review quantum chemical studies of protein...

15
Review Quantum chemical studies of protein structure Eric Oldfield * Department of Chemistry, University of Illinois at Urbana-Champaign, 600 South Mathews Avenue, Urbana, IL 61801, USA Quantum chemical methods now permit the prediction of many spectroscopic observables in proteins and related model systems, in addition to electrostatic properties, which are found to be in excellent accord with those determined from experiment. I discuss the developments over the past decade in these areas, including predictions of nuclear magnetic resonance chemical shifts, chemical shielding tensors, scalar couplings and hyperfine (contact) shifts, the isomer shifts and quadrupole splittings in Mo ¨ssbauer spectroscopy, molecular energies and conformations, as well as a range of electrostatic properties, such as charge densities, the curvatures, Laplacians and Hessians of the charge density, electrostatic potentials, electric field gradients and electrostatic field effects. The availability of structure/spectroscopic correlations from quantum chemistry provides a basis for using numerous spectroscopic observables in determining aspects of protein structure, in determining electrostatic properties which are not readily accessible from experiment, as well as giving additional confidence in the use of these techniques to investigate questions about chemical bonding and chemical reactions. Keywords: quantum mechanics; protein structure; protein electrostatics; nuclear magnetic resonance spectroscopy; Mo ¨ssbauer spectroscopy; chemical shifts 1. INTRODUCTION Quantum chemistry is a branch of chemistry involved with determining, computationally, the three-dimen- sional and electronic structures of molecules. In general terms, the basic problem is to solve the Schro ¨dinger equation, HjZEj, where H is the Hamiltonian operator, j is the wave function and E is the energy. Many spectroscopic properties (such as nuclear mag- netic resonance (NMR) chemical shifts and scalar couplings) are formal derivatives of the energy, whereas others (such as charge densities) are more directly related to j. Methods to evaluate E were first developed in the 1920s by Heisenberg, for the singlet and triplet states of the He atom (Heisenberg 1926), by Heitler & London (1927) for the H 2 molecule and by Pauling (1928a) for the hydrogen molecule ion (H 2 C ). The correct wave functions were found to correspond to sums of functions (resonance) since they yielded the lowest energies, and in the case of, for example, H 2 , this led to stable bond formation. The idea of resonance (hybridization) was then further developed by Pauling to explain the equivalence of the 2s and three 2p orbitals in methane (Pauling 1928b); then in 1932, Pauling described the double-bond character of amide groups, a key feature of protein structure, owing to resonance (Pauling 1932). Of course, at this time, the nature of protein ‘particles’ was quite obscure (Svedberg et al. 1939), but the observation of Bernal & Crowfoot (1934) of diffraction from a pepsin crystal helped to remove misconceptions about their colloidal nature, and giant molecules were then believed to exist. But what were their structures? 2. THE EARLY DAYS In the early 1950s, Pauling again used the theory of quantum mechanical (QM) resonance to develop structural models for the a-helix and parallel and anti-parallel pleated sheets in proteins, in which the planar nature of the peptide bond (and the electrostatic interactions between these structures) were essential ingredients (Pauling & Corey 1950, 1951), but still there were no high-resolution structures of proteins. Shortly after these proposals were published, Bernal visited Madras and the interests of a young G. N. Ramachandran were shifted from crystal physics to protein structure determination (Ramakrishnan 2001). With G. Kartha, a proposal for the structure of collagen, a triple helix, was published (Ramachandran & Kartha 1954, 1955), the details of which were objected to on the basis that there were too many steric hindrances (Rich & Crick 1955). This led Sasisekharan and Ramachandran to investigate this topic in more detail. With knowledge of Pauling’s postulate of the planar nature of the peptide bond, due to QM resonance, Ramachandran began the first compu- tational chemical studies of protein structure and stability—albeit armed with only a desk calculator (Ramakrishnan 2001). The peptide models used contained just two amino acids, but given the trans nature of the central peptide bond, could be well described by just two torsion angles, f i and j iK1 . All atoms were described as rigid spheres with radii Phil. Trans. R. Soc. B (2005) 360, 1347–1361 doi:10.1098/rstb.2003.1421 Published online 22 July 2004 * ([email protected]). Received 1 April 2003 Accepted 24 September 2003 1347 q 2004 The Royal Society

Upload: others

Post on 04-Oct-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Review Quantum chemical studies of protein structurefeh.scs.uiuc.edu/doc/papers/1201228224_278.pdfgroups, a key feature of protein structure, owing to resonance (Pauling 1932). Of

Phil. Trans. R. Soc. B (2005) 360, 1347–1361

doi:10.1098/rstb.2003.1421

Published online 22 July 2004

Review

Quantum chemical studies of protein structure

Eric Oldfield*

* (eo@c

ReceivedAccepted

Department of Chemistry, University of Illinois at Urbana-Champaign, 600 South Mathews Avenue,Urbana, IL 61801, USA

Quantum chemical methods now permit the prediction of many spectroscopic observables inproteins and related model systems, in addition to electrostatic properties, which are found to be inexcellent accord with those determined from experiment. I discuss the developments over the pastdecade in these areas, including predictions of nuclear magnetic resonance chemical shifts, chemicalshielding tensors, scalar couplings and hyperfine (contact) shifts, the isomer shifts and quadrupolesplittings in Mossbauer spectroscopy, molecular energies and conformations, as well as a range ofelectrostatic properties, such as charge densities, the curvatures, Laplacians and Hessians of thecharge density, electrostatic potentials, electric field gradients and electrostatic field effects. Theavailability of structure/spectroscopic correlations from quantum chemistry provides a basis for usingnumerous spectroscopic observables in determining aspects of protein structure, in determiningelectrostatic properties which are not readily accessible from experiment, as well as giving additionalconfidence in the use of these techniques to investigate questions about chemical bonding andchemical reactions.

Keywords: quantum mechanics; protein structure; protein electrostatics;nuclear magnetic resonance spectroscopy; Mossbauer spectroscopy; chemical shifts

1. INTRODUCTIONQuantum chemistry is a branch of chemistry involved

with determining, computationally, the three-dimen-

sional and electronic structures of molecules. In general

terms, the basic problem is to solve the Schrodinger

equation, HjZEj, where H is the Hamiltonian

operator, j is the wave function and E is the energy.

Many spectroscopic properties (such as nuclear mag-

netic resonance (NMR) chemical shifts and scalar

couplings) are formal derivatives of the energy, whereas

others (such as charge densities) are more directly

related to j. Methods to evaluate E were first developed

in the 1920s by Heisenberg, for the singlet and triplet

states of the He atom (Heisenberg 1926), by Heitler &

London (1927) for the H2 molecule and by Pauling

(1928a) for the hydrogen molecule ion (H2C). The

correct wave functions were found to correspond to

sums of functions (resonance) since they yielded the

lowest energies, and in the case of, for example, H2, this

led to stable bond formation. The idea of resonance

(hybridization) was then further developed by Pauling

to explain the equivalence of the 2s and three 2p

orbitals in methane (Pauling 1928b); then in 1932,

Pauling described the double-bond character of amide

groups, a key feature of protein structure, owing to

resonance (Pauling 1932). Of course, at this time, the

nature of protein ‘particles’ was quite obscure

(Svedberg et al. 1939), but the observation of Bernal &

Crowfoot (1934) of diffraction from a pepsin crystal

helped to remove misconceptions about their colloidal

had.scs.uiuc.edu).

1 April 200324 September 2003

1347

nature, and giant molecules were then believed to exist.But what were their structures?

2. THE EARLY DAYSIn the early 1950s, Pauling again used the theory ofquantum mechanical (QM) resonance to developstructural models for the a-helix and parallel andanti-parallel pleated sheets in proteins, in whichthe planar nature of the peptide bond (and theelectrostatic interactions between these structures)were essential ingredients (Pauling & Corey 1950,1951), but still there were no high-resolution structuresof proteins.

Shortly after these proposals were published, Bernalvisited Madras and the interests of a young G. N.Ramachandran were shifted from crystal physics toprotein structure determination (Ramakrishnan 2001).With G. Kartha, a proposal for the structure of collagen,a triple helix, was published (Ramachandran &Kartha 1954, 1955), the details of which wereobjected to on the basis that there were too many sterichindrances (Rich & Crick 1955). This led Sasisekharanand Ramachandran to investigate this topic in moredetail. With knowledge of Pauling’s postulate of theplanar nature of the peptide bond, due to QMresonance, Ramachandran began the first compu-tational chemical studies of protein structure andstability—albeit armed with only a desk calculator(Ramakrishnan 2001). The peptide models usedcontained just two amino acids, but given the transnature of the central peptide bond, could be welldescribed by just two torsion angles, fi and jiK1. Allatoms were described as rigid spheres with radii

q 2004 The Royal Society

Page 2: Review Quantum chemical studies of protein structurefeh.scs.uiuc.edu/doc/papers/1201228224_278.pdfgroups, a key feature of protein structure, owing to resonance (Pauling 1932). Of

1348 E. Oldfield Protein quantum chemistry

corresponding to their van der Waals radii and onlyhardcore repulsive interactions were considered.Nevertheless, the Ramachandran maps thus obtainedclearly indicated regions of stability in f, j space(Ramachandran et al. 1963) and successfully predictedthat two non-helical residues in myoglobin must beglycines, which proved to be the case (Ramakrishnan2001). With further improvements in the developmentof potential energy functions, such as the Buckinghamand Lennard–Jones potentials, the accuracy of suchRamachandran maps improved, although these poten-tials were still of an empirical nature.

In these early days, it was not possible to carry outelectronic QM structure calculations on molecules aslarge as peptides or even amino acids. The speed of anelectronic structure calculation strongly depends on thenumber of atoms in the system, or more precisely, thenumber of functions, basis functions, which are used todescribe the atomic orbitals that make up the molecularorbitals in the molecule. For Hartree–Fock self-consistent field (SCF) calculations, the length of acalculation varies as ca N3.5, where N is the number ofbasis functions, while for higher-level calculations thelength of a calculation can be as high as ca N 5–N 6.However, as computer speeds have increased, suchcalculations are now becoming tractable. In addition,recent developments in an alternative approach toelectronic structure calculations, density functionaltheory (DFT), are somewhat faster, scaling as ca N3

(Kohn et al. 1996). Armed with these tools, it is now arelatively straightforward matter to evaluate Rama-chandran maps for energies using first principlesmethods, and as discussed below, many other proper-ties of interest can now also be predicted. The results ofenergy calculations are particularly important sincethey can then be used to parameterize semi-empirical(molecular mechanics (MM)) approaches to structureprediction and refinement, and such derived‘force fields’ are now commonly used in X-raycrystallographic and NMR structural investigations.Moreover, such MM methods can be combinedwith QM methods to investigate very large systems,such as enzymes (Gogonea et al. 2001; Gao & Truhlar2002). But just how large a system can be investigatedby using purely quantum chemical methods? And whatcan we learn? Is this approach important? Cannot wenow just measure everything of interest?

3. THE ENERGYEvaluating the total eigenenergy of a molecule is thefirst critically important step to evaluating a wide rangeof properties, many of which (such as the chargedensity r(r), the charge density at the nucleus r(0) andthe electrostatic potential F(r)) cannot be readilydetermined experimentally. Likewise, many spectro-scopic properties, such as NMR chemical shifts, NMRhyperfine shifts, Mossbauer isomer shifts and quadru-pole splittings, quadrupole couplings in NMR andnuclear quadrupole resonance (NQR), can be readilymeasured, but without quantum chemistry they cannotbe interpreted in much detail. In addition, of course,there are no direct experimental measurements ofmolecular orbitals or electronic structure; rather, they

Phil. Trans. R. Soc. B (2004)

have to be deduced from an electronic wave functionwhose quality has to be assessed by its ability tosuccessfully predict a series of observables. By way ofexample, I consider first the energy of a proteinmolecule and how this is related to observablestructural properties.

In its simplest embodiment, the energy of a moleculehaving a given structure can be computed by usingvariational or SCF techniques in which coefficientsdescribing the composition of various atomic orbitalsare varied—the linear combination of atomic orbitalsmethod, such that the lowest eigenenergy for thesystem is obtained (Hehre et al. 1986). From such anelectronic structure calculation, it is then possible todeduce properties of interest, such as those outlinedabove. In most cases, it is not necessary to carry out acalculation on an entire protein, since the property ofinterest is essentially a local one, such as an NMRchemical shift. In other cases, however, this is not thecase. The most important and most obvious example ofthis exception is that of the structure of the proteinitself.

In recent work, it has been shown to be possible tocarry out electronic structure calculations on entireproteins by using ab initio techniques. The results ofboth single-point (single-geometry) energy calculationson the proteins endothelin, charybdotoxin andp53, and of ab initio Hartree–Fock geometry optimiz-ation calculations on crambin, a small hydrophobicprotein of 46 residues, have recently been reported(Challacombe & Schwegler 1997; Van Alsenoy et al.1998). The crambin study used a multiplicativeintegral approximation method, which when combinedwith the direct SCF approach results in linear scalingwith system size in building the Fock matrix, enablingthe investigation of very large systems in which manybody interactions are taken into account in a directfashion (Van Alsenoy et al. 1998). This approach isexpected to be more accurate than using force fieldsbased on the parameterization of two-body inter-actions, and should lead to direct evaluation of manyelectrostatic and energy derivative properties, whichcan be evaluated from the computed wave functions. Inthe case of crambin, 79 cycles of geometry optimizationwere carried out, and the resulting structure was foundto be remarkably close to that obtained crystallogra-phically, as shown in figure 1a. The results of thesetheoretical calculations revealed basically the samedeviations from planarity of the peptide groups as thoseseen crystallographically, even in the absence ofcorrelation (instantaneous electron–electron) effects,and the N–C(a)–C 0 three atom backbone bound angleswere also found to be highly correlated with thecrystallographic values—corresponding to only a 1.58r.m.s. deviation, as shown in figure 1b. Based on theseresults, it seems likely that quantum chemical geometryoptimization will play a significant role in structurerefinement in NMR and in crystallography in the nottoo distant future.

4. ELECTROSTATIC PROPERTIESEnergy is not a measurable quantity, although energydifferences and derivative properties are. Likewise,

Page 3: Review Quantum chemical studies of protein structurefeh.scs.uiuc.edu/doc/papers/1201228224_278.pdfgroups, a key feature of protein structure, owing to resonance (Pauling 1932). Of

120

118

116

114

112

110

108

106

1040 5 10 15 20 25 30 35 40 45

residue

N-C

a-C

(a) (b)

Figure 1. Comparison between experimental and ab initio computational results on crambin. (a) Superposition of thecrystallographic (pdblcnr) and HF/4-21G optimized heavy-atom framework of crambin. (b) Comparison of crystallographic(solid line; pdblcnr) and HF4-21G optimized N–C(a)–C 0 angles (dotted line) in crambin. Reprinted in part with permissionfrom Van Alsenoy et al. (1998). Copyright 1998 American Chemical Society.

Protein quantum chemistry E. Oldfield 1349

electrostatic properties, such as charges (or morespecifically, charge densities), electrostatic potentials,electrostatic fields and electric field gradients (EFGs)are not direct observables, although many such proper-ties can be deduced from experimental (NMR, diffrac-tion) results, at least for small molecule systems. Ifthese electrostatic properties can be successfully pre-dicted for systems such as amino acids and peptides, itseems reasonable to suppose that similar calculationson proteins will also yield accurate results.

The electric field, E, is a vector property and isdefined as

E ZKVf;

where f is the electrostatic potential, and from Gauss’sLaw, the divergence of the vector E is given by thescalar r, the charge density, as

V,E Zr

30

;

where 30 is the permittivity of free space, and it followsthat

ðV,VfÞZV2fZKr

30

;

where V2f is the Laplacian of the potential. The EFG,V, is defined as

V ZVE

and the gradient of the vector is a tensor of rank two. Toevaluate the electric potential, f, more commonlycalled the molecular electrostatic potential, an SCFcalculation is first carried out to determine the wavefunction j, with the electron probability density or thecharge density then being evaluated from jjj2. Themolecular electrostatic potential at a point can then beobtained by integrating over the charge density. Oncethe potential is known, electric fields (KVf) and fieldgradient tensors (VE) can be computed. With this briefintroduction, we next ask the questions: are thesemethods applicable to amino acids, peptides andproteins? Are the methods accurate? And what canwe learn about protein structure?

Phil. Trans. R. Soc. B (2004)

At present, there have been few reports of ab initiocalculations of r(r) or F(r) values in complete proteins.But for smaller systems (amino acids, peptides andsome protein prosthetic groups), numerous reports ofsuch calculations have been published. But how doesone know that these calculated results are accurate?The answer of course is that in each case the computedproperties need to be rigorously compared with theresults of experiments in which these properties areeither directly measured, or where the properties ofinterest are related in a known way to the experimentalproperty.

Arguably, the simplest property to extract from acomputed wave function is the charge density, r(r),since this is proportional to jjj2, the probability offinding an electron at a given position. This approachcan be used to compute three-dimensional plots ofcharge density isosurfaces, which can then becompared with r(r) surfaces extracted from X-raydiffraction data, since the scattering of an X-ray beam isa function of the electron distribution. Likewise,the electrostatic potential can be evaluated from thequantum chemical results by integrating overthe charge density

f1 Z

ðr2 dV2

4p30r12

;

where r2 is the charge density, dV2 is the volumeelement at point 2, and r12 is the distance betweenpoints 1 and 2, resulting in electrostatic potentialsurfaces F(r), which can again be compared withresults extracted from diffraction data as shown, forexample, in figure 2a–c, for L-asparagine$H2O. How-ever, these graphical comparisons are not satisfactoryfor making rigorous comparisons between computedand experimentally derived properties. One solution tothis problem is to use the atoms-in-molecules (AIM)theory method developed by Bader (1990). Thismethod uses the idea that associated with everychemical bond is a point, rb, at which the firstderivative of the charge density, r(r), is zero. At thisso-called bond critical point (BCP) there are three

Page 4: Review Quantum chemical studies of protein structurefeh.scs.uiuc.edu/doc/papers/1201228224_278.pdfgroups, a key feature of protein structure, owing to resonance (Pauling 1932). Of

30

20

10

0

–10

–20

–303020100–10–20–30

6

4

2

0

–26420–2

diffraction-derivedicosahedral Hessian tensors (eÅ–5)

diffraction-derivedicosahedral Hessian tensors (eÅ–5)

DFT

icos

ahed

ral H

essi

an te

nsor

s (e

Å–5

)

DFT

icos

ahed

ral H

essi

an te

nsor

s (e

Å–5

)

–0.07 +0.8

(d )

(a) (b) (c)

Φ(r)/au(e) ( f )

Figure 2. Electrostatics results for asparagine and charybdotoxin. (a–c) Molecular electrostatic potentials for asparagine, F(r),mapped from K0.07 to C0.8 e/4pe0a0, onto a 0.11 eAK3 isosurface of the electron density. (1 AZ1!10K10 m.) (a) X-raydiffraction. (b) HF/6-311CCG(3df,3pd) calculation. (c) B3LYP/6-311CCG(3df,3pd) calculation. (d ) Experimental (X-ray)versus theoretical icosahedral Hessian-of-r(rb) tensors for asparagine at BCPs. (e) As (d ) but at H-BCPs. ( f ) Electrostaticpotential surface for charybdotoxin. The light surface corresponds to a value of 0.0035 a.u., the dark surface K0.0035 a.u.Reprinted from Challacombe & Schwegler (1997) with permission. Reprinted in part with permission from Arnold et al.(2000a). Copyright 2000 American Chemical Society.

1350 E. Oldfield Protein quantum chemistry

non-zero principal curvatures in r(rb), two negativeand one positive. The curvature in the charge density ata BCP is described by a real symmetric tensor known asthe Hessian-of-r(r), H, which has nine elements of theform v2r/vrirj. When diagonalized, H is expressed bythree elements, l1–3, which correspond to curvatures inr(rb) along three principal axes. The Laplacian of thecharge or electron density at a BCP, V2r(rb), is simplythe sum of these three principal curvatures. In recentwork, it has been found that there is excellentagreement between li and V2r(r) values determinedvia quantum chemistry and those derived from X-raydiffraction data, for a series of amino acids and peptides(Dittrich et al. 2000; Arnold et al. 2000a; Li et al.2002). However, the principal curvatures in r(r)represent only the magnitudes of the Hessian tensors,H, not their orientations. Tensors are commonlyexpressed in a principal axis system (PAS) as threeprincipal values and the corresponding directioncosines with respect to a molecular reference frame.

Phil. Trans. R. Soc. B (2004)

Thus, to compare the orientations of such tensors,one must compare the angles between each PAS andthe molecular reference frame, and experimental errorsin individual Cartesian tensor elements must beindirectly translated into errors in degrees or radians.A comparison of tensors in an icosahedral represen-tation, however, is much more convenient, since boththe magnitudes and the orientations of the tensors arequantitatively evaluated at the same time by comparingthe six icosahedral tensor elements, c1–6. Also, the sixicosahedral elements are equally weighted in anycoordinate frame, making this representation ideal forfitting a least-squares line through theoretical-versus-experimental tensor data, an approach developedpreviously by Alderman and colleagues, for investi-gating chemical shielding tensors (Alderman et al.1993). Using this approach, there has been shown to bevery good agreement between theory and experimentfor both the magnitudes and the orientations of theseHessian tensors. For example, again investigating

Page 5: Review Quantum chemical studies of protein structurefeh.scs.uiuc.edu/doc/papers/1201228224_278.pdfgroups, a key feature of protein structure, owing to resonance (Pauling 1932). Of

Protein quantum chemistry E. Oldfield 1351

L-asparagine$H2O, a slope of 1.09 and R2Z0.96 for acomparison between the Hessian-of-r(r) tensors in theicosahedral representation at all heavy atom BCPs hasbeen demonstrated (Arnold et al. 2000a; figure 2d ).Even better accord is found at hydrogen bond BCPs(Arnold et al. 2000a; figure 2e). These results show thatr(r), li, V

2r(r) and Hessian-of-r(r) properties can nowall be accurately predicted using quantum chemicaltechniques. This helps validate their use on proteins,such as p53 and charybdotoxin (Challacombe &Schwegler 1997), and an interesting example of thisapproach is shown in figure 2f for F(r) for charybdo-toxin in which the positive F(r) surface thought to beresponsible for binding to the KC-channel can clearlybe seen (on top, light grey).

5. MOSSBAUER SPECTROSCOPY ANDELECTROSTATICSHow else might the results of quantum chemicalcalculations of charge densities be verified? Onepossibility is to investigate spectroscopic observablesthat are directly related to the charge density. The so-called isomer shift in Mossbauer spectroscopy is onesuch example. Mossbauer spectroscopy is a techniquewidely used to investigate the structures of metallopro-teins, metalloporphyrins and other model systems, andis a potentially powerful tool with which to deduce bothgeometric and electronic structures. 57Fe Mossbauerspectra have frequently been studied in proteins,and are typically dominated by two interactions: thequadrupole splitting, DEQ, which arises from the non-spherical nuclear charge distribution in theI*Z3/2 excited state in the presence of an EFG atthe 57Fe nucleus; and an isomer (or chemical) shift, dFe,which arises from differences in the charge density atthe nucleus between the absorber (A, the molecule orsystem of interest) and a reference compound (usuallya-Fe at 300 K). This interaction is given by

dFe ZEA KEFe

Z2p

3Z e2ðhR2i* K hR2iÞðjjð0Þj2A K jjð0Þj2FeÞ; (5.1)

where Z represents the atomic number of the nucleus ofinterest (iron) and R, R* are average nuclear radii of theground and excited states of 57Fe. Since jjð0Þj2Fe is aconstant, the isomer shift (from Fe) can be written as

dFe Za½rtotð0ÞKc; (5.2)

where a is the so-called calibration constant and rtot(0)is the computed charge density at the nucleus. In recentwork, quantum chemical methods have been used tocompute r(0) charge density values for a wide variety ofhaem proteins and model systems, and it has beenfound that the R2 values between experimental andcomputed dFe values are typically greater than 0.98,and r.m.s. errors between experimental and theoreticalisomer shifts are only ca 3–4% (Zhang et al. 2002a).These results are interesting since they mean that evencharge densities at the nucleus can now be quite welldescribed by using quantum chemistry. Moreover, theMossbauer isomer shift can begin to be used in

Phil. Trans. R. Soc. B (2004)

structure determination, since it is found to be highlysensitive to local geometry.

The other Mossbauer observable is the quadrupolesplitting, DEQ, which is related to the components ofthe EFG tensor (Vxx, Vyy, Vzz) at the nucleus, anotherelectrostatic property, as follows:

DEQ Z1

2eQVzz 1C

h2

3

1=2

;

where e is the electron charge, Q is the quadrupolemoment of the nuclear excited state, and the principalcomponents of the EFG tensor are labelled accordingto the convention: jVzzjO jVyyjO jVxxj with the asym-metry parameter, h, being given by

hZVxx KVyy

Vzz

:

That is, the Mossbauer quadrupole splitting is anotherdirect spectroscopic probe of electrostatics, since it isdirectly related to the gradient of the electric fieldvector, E. Once again, the availability of a good wavefunction enables an immediate evaluation of all ninecomponents of the EFG tensor, together with theirorientations in a molecular axis system (Zhang et al.2002b). The tensor is necessarily symmetric andtraceless (Laplace equation) so only three elements(and their orientations) are needed to fully describe thetensor. As might now be anticipated, there is found tobe excellent accord between theory and experiment forDEQ in a wide range of haem proteins and modelsystems when using an appropriate level of theory andlarge basis sets (typically, ca 500–1000 basis functions,for a haem model). For example, a theory versusexperiment correlation having an R2 of ca 0.98, a slopeof ca 1.12 and an error of ca 0.26 mm sK1, whenconsidering 23 different systems covering a5.63 mm sK1 range in DEQ is obtained, correspondingto a 4–5% overall error (Zhang et al. 2002b). Thus, 57FeMossbauer quadrupole splittings in proteins andrelated systems can now all be predicted quiteaccurately and provide another important test ofquantum chemical methods in predicting ‘electrostatic’properties.

6. EFGs AND NMR SPECTROSCOPYIt should also be noted that the EFG parameter is notrestricted to Mossbauer spectroscopy. All nuclei innon-spherically symmetric environments will have anassociated EFG, and if the nucleus of interest has anon-zero quadrupole moment, then there will be aninteraction, of strength (eq)(eQ)/h, where eqZVzz andeQ is the quadrupole moment of the nucleus. Inmagnetic resonance spectroscopy, if e2qQ/h is smallcompared with the nuclear Zeeman effect, there aretypically perturbations of the NMR line shape, fromwhich EFG tensor information can be determined,while for large e2qQ/h values, transitions betweenquadrupolar levels can often be observed directly, bymeans of pure NQR. These nuclear quadrupoleinteractions are generally difficult to measure inproteins, although they have been determined for2H (LiWang & Bax 1997) and 17O (Park et al. 1991),

Page 6: Review Quantum chemical studies of protein structurefeh.scs.uiuc.edu/doc/papers/1201228224_278.pdfgroups, a key feature of protein structure, owing to resonance (Pauling 1932). Of

1352 E. Oldfield Protein quantum chemistry

by NMR. But in amino acid or peptide crystals, bothtensor magnitudes and orientations have beenmeasured experimentally and provide another import-ant test of the accuracy of quantum chemicalcalculations. As an example, the 14N EFG tensorsin L-asparagine have been determined by NMR on asingle crystal sample (Naito & McDowell 1984). Suchsingle crystal studies yield both the magnitudes andthe orientations of the 14N EFG for both the amideand ammonium nitrogens, and both properties havebeen calculated by using DFT (Arnold et al. 2000a).Since the EFG is a tensor of rank two, the threeprincipal EFG tensor components for each site,together with their orientations (direction cosines),can be combined using the icosahedral representationto yield six icosahedral tensor elements for each site:a total of 12 observables. When the experimentalresults were compared with the results of largebasis DFT calculations, they were found to behighly correlated (R2Z0.94) with a slope of 1.14(Arnold et al. 2000a). Thus, both Mossbauer andNMR EFG observables can now be accuratelypredicted by using quantum chemistry, whichsuggests that in the future these properties may beused in protein structure determination, since againthey are very sensitive to local geometry, as describedbelow in more detail.

7. SPIN DENSITIES AND HYPERFINE SHIFTSIn what we have discussed so far, we have not explicitlyconsidered the case of paramagnetic species—molecules containing unpaired electron spins. Forparamagnetic systems, quantum chemical methodstypically describe these molecules as having what areessentially two subsets of interacting electrons andnuclei, each set having electrons with opposite spins(Hehre et al. 1986). The molecular orbitals obtainedhave either spin up (a) or spin down (b) electrons andconsequently two different sets of charge densities, r(a)and r(b), are obtained. Their sum gives the overallcharge density r(r) discussed above, but their differ-ence gives a new property, the so-called spin density

rab ZX

½jjai ð0Þj

2 K jjbi ð0Þj

2:

The spin density is an important property since, inNMR, the Fermi contact or hyperfine shift dFC isdirectly proportional to rab:

dFC Zm0m

2bg

2e ðSC1Þ

9kT,rab;

where m0 and mb are the nuclear and electron moments,ge is the Lande factor and S is the spin quantumnumber. The Fermi contact interaction is also ofimportance in mediating scalar or J-couplings inNMR, as well as hyperfine couplings in electronparamagnetic resonance (EPR). The Fermi contactshift in NMR can be very large—up to ca 6000 p.p.m.for 13C, and represents another major challenge fortheory to reproduce. In early work, 2H and 15N shifts ofamino acid residues bound to a FeS4 cluster in theprotein rubredoxin, covering a ca 800 p.p.m. range of‘chemical’ shifts, were quite accurately reproduced by

Phil. Trans. R. Soc. B (2004)

using DFT methods (Wilkens et al. 1998). Morerecently, the entire ca 6000 p.p.m. range of hyperfine(contact) shifts seen in a variety of haem proteins andmodel systems, having SZ1/2, 2 and 5/2, has beensuccessfully reproduced (Mao et al. 2002). The R2

values were again ca 0.99 (NZ39 compounds;p!0.0001) with a standard deviation of ca 100p.p.m. over the 6000 p.p.m. total shift range, asshown in figure 3a. These results give considerableconfidence in the use of quantum chemical methods topredict spectroscopic observables, even in paramag-netic proteins, and make it possible to begin to useNMR contact shifts in protein structure determination.In addition, they give added confidence in the use of theMO results to describe details of structure andbonding. For example, the spin density in a ferricmyoglobin model is found to have the distributionshown in figure 3b, in sharp contrast to that found inthe more unusual Fe(III) species, shown in figure 3c(Mao et al. 2002), while in an NO–haem system theunpaired electron is localized primarily in a dz

2 orbital(figure 3d). These representations also help visualizethe associated contact shifts, which depend on rab.

8. DERIVATIVE PROPERTIESSo far, discussion has focused on what might con-veniently be thought of as electrostatic properties: r(r),li, V

2r(r), the Hessian-of-r(r), F(r), VE, DEQ, dFe andrab, since they are all related to jjj2 or the chargedensity. There are, however, a second set of propertiesthat are, formally, energy derivatives:

Jij Zv2E

vmi vmj

; the Jcoupling tensor;

sij Zv2E

vmi vBj

; the NMR shielding tensor;

cij Zv2E

vBi vBj

; the susceptibility tensor;

Aijk Zv3E

vmi vBj vFk

; the dipole shielding polarizability;

where m is a nuclear magnetic moment, B is an externalmagnetic field and F is an external electric fieldcomponent. These, and other higher-order terms(hyper-polarizabilities), can all be evaluated by usingquantum chemical methods and in most cases com-mercial software packages are available to enable thenecessary calculations. Again, the question arises as tohow accurate the results are. Also, can we apply thesemethods to proteins—which at present, for practicalpurposes, means: are the interactions of short enoughrange that small molecular fragments can be used tocalculate the observables that are measuredexperimentally?

9. SCALAR OR J-COUPLINGS IN NMRSPECTROSCOPYThe scalar or J-coupling property is perhaps the mostwell known since it has been investigated since the

Page 7: Review Quantum chemical studies of protein structurefeh.scs.uiuc.edu/doc/papers/1201228224_278.pdfgroups, a key feature of protein structure, owing to resonance (Pauling 1932). Of

Protein quantum chemistry E. Oldfield 1353

1960s, where J-coupling calculations were pioneeredby Karplus (1996). With the development oftwo-dimensional NMR techniques in the 1980s, alarge body of experimental data on J-couplings inproteins became available, and this prompted the use ofquantum chemical methods to predict J-couplings inproteins. These calculations involve the evaluation ofdiamagnetic and paramagnetic spin–orbit contri-butions to J, in addition to paramagnetic Fermi contactand spin–dipole contributions (Helgaker et al. 2000).The use of DFT methods with hybrid exchangecorrelation functionals has met with considerablesuccess, and it has been possible to accuratelyreproduce all six 3J couplings, which involve the j

backbone torsion angle. Moreover, there was animprovement in the quality of the predictions whenthe effects of molecular motion were taken intoaccount, by using ‘snapshots’ from a moleculardynamics trajectory, on the protein ubiquitin (Caseet al. 2000).

In addition to these ‘classical’ J couplings, associatedwith protein backbone torsion angles, there has alsorecently been considerable interest in using quantumchemical methods to investigate hydrogen-bond scalarcouplings (Cornilescu et al. 1999), focusing onanswering the question: what is the nature of suchhydrogen bond scalar couplings? Do they representcovalent bonding? One way to investigate these ques-tions is to use AIM theory and to investigate thetopology of r(r) at the hydrogen BCP. The topology ofr(r) at a BCP is described by the real, symmetric,second-rank Hessian-of-r(r) tensor, and the trace ofthis tensor is related to the bond interaction energy by alocal expression of the virial theorem (Bader 1990):

TrðHessianÞZV2rðrÞZ ½2GðrÞCV ðrÞð4m=hK2Þ;

where G(r) is the kinetic energy density and V(r) is thepotential energy density (Bader 1990). Since G(r) mustalways be positive and V(r) must always be negative instable, bound, stationary states, the sign of V2r(r) at aBCP is determined by which energy density is in excessover the virial average of 2 : 1 kinetic-to-potentialenergy. A negative Laplacian reveals excess potentialenergy at the BCP, meaning that electronic charge isconcentrated into a bond. This is the case in all shared-electron (covalent) interactions. A positive BCPLaplacian reflects an excess of kinetic energy in abond, and a relative depletion of electronic chargealong a bond path. This is the case in all closed-shell(electrostatic) interactions. For every backbone hydro-gen bond examined, V2r(r) was found to be positiveand characteristic of a closed-shell or electrostaticinteraction (Arnold & Oldfield 2000). Although eachterm increased exponentially as j3hJNC 0j increased, thekinetic term was a slightly stronger exponential, andoutpaced the increase in the potential term. That is, thenet amount of excess kinetic energy density wasgreatest for the largest couplings. This could not bethe case if such couplings were dependent upon, orindicative of, covalent character in the bonds.

Although the kinetic energy term provides thedominant virial contribution, resulting in a netclosed-shell, electrostatic interaction, it is also

Phil. Trans. R. Soc. B (2004)

necessary to consider whether these interactionsrepresent a closed-shell limit by evaluating the totalenergy density, H(r), at the BCP (Cremer & Kraka1984)

HðrÞZGðrÞCV ðrÞ:

Bonds with any degree of covalent character (anyamount of potential energy stabilization resulting fromthe accumulation of charge in the internuclear region)must have a BCP H(r) that is less than zero. Quantumchemical results for the PGB1 protein backbonehydrogen bonds provided no evidence of partialcovalent character, H(r)O0 in all cases (Arnold &Oldfield 2000). Backbone hydrogen bonds thus appearto represent purely closed-shell, electrostaticinteractions.

If 3hJNC 0, does not arise from a partial covalentinteraction, then what permits the intermolecularcommunication between nuclear spins? The theoreticalresults indicated that the magnitude of 3hJNC 0, wasclearly related to the mutual penetration of non-bonding van der Waals shells; however, there was noconcentration of charge in the bonding region uponhydrogen bond formation. Thus, the electron densityin the hydrogen bond is simply the sum of isolateddonor and acceptor densities: the greater the van derWaals penetration, the greater the resulting summedelectron density and the greater the experimentallyobserved J-coupling. This description of 3hJNC 0 alsoprovides a simple physical explanation for the existenceof scalar couplings between nuclei that cannot possiblybe covalently bonded to one another. For example,Kimber et al. (1978) observed a field-independent19F– 19F coupling of 17G2 Hz in a (dihydrofolatereductase) DHFR–NADPH–methotrexate complexwhich had been prepared biosynthetically with 6-fluorotryptophan residues. With no crystal structureof DHFR available, these researchers postulated thattwo out of the four 6-fluorotryptophan residues mustbe situated such that their 19F nuclei are ca 3 A apart,based upon much earlier empirical observations of‘through space’ 19F– 19F J-couplings in small organicmolecules (Kimber et al. 1978). Subsequently, it wasfound that the folded protein brings the 6-positions ofTrp5 and Trp133 into extremely close proximity(figure 4a), and it was shown that sum-over-states(SOS)–DFT (Malkin et al. 1994) predicts a 19F– 19Fcoupling of ca 30 Hz between two (model) fluoro-methane molecules at this separation (Arnold et al.2000b). An AIM atomic interaction line was shown toexist between the two 19F nuclei, and the non-bondingvan der Waals shells of the monomers penetrate oneanother by DrZ0.47A along this line. Long-range(O3JFF), intramolecular 19F– 19F through-spaceJ-couplings in a wide range of systems are also welldescribed by SOS–DFT calculations, and small non-bonded dimer models such as (CH3F)2, give verysimilar results to covalently bonded models (Arnoldet al. 2000b). The extent of the van der Waalspenetration in several more fluoromethane dimers,whose F–C geometries were extracted from largerorganic molecules possessing such long-range JFF

couplings was then investigated, since it is well-established that the magnitudes of both JFF and

Page 8: Review Quantum chemical studies of protein structurefeh.scs.uiuc.edu/doc/papers/1201228224_278.pdfgroups, a key feature of protein structure, owing to resonance (Pauling 1932). Of

(6–19F)Trp5

(6–19F)Trp133

(a)200

150

100

50

00.2 0.4 0.6 0.8 1.0 1.2 0.4 0.6 0.8 1.0 1.2

6

4

2

0

–2

|JFF

| (H

z)

ln(|J

|/H

z)

mutual penetration of non-bondingvan der Waals shells (Å)

mutual penetration of non-bondingvan der Waals shells (Å)

(b) (c)

Figure 4. Through space J-couplings. (a) Structure of a [6-19F]Trp-labelled dihydrofolate reductase complex showing closeapproach between Trp6 and Trp133. (b) The exponential dependence of the magnitude of through-space 19F– 19F J-couplingsupon the mutual penetration of non-bonding van der Waals shells, S Dr. jJFFjZ2:6 exp½3:67 S Dr, R2Z0.94. (c) ln(jJj) versusS Dr. Filled circles, JFF; open circles, 3hJNC 0. Reprinted in part with permission from Arnold & Oldfield (2000) and Arnold et al.(2000b). Copyright 2000 American Chemical Society.

1000

0

–1000

–2000

–3000

–4000

–50001000–1000–2000–3000–4000–5000 0

(a)

experimental hyperfine shifts (p.p.m.)

Fe(TPP)(13CN)2

Mb13CN

(c)

(b) (d )

pred

icte

d hy

perf

ine

shif

ts (

p.p.

m.)

Figure 3. Spin density and hyperfine shifts in heme model systems and proteins. (a) Graph showing experimental NMRhyperfine shifts over the entire ca 6000 p.p.m. range seen in proteins and model systems plotted as a function of the hyperfineshifts computed by using the DFT/B3LYP method. Open squares represent the highly shifted resonances in a 13CN-labelledmetalloporphyrin and [13CN]K sperm whale cyanometmyoglobin. The line has a slope of 0.99 and an R2 value of 0.990. (b–d )Total spin density isosurfaces for: (b) [Fe(TMP)(1-MeIm)2]C; (c) [Fe(OEP)(t-BuNC)2]C; and (d ) Fe(OEP)(NO). Reprintedin part with permission from Mao et al. (2002) (a) and Zhang et al. (2002b) (b–d ). Copyright 2002 American Chemical Society.

1354 E. Oldfield Protein quantum chemistry

Phil. Trans. R. Soc. B (2004)

Page 9: Review Quantum chemical studies of protein structurefeh.scs.uiuc.edu/doc/papers/1201228224_278.pdfgroups, a key feature of protein structure, owing to resonance (Pauling 1932). Of

Protein quantum chemistry E. Oldfield 1355

3hJNC 0 increase exponentially with decreasing inter-nuclear separation (Cornilescu et al. 1999; Malloryet al. 2000), and figure 4b,c shows that thesemagnitudes are related to the mutual penetration ofnon-bonding van der Waals shells in the same manner.Both are of the exponential form jJjZA exp½B S Dr,where B is 3.36 for 3hJNC 0 and 3.67 for JFF. The fact thatIn[jJj] versus S Dr has a very similar slope for both typesof couplings strongly suggests that the through-spaceJFF coupling and the hydrogen bond 3hJNC 0 couplingare subject to the same inductive mechanism, andrequire neither an attractive electrostatic bond nor acovalent bond, only that two atomic surfaces contactone another in the intervening space between couplednuclei. Thus, both through bond as well as throughspace J-couplings can now be calculated using quan-tum chemical methods, with best results beingobtained when dynamics is incorporated into thecalculations (Case et al. 2000; Markwick et al. 2003).

10. THE CHEMICAL SHIFT IN NMRSPECTROSCOPYIn addition to these results, there has been muchprogress in predicting NMR chemical shifts in proteins.When a protein folds, there is a large range of chemicalshift non-equivalence generated. These effects werefirst noted in the 1960s for 1H (Dwek 1973) and in the1970s for 19F (Sykes & Weiner 1980) and 13C NMR(Allerhand et al. 1973). For the heavier elements (C, N,F), the effects can be very large—up to ca 10 p.p.m. for13C, 35 p.p.m. for 15N, and ca 20 p.p.m. for 19F.However, the successful prediction of folding-inducedchemical shifts has been reported only during the pastdecade. In early experimental work, it was found that13Ca chemical shifts in helical versus random coilpolypeptides were different, with the helical groupresonating ca 3 p.p.m. downfield from the random coilshift (Allerhand & Oldfield 1973), but only with thedevelopment of 2D NMR techniques could clearpatterns of Ca and Cb helix-sheet differences be seen(Spera & Bax 1991), which set the stage for ab initiocalculations of these properties, since they clearlyappeared to be dependent on f, j. There were,however, numerous questions or concerns at the outsetas to whether such calculations would be possible, sinceit was unclear whether solvation, dynamics andelectrostatics would also need to be incorporated intoa calculation or set of calculations, to reproduce theexperimental observations. By contrast, the obser-vation that helices and sheets had generally well-separated chemical shifts and the observation thatcrystal and solution shifts were similar (Cole et al.1988) made it unlikely that such effects woulddominate the observed shifts.

In early test calculations, the 13C NMR chemicalshifts in crystals of two amino acids: L-threonine andL-tyrosine were investigated (de Dios et al. 1994).The availability of tensor magnitude information fromthese crystalline samples provided an important test ofwhat might be considered a ‘basic’ test of chemicalshielding tensor predictions in model systems. It wasfound that both isotropic shifts as well as tensormagnitudes could be accurately predicted (de Dios

Phil. Trans. R. Soc. B (2004)

et al. 1994) by using Hartree–Fock theory combinedwith the gauge including atomic orbitals (GIAO)technique (Ditchfield 1974). And in the case ofL-threonine, where single crystal results were available,the tensor orientations could also be successfullypredicted. An exception was found with the carboxyltensors, but this could be corrected by including localelectrostatic field effects using a lattice of point charges,the charge field perturbation (CFP)–GIAO approach(de Dios et al. 1994).

Hartree–Fock methods were then also used topredict 13C NMR chemical shifts, for Ca and Cb, inproteins. An important first step here was to investigatehow large a fragment was needed to predict these 13Cshifts, since clearly the f, j effects of neighbouringresidues would need to be incorporated, based on theempirical observation that Ca, Cb shifts were a functionof these backbone torsion angles. After some testcalculations, N-formyl-amino acid amides wereselected, using a locally dense basis approach inwhich a larger number of basis functions is givento the atom of interest (Chesnut & Moore 1989).The coordinates of the atoms were based on proteinX-ray crystallographic data. The first sets of calcu-lations were uniform failures. Fortunately, it waspointed out by C. Jameson (private communication)that there might be significant errors in the crystallo-graphic bond lengths and bond angles used, so the useof geometry optimization (using an EFG force field)was next investigated, to ‘regularize’ bond lengths andthree-atom bond angles. This approach was successful,and using even these relatively small fragments andneglecting charge field and hydrogen bonding effects, itwas found to be possible to predict 13Ca, 13Cb NMRchemical shifts in a staphylococcal nuclease (SNase)protein with an r.m.s. error of ca 1 p.p.m. (de Dios et al.1993). The same techniques were also found to permit15N NMR chemical shift predictions in SNase, using asomewhat larger peptide fragment to take into accountthe conformational effects of the immediately preced-ing (iK1) residues which, based on empirical obser-vations (Le & Oldfield 1994), were expected to make asignificant contribution to shielding. The full ca 35p.p.m. range of 15N shifts in valine residues in SNasewas reproduced in the calculations, again with only asmall error on the individual shifts. For valine,threonine and isoleucine, the 15N shift ranges arelarger than with other residues, which can be attributedto a ‘g-gauche’ effect on shielding (Cheney & Grant1967), which varies with side-chain (c1) conformation.The ability to predict isotropic 13Ca shifts in proteinsand both isotropic and anisotropic shielding tensorinformation in amino acids together with 15N isotropicshifts in proteins (de Dios et al. 1993; Le & Oldfield1996) then prompted a determination of completeisotropic and anisotropic shielding tensor informationfor all amino acids (except proline) on Ramachandranf, j surfaces (Sun et al. 2002), as shown in, forexample, figure 5a,b. The results obtained wereinteresting in that, for Ca, sheet residues were clearlypredicted to resonate upfield (figure 5a), but thebreadths of the tensors for both helical and sheetgeometries were very similar, ca 34 p.p.m., for all aminoacids except valine, threonine and isoleucine. For these

Page 10: Review Quantum chemical studies of protein structurefeh.scs.uiuc.edu/doc/papers/1201228224_278.pdfgroups, a key feature of protein structure, owing to resonance (Pauling 1932). Of

+142

+140

+142+144

+146+148

+148+146

+144

+142 +140

+138

+140

+14

+1

+20+25

+45+40

+35+30

+15

+25

+50+45+40+35

+30 +15

+25+20

+40+35+30

+20 +25 +30

+20

+25

+30

+25

+30+3

+10+15

+30

+25

+35

+30+25

+5

+10

+20+25

+35 +30

+20

+15

+10 +5 +0+–5

+0

+5

+10

+10

+15

+5

+10

+15

+15

150

100

50

0

–50

–100

–150

15010050–50–100–150 0 15010050–50–100–150 0 15010050–50–100–150 0

Φ (deg) Φ (deg) Φ (deg)

s33

s11s22

s33

s11

s22

s33

s11 s22

s33

s11 s22

∆s *

Ψ (d

eg)

(a) (b) (c)

(d ) (e) ( f ) (g)

sisoΩ

Figure 5. Computed isotropic shielding and anisotropic shielding tensors and their orientations for 13Ca in N-formyl alanineamide: (a) siso; (b) UZs33Ks11; (c) Ds*. (d ) Helical 13Ca tensor orientation in lysine. (e) Sheet 13Ca tensor orientation in lysine.( f ) Helical 13Cb tensor orientation. (g) Sheet Cb tensor orientation. Reprinted in part with permission from Sun et al. (2002).Copyright 2002 American Chemical Society.

1356 E. Oldfield Protein quantum chemistry

residues, the tensor breadths were generally muchsmaller, ca 22 p.p.m., in helical geometries. This effectappears to be related to the presence of b-branchedside-chains, reminiscent of the g-gauche effect seen inthe case of the 15N NMR shifts of these same threeamino acids. The results of these calculations were thenverified by using solid-state NMR techniques to deducethe magnitudes of the tensor elements in Ala, Val andLeu peptides (Havlin et al. 2001), and the Ca tensororientations, shown in figure 5d,e, have now beenconfirmed experimentally in several amino acids andpeptides by other groups (Yao & Hong 2002; Chekme-nev et al. 2002).

These computational results were at first somewhatsurprising in that, in solution NMR, it had been foundthat a chemical shift anisotropy (CSA), Ds*, had beenfound to vary very considerably between helical andsheet geometries, with Ds*z6 p.p.m. in helices andDs*z27 p.p.m. in sheets (Tjandra & Bax 1997).However, this CSA, determined by dipole–dipole/CSAinterference techniques, is not the same as thatdetermined theoretically or by solid-state NMR sinceit is convoluted with tensor orientation information.In particular, the s22 element of the 13Ca tensor isoriented approximately along the C–H bond vector inhelical geometries (figure 5d ) while s11 is orientedapproximately along the C–H bond vector in sheetgeometries (Walling et al. 1997; figure 5e). It is,however, a simple matter to effect the appropriatetransformation, and typical Ds*(f, j) surface results

Phil. Trans. R. Soc. B (2004)

are shown in figure 5c. There was found to be generallygood agreement between experimental and theoreticalDs* values (Case 1998; Sun et al. 2002), with Ds*z0p.p.m. being found for ideal helical residues and Ds* ofca 25 p.p.m. being predicted for ideal sheet residues.Interestingly, for Cb, there is essentially no change intensor orientation between helices and sheets, as shownin figure 5f,g.

The experimental Ds* results for Ca and thetheoretical predictions were all found to be highlycorrelated, as shown in figure 6a. Moreover, the errorsthat were observed were found to be a function of thequality of the (crystallographic) structure used to setthe f,j values in the calculations, as can be seen infigure 6b. Future work on geometry optimization, asoutlined above, may reduce these errors. The GIAOapproach was also found to be capable of accuratelypredicting 13C side-chain shifts, as shown in figure 6cfor Cb, Cg1, Cg2, Cd (and Ca) for isoleucine side-chains, where again (g-gauche) torsional effects makemajor contributions to shielding (Sun et al. 2002).

In most of the work described above, chemical shiftswere calculated using individual or specific molecules.A much faster method is to use a database approach,which relies on additivity. Here, Case and coworkers(Sitkoff & Case 1997; Xu & Case 2001, 2002) havebeen very successful in producing a fast algorithm forquite accurately predicting isotropic 13C and 15N shifts,based on a large database of chemical shifts computedusing DFT. This prediction algorithm (‘SHIFTS’) is

Page 11: Review Quantum chemical studies of protein structurefeh.scs.uiuc.edu/doc/papers/1201228224_278.pdfgroups, a key feature of protein structure, owing to resonance (Pauling 1932). Of

50

40

30

20

10

0

–10

20

15

10

5

0

190

180

170

160

150

140

13050403020100–10 0 2 4 6 8 10 0 20 30 40 50 60 70

* r.m.s.d. ∆σ* (p.p.m.)

∆σ th

eory

(p.p

.m.)

chemical shift ssNMR (p.p.m.)

perc

enta

ge o

f re

sidu

es o

ut o

fm

ost f

avou

red

regi

ons

shie

ldin

g X-r

ay (

p.p.

m.)

1D3Z

1UBI

1CDL

1UBQ

2BBM

∆σExp (p.p.m.)

(a) (b) (c)

Figure 6. Experimental versus theoretical shift/shielding results in proteins. (a) Correlations between experimental andcomputed 13C Ds* values in ubiquitin using PDB structure 1D3Z and individual amino acid Ds* surfaces; (b) Relation betweenDs* error (r.m.s. error between the experimental Ds* and that computed from the structure indicated) and the structure quality(percentage residues outside the most favoured regions of the protein structure). Structure used: ubiquitin, 1UBI, 1UBQ and1D3Z; calmodulin, 1CDL and 2BBM. (c) Correlation between experimental isotropic chemical shifts, d, in p.p.m. (from TMS)and the computed absolute shieldings, s, for Ca, Cb, Cg1, Cg2 and Cd in Ile-18 and Ile-19 in the bovine pancreatic trypsininhibitor. Experimental solid state shift/X-ray predicted shieldings; slope: K0.84, y-intercept: 192.1 p.p.m., R2Z1.00,r.m.s.d.Z1.1 p.p.m. Reprinted in part with permission from Sun et al. (2002). Copyright 2002 American Chemical Society.

Protein quantum chemistry E. Oldfield 1357

expected to be of great use in NMR peak assignments,as well as in structure refinement.

Finally, in the area of solution 13C, 15N NMRchemical shift calculations, there have been importantadvances made by Vila et al. (2003) in predicting‘random coil’ shifts, or more precisely Boltzmann-averaged values of the chemical shifts of a series oftetra-peptides. This is an exciting new developmentsince it potentially opens up the use of NMR chemicalshifts and quantum chemistry for the investigation ofless well-structured regions in proteins (Vila et al. 2003).

11. FLUORINE NMR CHEMICAL SHIFTS ANDELECTROSTATICSA rather challenging task is posed by 19F NMRchemical shifts in proteins. These had been investigatedexperimentally for many years, but unlike the situationwith 13C and 15N NMR shifts, it seemed unlikely thatthe 19F NMR chemical shifts of, for example, afluorophenylalanine residue, would be related to f, j orc1, c2 conformational effects. It was found to be possibleto accurately predict 19F chemical shifts in manychemically different fluorobenzenes (de Dios & Oldfield1994), but what would cause a large range of chemicalshift non-equivalence in a protein, due to folding?

One likely possibility appeared to be that 19F NMRchemical shifts in proteins might be dominated byelectrostatic field effects, an idea which was based onearlier observations that 13C and 17O NMR shifts ofCO bound to different haem proteins, the C17O EFGand the CO infrared vibrational frequencies were allrelated (Park et al. 1991). Moreover, they could bepredicted by using an electrical polarization model(Augspurger et al. 1991, 1993a). To test this idea, weinvestigated how the 19F NMR chemical shifts influorobenzene (a phenylalanine model) might beperturbed by an electrostatic field, generated byclusters of HF molecules or point charges/point dipoles(the protein), using supermolecule cluster calculations,and the CFP method outlined above for the aminoacids threonine and tyrosine (de Dios & Oldfield1993). The shift effects were found to be large at

Phil. Trans. R. Soc. B (2004)

plausible intermolecular distances, and the results ofthe supermolecule and point charge/point dipolecalculations agreed. In addition, a third method wasdeveloped, in which the derivatives of the shieldingtensor with respect to the field ð ~AÞ and field gradient ð ~BÞwere evaluated, by using derivative Hartree–Focktheory (Augspurger & Dykstra 1991; Augspurgeret al. 1992). For example, the response of the shift toa uniform field is given by

~Aab;g Zvsab

vFg

;

where ~A is called the dipole shielding polarizabilitytensor. That is, ~A is the formal derivative of s withrespect to a uniform field, while ~B is the formalderivative of s with respect to the non-uniform fieldor field gradient components. Hyperpolarizabilities(v2/vF2) and mixed derivatives (v2/vF vV ) were alsoevaluated (Augspurger et al. 1993b), but based on theresults of model calculations, only the ~A and ~B tensorswere found to be significant. The results of these‘multipole shielding polarizability’ calculations on thefluorobenzene–HF clusters were found to agree withthose obtained from supermolecule or CFP techniques(Augspurger et al. 1993b; de Dios & Oldfield 1993).

What this means for proteins is that the electrostaticcontributions to the 19F NMR chemical shifts can becalculated by simply multiplying the ~A and ~B tensors bythe electrostatic fields and field gradients in the protein.To evaluate the fields in a protein, the ENZYMIX

program was used. This involves use of a pointdipole–Langevin dipole method to evaluate the uni-form field component, F, and by finite difference, thefield gradient, V (Pearson et al. 1993). This iscomputationally lengthy, since solvation as well asdynamical effects need to be taken into account, at leastif solvent-exposed residues are to be considered, whichnecessitates a molecular dynamics (MD) approach.That is, the fields (F ) and field gradients (V ) arecalculated over an MD trajectory, then the uniform andnon-uniform electrostatic field contributions to shield-ing can be evaluated given knowledge of the dipole and

Page 12: Review Quantum chemical studies of protein structurefeh.scs.uiuc.edu/doc/papers/1201228224_278.pdfgroups, a key feature of protein structure, owing to resonance (Pauling 1932). Of

1358 E. Oldfield Protein quantum chemistry

quadrupole shielding polarizability. Perhaps surpris-ingly, the method was found to give an excellentdescription of the 19F NMR chemical shifts of the five[5-19F]Trp residues in a galactose binding protein(Pearson et al. 1993) as well as quite good accord withthe experimental [4-19F]Phe residue shifts in myoglo-bin and haemoglobin (Pearson et al. 1997). In the caseof the galactose binding protein, good accord was alsofound with the CFP method for buried residues, butthe slope of the correlation was worse since thedielectric constant (in the protein) was not consideredin the calculations (de Dios et al. 1993). The effects ofelectrostatic interactions can also be extremely wellmodelled by use of hybrid QM–MM techniques, asdemonstrated by Cui & Karplus (2000), who appliedthis approach to a wide number of systems, includinghaem protein models. It seems likely that the QM–MMtechnique will be of great future importance, especiallywhen combined with the Car–Parrinello MD technique(Gervasio et al. 2003), enabling the incorporation ofboth inter-residue interactions as well as dynamics, intoproperty calculations.

12. METAL ION NMRAnotherNMRchallenge forquantumchemistrywas to tryto predict the NMR shifts and shift (or shielding) tensorsfor a transition metal in a protein, a difficult problem giventhe large number of electrons in such metals. Inearly work,several groups investigated experimentally the 57Fe(IZ1/2) NMR chemical shifts in the proteins myoglobin(Mb) and ferrocytochrome c (Baltzer et al. 1985; Baltzer1987; Chung et al. 1990), and there was found to be anextremely large (ca 3000 p.p.m.) chemical shift range. Inaddition, relaxation measurements enabled the determi-nation of the tensor spans, Ds, for MbCO, MbRNC andcyt c. Early investigations of Rh(I) NMR chemical shifts inorganometallic compounds using Hartree–Fock methodswere unsuccessful, but good correlations were later foundwhen using DFT techniques, especially when usingthe hybrid exchange/correlation functional B3LYP(Donkervoort et al. 1999). This approach was thusadapted to investigate 57Fe and 59Co shifts and shifttensors in a series of model systems and in metalloproteinsthemselves (Godbout & Oldfield 1997; Godbout et al.1998). Again, perhaps surprisingly, it was found to bepossible to predict both the isotropic and anisotropic shift(shielding) tensor properties, when using the DFT/B3LYP approach, and suitably large basis sets. Initialresults with a protein, carbonmonoxymyoglobin, werevery poor, but this was found to be the result of using anearly crystallographic structure in which the CO ligandbound to Fe appeared to adopt a bent geometry. Use of anupright geometry, based on crystallographic investigationsof a small haem–CO complex, corrected this problem andled to the idea, as described below, that NMR shifts mightbe used to accurately determine at least some aspects ofprotein structure.

13. TOWARDS STRUCTURE DETERMINATIONThe results outlined above indicate that it is nowpossible to predict many spectroscopic observables inproteins and model systems; dFe and DEQ in Mossbauer

Phil. Trans. R. Soc. B (2004)

spectroscopy; d(13C, 15N, 19F, 57Fe) and Dd inproteins, peptides and other model systems; dNMR

hyperfine shifts in proteins and model systems; and VE(the EFG) in amino acids and haems. There is,therefore, a correlation between ‘structure’ and‘spectra’. In what we have described so far, we havetypically gone in the ‘forward’ direction: structure tospectra. But what might be more interesting would beto go in the reverse direction: spectra to structure. Atleast at the empirical level, this is not new since, forexample, J-couplings have been used for many years asstructure restraints just using empirical parameteriza-tions of the Karplus relation. However, with quantumchemistry there might be more possibilities using, forexample, chemical shifts deduced from Ramachandranshielding surfaces or indeed, ab initio structure deter-minations based in part on NMR shifts, or chemicalshift assisted approaches to ab initio folding.

Two simple examples of de novo approaches tostructure determination are given in which the struc-ture to spectral relation is inverted analytically. Assumea property, P, is a function of two independentvariables, a; b : PZPða; bÞ. For example, P might be aCa chemical shift and aZj, bZf, the backbone torsionangles. For a given a, b, the value of the property can bederived from the property surface P ZP(a, b) and theprobability, Z, that a, b represent the actual exper-imental values will be

eKðPexptKPða;bÞÞ2=W ;

where Pexpt is the experimental value of the property,P(a, b) is the calculated value of the property for thegiven a, b values and W is a search width parameterrelated to the r.m.s. error in the property prediction.When PexptZP(a, b), ZZ 1 and the a, b values chosenare possible solutions. Of course, the actual form ofP(a, b) may yield multiple a, b solutions yielding ZZ1,in which case it is necessary to use more than oneproperty to deduce unique values fora and b. In this case,it is possible to use conditional probability such that

ZN ZPZi ;

where there are N observables. In an early application ofthis idea, isotropic and anisotropic 13Ca NMR shiftinformation was used to predict the backbone f, j

torsion angles in a tripeptide with an error of between 58and 108 (Heller et al. 1997), and in another application,d(13C), Dd(13C), d(17O), VE(17O), d(57Fe) and DEQ

(57Fe) Z surface information was used to predict theFe–CO ligand tilt and bend angles for CO bound toFe(II) in myoglobin (McMahon et al. 1998). The natureof this ligand bonding had been controversial for manyyears, with early crystallographic and indeed otherresults suggesting a highly bent (ca 1208) ligand-bindinggeometry. Using the combination of NMR and Moss-bauer spectroscopic property information, a ligand tilt(t)Z48 and a bend (b)Z78were found (McMahon et al.1998). After these results were published, the X-raystructure was re-determined using synchrotron radi-ation (Kachalova et al. 1999). The new results obtainedshowed only a ca 0.608 variance between the Z-surfacepredictions, based on quantum chemistry, and the newcrystal structure. In a similar vein, we have recentlyinvestigated the geometry of NO bound to haemoglobin

Page 13: Review Quantum chemical studies of protein structurefeh.scs.uiuc.edu/doc/papers/1201228224_278.pdfgroups, a key feature of protein structure, owing to resonance (Pauling 1932). Of

Protein quantum chemistry E. Oldfield 1359

and myoglobin. These results again vary considerablyfrom some crystallographic structures, but enable theprediction of both dFe and DEQ in the Mossbauerspectra, as well as the electron paramagnetic resonanceobservables, and are essentially the same as thoseobtained by using quantum chemical geometry optim-ization on a large molecular cluster (Y. Zhang, W.Gossman and E. Oldfield, unpublished data).

14. CONCLUSIONS AND FUTURE PROSPECTSThe results we have described above cover develop-ments in predicting a wide variety of spectroscopic andother observable properties in proteins and modelsystems, which have been carried out over the pastdecade. At perhaps the most basic level, ab initio SCFenergy calculations have been successful in reproducingthe fine details of protein structure seen in very-high-resolution crystal structures. In the future, it thereforeseems likely that it will be possible to directlyincorporate quantum chemical results in structurerefinement. From the associated ground state wavefunctions, most electrostatic properties (r(r), li,V

2r(r),Hessian-of-r(r), F(r), VE) can be derived. All theseproperties should be accessible in proteins. From aspectroscopic standpoint, it is now also possible toaccurately predict the Mossbauer observables asfollows: dFe, the isomer shift (from r(0)), and DEQ,the quadrupole splitting (fromVZZ, the z-component ofthe EFG tensor). This opens the way to using theseobservables to determine local geometries, and alsogives more confidence in using MO methods to probedetails of metal–ligand bonding. In the area of NMRspectroscopy, it is now possible to predict 13C, 15N, 19Fand even 57Fe chemical shifts and shift (shielding)tensors, as well as their orientations, in addition toenabling detailed studies of J-couplings, of interest inthe context of structure refinement and hydrogenbonding. And for paramagnetic systems, it has beenfound to be possible to quite accurately predict manyNMR contact shifts ab initio, which for haem proteinsencompass a ca 6000 p.p.m. shift range. Already, thestructure to spectra nexus is being used to makestructure refinements, and in the future it seems likelythat NMR chemical shifts may play an even moreimportant role in structure determination, especially innon-crystalline proteins. The success of quantumchemical methods in predicting such a wide range ofproperties also gives added confidence in using quan-tum chemical methods to investigate reaction mechan-isms (Gogonea et al. 2001; Gao & Truhlar 2002) which,since they involve bond rearrangements, must necess-arily be treated, at least in part, quantum mechanically.Hopefully, over the next few years, these two ratherdistinct areas: structural/spectroscopic (large basis/small fragment/property calculations) and enzymemechanism (small basis—QM–MM, entire protein) ofquantum chemistry will merge, and it will be possible touse one method to determine or refine structures, tomake property calculations, and to investigate reactionmechanisms and their inhibition. For example, at thelevel of using NMR chemical shift information, itshould be possible to probe drug–protein interactions—a potentially ‘useful’ application of quantum chemical

Phil. Trans. R. Soc. B (2004)

technology. It should also be possible to use QMmethods in a more direct manner in drug design, ratherthan simply using force-field-based methods to makestructure–activity correlations, by more accuratelycalculating interaction energies, and by using AIMtheory (O’Brien & Popelier 2002). As with most thingscomputational, it is just a matter of time.

The author thanks A. C. de Dios and C. Jameson for theirearly contributions to this work and Y. Zhang and H. Sun forproviding the figures. This work was supported by the UnitedStates Public Health Service (National Institutes of Healthgrants EB00271 and GM50694).

REFERENCESAlderman, D. W., Sherwood, M. H. & Grant, D. M. 1993

Comparing, modeling, and assigning chemical-shift ten-sors in the Cartesian, irreducible spherical, and icosahe-dral representations. J. Magn. Res. Ser. A 101, 188–197.

Allerhand, A. & Oldfield, E. 1973 Determination ofrotational mobilities of backbone and side chain carbonsof poly(g-benzyl L-glutamate) in the helical and random-coil states from measurements of carbon-13 relaxationtimes and nuclear Overhauser enhancements. Biochemistry12, 3428–3433.

Allerhand, A., Childers, R. F. & Oldfield, E. 1973 Natural-abundance carbon-13 nuclear magnetic resonance studiesin 20 min sample tubes. Observation of numerous single-carbon resonances of hen egg-white lysozyme. Biochem-istry 12, 1335–1341.

Arnold, W. D. & Oldfield, E. 2000 The chemical nature ofhydrogen bonding in proteins via NMR: J-couplings,chemical shifts, and AIM theory. J. Am. Chem. Soc. 122,12 835–12 841.

Arnold, W. D., Sanders, L. K., McMahon, M. T., Volkov,R. V., Wu, G., Coppens, P., Wilson, S. R., Godbout, N. &Oldfield, E. 2000a Experimental Hartree–Fock, anddensity functional theory investigations of the chargedensity, dipole moment, electrostatic potential, andelectric field gradients in L-asparagine monohydrate.J. Am. Chem. Soc. 122, 4708–4717.

Arnold, W. D., Mao, J. H., Sun, H. H. & Oldfield, E. 2000bComputation of through-space F-19–F-19 scalar coup-lings via density functional theory. J. Am. Chem. Soc. 122,12 164–12 168.

Augspurger, J. D. & Dykstra, C. E. 1991 Electromagneticproperties of molecules from a uniform procedure fordifferentiation of molecular wave-functions to high-order.J. Phys. Chem. 95, 9230–9238.

Augspurger, J. D., Dykstra, C. E. & Oldfield, E. 1991Correlation of carbon-13 and oxygen-17 chemical shiftsand the vibrational frequency of electrically perturbedcarbon monoxide: a possible model for distal ligand effectsin carbonmonoxyheme proteins. J. Am. Chem. Soc. 113,2447–2451.

Augspurger, J. D., Pearson, J. G., Oldfield, E., Dykstra, C. E.,Park, K. D. & Schwartz, D. 1992 Chemical-shift ranges inproteins. J. Magn. Reson. 100, 342–357.

Augspurger, J. D., Dykstra, C. E., Oldfield, E. & Pearson, J. G.1993a Intra- and intermolecular electrical effects on nuclearmagnetic resonance, nuclear quadrupole resonance andinfra-red spectroscopic parameters from ab initio calculationand experiment: from CO to proteins. In Nuclear magneticshieldings and molecular structure, NATO ASI Series C,vol. 386 (ed. J. A. Tossell), pp. 75–94. Boston, MA: Kluwer.

Augspurger, J. D., de Dios, A. C., Oldfield, E. & Dykstra,C. E. 1993b Computing nuclear magnetic resonancechemical shielding in large systems via multipole shieldingpolarizabilities. Chem. Phys. Lett. 213, 211–216.

Page 14: Review Quantum chemical studies of protein structurefeh.scs.uiuc.edu/doc/papers/1201228224_278.pdfgroups, a key feature of protein structure, owing to resonance (Pauling 1932). Of

1360 E. Oldfield Protein quantum chemistry

Bader, R. F. W. 1990 Atoms in molecules—a quantum theory.

Oxford University Press.

Baltzer, L. 1987 57Fe NMR of heme proteins: the chemical

shift anisotropy of ferrocytochrome-c. J. Am. Chem. Soc.

109, 3479–3481.

Baltzer, L., Becker, E. D., Tschudin, R. G. & Gansow, O. A.

1985 57Fe NMR spectroscopy of heme proteins: chemical

shift anisotropy and relaxation parameters of carbonyl-

myoglobin. J. Chem. Soc. Chem. Commun., 1040–1041.

Bernal, J. D. & Crowfoot, D. 1934 X-ray photographs of

crystalline pepsin. Nature 794, 133–134.

Case, D. A. 1998 The use of chemical shifts and their

anisotropies in biomolecule structure determination. Curr.

Opin. Struct. Biol. 8, 624–630.

Case, D. A., Scheurer, C. & Bruschweiler, R. 2000 Static and

dynamic effects on vicinal scalar J-couplings in proteins

and peptides: a MD/DFTanalysis. J. Am. Chem. Soc. 122,

10 390–10 397.

Challacombe, M. & Schwegler, E. 1997 Linear scaling compu-

tation of the Fock matrix. J. Chem. Phys. 106, 5526–5536.

Chekmenev, E. Y., Xu, R. Z., Mashuta, M. S. & Wittebort,

R. J. 2002 Glycyl Ca chemical shielding in tripeptides:

measurement by solid-state NMR and correlation with X-

ray structure and theory. J. Am. Chem. Soc. 124,

11 894–11 899.

Cheney, B. V. & Grant, D. M. 1967 Carbon-13 magnetic

resonance. VIII. The theory of carbon-13 chemical shifts

applied to saturated hydrocarbons. J. Am. Chem. Soc. 89,

5319–5327.

Chesnut, D. B. & Moore, K. D. 1989 Locally dense basis-sets

for chemical-shift calculations. J. Comput. Chem. 10,

648–659.

Chung, J., Lee, H. C. & Oldfield, E. 1990 Iron-57 nuclear

magnetic resonance spectroscopic study of alkyl isocya-

nide myoglobins and a comparison of the 57Fe chemical

shift anisotropies of alkyl isocyanide myoglobins, carbon-

monoxy-myoglobin, ferrocytochrome c and [57Fe(bipy)3]-

X2$5H2O (xZCl, Br, I). J. Magn. Res. 90, 148–157.

Cole, H. B. R., Sparks, S. W. & Torchia, D. A. 1988

Comparison of the solution and crystal structures of

staphylococcal nuclease with 13C and 15N chemical shifts

used as structural fingerprints. Proc. Natl Acad. Sci. USA

85, 6362–6365.

Cornilescu, G., Hu, J. S. & Bax, A. 1999 Identification of the

hydrogen bonding network in a protein by scalar

couplings. J. Am. Chem. Soc. 121, 2949–2950.

Cremer, D. & Kraka, E. 1984 A description of the chemical

bond in terms of local properties of electron density and

energy. Croat. Chem. Acta 57, 1259–1281.

Cui, Q. & Karplus, M. 2000 Molecular properties from

combined QM/MM methods. 2. Chemical shifts in large

molecules. J. Phys. Chem. B 104, 3721–3743.

de Dios, A. C. & Oldfield, E. 1993 Methods for computing

nuclear magnetic resonance chemical shielding in large

systems. Multiple cluster and charge field approaches.

Chem. Phys. Lett. 205, 108–116.

de Dios, A. C. & Oldfield, E. 1994 Evaluating 19F chemical

shielding in fluorobenzenes: implications for chemical

shifts in proteins. J. Am. Chem. Soc. 116, 7453–7454.

de Dios, A. C., Pearson, J. G. & Oldfield, E. 1993 Secondary

and tertiary structural effects on protein NMR chemical

shifts: an ab initio approach. Science 260, 1491–1496.

de Dios, A. C., Laws, D. D. & Oldfield, E. 1994 Predicting

carbon-13 nuclear magnetic resonance chemical shielding

tensors in zwitterionic L-threonine and DL-tyrosine via

quantum chemistry. J. Am. Chem. Soc. 116, 7784–7786.

Ditchfield, R. 1974 Self-consistent perturbation theory of

diamagnetism. I. A gauge-invariant LCAO method for

N.M.R. chemical shifts. Mol. Phys. 27, 789–807.

Phil. Trans. R. Soc. B (2004)

Dittrich, B., Flaig, R., Koritsanszky, T., Krane, H. G.,Morgenroth, W. & Luger, P. 2000 Topological propertiesof the peptide bond in glycyl-L-threonine dihydrate basedon a fast synchrotron/CCD-diffraction experiment at100 K. Chem. A Eur. J. 6, 2582–2589.

Donkervoort, J. G., Buhl, M., Ernsting, J. M. & Elsevier, C. J.1999 Steric and electronic effects on the Rh-103 NMRchemical shifts of Rh-I(cyclooctadiene) compounds bear-ing N-donor ligands. Chem. Eur. J., 27–33.

Dwek, R. A. 1973 Nuclear magnetic resonance (N.M.R.) in bio-chemistry. Oxford: Clarendon Press.

Gao, J. & Truhlar, D. G. 2002 Quantum mechanical methodsfor enzyme kinetics. Annu. Rev. Phys. Chem. 53, 467–505.

Gervasio, F. L., Schettino, V., Mangani, S., Krack, M.,Carloni, P. & Parrinello, M. 2003 Influence of outer-shellmetal ligands on the structural and electronic properties ofhorse liver alcohol dehydrogenase zinc active site. J. Phys.Chem. B 107, 6886–6892.

Godbout, N. & Oldfield, E. 1997 Density functional theorystudy of cobalt-59 nuclear magnetic resonance chemicalshifts and shielding tensor elements in Co(III) complexes.J. Am. Chem. Soc. 119, 8065–8069.

Godbout, N., Havlin, R., Salzmann, R., Debrunner, P. G. &Oldfield, E. 1998 Iron-57 NMR chemical shifts andMossbauer quadrupole splittings in metalloporphyrins,ferrocytochrome c, and myoglobins: a density functionaltheory investigation. J. Phys. Chem. A 102, 2342–2350.

Gogonea, V., Suarez, D., van der Vaart, A. & Metz, K. W.2001 New developments in applying quantum mechanicsto proteins. Curr. Opin. Struct. Biol. 11, 217–223.

Havlin, R. H., Laws, D. D., Bitter, H.-M.L., Sanders, L. K.,Sun, H., Grimley, J. S., Wemmer, D. E., Pines, A. &Old-field, E. 2001 An experimental and theoreticalinvestigation of the chemical shielding tensors of 13Ca ofalanine, valine, and leucine residues in solid peptides and inproteins in solution. J.Am.Chem.Soc.123, 10 362–10 369.

Hehre, W. J., Radom, L., von Schleyer, P. R. & Pople, J. A.1986 Ab initio molecular orbital theory. New York: Wiley.

Heisenberg, W. 1926 Mehrkoperproblem und resonanz inder quantenmechanik. Zeitschr. Phys. 38, 411–426.

Heitler, W. & London, F. 1927 Wechselwirkung neutraleratome und homoopolare bindung nach der quantenme-chanik. Zeitschr. Phys. 44, 455–472.

Helgaker, T., Watson, M. & Handy, N. C. 2000 Analyticalcalculation of nuclear magnetic resonance indirect spin–spin coupling constants at the generalized gradientapproximation and hybrid levels of density-functionaltheory. J. Chem. Phys. 113, 9402–9409.

Heller, J., Laws, D. D., Tomaselli, M., King, D. S., Wemmer,D. E., Pines, A., Havlin, R. H. & Oldfield, E. 1997Determination of dihedral angles in peptides throughexperimental and theoretical studies of a-carbon chemicalshielding tensors. J. Am. Chem. Soc. 119, 7827–7831.

Kachalova, G. S., Popov, A. N. & Bartunik, H. D. 1999 Asteric mechanism for inhibition of CO binding to hemeproteins. Science 284, 473–476.

Karplus, M. 1996 Theory of vicinal coupling. In Encyclopediaof nuclear magnetic resonance (ed. D. Grant & R. Harris),pp. 420–422. New York: Wiley.

Kimber, B. J., Feeney, J., Roberts, G. C. K., Birdsall, B.,Griffiths, D. V., Burgen, A. S. V. & Sykes, B. D. 1978Proximity of two tryptophan residues in dihydrofolatereductase determined by fluorine-19 NMR. Nature 271,184–185.

Kohn, W., Becke, A. D. & Parr, R. G. 1996 Densityfunctional theory of electronic structure. J. Phys. Chem.100, 12 974–12 980.

Le, H. B. & Oldfield, E. 1994 Correlation between 15N NMRchemical shifts in proteins and secondary structure.J. Biomol. NMR 4, 341–348.

Page 15: Review Quantum chemical studies of protein structurefeh.scs.uiuc.edu/doc/papers/1201228224_278.pdfgroups, a key feature of protein structure, owing to resonance (Pauling 1932). Of

Protein quantum chemistry E. Oldfield 1361

Le, H. B. & Oldfield, E. 1996 Ab initio studies of amide 15N

chemical shifts in dipeptides: applications to protein NMR

spectroscopy. J. Phys. Chem. 100, 16 423–16 428.

Li, X., Wu, G., Abramov, Y. A., Volkov, A. V. & Coppens, P.

2002 Application of charge density methods to a protein

model compound: calculation of Coulombic intermole-

cular interaction energies from the experimental charge

density. Proc. Natl Acad. Sci. USA 99, 12 132–12 137.

LiWang, A. C. & Bax, A. 1997 Solution NMR characterization

of hydrogen bonds in a protein by indirect measurement of

deuterium quadrupole couplings. J.Magn. Res. 127, 54–64.

McMahon, M. T., de Dios, A. C., Godbout, N., Salzmann,

R., Laws, D. D., Le, H., Havlin, R. H. & Oldfield, E. 1998

An experimental and quantum chemical investigation of

CO binding to heme proteins and model systems: a unified

model based on 13C, 17O, and 57Fe nuclear magnetic

resonance and 57Fe Mossbauer and infrared spectro-

scopies. J. Am. Chem. Soc. 120, 4784–4797.

Malkin, V. G., Malkina, O. L., Casida, M. E. & Salahub,

D. R. 1994 Nuclear magnetic resonance shielding tensors

calculated with a sum-over-states density functional

perturbation theory. J. Am. Chem. Soc. 116, 5898–5908.

Mallory, F. B. et al. 2000 Nuclear spin–spin coupling via

nonbonded interactions. 8. The distance dependence of

through-space fluorine–fluorine coupling. J. Am. Chem.

Soc. 122, 4108–4116.

Mao, J., Zhang, Y. & Oldfield, E. 2002 Nuclear magnetic

resonance shifts in paramagnetic metalloporphyrins and

metalloproteins. J. Am. Chem. Soc. 124, 13 911–13 920.

Markwick, P. R. L., Sprangers, R. & Sattler, M. 2003

Dynamic effects on J-couplings across hydrogen bonds in

proteins. J. Am. Chem. Soc. 125, 644–645.

Naito, A. & McDowell, C. A. 1984 Determination of the 14N

quadrupole coupling tensors and the 13C chemical

shielding tensors in a single-crystal of L-asparagine

monohydrate. J. Chem. Phys. 81, 4795–4803.

O’Brien, S. E. & Popelier, P. L. A. 2002 Quantum topological

molecular similarity. Part 4. A QSAR study of cell growth

inhibitory properties of substituted (E)-1-phenylbut-1-en-

3-ones. J. Chem. Soc. Perkin Trans. 2, 478–483.

Park, K. D., Guo, K., Adebodun, F., Chiu, M. L., Sligar,

S. G. & Oldfield, E. 1991 Distal and proximal ligand

interactions in heme–proteins: correlations between C–O

and Fe–C vibrational frequencies, O-17 and C-13 nuclear

magnetic resonance chemical shifts and O-17 nuclear

quadrupole coupling constants in C17O-labeled and13CO-labeled species. Biochemistry 30, 2333–2347.

Pauling, L. 1928a The application of the quantum mechanics

to the structure of the hydrogen molecule and hydrogen

molecule-ion and to related problems. Chem. Rev. 5,

173–213.

Pauling, L. 1928b The shared-electron chemical bond. Proc.

Natl Acad. Sci. USA 14, 359–362.

Pauling, L. 1932 Interatomic distances in covalent molecules

and resonance between two or more Lewis electronic

structures. Proc. Natl Acad. Sci. USA 18, 293–297.

Pauling, L. & Corey, R. B. 1950 Two hydrogen-bonded spiral

configurations of the polypeptide chain. J. Am. Chem. Soc.

72, 5349.

Pauling, L. & Corey, R. B. 1951 Configurations of polypeptide

chains with favored orientations around single bonds: two

new pleated sheets. Proc. Natl Acad. Sci. USA 37, 729–740.

Pearson, J. G., Oldfield, E., Lee, F. S. & Warshel, A. 1993

Chemical shifts in proteins: a shielding trajectory analysis of

the fluorine nuclear magnetic resonance spectrum of the

Escherichia coli galactose binding protein using a multipole

shielding polarizability-local reaction field-molecular

dynamics approach. J. Am. Chem. Soc. 115, 6851–6862.

Phil. Trans. R. Soc. B (2004)

Pearson, J. G., Montez, B., Le, H. B., Oldfield, E., Chien,E. Y. T. & Sligar, S. G. 1997 Assignment and analysis offluorine nuclear magnetic resonance spectra of 4-fluoro-tryptophan myoglobins and hemoglobins. Biochemistry 36,3590–3599.

Ramachandran, G. N. & Kartha, G. 1954 Structure ofcollagen. Nature 174, 269–270.

Ramachandran, G. N. & Kartha, G. 1955 Structure ofcollagen. Nature 176, 593–595.

Ramachandran, G. N., Ramakrishnan, C. & Sasisekharan, V.1963 Stereochemistry of polypeptide chain configur-ations. J. Mol. Biol. 7, 95–99.

Ramakrishnan, C. 2001 Remembrances of Professor G. N.Ramachandran. Protein Sci. 10, 1689–1693.

Rich, A. & Crick, F. H. C. 1955 The structure of collagen.Nature 176, 915–916.

Sitkoff, D. & Case, D. A. 1997 Density functional calcu-lations of proton chemical shifts in model peptidcs. J. Am.Chem. Soc. 119, 12 262–12 273.

Spera, S. & Bax, A. 1991 Empirical correlation betweenprotein backbone conformation and Ca and Cb 13Cnuclear magnetic resonance chemical shifts. J. Am.Chem. Soc. 113, 5490–5492.

Sun, H. H., Sanders, L. K. & Oldfield, E. 2002 Carbon-13NMR shielding in the twenty common amino acids:comparisons with experimental results. J. Am. Chem. Soc.124, 5486–5495.

Svedberg, T. et al. 1939 A discussion on the protein molecule.Proc. R. Soc. B 846, 1–40.

Sykes, B. D. & Weiner, J. H. 1980 Biosynthesis and 19F NMRcharacterization of fluoroamino acid containing proteins.In Magnetic resonance in biology (ed. J. S. Cohen),pp. 171–194. New York: Wiley.

Tjandra, N. & Bax, A. 1997 Large variations of 13Ca chemicalshift anisotropy in proteins correlate with secondarystructure. J. Am. Chem. Soc. 119, 9576–9577.

Van Alsenoy, C., Yu, C.-H., Peeters, A., Martin, J. M. L. &Schafer, L. 1998 Ab initio geometry determinations ofproteins. 1. Crambin. J. Phys. Chem. 102, 2246–2251.

Vila, J. A., Baldoni, H. A., Ripoll, D. R. & Scheraga, H. A.2003 Unblocked statistical-coil tetrapeptides in aqueoussolution: quantum-chemical computation of the 13CNMR chemical shifts. J. Biomol. NMR 26, 113–130.

Walling, A. E., Pargas, R. E. & de Dios, A. C. 1997 Chemicalshift tensors in peptides: a quantum mechanical study.J. Phys. Chem. A 101, 7299–7303.

Wilkens, S. J., Xia, B., Weinhold, F., Markley, J. L. & Westler,W. M. 1998 NMR investigations of Clostridium pasteur-ianum rubredoxin. Origin of hyperfine H-1, H-2, C-13and N-15 NMR chemical shifts in iron–sulfur proteins asdetermined by comparison of experimental data withhybrid density functional calculations. J. Am. Chem. Soc.120, 4806–4814.

Xu, X. P. & Case, D. A. 2001 Automated prediction of 15N,13Ca, Cb and 13C 0 chemical shifts in proteins using adensity functional database. J. Biomol. NMR 21, 321–333.

Xu, X. P. & Case, D. A. 2002 Probing multiple effects on 15N,13Ca, 13Cb, and 13C 0 chemical shifts in peptides usingdensity functional theory. Biopolymers 65, 408–423.

Yao, X. L. & Hong, M. 2002 Determination of Ca chemicalshift tensor orientation in peptides by dipolar-modulatedchemical shift recoupling NMR spectroscopy. J. Am.Chem. Soc. 124, 2730–2738.

Zhang, Y., Mao, J. H. & Oldfield, E. 2002a Fe-57 Mossbauerisomer shifts of heme protein model systems: electronicstructure calculations. J. Am. Chem. Soc. 124, 7829–7839.

Zhang, Y., Mao, J. H. & Godbout, N. 2002b Mossbauerquadrupole splittings and electronic structure in hemeproteins and model systems: a density functional theoryinvestigation. J. Am. Chem. Soc. 124, 13 921–13 930.