.62, 2.68
`Glu 44
`184.6
`2.44, 2.14
`Glu 48
`186.6
`2 53, 2.50
`Gln 50
`186.6
`2 28, 2.47
`Gln 62
`186.6
`2.45, 2.45
`Gln 98
`a Assignments are presented only for those resonances that could
`be unambiguously identified from the 2D H(CA)CO spectrum.
`
`Asp/Asn
`Asp 15
`Asp 26
`Asn 59
`Asn 63
`Asn 83
`Asn 106
`
`13CO
`180.3
`173.7
`177.5
`179.1
`180.3
`178.1
`
`residues and the 13CO-Hγ correlations of the glutamic acid
`and glutamine residues.
`The 13CO chemical shifts for resonances that could be
`unambiguously assigned from the H(CA)CO spectrum are
`shown in Table 1. The cross peaks for the majority of the
`11 aspartate residues are heavily overlapped in the region
`(1H ) 2.6-2.9 ppm and 13C ) 180-182 ppm). The cross
`peaks due to Asp 26 are readily identifiable, since the C(cid:226)H
`resonances are distinctive (Dyson et al., 1989), and the
`13CO resonance is upfield-shifted, as expected for a proto-
`nated carboxyl carbon (Oda et al., 1994). However, the cross
`peaks are of lower intensity than those for the other carboxyl
`groups of the molecule, presumably reflecting the unique
`position of the Asp 26 side chain, deeply buried in the
`hydrophobic cavity behind the active site (Jeng et al., 1994).
`
`pH
`
`5.92
`
`6.49
`
`7.00
`
`7.29
`
`7.61
`
`0
`i-\-_-_-_-_-_-.:._r_-_-_-_-_-.:..•_-_-_-_-_-.:._•_-_-_-_-_-.:._.,.._-_~f 174.1
`~ 173.5
`
`~ 173.6
`
`~=====i:::::::::i:::::::::i:::::::::i::::::::'.~ 174.0
`
`~ 174.2
`
`a
`
`1-;-:::::!::::::::i::::::r::::::!:::!_t 174.7
`e F174.5
`~
`~=====i::::=====i::::=====i::::====:;;:r:=~~ 175.0
`t174.9
`lft,,,_
`D
`U F
`~=====i::::=====i::::=====i::::=====i::::==,.L 175.4 ppm
`
`13co
`
`8.521'
`
`8.08 It, ====::!o=====!:====::!====~:::o!:=~[:::
`0
`0 ~175.3
`·t=====:::=====:::=====:::=====:::=~ .... f 175.8
`0
`0 ~ 175.5
`9.021'
`9.71 I o
`,
`
`~1H~
`
`~ 175.7
`
`r· - - - - - . - - - - . - - -~ __ &_o~__Jf 116.2
`3.0
`2.8
`2.6
`2.4
`2.2
`1H~ ppm
`FIGURE 3: Portions of the 500 MHz modified two-dimensional
`H(CA)CO spectrum of reduced thioredoxin at a number of pH
`values between 5.92 and 9.71, showing the cross peaks assigned
`to the Asp 26 C(cid:226)H-13CO connectivities. These spectra are in
`general plotted at a lower contour level than that in Figure 1.
`
`All of the other carboxyl side chains and many of the
`carboxamides (Asn and Gln) are present on the surface of
`the molecule, where they are in many cases in free rotation
`even at the CR-C(cid:226) bond (Chandrasekhar et al., 1994; Jeng
`et al., 1994). This would give rise to relatively narrow line
`widths for the carboxyl carbons. For Asp 26, however, the
`side chain is fixed in the interior of the molecule, so the
`correlation time is strictly controlled by the overall tumbling
`of the molecule. It is significant that the Asp 26 cross peaks
`have a tendency to disappear in older samples and at the
`extremes of the pH range, where a small amount of
`aggregation (at pH < 6) or unfolding (at pH > 9.5) is more
`likely to occur. The concomitant
`increase in average
`to broaden the 13CO
`correlation time is then sufficient
`resonance beyond detection. A similar lowering of the cross-
`peak intensity is observed for those carboxamide residues
`that the solution structure of Trx-(SH)2 indicates are buried,
`for example, Asn 59 and Gln 98.
`The cross peaks corresponding to the Asp 26 C(cid:226)H-13CO
`are shown in Figure 3 for all pHs except 5.7 and 10.0, plotted
`at a lower contour level than for Figure 2. The pH-dependent
`behavior of the 13CO and 1H(cid:226) resonances is shown in Figure
`4 and the pKas are shown in Table 2. The magnitude of the
`shift in the 13C chemical shift, 2.2 ppm, is comparable to
`those observed for the titrations of aspartates in ribonuclease
`H1 (Oda et al., 1994).
`It is immediately obvious that the
`pKa for the transition in chemical shift between pH 5.7 and
`pH 10 is quite similar for the 13CO chemical shifts and for
`those of the H(cid:226)2 resonances [stereospecific assignments
`
`Bausch Health Ireland Exhibit 2043, Page 3 of 6
`Mylan v. Bausch Health Ireland - IPR2022-00722
`
`

`

`4 Biochemishy, Vol. 35, No. 1, 1996
`
`Accelerated Publications
`
`i3co
`
`•
`
`•
`
`177
`
`176
`
`175
`
`174
`
`173
`2.90 0
`•
`
`2.80
`
`E
`a.
`-9;
`;:::
`i:
`(J)
`
`0
`
`0
`
`of a titration at pK. 2'.: 9 as suggested by Wilson et al. (1995).
`A mnnber of attempts have been made to localize tile pK.s
`of the titrating groups in the active site of reduced thio(cid:173)
`redoxin. Early cheniical modification studies indicated that
`the pK. of one of the cysteine thiols in reduced thioredoxin
`was ~6.8 and the other ~9 (K.allis & Holmgren, 1980). A
`later NMR study (Dyson et al., 1991) interpreted a large
`volume of 1H titration data in tenns of only two pK.s in the
`active site region, although it was obvious that three titrating
`groups (Asp 26, Cys 32, and Cys 35) were present. This
`approach was justified by the inference that two of the pK.s
`(Asp 26 and Cys 32) appeared to have very similar pK.s, in
`the vicinity of 7. A higher pK. of 8.4 was foru1d for Cy s
`35, closer to nonnal values for cysteine pK.s in proteins.
`However, tl1e pH range over which these measurements were
`taken was smaller, at least at the high-pH range, than later
`studies: it appears that the pK. obtained for Cys 35 was
`ru1derdetennined. Later work using specifically (Wilson et
`al., 1995) and seniispecifically labeled thioredoxin (Jeng et
`al., 1995) indicated that the higher pK. in fact more closely
`appruadu::s 9.5, as originally fuuml u y Kallis aml Hulmgn::u
`(1981) . A Raman study (Li et al., 1993) gave pK.s of 7.1
`and 7.9 for the two cysteine thiols, wiili the behavior of the
`complex Raman SH stretching band indicating that the two
`cysteine pK.s were below pH 8.2. The same work indicated
`that all of the carboxyl groups in the molecule were
`completely deprotonated below pH 8.
`TI1e question of inte1pretation of these data hinges on the
`identification of the titrations of tlrree groups, Asp 26, Cys
`32, and Cys 35. The 13C NMR indicates two pK.s for the
`cysteines of 7.5 and 9.5 (Jeng et al., 1995), while the pK. of
`Asp 26 is shown by the present work to be 7.5. The sin1plest
`explanation is that of Jeng et al. (1995): the two cysteine
`pK.s are 7.5 and 9.5 and the Asp 26 pK. is also 7.5. By
`contrast, tl1e interpretation of Wilson et al. (1995) is more
`complex. The higher of the two pK.s seen in ilie 13C
`experiments on tile cysteines is ascribed to ilie titration of
`Asp 26. Even without the direct evidence presented in tliis
`paper on the Asp 26 pK., several facts argue against this
`inte1pretation. First, while proton cheniical sliifts can be
`influenced by a number of titrating groups (Dyson et al.,
`1991 ), tl1e influence of titrating groups otl1er than the
`i1mnediately adjacent one on the 13C chemical shift of the
`adjacent carbon atom should be small: the major influence
`on the nCJl carbon cheniical shifts of the cysteines should
`therefore be the titrations of the cysteine thiols tl1emselves.
`Second, a significantly greater change in 13Cfl cheniical shift
`is seen for the pK. 9.5 transition for Cys 32 than for Cys 35,
`exactly the opposite of what would be expected from the
`relative proxiniity of the two cysteines to ilie Asp 26 side
`chain. Tiiird, ilie stmctrues ofTrx-(SH)2 and Trx-S2 are the
`same (Jeng et al., 1994): only the most subtle changes in
`dynaniics (Stone et al., 1993) and hydrogen exchange (Jeng
`& Dyson, 1995) reveal any differences between the two
`f01ms at all. It is therefore at least plausible that the pH(cid:173)
`dependent behavior of Asp 26 should be the same in ilie
`two fonns of the protein, even though the charge balance in
`the active site is more negative at high pH in Trx-(SH)2 than
`in Trx-S2. In addition, the same differences in dynaniics
`and hydrogen exchange indicate that the backbone mobility
`in the region of tl1e active site is actually sigiiificantly greater
`in Trx-(SH)2: we would therefore expect the Asp 26 pK., if
`different from that ofTrx-82, to be removed toward the lower
`
`~ 2 .70
`.E
`.c u
`
`Q)
`
`2.60
`2.30
`
`2.20
`
`1H~3
`
`•
`
`0
`
`0
`
`0
`
`0
`
`0 00
`
`0 •
`1H~2
`2.10-- ~ - - - - -~ - -- -~
`5.5
`6.5
`7.5
`10.5
`8.5
`9.5
`
`pH
`FIGURE 4: Plot of the chemical shift of the resonances shown in
`Figtll'e 2 as a ftmction of pH. Solid lines are ctu-ves fitted to the
`data (e ) using the method of least squru·es to these points. Also
`included are data points obtained previously from 1H NMR spectra
`(0) (Dyson et al., 1991). The points at the high-pH extremum
`appeat· to be influenced by another pH transition and were not
`included in the fit.
`
`Table 2: Measured pK, Values for the Asp 26 Carboxyl Group
`
`atom
`Asp 26 13C
`Asp26 1IJ/32
`
`M(ppm)
`pK,
`source
`this work
`2.2
`7.3
`this work
`(78Y,
`(003Y,
`7.3
`Dyson et al., 1991
`0.04
`this work
`0.21
`7.5
`Dyson et al., 1991
`0.22
`7.3
`• The pK, value obtained for Asp 26 1Hf32 is not well determined
`due to the small size of the chemical shift change with pH (,M).
`
`Asp 26 1IJ/33
`
`according to Chandrasekhar et al. ( 1994)]. The H/33
`resonance undergoes ve1y little change over this pH range.
`At pH values greater than 9 there appears to be another pH(cid:173)
`dependent process occruring. TI1e magnitude and abruptness
`of the change indicates that it is probably not related to the
`cysteine pK. at 9.5. There are two possible explanations
`for this- changes due to general tmfolding of the molecule
`or deprotonation of another nearby group such as the Lys
`57 E-amino group. Also included in Figure 4 are the data
`points from the previous 1H NMR study of reduced tliio(cid:173)
`redoxin (Dyson et al., 1991). TI1e chemical shift values at
`any given pH differ slightly between the two data sets,
`possibly due to ilie 10° temperatrue difference between the
`two sets of measurements. However, ilie pH-dependent
`behavior of the resonances in the 1H COSY spectra closely
`parallels that observed for the 13C- 1H HSQC measurements,
`further evidence that the low-intensity cross peaks observed
`in the latter expe1iment are indeed those of Asp 26. Most
`significantly, Figure 4 shows a pronoru1ced single titration
`in the 13CO for Asp 26, with a pK. of 7.3. There is no sign
`
`Bausch Health Ireland Exhibit 2043, Page 4 of 6
`Mylan v. Bausch Health Ireland - IPR2022-00722
`
`

`

`Accelerated Publications
`\
`
`•
`
`,os
`
`t ,.
`t ~ :· 110
`· . ~
`, •• 1 ... e,o ,. ..
`• •~ I
`I ~ f
`
`.. , ... ,,, . ~ 15N
`•• __. ...
`'
`
`OU~
`
`~
`
`115
`
`" '
`
`~
`
`,
`
`f
`
`c,I
`,

`
`120
`
`ppm
`
`t
`
`125
`
`130
`
`'• •
`Wl
`.,
`
`•
`
`~
`\.1'
`
`' I
`-· , u•
`' .,.
`
`I ..,
`
`10.0
`
`..
`
`•
`90
`8 0
`1H ppm
`FIGURE 5: Superposition of the 500 MHz 15N-1H HSQC spectra
`of wild-type Trx-(SH)2 (black) at pH 6.49, 298 K, and D26A mutant
`Trx-(SH)2 (red) at pH 6.51, 298 K, showing the extent of the
`changes in the chemical shifts in the vicinity of the active site as
`a result of the mutation.
`
`7.0
`
`values more characteristic of solvent-exposed carboxyl
`groups, rather than higher by 2 units in Trx-(SH)2, as
`suggested by Wilson et al. (1993).
`The interpretation of Wilson et al. (1995) relies on two
`lines of evidence: changes in cysteine pKas in mutant
`proteins and Raman spectra that indicate that the cysteine
`pKas are 7.1 and 7.9 (Li et al., 1993). However, they ignore
`other Raman data in the same paper that indicate that the
`titrations of all of the carboxyl groups (presumably including
`that of Asp 26) are complete below pH 8.0, and the fact that
`neither of the NMR titration studies (Wilson et al., 1995;
`Jeng et al., 1995) shows any indication of a pKa of 7.9 for
`the cysteine 13C(cid:226) titrations. The major evidence cited by
`Wilson et al. (1995) in favor of a high pKa (>9) for the
`buried Asp 26 carboxyl group is a comparison of the results
`for wild-type reduced thioredoxin with those obtained for a
`mutant in which the aspartate side chain is replaced by an
`alanine (D26A). This mutant has been the subject of a large
`amount of experimental research (Langsetmo et al., 1990,
`1991a,b), and we have also undertaken extensive NMR and
`biochemical studies of this and other mutants in the active
`site region (Dyson et al., 1994; H. J. Dyson and A.
`Holmgren, manuscripts in preparation). Both the study of
`Wilson et al. (1995) and our own measurements indicate
`that the pKas of the two cysteine thiols in the mutant Trx-
`(SH)2 are between 7.5 and 8.0. We have made complete
`assignments of the 1H, 13C, and 15N NMR spectra of the
`D26A mutant using a combination of two- and three-
`dimensional homo- and heteronuclear spectra. These studies
`reveal that while the majority of the protein is relatively
`unaffected by the mutation, the structure in the active site
`region is greatly perturbed, as shown by large chemical shift
`differences between mutant and wild-type Trx-(SH)2 at sites
`far removed in the sequence from the immediate area of the
`mutation. This is illustrated in Figure 5, which shows a
`comparison of the wild-type and D26A mutant 15N HSQC
`
`Biochemistry, Vol. 35, No. 1, 1996 5
`
`spectra. A significant rearrangement of the hydrophobic
`pocket where the Asp 26 side chain resides in the wild-type
`protein is consistent with the observed increase in thermo-
`dynamic stability of the mutant protein (Langsetmo et al.,
`1991b), although the stability of the protein to pHs lower
`than 6.0 is decreased (H. J. Dyson, unpublished observa-
`tions). In view of the structural differences apparent between
`the mutant and wild-type proteins,
`it
`is not valid to
`extrapolate from the behavior of the mutant
`to make
`conclusions about the behavior of the wild-type protein: the
`pKas of the two cysteine thiols are most probably changed
`by a rearrangement of the active site structure in the mutant.
`In addition, the lowering of the pKa of Cys 35 from 9.5 in
`wild-type to 7.5-8 in the mutant would appear to be a logical
`consequence of the removal of the negatively charged Asp
`26 from its local environment.
`A change in the pKa of Asp 26 to 8.3 is observed in the
`double mutant C32S/C35S (Dyson et al., 1994). A pKa
`increase upon removal of the two potential negative charges
`on the cysteines is apparently paradoxical and was another
`of the justifications of Wilson et al. (1995) for their
`to the >9.0 titration. By this
`assignment of the pKa
`argument, the high pKa of Asp 26 in the wild-type protein
`is reduced in the mutant by the removal of the negative
`charges. Once again, extrapolation of mutant data to make
`conclusions about the wild-type protein is simplistic: other
`changes occur than a simple charge removalsthe group that
`has replaced the charged group must be considered, as well
`as the structural rearrangements caused by the mutation in
`the hydrophobic pocket where the Asp 26 carboxyl
`is
`situated.
`We have shown unequivocally that the Asp 26 pKa in wild-
`type reduced thioredoxin closely resembles that observed for
`the oxidized form of the protein. This is consistent with
`the high degree of similarity between the two structures (Jeng
`et al., 1994) and with a mechanism that includes a shared
`proton between two thiols as a means both of stabilization
`of the reactive thiolate form of the Cys 32 side chain and of
`promoting complex formation between reduced thioredoxin
`and protein substrates (Jeng et al., 1995). Both the oxidative
`and reductive reactions of thioredoxin are severely impaired
`in the D26A mutant, mainly in the slowing of reaction rates
`(A. Holmgren, H. J. Dyson, and I. Slaby, manuscript in
`preparation). The buried aspartate, with its pKa poised at
`the lower of the two cysteine thiol pKas, is clearly involved
`in control of the efficient proton transfers at the active site
`during the reactions of thioredoxin, probably by serving as
`a proton source and proton sink removed from the exterior
`solvent after the formation of the hydrophobic complex
`between thioredoxin and its substrate.
`
`ACKNOWLEDGMENT
`
`We thank Dr. Peter Wright for continuing helpful discus-
`sions and for a critical reading of the manuscript, Professor
`Arne Holmgren for providing the 13C-labeled protein, and
`Stefan Prytulla, Jian Yao, and John Chung for helpful
`discussions on the NMR experiments.
`
`REFERENCES
`
`Chandrasekhar, K., Krause, G., Holmgren, A., & Dyson, H. J.
`(1991) FEBS Lett. 284, 178-183.
`Chandrasekhar, K., Campbell, A. P., Jeng, M.-F., Holmgren, A.,
`& Dyson, H. J. (1994) J. Biomol. NMR 4, 411-432.
`
`Bausch Health Ireland Exhibit 2043, Page 5 of 6
`Mylan v. Bausch Health Ireland - IPR2022-00722
`
`

`

`6 Biochemistry, Vol. 35, No. 1, 1996
`
`Accelerated Publications
`
`Dyson, H. J., Holmgren, A., & Wright, P. E. (1989) Biochemistry
`28, 7074-7087.
`Dyson, H. J., Tennant, L. L., & Holmgren, A. (1991) Biochemistry
`30, 4262-4268.
`Dyson, H. J., Jeng, M.-F., Model, P., & Holmgren, A. (1994) FEBS
`Lett. 339, 11-17.
`Eklund, H., Gleason, F. K., & Holmgren, A. (1991) Proteins 11,
`13-28.
`Jeng, M.-F., & Dyson, H. J. (1995) Biochemistry 34, 611-619.
`Jeng, M.-F., Campbell, A. P., Begley, T., Holmgren, A., Case, D.
`A., Wright, P. E., & Dyson, H. J. (1994) Structure 2, 853-868.
`Jeng, M.-F., Holmgren, A., & Dyson, H. J. (1995) Biochemistry
`34, 10101-10105.
`Kallis, G. B., & Holmgren, A. (1980) J. Biol. Chem. 255, 10261-
`10265.
`Katti, S. K., LeMaster, D. M., & Eklund, H. (1990) J. Mol. Biol.
`212, 167-184.
`Kay, L. E., Ikura, M., Tschudin, R., & Bax, A. (1990) J. Magn.
`Reson. 88, 496-514.
`Langsetmo, K., Sung, Y.-C., Fuchs, J. A., & Woodward, C. (1990)
`in Current Research in Protein Chemistry: Techniques, Struc-
`ture, and Function (Villafranca, J. J., Ed.) pp 449-456,
`Academic Press, San Diego.
`
`Langsetmo, K., Fuchs, J. A., & Woodward, C. (1991a) Biochemistry
`30, 7603-7609.
`Langsetmo, K., Fuchs, J. A., Woodward, C., & Sharp, K. (1991b)
`Biochemistry 30, 7609-7614.
`Li, H., Hanson, C., Fuchs, J. A., Woodward, C., & Thomas, G. J.,
`Jr. (1993) Biochemistry 32, 5800-5808.
`Oda, Y., Toshio, Y., Nagayama, K., Kanaya, S., Kuroda, Y., &
`Nakamura, H. (1994) Biochemistry 33, 5275-5284.
`Qin, J., Clore, G. M., & Gronenborn, A. M. (1994) Structure 2,
`503-522.
`Stone, M. J., Chandrasekhar, K., Holmgren, A., Wright, P. E., &
`Dyson, H. J. (1993) Biochemistry 32, 426-435.
`Wilson, N. A., Barbar, E., Fuchs, J. A., & Woodward, C. (1995)
`Biochemistry 34, 8931-8939.
`Wishart, D. S., Bigam, C. G., Yao, J., Abildgaard, F., Dyson, H.
`J., Oldfield, E., Markley, J. L., & Sykes, B. D. (1995) J. Biomol.
`NMR 6, 135-140.
`Yamazaki, T., Yoshida, M., & Nagayama, K. (1993) Biochemistry
`32, 5656-5669.
`BI952404N
`
`Bausch Health Ireland Exhibit 2043, Page 6 of 6
`Mylan v. Bausch Health Ireland - IPR2022-00722
`
`

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