throbber
Andrei L. Lomize
`Henry I. Mosberg
`College of Pharmacy,
`University of Michigan,
`Ann Arbor, Michigan 48109
`
`Received 5 December 1996;
`accepted 19 February 1997
`
`Thermodynamic Model of
`Secondary Structure for
`a-Helical Peptides
`and Proteins
`
`Abstract: A thermodynamic model describing formation of a-helices by peptides and proteins
`in the absence of specific tertiary interactions has been developed. The model combines free
`energy terms defining a-helix stability in aqueous solution and terms describing immersion of
`every helix or fragment of coil into a micelle or a nonpolar droplet created by the rest of
`protein to calculate averaged or lowest energy partitioning of the peptide chain into helical
`and coil fragments. The a-helix energy in water was calculated with parameters derived from
`peptide substitution and protein engineering data and using estimates of nonpolar contact
`areas between side chains. The energy of nonspecific hydrophobic interactions was estimated
`considering each a-helix or fragment of coil as freely floating in the spherical micelle or
`droplet, and using water/cyclohexane (for micelles) or adjustable (for proteins) side-chain
`transfer energies. The model was verified for 96 and 36 peptides studied by 1H-nmr spectroscopy
`in aqueous solution and in the presence of micelles, respectively ([set 1] and [set 2]) and for
`30 mostly a-helical globular proteins ([set 3]). For peptides, the experimental helix locations
`were identified from the published medium-range nuclear Overhauser effects detected by 1H-
`nmr spectroscopy. For sets 1, 2, and 3, respectively, 93, 100, and 97% of helices were identified
`with average errors in calculation of helix boundaries of 1.3, 2.0, and 4.1 residues per helix
`and an average percentage of correctly calculated helix–coil states of 93, 89, and 81%,
`respectively. Analysis of adjustable parameters of the model (the entropy and enthalpy of the
`helix–coil transition, the transfer energy of the helix backbone, and parameters of the bound
`coil), determined by minimization of the average helix boundary deviation for each set of
`peptides or proteins, demonstrates that, unlike micelles, the interior of the effective protein
`droplet has solubility characteristics different from that for cyclohexane, does not bind frag-
`ments of coil, and lacks interfacial area. q 1997 John Wiley & Sons, Inc. Biopoly 42: 239–
`269, 1997
`
`Keywords: a-helix stability; secondary structure prediction; micelles; protein folding
`
`INTRODUCTION
`
`There are two types of theoretical approaches to
`the protein folding problem. Approaches originating
`from conformational analysis of peptides and poly-
`mer physics consider a protein molecule as a long
`
`polymer chain, the energy of which must be mini-
`mized by searching in the space of torsion angles, 1–3
`or by using simplified lattice models.4 – 6 An alterna-
`tive way of looking at the problem is to represent
`a protein as a system of secondary structure ele-
`ments, 7 – 12 as in every publication describing three-
`
`Correspondence to: Henry I. Mosberg
`Contract grant sponsor: National Institutes of Health
`Contract grant numbers: DA03910 and DA00118
`q 1997 John Wiley & Sons, Inc.
`
`CCC 0006-3525/97/020239-31
`
`239
`
`/
`
`8K26$$5476
`
`06-17-97 09:16:43
`
`bpa
`
`W: Biopolymers
`
`5476
`
`CFAD Exhibit 1019
`
`1
`
`

`
`240
`
`Lomize and Mosberg
`
`dimensional structures of specific proteins. Only a-
`helices, b-sheets, or short covalently bridged cycles
`(as in conotoxins or in metallothioneins) can be
`stable enough to serve as nucleations initiating pro-
`tein folding, and therefore they are present in 3D
`structures of all known proteins. Cooperative forma-
`tion of backbone hydrogen bonds in a-helices and
`b-sheets provides their high intrinsic stability, and
`simultaneously, burial of the polar main chain,
`which gives an additional energy gain when the
`amphiphilic secondary structure elements aggregate
`with each other, creating the nonpolar protein core.
`A simultaneous or stepwise formation of the sec-
`ondary structure frameworks by the hydropho-
`bically collapsed peptide chain, which is usually
`supplemented by covalent cross-linking in small
`proteins, has been directly demonstrated in experi-
`mental studies of protein folding.13 – 17 In terms of
`secondary structure, the protein folding process can
`be represented as a sequence of the following
`events: (1) formation of a-helices and b-sheets by
`the collapsed peptide chain, (2) assembly of the
`regular secondary structure elements into the protein
`core, and (3) joining of nonregular loops and the
`less stable ‘‘peripheral’’ helices and b-strands to the
`core and the association of independently formed
`domains. A theory of protein self-organization must
`reproduce all these events to calculate the protein
`3D structure.
`Formation of a-helices depends on various fac-
`tors that can be studied separately by considering
`the following, increasingly complicated situations:
`(1) small linear peptides in aqueous solution, where
`stability of each helix depends only on interactions
`between its own residues (Figure 1a); (2) peptide–
`micelle complexes, where each helix is stabilized by
`a combination of the intrahelical and hydrophobic
`interactions with the micelle (Figure 1b); and (3)
`proteins, in which helices are stabilized by specific
`tertiary interactions along with intrahelical and non-
`specific hydrophobic ones (Figure 1c; we denote as
`‘‘specific’’ the interactions between atoms or groups
`that must be described by pairwise potentials, and as
`‘‘nonspecific’’ the interactions of individual groups
`with a medium or averaged surrounding which can
`be described by transfer energies). The helix–coil
`transition is usually treated by Lifson–Roig and
`Zimm–Bragg theories.18 However, even with essen-
`tial modifications, 19 – 22
`these theories and other
`models 23,24 deal only with intrahelical interactions
`(i.e., they describe formation of individual helices
`in water, Figure 1a), or can be modified for the
`specific case of dimeric coiled coils.25 The goal of
`the present work is to develop a thermodynamic
`
`FIGURE 1 Three models of a-helix formation (the
`helices are shown as rectangles, solid circles are hy-
`drophobic side chains): (a) ‘‘peptide in aqueous solu-
`tion’’ (there are only specific interactions between resi-
`dues within each a-helix); (b) ‘‘peptide in complex with
`a micelle’’ (there are specific intrahelical and nonspecific
`hydrophobic interactions of every a-helix with the mi-
`celle) —coil fragments may compete with helices for
`binding with the micelle; (c) ‘‘droplet-like protein’’
`model (each helix and coil fragment floats in the liquid-
`like nonpolar spherical droplet created by the rest of pro-
`tein).
`
`model of a-helix formation that would be applicable
`for micelle-bound peptides and for proteins (Figure
`1b,c). The model, also for the first time, reproduces
`locations of the a-helices identified from medium-
`range NOEs in a representative set of peptides, in-
`stead of using average a-helicities derived from CD
`spectroscopy data, or qualitative comparisons with
`chemical shifts of C aH protons, as in the previous
`theoretical studies of peptides in aqueous solu-
`tion.23,24
`The model can be briefly outlined as follows.
`
`/ 8K26$$5476
`
`06-17-97 09:16:43
`
`bpa
`
`W: Biopolymers
`
`5476
`
`2
`
`

`
`a-Helical Peptides and Proteins
`
`241
`
`proach for hydrophobic interactions between side
`chains in helices and a slightly different parametri-
`zation of some other interactions.
`For a peptide in the micelle bound state (Figure
`1b) DGl, the free energy of its bound helix–coil
`partition l relative to a coil in aqueous solution, can
`be given by
`DGl (cid:129) (Eel 0 TDSimm)
`/ (cid:229)
`DG a(ki , mi ) / (cid:229)
`
`DG coil(kj, mj)
`
`(2)
`
`i
`
`j
`
`where Eel is the peptide-micelle electrostatic interac-
`tion energy, DSimm is the immobilization entropy of
`the peptide, 27 and the two sums in this equation are
`free energy changes for bound a-helical and coil
`fragments of m residues starting from residue k.
`Equation (2) can be simplified assuming, first, that
`the equilibrium is strongly shifted toward the bound
`peptide form, so that only bound helix–coil parti-
`tions need be considered, and second, that the total
`energy of electrostatic interactions of charged pep-
`tide groups with the micelle does not depend on the
`secondary structure of the peptide. Then the (Eel
`0 TDSimm) term, which is of crucial importance for
`peptide–micelle binding, can be considered to be
`a constant for all bound helix–coil partitions and
`subtracted in calculations of their relative energies.
`The energies of individual helices are additive
`[as in Eqs. (1) and (2)] when the helices do not
`interact with each other, i.e., for monomeric pep-
`tides lacking tertiary structure (Figure 1a,b), but
`the situation is more complicated in the presence of
`specific tertiary interactions. However, if the tertiary
`interactions are reduced, as in molten globules and
`in the intermediate and transition protein folding
`states, 14,28 – 31 the additivity approximation for helix
`energies can be applied. In a fluctuating compact
`state, each a-helix can be considered as floating
`in a dynamically averaged interior of a nonpolar
`spherical droplet created by the rest of the protein
`(Figure 1c) and stabilized independently of other
`helices by intrahelical and nonspecific hydrophobic
`interactions, similar to the micelle-bound peptides.
`Then, the energies of individual helices and coil
`fragments in a protein can also be simply summed:
`DGl (cid:129) (cid:229)
`
`i
`
`DG a(ki , mi )
`/ (cid:229)
`DG coil(kj, mj) / DG *
`
`(3)
`
`j
`
`where the energies of the bound helices and coil
`
`FIGURE 2 Helix–coil partitions as conformational
`states of the peptide chain. The coil in aqueous solution
`serves as a reference state with zero energy. The helices
`A, B, and C, shown as rectangles, with DG (cid:155) 0, are
`more stable than the coil. The helices compete with each
`other, and partition 1 consisting of two (A / B) helices
`can be of lower energy than partition 2 containing only
`helix C overlapped with A and B, even if helix C has
`lower energy than either of the individual A and B helices.
`Helix D (DG (cid:156) 0) is less stable than coil but may be
`detected spectroscopically. Partitions 1–4 are in equilib-
`rium with each other and all may contribute to observed
`parameters of nmr and CD spectra.
`
`Each partition of a peptide into helix and coil frag-
`ments (Figure 2) can be considered as a molecular
`conformational state defined by the variables N, k1,
`m1, . . . , ki , mi , . . . , kN, and mN, where N is the
`number of helices in the molecule, and where ki
`and mi (i (cid:129) 1, 2, . . . , N) represent the number of
`the first residue and the length, respectively, for
`each helix. Like coil or folded protein states, each
`helix–coil partition is an ensemble of conformers
`defined by torsion angles w, c, x, and, judging from
`molecular dynamics simulations, 26 interconversions
`of the partitions, i.e., lengthening, shortening, or
`breaking of helices, are slower than rotations of side
`chains and coil fluctuations.
`For a peptide in aqueous solution (Figure 1a),
`the unfolding free energy, DGl, of helix-coil parti-
`tion l can be written as the sum of the helix–coil
`free energy differences, DG a(ki , mi ), for all indi-
`vidual a-helices from the partition:
`
`Nl
`
`DGl (cid:129) (cid:229)
`i(cid:129)1
`
`DG a(ki , mi )
`
`(1)
`
`where Nl is the number of helices in partition l. The
`energies of individual helices in water, DG a(ki ,
`mi ), were calculated here with parameters derived
`from peptide substitution and protein engineering
`data, similar to that in the work of Munoz and Ser-
`rano, 23,24 but using a more physically justified ap-
`
`/
`
`8K26$$5476
`
`06-17-97 09:16:43
`
`bpa
`
`W: Biopolymers
`
`5476
`
`3
`
`

`
`Free Energy of a-Helix
`in Aqueous Solution
`
`The helix–coil free energy difference, DG a(k, m), for
`a fragment of peptide chain of m residues, starting from
`residue k, can be divided into the contribution of main-
`chain interactions ( DG mch), which is the free energy
`difference for the ‘‘host’’ polyAla peptide, the interac-
`tions of side chains with the helix backbone ( DG sch
`int )
`that describes free energy changes associated with re-
`placement of the host Ala C bH3 group by other side
`chains, 34,35 the hydrogen-bonding and electrostatic inter-
`36 –38
`actions between polar side chains in water, DG sch
`hb
`and the hydrophobic interactions of side-chains DG sch
`pho
`(Refs. 39 and 40):
`
`
`
`pho (4)hb / DG schDG a(ki , mi ) (cid:129) DG mch / DG schint / DG sch
`
`
`
`
`
`242
`
`Lomize and Mosberg
`
`segments can be calculated similar to that for mi-
`celle-bound peptides, and the DG * term arises from
`loss of entropy by aggregating helices and is as-
`sumed to be a constant, independent of the helix–
`coil partition. Then, the relative energies of the he-
`lix–coil partitions can be approximated by the first
`two sums from this equation, which differ from un-
`folding free energies by the term DG *.
`All possible helix–coil partitions are in equilib-
`rium with each other (Figure 2), including single
`helices, which are less stable than coil, but still
`detectable spectroscopically. This situation can be
`treated using Boltzmann averaging of the parti-
`tions 32 to calculate local a-helicities that can be
`compared with spectroscopically observed parame-
`ters. The number of the possible partitions grows
`rapidly with the chain length, which makes such
`calculations impossible for proteins. However, we
`show here that even the single lowest energy helix–
`coil partition (Figure 2) can satisfactorily reproduce
`experimentally observed locations of the helices,
`which are additionally stabilized by hydrophobic
`interactions with the micelles or with the rest of the
`protein. If the helix energies are additive, the search
`for the lowest free energy helix–coil partition (i.e.,
`the global energy minimization with respect to the
`N, k1, m1, k2, m2, . . . , kN, mN variables) can be
`easily performed using the dynamic programming
`algorithm.33
`
`METHODS
`
`The computational procedure implemented here in the
`program FRAMEWORK consists of the following steps:
`(1) Calculation of a-helix and bound coil energies for
`each fragment of the molecule, depending on the chosen
`model [‘‘peptide in aqueous solution,’’ ‘‘peptide in mi-
`celle,’’ or ‘‘droplet-like protein’’; Eqs. (1) – (20)]. (2)
`Boltzmann averaging of helix–coil partitions to calculate
`the local a-helicities of every tripeptide fragment of the
`molecule [Eqs. (21) and (22)] or search for the lowest
`energy helix–coil partition [Eqs. (23) and (24). (3)
`Minimization of the average deviation of calculated and
`experimental boundaries of a-helices [Eq. (25)] with
`respect to several adjustable parameters of the model.
`The average helix boundary deviation [Eq. (25)] was
`implemented, since the widely used percentage of cor-
`rectly calculated secondary structure states ( a, b, or non-
`regular) does not properly reflect success or failure of a
`prediction algorithm: a wrong prediction that sperm
`whale myoglobin, for example, is a single long helix
`would have a ‘‘success’’ rate as high as 89%, while the
`correct identification of all myoglobin helices with a
`small ((cid:130) 10%) error in the ends of each helix would
`produce the same success rate.
`
`Main-Chain Interactions. The helix–coil free energy
`difference for the host polyAla peptide is given by
`
`DG mch(ki , mi ) (cid:129) (mi 0 2)DH 0 miTDS
`
`(5)
`
`where DH is the enthalpy of the hydrogen-bonding inter-
`action between two peptide groups in the a-helix, and
`DS is the conformational entropy change per residue dur-
`ing the helix–coil transition.34 The DH and DS contribu-
`tions measured by Hermans 41 and Scholtz et al.42 are
`considered here as adjustable parameters of the model
`and must be determined independently by fit of calculated
`and experimentally identified positions of a-helices in
`peptides.
`
`Side-Chain–Main-Chain Interactions. The energy
`of interaction between side-chains and the a-helix back-
`bone DG sch
`int was calculated as the sum of corresponding
`published free energy differences DDG sch
`, measured by
`i
`replacing the host Ala residue in model peptides and
`proteins:
`
`int (cid:129) (cid:229)k/m01
`
`DG sch
`
`i(cid:129)k01
`
`DDG sch
`i
`
`(6)
`
`where the replacement energies DDG sch
`depend on the
`i
`type of side chain i and its position within the a-helix or
`nearby: the energies can be different in the middle of the
`a-helix and near its termini, in positions denoted as N *-
`Ncap-N1-N2-N3-rrr-C3-C2-C1-Ccap-C *. The corre-
`sponding a-helix propensities (DDG sch) measured for
`different peptides and proteins are not perfectly mutually
`consistent, and some of them reproduce the nmr-detected
`peptide helices more satisfactorily than others. Attempts
`to reproduce the peptide helices led to the parametrization
`and interpretation of the published a-helix propensity
`data described below.
`
`/ 8K26$$5476
`
`06-17-97 09:16:43
`
`bpa
`
`W: Biopolymers
`
`5476
`
`4
`
`

`
`Middle Helix, C-Turn, C-Cap, and N-Turn Posi-
`tions. Because of the two-state behavior of proteins, the
`corresponding protein engineering scales were derived
`directly from thermodynamic measurements, while the
`corresponding energies for peptides have been obtained
`by using theories of the helix–coil transition. Remark-
`ably, the averaging of two protein engineering scales
`measured in the middle helix positions (for a-helical di-
`mers 43 and 44 site of T4 lysozyme 44) gives a set of
`DDG sch values that is nearly identical (the correlation
`coefficient is 0.98) to the scale independently developed
`for 10 residues by Lyu et al.45 using the model ‘‘EXK’’
`peptide. The ‘‘EXK’’ peptide, which is stabilized by nu-
`merous ionic pairs and by the N-capping motif, also has
`a protein-like two state behavior, as can be seen from the
`similar DDG sch energies calculated using two-state and
`multistate models from CD data.45 Thus, all these three
`middle-helix scales are consistent and can simply be aver-
`aged to reduce the experimental errors. The correspond-
`ing average DDG sch values used here (Table I) are close
`to the AGADIR scale 23,24 for all but Pro and Gly residues,
`and to the scale of Chakrabatty et al.53 for all residues,
`except Val, Phe, Trp, Pro, and Gly.
`In the helix C-turn (C2 position), 46 the experimental
`DDG sch energies are different: they are larger than in the
`middle of the a-helix by 0.3–0.5 kcal/mol for aromatic
`Trp, Phe, and Tyr residues and Cys, by (cid:130) 0.4 kcal/mol
`for b-branched Ile and Val side chains, by 0.1–0.2 kcal/
`mol for linear side chains containing a C gH2 group (Leu,
`Met, Glu, Gln), and are unchanged for Gly and the short
`polar Ser and Asn side chains (Table I). These energeti-
`cally unfavorable effects probably arise from shielding
`of unpaired carbonyls at the C-terminus of the a-helix
`by the g substituents of the side chains and the larger
`accessibility of the nonpolar g substituents themselves in
`the C-turn, compared to that in the middle of the a-helix.
`If the C2 side chain has a trans orientation (x 1 (cid:130) 1807),
`its g-methyl group or aromatic ring (of Phe, for example)
`reduces accessibility of the closest (C2) free C|O main
`chain oxygen by 26 or 36%, respectively, while the acces-
`sibilities of the nonpolar g-methyl or aromatic ring them-
`selves are increased by (cid:130) 11 A˚ 2 (the equivalent transfer
`energy is (cid:130) /0.2 kcal/mol) compared to that in the
`middle of a-helix. At the same time, the accessibilities
`of the C|O groups and side chains are not affected in
`the C-turn if the side chains have gauche orientations
`(x 1 (cid:130) 0607). As discussed below, this solvation effect
`changes preferred conformations of side chains in the C-
`turn from trans to gauche.
`The destabilization in C-turn positions is less for Lys
`and Arg compared with other residues with linear chains,
`and for His compared to other aromatic residues (Table
`I), probably because of small ((cid:130) 00.2 kcal/mol) electro-
`static attractions between the positively charged side
`chains and the helix dipole (as a result, the DDG sch ener-
`gies of Lys and Arg in the middle of the helix and C-
`turn are identical, Table I). Repulsions of the Asp side
`chain with the helix dipole increases its DDG sch energy
`
`a-Helical Peptides and Proteins
`
`243
`
`by (cid:130) /0.2 kcal/mol in C-turn positions compared to
`middle helix positions (Table I). The influence of electro-
`static interactions is smaller for the Glu residue ((cid:130) /0.1
`kcal/mol), because its longer, flexible side chain can
`move away from the helix reducing the electrostatic re-
`pulsion. The electrostatic interactions of side chains at
`the C-terminus of the helix are weaker than at the N-
`terminus (00.6 to 00.9 kcal/mol 54) because the interac-
`tions depend on the spatial position of the charged groups
`relative to the helix dipole. The C a-C b bonds of side
`chains are tilted relative to the helix axis and directed
`toward the helix N-terminus. As a result, in the N-turn,
`the COO0 groups of Asp and Glu side chains are situated
`close to the helix dipole axis, near unsatisfied local di-
`poles of backbone NH groups, and may even form hydro-
`gen bonds with them, while the positively charged side
`chains in the C-turn are far from the helix dipole axis.
`However, when His, Lys, or Arg residues occupy the C-
`cap position and their w and c angles are in the left-
`handed helix area of the Ramachandran map (the struc-
`tural motif of His 18 in barnase), the positively charged
`side chains are brought into the same position relative to
`the helix dipole as the negatively charged side chains in
`the N-turn: they are situated near the helix axis and can
`form hydrogen bonds with the main chain C|O groups,
`thus producing stronger electrostatic interactions: (cid:130) 00.6
`kcal/mol.55 Stabilization of a-helices by positively
`charged side chains, observed for model peptides, 56 may
`arise chiefly from this C-capping interaction. No special
`contributions for electrostatic interactions in the C-turn
`were used since they are already included in the C-turn
`DDG sch energies, and an average energy of electrostatic
`interactions for His, Lys, and Arg residues in the C-
`cap position was considered as an adjustable parameter,
`whose optimum value was found to be 00.4 kcal/mol. No
`other contributions were used for C-cap residues because
`experimental data here are contradictory: some studies 57
`clearly demonstrate the significance of the C-capping in-
`teractions, especially for Asn residues, while others 48
`show that these interactions are negligible.
`In N-turn (N1-N3) positions, a small (00.2 kcal/mol)
`correction of the middle helix scale was applied for the
`short polar Ser, Thr, and Asn residues and for Gly based
`on results of Serrano et al.58 The DDG sch of Pro in N2 and
`N3 positions was reduced to 1 kcal/mol, 59,60 since the Pro
`side chain in the N turn of the a-helix causes steric hin-
`drances with the preceding residue but does not produce an
`energetically unfavorable kink in the a-helix (this correlates
`with the much higher statistical occurrence of Pro in N-turn
`compared to middle helix positions).61
`The pH dependence of all electrostatic contributions
`and pKs for charged side chains were taken into account
`as in the work of Munoz and Serrano.24 Energies of elec-
`trostatic interactions of completely ionized side chains in
`N-turn positions with the helix dipole were considered
`as adjustable parameters, and their optimum values were
`00.9 kcal/mol for Asp and Glu in the N1 and N2 posi-
`tions (the ‘‘capping box’’ N3 residues were treated sepa-
`
`/
`
`8K26$$5476
`
`06-17-97 09:16:43
`
`bpa
`
`W: Biopolymers
`
`5476
`
`5
`
`

`
`244
`
`Lomize and Mosberg
`
`Table I a-Helix ‘‘Propensities’’ (DDGsch) and Transfer Energies (DG sch
`tr ) of Side Chains
`
`DDG sch (kcal/mol)
`
`DG sch
`tr
`
`(kcal/mol)
`
`Residue
`
`Middle
`Helixa
`
`C-turn
`Positionsb
`
`Leu
`Ile
`Val
`Phe
`Trp
`Met
`Pro
`Ala
`Cys
`Tyr
`Gly
`Thr
`Ser
`His
`Lys
`Gln
`Glu
`Asn
`Asp
`Arg
`
`0.14
`0.35
`0.46
`0.37
`0.35
`0.20
`3.40
`0.0
`0.49
`0.42
`0.86
`0.54
`0.43
`0.55
`0.17
`0.30
`0.46
`0.63
`0.53
`0.14
`
`0.35
`0.81
`0.88
`0.69
`0.84
`0.31
`3.40
`0.0
`0.82
`0.82
`0.91
`0.79
`0.41
`0.78
`0.19
`0.48
`0.55
`0.66
`0.71
`0.14
`
`N-Cap
`Positionsc
`00.17
`00.10
`0.01
`00.17
`00.41
`00.06
`00.06
`0.0
`00.08
`00.25
`00.34
`00.19
`00.34
`00.17
`0.06
`0.22
`00.17
`00.52
`00.51
`0.0
`
`Cyclohexane
`Coild
`03.91
`03.76
`02.89
`02.00
`01.63
`01.79
`02.23
`00.94
`00.89
`0.86
`00.29
`3.32
`3.99
`5.29
`6.38
`6.37
`7.53
`7.23
`9.39
`15.75
`
`Cyclohexane
`a-Helixd
`03.01
`02.63
`01.97
`01.99
`01.58
`01.36
`01.00
`00.63
`00.57
`0.85
`00.25
`2.38
`3.06
`5.82
`7.25
`7.03
`8.23
`7.71
`9.81
`16.49
`
`Protein
`a-Helixe
`02.10
`01.89
`01.40
`02.03
`01.68
`01.47
`00.49
`00.21
`01.68
`00.63
`00.14
`0.00
`2.10
`5.8
`7.2
`7.0
`8.2
`7.7
`9.8
`16.5
`
`a Average parameters for the host a-helix dimers,43 EXK model peptide,45 and T4 lysozyme44 (the data for the 44 and 131 sites
`were averaged).
`b Data of Horovitz et al.46
`c Calculated using two-state helix–coil approximation from data of Chakrabatty et al.47 and Doig and Baldwin48 (peptide of 17
`residues with uncharged N-terminus); pH dependencies of the energies were not taken into account: the data are for charged Glu and
`Asp and uncharged His and Cys residues (pH (cid:129) 7). The N-capping energies for Asn, Ser, Thr, and Gly residues in the ‘‘capping box’’
`combination (when Glu residue occupies N3 position in helix) for peptides were 00.58, 00.74, 00.59, and 00.25 kcal/mol, respectively,
`and the optimized N-capping energies for proteins for the Asp, Asn, Ser, and Thr residues were 00.8, 00.7, 00.5, and 00.4 kcal/mol,
`respectively, with any residue occupying the N3 position.
`d Mole fraction based water–cyclohexane transfer energies49 corrected for burial of side-chain analogues in a-helix and coil (as
`described below); for Ser and Thr forming hydrogen bonds with backbone of their own a-helix,50 the transfer energies were corrected
`by 01.5 kcal/mol, the energy of a buried hydrogen bond in proteins.51,52
`e Parameters obtained by minimization of helix boundary deviation with respect to side-chain transfer energies for a set of 30
`proteins. For hydrophilic side chains (His and subsequent residues in the table), the energies could not be defined by this adjustment
`and correspond to the cyclohexane scale.
`
`rately), and /0.5 kcal/mol for His, Lys, and Arg in the
`N1, N2, and N3 positions, close to the 0.6–0.9 kcal/mol
`estimated by mutagenesis.54 DDG sch of Glu, Asp, and
`Gln residues in the capping box 62 N3 position (00.40,
`00.11, and 00.09 kcal/mol, respectively) were calcu-
`lated using as the first approximation the all-or-none two-
`state model from CD data.63
`
`N-Cap and N * and C * Positions. In contrast to the
`middle helix positions, the N-capping energies identified
`by protein engineering are highly variable: 02.2 to 00.4
`kcal/mol for Asn, 02.1 to 00.9 kcal/mol for Ser, and
`00.6 to 0.1 for Gly in the capping box motif.58,64–66 This
`can be explained by ‘‘context-dependent factors’’: unlike
`residues in the a-helix, the replaced N-cap residues have
`
`different main chain w and c angles, 64 and form extra
`hydrogen bonds with sequentially distant residues of the
`protein and with bound water molecules, as observed, for
`example, in substitution sites Thr 59-Glu 62 of T4 lyso-
`zyme 64 and Ser 31-Glu 34 of chymotrypsin inhibitor II.66
`The DDG sch energies for the N-cap position applied here
`(Table I) were calculated using the two-state model from
`CD data of Chakrabatty et al.47 and Doig and Baldwin, 48
`and for the special case of the capping box (Glu residue
`occupies N3 position), from data of Zhou et al.63 for Gly,
`Asn, and Ser residues. These two-state peptide energies
`(00.58, 00.74, and 00.25 kcal/mol, for Asn, Ser, and
`Gly, respectively, with the capping box motif, for exam-
`ple) are close to the lower limits estimated by protein
`engineering.
`
`/ 8K26$$5476
`
`06-17-97 09:16:43
`
`bpa
`
`W: Biopolymers
`
`5476
`
`6
`
`

`
`a-Helical Peptides and Proteins
`
`245
`
`Table II The Replacement Energies (DDGsch, kcal/mol) of Nonpolar Side Chains in the N* Positiona
`
`N* Position Residue
`
`Val
`Ile
`Leu
`Met
`Phed
`
`Intrinsic Contributionb
`00.2
`00.3
`00.4
`00.6
`0.0
`
`Energy of Interaction with
`Leu in N4 Positionc
`00.4
`00.4
`00.6
`00.5
`00.3
`
`Energy of Interaction with
`Leu in N7 Positionc
`
`0.0
`00.1
`00.2
`0.0
`0.0
`
`a The energies, relative to the reference Ala-containing peptide, were calculated using the all-or-none two state model (as described
`below) from u222 ellipticities for a series of model peptides published by Munoz and Serrano.72
`b The intrinsic helix-stabilizing contribution of bulky N* aliphatic residues (Ile, Leu, Val, or Met) was observed even when the N4
`residue is Ala, probably because of hydrophobic interactions between the N* side chain and CbH2(3) groups of N3 and N4 residues,72
`similar to helix-stabilizing interactions of nonpolar N-cap residues.48
`c The same energies were applied when Ile or Met occupied the N4 and N7 positions.
`d The same energies were used for Tyr and Trp residues.
`
`A further helix-stabilizing contribution arises from hy-
`drophobic interactions of flanking N * and C * residues
`with the a-helix.67 The hydrophobic contact between side
`chains in the N * and N4 positions, the ‘‘hydrophobic-
`staple’’ motif, 68 can be detected by nuclear Overhauser
`effects (NOEs) between the side chains in peptide a-
`helices.63,69–71 The corresponding contributions to helix
`energy (Table II) were estimated using the two-state
`model from u 222 ellipticities published by Munoz and
`Serrano 72 for a series of model peptides, and it was as-
`sumed that the ‘‘hydrophobic staple’’ motif can exist only
`in combination with characteristic N-capping residues
`(Ser, Thr, Asn, Asp, or Gly). These energies are smaller
`by 0.1–0.4 kcal/mol than was estimated using the AGA-
`DIR program.72
`A similar hydrophobic interaction at the C-terminus
`of the a-helix between the bulky C * and C4 or C1 side
`chain when the C-cap residue is Gly, 73 the ‘‘Schellman
`motif,’’ is very common for proteins, 61,74 and has also
`been found in crystal structures of a-helical peptides 75
`and detected by nmr spectroscopy for peptides in the
`presence of SDS micelles 76–79 or trifluoroethanol.71,80 The
`Schellman motif is only marginally stable in water: no
`NOEs between the C * and C1 or C4 side chains were
`observed in a model peptide, though analysis of CD spec-
`tra indicated that the interaction does contribute a little
`to a-helix stability.80 The optimum adjustable energy for
`the Schellman motif (considered as a single parameter
`for all combinations of bulky Val, Phe, Tyr, Lys, Arg,
`Leu, Ile, Met, or Trp residues, but excluding contacts of
`two positively charged side chains) was identified as
`00.3 kcal/mol.
`
`Interactions Between Side Chains in the a-He-
`lix. H-Bonding and Electrostatic Interactions of Side
`Chains. The helix-stabilizing interactions of polar side
`chains in water, arising from their hydrogen bonding and
`electrostatic attraction, have been investigated using sev-
`eral model peptides 36–38,81,82 and range up to 00.5 kcal/
`
`mol. Based on the published data, the following estima-
`tions of the interaction energies with completely ionized
`side chains were used in the present work: Glu i–Lys,
`Arg i{4: 00.5 kcal/mol; Asp i–Lys, Arg i{3, Asp i–Arg i/4,
`Gln i–Asn, Asp, Glu i/4; and Glu i–Asn i/4, and Lys i–
`Asp i/4: 00.4 kcal/mol; Asp i–Arg, Lys i/3: 00.3 kcal/
`mol. All these pairs are present in a-helical proteins. Two
`more hydrogen bonding side-chain pairs were taken into
`account with a tentative assigned energy of 00.5 kcal/
`mol: the Glu, Asp i–Trp i/4 pair present as Glu 136-Trp 140
`in colicin (1col, the four-letter codes indicate names of
`Protein Data Bank files 83), and Asp 87–Trp 91 in interleu-
`kin 4 (1rcb; the Trp residue is in the last turn of each of
`the helices and has x 1 (cid:130) 0607), and the Ser i–Gln i/1
`pair that can be present only at the C-terminus of the a-
`helix (as Ser 245–Glu 246 of thermolysin, 2tmn), or imme-
`diately preceding a Pro-induced helix kink (as Ser 215–
`Gln 216 in glucoamylase, 3gly, and Ser 21–Gln 22 in cyto-
`chrome c *, 2ccy).
`The total contribution of the interactions between the
`polar side chains was calculated assuming additivity of
`their pairwise DDG sch
`ij energies:
`
`hb (cid:129) (cid:229)k/m04
`
`DG sch
`
`i(cid:129)k
`
`DDG sch
`ij,hb
`
`(7)
`
`j(cid:129)i/3,4
`
`This additivity approximation means that each side
`chain i interacts simultaneously with all surrounding side
`chains (in i 0 4 and i / 4 positions, for example) rather
`than adopting any fixed orientation in solution. This situa-
`tion can be expected for long flexible Lys, Arg, Glu, and
`Gln side chains, but not for aromatic side chains since
`they have preferred x 1 orientations in the a-helix, as
`discussed below, and therefore do not interact simultane-
`ously and equally well with other side chains in opposite
`i / 4 and i 0 4 directions.
`Hydrophobic Interactions Between ‘‘Rotationally
`Frozen’’ Side Chains. The energies of hydrophobic
`intera

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