throbber
The effect of net charge on the solubility,
`activity, and stability of ribonuclease Sa
`
`KEVIN L. SHAW,1,4 GERALD R. GRIMSLEY,1 GENNADY I. YAKOVLEV,2
`ALEXANDER A. MAKAROV,2 AND C. NICK PACE1,3
`1Department of Medical Biochemistry and Genetics, Texas A&M University, College Station, Texas 77843, USA
`2Engelhardt Institute of Molecular Biology, Moscow 119991, Russia
`3Department of Biochemistry and Biophysics and The Center for Advanced Biomolecular Research, Texas A&M
`University, College Station, Texas 77843, USA
`(RECEIVED January 28, 2001; FINAL REVISION March 14, 2001; ACCEPTED March 20, 2001)
`
`Abstract
`
`The net charge and isoelectric pH (pI) of a protein depend on the content of ionizable groups and their pK
`values. Ribonuclease Sa (RNase Sa) is an acidic protein with a pI ⳱ 3.5 that contains no Lys residues. By
`replacing Asp and Glu residues on the surface of RNase Sa with Lys residues, we have created a 3K variant
`(D1K, D17K, E41K) with a pI ⳱ 6.4 and a 5K variant (3K + D25K, E74K) with a pI ⳱ 10.2. We show that
`pI values estimated using pK values based on model compound data can be in error by >1 pH unit, and
`suggest how the estimation can be improved. For RNase Sa and the 3K and 5K variants, the solubility,
`activity, and stability have been measured as a function of pH. We find that the pH of minimum solubility
`varies with the pI of the protein, but that the pH of maximum activity and the pH of maximum stability do
`not.
`Keywords: Ribonuclease Sa; isoelectric pH; net charge; electrostatic interactions; protein solubility; en-
`zyme activity; protein stability
`
`The net charge on a protein at any given pH is determined
`by the pK values (pKs) of the ionizable groups (Tanford
`1962). The net charge on a protein is zero at the isoelectric
`point (pI), positive at pHs below the pI, and negative at pHs
`above the pI. Ribonuclease Sa (RNase Sa) has 7 Asp, 5 Glu,
`2 His, 0 Lys, and 5 Arg residues. Consequently, the wild-
`type protein has an excess of acidic residues giving it a
`pI ⳱ 3.5 and a net charge ∼−7 at pH 7 (Hebert et al. 1997).
`We have previously reported studies of the thermodynamics
`of folding and conformational stability of RNase Sa (Pace et
`al. 1998) and several mutants (Hebert et al. 1998; Grimsley
`et al. 1999; Pace et al. 2000). In this study, we have replaced
`Asp and Glu residues with Lys residues singly and in com-
`
`Reprint requests to: C. Nick Pace, Department of Medical Biochemistry
`and Genetics, Texas A&M University, College Station, Texas 77843–
`1114, USA; e-mail: nickpace@tamu.edu; fax: 979-847-9481.
`4Present address: Department of Biology, Grove City College, Grove
`City, Pennsylvania 16127, USA.
`Article and publication are at www.proteinscience.org/cgi/doi/10.1110/
`ps.440101.
`
`bination to produce variants with different pIs. The residues
`selected for replacement, Asp 1, Asp 17, Asp 25, Glu 41,
`and Glu 74, are shown in Figure 1. All of these residues are
`well exposed to solvent and do not form ion pairs or hy-
`drogen bonds. We report here the experimentally deter-
`mined isoelectric points for RNase Sa (0K) and the 2K
`(D17K + E41K), 3K (2K + D1K), 4K (3K + D25K), and 5K
`(4K + E74K) variants, and we compare these values to the
`calculated isoelectric points. The addition of five lysines is
`expected to raise the pI to above 10 and reverse the net
`charge from −7 to +3 at pH 7. The ability to predict the pI
`of a protein is useful for developing methods to purify pro-
`teins by ion exchange chromatography (Janson andRyden
`1989). We also report the dependence of solubility on pH
`for RNase Sa and the 3K and 5K variants. Theory predicts
`that the solubility of a protein will be minimal near the pI
`(Tanford 1961), and we test this in a system where the
`minimum possible changes in structure are used to vary the
`pI. Protein solubility is important in diseases such as Al-
`zheimer’s disease (Kaytor andWarren 1999), in the devel-
`
`1206
`
`Protein Science (2001), 10:1206–1215. Published by Cold Spring Harbor Laboratory Press. Copyright © 2001 The Protein Society
`
`MYLAN INST. EXHIBIT 1099 PAGE 1
`
`MYLAN INST. EXHIBIT 1099 PAGE 1
`
`

`

`Effect of net charge on the properties of RNase Sa
`
`Fig. 1. Ribbon diagram of RNase Sa. The acidic residues changed to lysines are indicated in ball and stick representation. Also given
`is the percent solvent exposure of the side chain and the oxygens in the carboxyl groups estimated by the method of Lee and Richards
`( 1971). The figure was generated with the program MOLSCRIPT (Kraulis 1991). The PDB identifier for RNase Sa is 1RGG (Sevcik
`et al. 1996).
`
`opment of recombinant proteins for treating diseases such as
`fast-acting Lys-Pro insulin (Bakaysa et al. 1996), and to
`many biochemists, especially X-ray crystallographers
`(McPherson 1998) and NMR spectroscopists (Harris 1986).
`We also report the enzyme activity as a function of pH for
`RNase Sa and the 3K and 5K variants. This is of interest
`because the natural substrate for the ribonucleases is nega-
`tively charged and the net charge on the enzyme might
`influence the steady-state enzyme kinetics. Finally, we re-
`port the pH dependence of the thermodynamics of folding
`and the conformational stability of RNase Sa and the 3K
`and 5K variants. The pH dependence of protein stability
`depends both on the net charge of the protein and on the
`difference in pKs of the ionizable groups between the folded
`and unfolded states. Our system may allow us to assess the
`relative importance of these two effects. There is wide-
`spread interest in the pH dependence of protein stability
`(Tanford 1970; Matthew andRichards 1982; Anderson et al.
`1990; Pace et al. 1990; Dao-Pin et al. 1991; Yang andHonig
`1993; Antosiewicz et al. 1994; Barrick et al. 1994; Tan et al.
`
`1995; Schaefer et al. 1998; Warwicker 1999; Pace et al.
`2000; Whitten andGarcia-Moreno 2000).
`
`Results
`
`The measured and calculated pI values for RNase Sa and the
`2K, 3K, 4K, and 5K variants are given in Table 1. The
`difference between the measured and calculated pI values
`will be discussed below.
`The solubilities of RNase Sa and the 3K and 5K variants
`are given in Table 2, and the pH dependences are shown in
`Figure 2. It is clear that the solubility minimum shifts with
`the pI of the protein from near pH 3.5 for RNase Sa to >pH 9
`for the 5K variant. Despite this, the minimum solubility of
`RNase Sa and the 5K variant is about the same, ∼2 mg
`mL−1. In contrast, the 3K variant is less soluble, having a
`minimum solubility of 0.5 mg mL−1 near pH 7.
`The steady-state kinetic parameters KM and kcat for the
`hydrolysis of poly inosinic acid (poly(I)) were determined
`as a function of pH for RNase Sa and the 3K and 5K
`
`www.proteinscience.org
`
`1207
`
`MYLAN INST. EXHIBIT 1099 PAGE 2
`
`MYLAN INST. EXHIBIT 1099 PAGE 2
`
`

`

`Shaw et al.
`
`Table 1. Isoelectric points of RNase Sa and the 2K, 3K, 4K,
`and 5K variants
`
`Sa variant
`
`Experimental pIa
`
`Calculated pIb
`
`Wild type
`2K
`3K
`4K
`5K
`
`3.5
`4.6
`6.4
`8.4
`10.2
`
`4.4 (3.9)
`5.3 (4.8)
`6.6 (6.5)
`8.2 (8.2)
`9.1 (9.1)
`
`Net charge
`at pH 7c
`
`−7
`−3
`−1
`+1
`+3
`
`a pIs of the wild-type and the 2K, 3K, and 4K variants were measured using
`isoelectric focusing gels and are accurate to ≈0.1 pH units. The pI of the 5K
`variant was estimated by determining the pH of immobility using continu-
`ous polyacrylamide gel electrophoresis and buffer systems recommended
`by McLellan (1982). This method is accurate to ≈0.3 pH units.
`b Obtained by calculating the titration curve of the protein using pK values
`for the amino acid side chains and the ␣-COOH and ␣-NH3
`+ termini
`determined from peptide/model compound studies: Asp ⳱ 4.1, Glu ⳱ 4.5
`(Nozaki and Tanford, 1967); His ⳱ 6.6 (McNutt et al. 1990); Tyr ⳱ 9.6,
`Lys ⳱ 10.4, Arg ⳱ 12.5, ␣-COOH ⳱ 3.6, ␣-NH3
`+ ⳱ 7.8 (Tanford 1968).
`The values in parentheses were obtained using the average pKs for the side
`chains of Asp, Glu, and His measured in folded proteins: Asp ⳱ 3.4 ± 1.1,
`Glu ⳱ 4.1 ± 0.8 and His ⳱ 6.5 ± 0.9. The Asp and Glu values are based
`on >200 measured pK values in 24 proteins and they were compiled and
`kindly provided to us by William Forsyth and Andy Robertson at the
`University of Iowa. The His values are based on 58 measured pK values in
`21 proteins and they were compiled and kindly proved by Steve Edgcomb
`and Kip Murphy at the University of Iowa.
`c Calculated using the pK values from the model compound data in the
`footnote above. The net change is rounded to the nearest integer.
`
`variants. The results are given in Table 3. The pH depen-
`dence of log (kcat/KM) is shown in Figure 3. The pH of
`maximum activity is ∼6.5 for RNase Sa and the 3K variant,
`and ∼7.0 for the 5K variant. The 3K variant has somewhat
`less activity than wild-type RNase Sa, whereas the 5K vari-
`ant is significantly less active than both. At the pH of op-
`timum activity, (kcat/KM) is reduced from 1187 (mM sec)−1
`for wild-type RNase Sa to 105 (mM sec)−1 for the 5K vari-
`ant. The decrease in activity for both variants results entirely
`from a decrease in kcat.
`
`Table 2. Solubility as a function of pH for RNase Sa and the
`3K and 5K variantsa
`
`Wild type
`
`3K
`
`pH
`
`2.3
`2.9
`3.6
`4.0
`4.5
`4.8
`5.0
`5.2
`5.4
`
`S (mg mL−1)
`
`11.0
`3.7
`2.2
`2.7
`4.3
`7.6
`9.8
`17.7
`23.3
`
`pH
`
`5.4
`5.7
`5.9
`6.5
`7.3
`7.7
`8.3
`9.3
`
`S (mg mL−1)
`
`4.3
`2.7
`1.7
`0.6
`0.5
`0.6
`0.7
`1.7
`
`5K
`
`S (mg mL−1)
`
`7.8
`8.5
`5.9
`4.6
`4.1
`3.2
`2.3
`5.3
`
`pH
`
`6.7
`6.7
`7.4
`7.9
`8.5
`8.8
`9.3
`10.2
`
`a Measured at 25°C in the presence of 10 mM buffer as described in
`Materials and Methods.
`
`1208
`
`Protein Science, vol. 10
`
`Fig. 2. Solubility of RNase Sa (␱) and the 3K (䊉) and 5K (⌬) variants as
`a function of pH. The lines have no theoretical significance.
`
`Thermal denaturation curves were determined by moni-
`toring the circular dichroism at 234 nm. The curves for the
`variants are similar to those previously shown for wild-type
`RNase Sa (Pace et al. 1998). All curves were analyzed by
`assuming a two-state mechanism. The analysis yields the
`melting temperature, Tm, and the van’t Hoff enthalpy
`change at Tm, ⌬Hm. These parameters and the differences in
`conformational stability, ⌬(⌬G), at pH 7 in 30 mM MOPS
`are given in Table 4. We also show the calculated ⌬⌬G
`values obtained by summing the appropriate ⌬⌬G values of
`the single lysine variants. The good agreement between the
`measured and calculated ⌬⌬G values shows that the muta-
`tional effects are additive.
`The pH dependence of the conformational stability of
`RNase Sa and the 3K and 5K variants was also studied by
`thermal denaturation. Thermal denaturation curves were de-
`termined as a function of pH from pH 2 to pH 10 and
`analyzed by assuming a two-state mechanism as previously
`described (Pace et al. 1998). The thermodynamic param-
`eters are summarized in Table 5. The free energy change for
`folding at any temperature, ⌬G(T), can be calculated using
`the modified Gibbs-Helmholtz equation,
`
`⌬G(T) ⳱ ⌬Hm(1 − T/Tm) +⌬C p[(T − Tm) − Tln(T/Tm)].
`(1)
`
`We use a value of 1.52 kcal mole−1 K−1 for ⌬Cp in
`Equation 1 to calculate ⌬G(25°C), based on our earlier stud-
`ies of the conformational stability of RNase Sa (Pace et al.
`1998). Using the results in Table 5 between pH 2 and pH 5,
`where it is not necessary to correct for the ⌬H for buffer
`ionization, gives ⌬Cp ⳱ 1.31 ± 0.18 kcal mole−1 K−1. This
`lower value of ⌬Cp would increase the ⌬G values by, at
`most, 4%. Changes in amino acid side chains on the exterior
`
`MYLAN INST. EXHIBIT 1099 PAGE 3
`
`MYLAN INST. EXHIBIT 1099 PAGE 3
`
`

`

`Table 3. Kinetic parameters characterizing the cleavage of
`poly(I) as a function of pH for RNase Sa and the 3K and
`5K variantsa
`
`Wild type
`
`3K
`
`5K
`
`pH
`
`4.02
`4.47
`5.05
`5.60
`6.05
`6.50
`7.06
`7.50
`8.01
`8.50
`4.00
`4.50
`5.00
`5.50
`6.00
`6.50
`7.06
`7.48
`8.00
`8.48
`4.47
`5.04
`5.60
`6.05
`6.50
`7.06
`7.50
`8.10
`8.50
`
`kcat(s−1)
`
`KM (mM)
`
`kcat/KM (mM s)−
`
`2.2
`14.2
`59
`131
`175
`178
`146
`95
`42.1
`5.6
`1.5
`7.5
`24
`70
`98
`112
`88
`51
`26
`8
`0.3
`1.2
`4.7
`7.7
`11.1
`10.5
`6.7
`3.5
`1.3
`
`0.25
`0.24
`0.21
`0.18
`0.16
`0.15
`0.19
`0.26
`0.29
`0.34
`0.75
`0.57
`0.33
`0.26
`0.18
`0.14
`0.14
`0.19
`0.20
`0.24
`1.10
`0.64
`0.47
`0.23
`0.13
`0.10
`0.10
`0.11
`0.14
`
`8.8
`59
`281
`728
`1094
`1187
`768
`365
`145
`16
`3.3
`13
`73
`269
`544
`800
`629
`268
`130
`33
`0.3
`1.9
`10
`34
`85
`105
`67
`32
`9.3
`
`a Measured at 25°C in the presence of 0.05 M Tris, 0.1 M potassium
`chloride and 0.05 M sodium acetate as described in Materials and Methods.
`
`of a protein are expected to have only a small effect on ⌬Cp
`because the side chains will be largely exposed to solvent in
`both the folded and unfolded conformations. Therefore we
`decided to use the more accurately determined value of 1.52
`kcal mole−1 K−1. Note that all three proteins are maximally
`stable near pH 5, with stabilities (kcal mole−1) of 7.6 for
`RNase Sa, 5.4 for 3K RNase Sa, and 5.6 for 5K RNase Sa.
`
`Discussion
`
`Isoelectric points of RNase Sa
`and the charge reversal variants
`
`Knowing the pI of a protein is helpful to protein chemists
`developing purification schemes or performing experiments
`in which the protein’s solubility is a crucial factor. (As
`discussed below, proteins tend to be the least soluble near
`their pI.) The pI is usually determined experimentally by
`isoelectric focusing (Righetti et al. 1981); however, one can
`estimate a protein’s pI by calculation if the amino acid
`
`Effect of net charge on the properties of RNase Sa
`
`Fig. 3. pH dependence of log (kcat/KM) for the cleavage of poly(I) by
`RNase Sa (␱) and the 3K (䊉) and 5K (⌬) variants.
`
`composition and the pKs of the ionizable groups are known.
`Usually the pKs of the groups are not known, so values
`obtained from model compound studies are used for the
`calculation. Because these pKs may differ significantly
`from the pKs of groups in the folded protein, calculated pIs
`often disagree with experimentally measured pIs. A number
`of methods have been proposed for the theoretical determi-
`nation of the pIs of proteins (Sillero andRibeiro 1989; Pa-
`trickios andYamasaki 1995). Typically, these methods give
`results that are within ±1 pH unit of the experimental pI.
`In Table 1 we give the experimentally determined pIs for
`RNase Sa and the 2K, 3K, 4K, and 5K variants. The value
`of 3.5 obtained for RNase Sa is in exact agreement with our
`previously reported value (Hebert et al. 1997). Thus, RNase
`Sa is an acidic ribonuclease like RNase T1, which has a pI
`of 3.8 (Hebert et al. 1997). In contrast, barnase, another
`well-studied microbial ribonuclease, is a basic protein with
`a pI of 9.2 (Bastyns et al. 1996). Replacing three acidic
`residues with lysines (3K) raises the pI close to neutrality,
`and replacing five acidic residues with lysines (5K) raises
`the pI above 10. By reversing five charges on RNase Sa, we
`have changed it from one of the most acidic proteins to one
`of the most basic proteins. This is shown in Table 6, where
`we give the experimentally determined pIs and calculated
`pIs for a number of well-studied proteins.
`When the pIs are calculated for RNase Sa and the 3K and
`5K variants using pK values taken from model compound
`studies (Table 1), the pI is overestimated for RNase Sa and
`the 2K and 3K variants and underestimated for the 4K and
`5K variants. When the calculation is performed using the
`average pKs for Asp, Glu, and His measured in folded pro-
`teins, the agreement between measured and calculated im-
`proves for RNase Sa and the 2K and 3K variants. The pKs
`of the ionizable groups of RNase Sa have been measured by
`
`www.proteinscience.org
`
`1209
`
`MYLAN INST. EXHIBIT 1099 PAGE 4
`
`MYLAN INST. EXHIBIT 1099 PAGE 4
`
`

`

`Shaw et al.
`
`Table 4. Parameters characterizing the thermal unfolding of RNase Sa and the
`charge reversal variants at pH 7a
`
`Sa variant
`
`Wild type
`D1K
`D17K
`D25K
`E41K
`E74K
`2K (E41K,D17K)
`3K (E41K,D17K,D1K)
`4K (E41K,D17K,D1K,D25K)
`5K (E41K,D17K,D1K,D25K,E74K)
`
`b
`Tm
`(°C)
`
`47.2
`48.7
`43.3
`50.2
`42.9
`51.1
`39.5
`40.4
`44.4
`46.8
`
`⌬Hm
`c
`(kcal mole−1)
`
`⌬⌬Gd
`(kcal mole−1)
`
`⌬⌬Ge
`(kcal mole−1)
`
`91
`89
`90
`93
`92
`94
`83
`78
`84
`82
`
`—
`0.4
`−1.1
`0.9
`−1.2
`1.1
`−2.2
`−1.9
`−0.8
`−0.1
`
`—
`—
`—
`—
`—
`—
`−2.3
`−1.9
`−1.0
`0.1
`
`a pH 7, 30 mM MOPS.
`b Midpoint of the thermal unfolding curve. The standard deviation is ±0.3°C.
`c Enthalpy change at Tm. The standard deviation is ±5%.
`⳱ ⌬Tm × ⌬Sm (wild type). See Becktel and Schellman (1987) or Pace and Scholtz
`d ⌬⌬Gmeas
`(1997) for a discussion of this method. A positive sign indicates an increase in stability.
`e Calculated by summing the appropriate ⌬⌬G values of the single mutant variants.
`
`NMR (D. Laurents, pers. comm. Similar measurements on
`the 5K variant are in progress.) When these values are used
`in the calculation, the pI ⳱ 3.7, in good agreement with the
`measured value. Similarly, if the pK values measured for
`RNase Sa are used for the 5K variant,
`the calculated
`pI ⳱ 10.1, again in excellent agreement with the measured
`pI. The calculated pI for the 5K variant in Table 1 is low
`because RNase Sa has eight Tyr that are largely buried and
`they all have pKs >11. This is much higher than the
`pK ⳱ 9.6 that is assumed for tyrosine -OH groups on the
`basis of model compound studies. This may be true with
`other proteins because Tyr side chains are 76% buried on
`average in a large sample of folded proteins (Lesser and
`Rose 1990).
`The fact that the pKs of the ionizable groups in model
`compounds may differ significantly from the pKs of the
`same groups in folded proteins is reflected by the calculated
`and measured pI values shown in Table 6. In general, acidic
`proteins have their pIs overestimated and basic proteins
`have their pIs underestimated when calculated using pKs
`from model compound studies. This is expected. For Asp
`side chains, a pK ⳱ 4.1 is observed with uncharged model
`compounds, but a pK ⳱ 3.4 is the average value observed
`in proteins (See the footnotes to Table 1). This results
`mainly because proteins have a net positive charge in the
`region where Asp and Glu carboxyls titrate, and this will
`tend to lower their pKs relative to those measured in un-
`charged model peptides (Antosiewicz et al. 1996). The con-
`verse will be true for basic proteins, where the proteins will
`have a net negative charge above the pI that will tend to
`raise the pKs of the ionizable groups. In addition to net
`charge, other environmental effects can change pKs. For
`example, we have shown that the completely buried and
`hydrogen bonded carboxyl of Asp 76 in RNase T1 has a pK
`
`1210
`
`Protein Science, vol. 10
`
`Table 5. Parameters characterizing the thermal unfolding of
`RNase Sa and the 3K and 5K variants between pH 2 and 10
`
`Sa variant
`
`Wild type
`
`3K
`
`5K
`
`pH
`
`2
`3.3
`5
`7
`8.3
`9
`10
`2
`3.3
`4
`5
`6
`7
`8.3
`9
`10
`2
`3
`4
`5
`6
`7
`8
`9
`10
`
`a
`Tm
`(°C)
`
`27.3
`42
`54.2
`47.2
`39.5
`35.4
`30.1
`27.9
`44.4
`50.8
`49.4
`50.0
`40.7
`31.9
`29.1
`22.3
`25.6
`41.4
`52.4
`53.3
`51.3
`47.3
`44.2
`42.7
`36.4
`
`⌬Hm
`b
`(kcal
`mole−1)
`
`⌬Sm
`c
`(cal K−1
`mole−1)
`
`⌬G (25°C)d
`(kcal
`mole−1)
`
`56
`85
`108
`91
`92
`78
`72
`60
`83
`89
`91
`84
`77
`70
`62
`50
`62
`77
`85
`87
`85
`82
`76
`72
`59
`
`186
`270
`330
`284
`294
`253
`237
`199
`261
`275
`282
`260
`245
`230
`205
`169
`208
`245
`261
`267
`262
`256
`240
`228
`191
`
`0.42
`3.88
`7.59
`5.11
`3.75
`2.36
`1.15
`0.56
`4.15
`5.48
`5.44
`4.99
`3.24
`1.46
`0.80
`−0.48
`0.12
`3.31
`5.35
`5.62
`5.22
`4.50
`3.70
`3.50
`1.85
`
`a Midpoint of the thermal unfolding curve. The standard deviation is
`±0.3°C.
`b Enthalpy change at Tm. The standard deviation is ±5%.
`⳱ ⌬Hm/Tm.
`c ⌬Sm
`d Calculated with Equation 1 using the Tm and ⌬Hm values in this table and
`a value of 1.52 kcal K−1 mole−1 for ⌬Cp (Pace et al. 1998).
`
`MYLAN INST. EXHIBIT 1099 PAGE 5
`
`MYLAN INST. EXHIBIT 1099 PAGE 5
`
`

`

`Table 6. pH of maximum stability, pHmax and isoelectric points
`of some well-studied proteins
`
`Protein
`
`Pepsin
`RNase Sa
`RNase T1
`␤-Lactoglobulin
`RNase Sa2
`RNase Sa (3K)
`RNase Sa3
`Whale myoglobin
`RNase Ba
`Horse cytochrome C
`RNase A
`Chymotrypsinogen A
`Staph nuclease
`RNase Sa (5K)
`Human lysozyme
`T4 lysozyme
`BPTI
`Hen lysozyme
`
`pHmax
`
`—
`∼5.0c,d
`4–5
`—
`∼4.5d
`4–5c
`∼5.5d
`∼6.5g
`5–6h
`∼7.5j
`7–9f
`—
`∼9.0k
`∼5.0c
`∼4.5l
`∼5.5m
`∼5.0o
`5–7p
`
`Experimental pI
`
`Calculated pIa
`
`2.9b
`3.5c,e
`3.8e
`5.1b
`5.3e
`6.4c
`7.2e
`8.0b
`9.2i
`9.4b
`9.6b
`9.6b
`>10.0k
`10.2c
`10.5b
`>10.5n
`10.6b
`11.2b
`
`3.2
`4.4
`4.3
`4.8
`7.4
`6.6
`6.2
`8.4
`9.0
`9.7
`9.3
`9.4
`9.7
`9.1
`10.3
`10.0
`10.2
`10.5
`
`a Calculated using pK values determined from model compound studies as
`described in footnote b of Table 1.
`b From Table 3 of Patrickios and Yamasaki (1995).
`c From this paper.
`d From Pace et al. (1998).
`e From Hebert et al. (1997).
`f From Pace et al. (1990).
`g From Puett (1973).
`h From Pace et al. (1992).
`i From Bastyns et al. (1996).
`j From Walter Englander, personal communication.
`k From Whitten and Garcia-Moreno (2000).
`l From Takano et al. (2000).
`m From Dao-pin et al. (1991).
`n From Becktel and Base (1987).
`o From Makhatadze et al. (1993).
`p From Pfeil and Privalov (1976) and Schaefer et al. (1998).
`
`∼0.5 (Giletto andPace 1999), but that the carboxyl group
`introduced in N44D in RNase T1 has a pK ∼5.7 (Hebert et
`al. 1998). Thus, it is always best to measure rather than
`estimate the pI of a protein unless the measured pKs of the
`ionizable groups are available.
`
`Solubility of the RNase Sa and the 3K and 5K variants
`
`Protein solubility has become increasingly important as it
`has become evident that many diseases result from the in-
`solubility of proteins. For example, a genetic variant of
`␣1-antitrypsin forms inclusion bodies in the liver and this
`can lead to early onset emphysema (Carrell andGooptu
`1998). In addition, the plaques found in patients suffering
`form Alzheimer’s disease are formed by the precipitation of
`A␤
`1–42, a peptide generated by the inappropriate cleavage of
`amyloid precursor protein (Kaytor andWarren 1999). Solu-
`bility can also be important in treating diseases. The devel-
`opment of a fast-acting form of insulin (Lys-Pro insulin)
`
`Effect of net charge on the properties of RNase Sa
`
`was prompted by the fact that formulations of human insulin
`were not solubilized rapidly enough after they were injected
`(Bakaysa et al. 1996). Solubility is also important in both of
`the primary techniques used to determine the structure of
`proteins: X-ray crystallography (McPherson 1998) and
`NMR (Harris, 1986).
`The relationship between protein solubility and pH has a
`long history in protein chemistry (Cohn andEdsall 1943;
`Tanford 1961; Arakawa andTimasheff 1985; Ries-Kautt
`and Ducruix 1997). To a first approximation, the solubility
`of a protein is proportional to the square of the net charge on
`the protein (Tanford 1961). Consequently, proteins are ex-
`pected to be least soluble near their isoelectric points, and
`the solubility should increase as the pH is raised or lowered
`as the magnitude of the net charge increases. Our results
`support this. Figure 3 shows that the minimum solubility of
`RNase Sa and the 3K and 5K variants occurs within a pH
`unit of the pIs and the solubility increases markedly at both
`higher and lower pH.
`
`Enzyme activity of RNase Sa
`and the 3K and 5K variants
`
`Our hope was that the charge reversal mutants would retain
`most of their enzyme activity. Of the five Glu and Asp
`residues that were considered (Fig. 1), only Glu 41 appeared
`to be important to substrate binding or catalysis. Glu 41 in
`RNase Sa is conserved in all of the microbial ribonucleases
`where it forms hydrogen bonds to the guanine base of the
`nucleotide (Sevcik et al. 1990; Hebert et al. 1997). Thus, it
`is important to both specificity and binding. Previous stud-
`ies have shown that the equivalent residues in barnase (Glu
`60) and RNase T1 (Glu 46) were important to the activity.
`The value of kcat/Km for RNA hydrolysis is reduced by 30%
`in the E60Q mutant of barnase (Bastyns et al. 1994) and by
`95% for the E46Q mutant of RNase T1 (Steyaert et al.
`1991). It seemed likely that replacing Glu with Lys would
`be much more detrimental to the enzyme activity than re-
`placing Glu with Gln. Thus, it was surprising that kcat/Km
`for the 3K variant is only reduced by ∼33%, and kcat/Km for
`the 5K variant is only reduced by 90% (Table 3). In both of
`these mutants, Glu 41 has been replaced by Lys. It is even
`more surprising that the decrease in activity is entirely
`caused by a decrease in kcat. Based on the Km value, these
`mutants appear to bind substrate better than wild-type
`RNase Sa. Perhaps, the loss of the hydrogen bond to the
`inosine base in E41K RNase Sa is compensated for by more
`favorable electrostatic interactions, because the net charge is
`∼−6 for RNase Sa, ∼1 for the 3K variant, and ∼+4 for the 5K
`variant near pH 7 (using the measured pKs of all charged
`groups). The activity of the 5K variant is lower that that of
`RNase Sa and the 3K variant over the pH range pH 4– pH
`8.5, but the pH of maximum activity is similar: ∼6.5 for
`RNase Sa and 3K variant, and ∼7 for the 5K variant (Fig. 3).
`
`www.proteinscience.org
`
`1211
`
`MYLAN INST. EXHIBIT 1099 PAGE 6
`
`MYLAN INST. EXHIBIT 1099 PAGE 6
`
`

`

`Shaw et al.
`
`Thus, unlike solubility, the pH dependence of the enzyme
`activity does not vary significantly with the pI of the en-
`zyme. This is not surprising because the pKs of the ioniz-
`able groups on the enzyme actually involved in binding and
`catalysis will influence the pH dependence to a greater ex-
`tent than the net charge on the protein (Fersht 1999).
`
`problem to analyze. Note that even though RNase Sa is a
`very acidic protein (net charge ∼−6 at pH 7) and the 5K
`variant a very basic protein (net charge ∼+4 at pH 7) they
`have the same stability at pH 7. In contrast, the 3K variant
`is about 2 kcal mole−1 less stable than both RNase Sa and
`the 5K variant.
`
`Conformational stability of the single
`charge reversal mutants at pH 7
`
`pH dependence of the conformational stability
`of RNase Sa and the 3K and 5K variants
`
`As shown in Figure 1, the Glu and Asp residues selected for
`charge reversal are highly accessible to solvent and not
`hydrogen bonded. In addition, when Lys residues are sub-
`stituted for the Asp and Glu residues, calculations show that
`+ group of the Lys residues should still be too far
`the -NH3
`from any negative charges to form an ion pair. Thus, the
`main effect of these charge reversals on stability should be
`through a change in the long-range electrostatic interactions
`with the rest of the charged groups in RNase Sa. The actual
`net charge on RNase Sa at pH 7 will be ∼−6 based on the
`measured pKs of all charged groups (Laurents et al. un-
`publ.). Consequently, replacing a negative charge with a
`positive charge is expected to lead to an increase in stability.
`This is supported by calculations using Coulomb’s law and
`a dielectric constant of 80 to sum up the electrostatic inter-
`actions. If we assume that electrostatic interactions are neg-
`ligible in the unfolded states, then the predicted increases in
`stability are (kcal/mole): 2.9 for D1K, 3.2 for D17K, 2.5 for
`D25K, 1.2 for E41K, and 2.3 for E74K. The results in Table
`4 show that, in fact, much smaller increases in stability are
`observed for three of the single variants and that a decrease
`in stability is observed for the other two. We have discussed
`some of these data and related results in two previous papers
`and reached the following conclusions. First, protein stabil-
`ity can sometimes be increased by improving the electro-
`static interactions among charged groups on the surface of a
`protein (Grimsley et al. 1999). Second, the discrepancy be-
`tween the observed and calculated changes in stability leads
`to the surprising suggestion that the free energy of the de-
`natured state may be decreased more than that of the native
`state by these single charge reversal mutations (Pace et al.
`2000). This makes it difficult to predict which charge re-
`versals will be most likely to increase the stability of a
`protein.
`
`Conformational stability of
`2K, 3K, 4K, and 5K at pH 7
`The observed ⌬⌬G values of the multiple mutants can be
`approximated remarkably well by summing the appropriate
`⌬(⌬G) values of the single mutants (Table 4). At first
`glance, this seems surprising. However, at pH 7 each charge
`is interacting with ∼19 other charges in both the folded and
`unfolded states of the protein so this is clearly a difficult
`
`1212
`
`Protein Science, vol. 10
`
`In 1924, a model for the electrostatic interactions of a pro-
`tein was proposed that smeared the net charge over the
`surface of a sphere with a dielectric constant, D, that was
`impenetrable to solvent (Linderstrom-Lang 1924). This
`model predicts that proteins should be most stable near their
`pIs, where their net charge is zero, because unfavorable
`electrostatic interactions resulting from either an excess of
`positive or negative charges will be minimized. Fig. 4
`shows the conformational stabilities estimated from thermal
`denaturation curves, ⌬G(25°C), as a function of pH for
`RNase Sa and the 3K and 5K variants. The maximum sta-
`bility for all three proteins is near pH 5, and clearly does not
`shift with the pI of the protein. At pH 5, RNase Sa and the
`3K and 5K variant have approximate net charges of −4, +2,
`and +5, respectively. Thus, the 3K variant is the least stable
`of the three proteins even though it has the net charge clos-
`est to 0 at this pH. Table 6 gives the pH of maximum
`stability, pHmax, for many of the proteins listed. It is inter-
`esting that the pHmax for eight of the 15 proteins is >2 pH
`units from their experimental pIs. These results show clearly
`that the Linderstrom-Lang model cannot account for the pH
`dependence of the stability of these proteins.
`The more important factor that determines the pH depen-
`dence of protein stability is the difference in the pK values
`of the ionizable groups in the folded and unfolded confor-
`mations of a protein. This subject has been discussed by
`several authors (Tanford 1962, 1970; Matthew and Gurd
`1986; Yang and Honig 1993; Antosiewicz et al. 1994; El-
`cock 1998; Schaefer et al. 1998; Warwicker, 1999). Experi-
`mental studies of the pH dependence of stability have been
`reported for RNase A (Pace et al. 1990), RNase T1 (McNutt
`et al. 1990; Pace et al. 1990; Hu et al. 1992; Yu et al. 1994),
`T4 Lysozyme (Anderson et al. 1990; Dao-Pin et al. 1991),
`barnase (Pace et al. 1992; Makarov et al. 1993; Oliveberg et
`al. 1995), ovomucoid third domain (Swint-Kruse andRob-
`ertson, 1995), chymotrypsin inhibitor 2 (Tan et al. 1995),
`N-terminal domain of L9 (Kuhlman et al. 1999), and staph-
`ylococcal nuclease (Whitten andGarcia-Moreno 2000).
`As shown in Figure 4, the maximum stability for RNase
`Sa and the 3K and 5K variants occurs near pH 5. We will
`first discuss the marked decrease in stability between pH 5
`and 2. An analysis of this region shows that the three pro-
`teins take up ∼2 protons when they unfold. This shows that
`some of the groups in the unfolded protein have higher pKs
`
`MYLAN INST. EXHIBIT 1099 PAGE 7
`
`MYLAN INST. EXHIBIT 1099 PAGE 7
`
`

`

`Effect of net charge on the properties of RNase Sa
`
`than the same groups in the folded protein. The only groups
`that titrate in this pH range are the carboxyl groups. The pKs
`of carboxyl groups in folded proteins are generally consid-
`erably lower than observed in uncharged peptide models.
`The average pKs in folded proteins are less than in peptide
`models by 0.7 for Asp, 0.4 for Glu, and 0.1 for His (See
`footnote b in Table 1). The fact that the effect is larger for
`Asp than Glu reflects the fact that this is caused, in part, by
`the large positive charge that develops on proteins at low pH
`where the carboxyls titrate (Antosiewicz et al. 1996). In the
`unfolded protein, the pKs will be higher because the mol-
`ecule expands and the effective dielectric constant
`in-
`creases. However, it is clear that the carboxyl pKs will still
`be less than in peptide models (Elcock 1998). Only in 6M
`GdnHCl or 8M urea in the presence of 1M salt do the pKs
`in unfolded proteins approach the pKs observed in model
`compounds (Whitten and Garcia-Moreno 2000).
`RNase Sa has 13 carboxyls, and all but three have lower
`pKs than in model compounds (D. Laurents, unpubl.). The
`3K and 5K variants have fewer carboxyls than does RNase
`Sa, but the carboxyls with the lowest pKs have not been
`removed. In addition, the 3K and 5K variants have greater
`positive charges at low pH than RNase Sa, so the pKs of the
`carboxyl groups are expected to be even lower. Conse-
`quently, the pH dependence is quite similar for RNase Sa
`and the 3K and 5K variants at low pH and can be explained
`by the difference in pKs between the groups in the folded
`and unfolded states of the protein.
`Above pH 5, the dependence on pH is less and ∼0.8
`protons are released when RNase Sa or the 3K or 5K vari-
`ants unfold. Again, this is expected based on the pKs that
`we have measured in RNase Sa. Both the ␣-amino group
`and the side chain of His 53 have considerably higher pKs
`in folded RNase Sa than in model compounds and can easily
`account for the pH dependence observed above pH 5. We
`will attempt to account completely for the pH dependence
`of RNase Sa and the 5K variant when we have completed
`measurements of the pKs in the folded and unfolded states
`of these proteins.
`
`Materials and methods
`
`Glycine, diglycine, sodium acetate, MES (2–[N-morpholino]eth-
`anesulf

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