16.01.2013 Views

v - ISMAR

v - ISMAR

v - ISMAR

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

BULLETIN OF MAGNETIC RESONANCE<br />

The Quarterly Review Journal of the<br />

International Sodiety of Magnetic Resonance<br />

VOLUME 14 October 1992 NUMBERS 1-4<br />

,\ Proceedings<br />

*' International<br />

Society of ;<br />

; Magnetic<br />

Resonance<br />

Xlth,<br />

- .Meeting<br />

•>• 4<br />

1- a<br />

VF:<br />

]\n\ 19 :i<br />

\ \NC()U\ TR<br />

BC,C\N\O\


BULLETIN OF MAGNETIC RESONANCE<br />

The Quarterly Review Journal of the<br />

International Society of Magnetic Resonance<br />

Editor:<br />

DAVID G. GORENSTEIN<br />

Department of Chemistry<br />

Purdue University<br />

West Lafayette, IN 47907 USA.<br />

Fax: 317-494-0230<br />

INTERNET:david@chem.purdue.edu<br />

Editorial Board:<br />

E.R.ANDREW LAWRENCE BERLINER ROBERT BLINC<br />

University of Florida Ohio State University E. Kardelj University of Ljubljana<br />

Gainesville, Florida, U.S.A. Columbus, Ohio, U.S.A. Ljubljana, Yugoslavia<br />

H.CHIHARA GARETH R. EATON DANIEL FIAT<br />

Osaka University University of Denver University of Illinois at Chicago<br />

Toyonaka, Japan Denver, Colorado, U.S.A. Chicago, Illinois, U.S.A.<br />

SHIZUO FUJIWARA DAVID GRANT ALEXANDER PINES<br />

University of Tokyo University of Utah University of California<br />

Bunkyo-Ku, Tokyo, Japan Salt Lake City, Utah, U.S.A. Berkeley, California, U.S.A.<br />

MIK PINTAR CHARLES P. POOLE, JR. BRIAN SYKES<br />

University of Waterloo University of South Carolina University of Alberta<br />

Waterloo, Ontario, Canada Columbia, South Carolina, U.S.A. Edmonton, Alberta, Canada<br />

The Bulletin of Magnetic Resonance is a quarterly review journal by the International Society of<br />

Magnetic Resonance. Reviews cover all parts of the broad field of magnetic resonance, viz.. the<br />

theory and practice of nuclear magnetic resonance, electron paramagnetic resonance, and nuclear<br />

quadrupole resonance spectroscopy including applications in physics, chemistry, biology, and<br />

medicine. The BULLETIN also acts as a house journal for the International Society of Magnetic<br />

Resonance.<br />

CODEN: BUMRDT ISSN: 0163-559X<br />

Bulletin of Magnetic Resonance, The Quarterly Journal of International Society of Magnetic<br />

Resonance. 1992 copyright by.the International Society of Magnetic Resonance. Rates: Libraries<br />

and non-<strong>ISMAR</strong> members $80.00, members of JSMAR, $25.00. All subscriptions are for a volume<br />

year. All rights reserved. No part of this journal may be reproduced in any form for any purpose or by<br />

any means, abstracted, or entered into any data base, electronic or otherwise, without specific<br />

permission in writing from the publisher.


J. ANGLISTER<br />

Israel<br />

R. BLINC<br />

Yugoslavia<br />

P.T. CALLAGHAN<br />

New Zealand<br />

L.G.CONTI<br />

Italy<br />

E.L.HAHN<br />

U.S.A.<br />

M.J.R. HOCH<br />

5. Africa<br />

P.C. LAUTERBUR<br />

U.S.A.<br />

M.MEHRING<br />

Germany<br />

C.P: POOLE<br />

U.S.A.<br />

G.C.K. ROBERTS<br />

England<br />

N.V. VUGMAN<br />

Brazil<br />

C.S. YANNONI<br />

U.S.A.<br />

Council of the International Society of Magnetic Resonance<br />

President: R. FREEMAN, England<br />

Vice-President: A. PINES, U.S.A.<br />

Founding Chairman: D. FIAT, U.S.A.<br />

Secretary-General: R.K. HARRIS, England<br />

Treasurer: R.R. VOLD, U.S.A.<br />

Past President: C.P. SLICHTER, U.S.A.<br />

E.D. BECKER<br />

U.S.A.<br />

G. BODENHAUSEN<br />

Switzerland<br />

H.CHIHARA<br />

Japan<br />

R. DESLAURIERS<br />

Canada<br />

KM. HAUSSER<br />

Germany<br />

C.L. KHETRAPAL<br />

India<br />

E. LIPPMAA<br />

Estonia<br />

H. PFEIFER<br />

Germany<br />

M. PUNKINEN<br />

Finland<br />

P. SERVOZ-GAVESf<br />

France<br />

J.S. WAUGH<br />

U.S.A.<br />

MR. BENDALL<br />

Australia<br />

W.S. BREY<br />

U.S.A.<br />

S. CLOUGH<br />

England<br />

R.R. ERNST<br />

Switzerland<br />

J.W. HENNEL<br />

Poland<br />

VJ. KOWALEWSKI<br />

Argentina<br />

B. MARAVIGLIA<br />

Italy<br />

M.M. PINTAR<br />

Canada<br />

L.W. REEVES<br />

Canada<br />

J. STANKOWSKI<br />

Poland<br />

K. WUTHRICH<br />

Switzerland<br />

The aims of the International Society of Magnetic Resonance are to advance and diffuse knowledge<br />

of magnetic resonance and its applications in physics, chemistry, biology, and medicine, and to<br />

encourage and develop international contacts between scientists.<br />

The Society sponsors international meetings and schools in magnetic resonance and its applications<br />

and publishes the quarterly review journal. The Bulletin of Magnetic Resonance, the house journal of<br />

<strong>ISMAR</strong>.<br />

The annual fee for <strong>ISMAR</strong> membership is $20 plus $25 for a member subscription to the Bulletin of<br />

Magnetic Resonance.<br />

Send subscription to: International Society of Magnetic Resonance<br />

Professor Regitze R. Void, Treasurer<br />

Department of Chemistry, 0342<br />

University of California, San Diego<br />

9500 Gilman Drive<br />

La Jolla, CA 92093-0342<br />

(619) 534-0200; FAX (619) 534-7042<br />

Bitnet: rrvold@ucsd.bitnet


Vol. 14, No. 1-4<br />

ORGANIZING COMMITTEE<br />

for<br />

<strong>ISMAR</strong> 92<br />

C. Fyfe, Chairman<br />

Department of Chemistry<br />

University of British Columbia<br />

Vancouver, British Columbia<br />

CANADA<br />

R. Andersen<br />

Department of Chemistry<br />

University of British Columbia<br />

Vancouver, British Columbia<br />

CANADA<br />

G. S. Bates, Treasurer<br />

Department of Chemistry<br />

University of British Columbia<br />

Vancouver, British Columbia<br />

CANADA<br />

M. Bloom<br />

Department of Physics<br />

University of British Columbia<br />

Vancouver, British Columbia<br />

CANADA<br />

E. E. Burnell<br />

Department of Chemistry<br />

University of British Columbia<br />

Vancouver, British Columbia<br />

CANADA<br />

G. Drobny<br />

Department of Chemistry<br />

University of Washington<br />

Seattle, Washington<br />

USA<br />

I. Gay<br />

Department of Chemistry<br />

Simon Fraser University<br />

Burnaby, British Columbia<br />

CANADA<br />

F. G. Herring<br />

Department of Chemistry<br />

University of British Columbia<br />

Vancouver, British Columbia<br />

CANADA<br />

CO-EDITORS OF THE PROCEEDINGS<br />

D. G. Gorenstein<br />

Department of Chemistry<br />

Purdue University<br />

West Lafayette, Indiana<br />

USA<br />

C. Fyfe<br />

Department of Chemistry<br />

University of British Columbia<br />

Vancouver, British Columbia<br />

CANADA


4 Bulletin of Magnetic Resonance<br />

ACKNOWLEDGMENTS<br />

The Organizing Committee of the Xlth conference of the International Society of Magnetic Resonance<br />

gratefully acknowledges the financial support of the following organizations:<br />

B.P. Research, pic<br />

Bio-Mega, Inc.<br />

Bio Rad - Sadtler Division<br />

Bruker Spectrospin Canada Ltd.<br />

Bulletin of Magnetic Resonance<br />

Chemagnetics/Otsuka Electronics<br />

Dean of Graduate Studies, U.B.C.<br />

Dean of Science, U.B.C.<br />

Doty Scientific, Inc.<br />

Dow Chemical Co.<br />

ICON (Isotope) Services Inc.<br />

JEOL USA, Inc.<br />

John Wiley & Sons<br />

Molecular Simulations<br />

Oxford Instruments<br />

Pfizer<br />

Quadra Logic Technologies, Inc.<br />

Varian NMR Instruments<br />

Special thanks are extended to the National Sciences and Engineering Research Council of Canada (NSERC)<br />

for its support through the award of a Conference Grant to <strong>ISMAR</strong> 92. We are also grateful to the Department<br />

of Chemistry at the University of British Columbia for its support.


Vol. 14, No. 1-4 5<br />

Contents<br />

Proceedings of the Xlth Meeting of the<br />

International Society of Magnetic Resonance<br />

July 19 - 24, 1992<br />

Vancouver, B. C. Canada<br />

Solid-State Polarization-Transfer Experiments Involving Quadrupolar Nuclei,<br />

K. T. Mueller, C. A. Fyfe, H. Grondey, K. C. Wong-Moon and T. Markus 9<br />

Magnetic Resonance Evidence for Superconductivity in a Semimetal, I. P. Goudemond,<br />

G. J. Hill and M. J. R. Hoch 12<br />

NMR Spectroscopy in Cardiac Surgery, R. Deslauriers, S. Lareau, R. S. Labow,<br />

W. J. Keon, G-H. Tian, A. L. Panos, C. A. M. Barrozo, I. S. Ali, O. Al-Nowaiser and T. A. Salerno 15<br />

Topology and Spin Alignment in Organic High-Spin Molecules, Y. Teki, K. Sato,<br />

M. Okamoto, A. Yamashita, Y. Yamaguchi, T. Takui, T. Kinoshita and K. Itoh 24<br />

New Developments in Pulsed Electron Paramagnetic Resonance: Relaxation Mechanisms<br />

of Nitroxide Spin Labels, C. Mailer, B. H. Robinson and D. A. Haas 30<br />

New Developments in Pulsed Electron Paramagnetic Resonance: Direct Measurement<br />

of Rotational Correlation Times from Decay Curves, D. A. Haas, C. Mailer,<br />

T. Sugano and B. H. Robinson 35<br />

Non-Linear Effects in Standard 2D NOE Experiments in Coupled Spin Systems,<br />

R. C. R. Grace and A. Kumar 42<br />

Deriving Structures from 2D NMR. A Method for Denning the Conformation of<br />

a Protein Adsorbed to Surfaces, D. A. Keire and D. G. Gorenstein 57<br />

3D-Structure Determination of Flavoridin in Solution: New Computational Strategy<br />

for Disulfide-Bridge Mapping, H. Senn, W. Klaus and P. Gerber 64<br />

NMR Approaches to Large Proteins: trp Repressor and Chloramphenicol Acetyltransferase,<br />

L.-Y. Lian, J. P. Derrick, V. Ramesh, R. O. Frederick, S. E. H. Syed and<br />

G. C. K. Roberts 65<br />

2D NMR Study of Drug-Protein Interactions: Ethidium Bromide - Neocarzinostatin<br />

Complex, S. Mohanty, L. C. Sieker and.G. P. Drobny 68<br />

Quantitative Analysis in Multi-Dimensional Transferred NOE Experiments: Improved<br />

Spectral Acquisition and Processing, F. Ni 75<br />

Time-Resolved Solid-State NMR: Small Molecules and Enzymes in Rapidly Frozen<br />

Solution, J. N. S. Evans, R. J. Appleyard and W. Shuttleworth 81<br />

Coupled Methyl Groups in Dimethyl Sulphide, M. R. Johnson, S. Clough, A. J. Horsewill<br />

and I. B. I. Tomsah 86


Bulletin of Magnetic Resonance<br />

NMR Relaxation Studies of Microdynamics in Chloroaluminate Melts, P. A. Shaw,<br />

W. R. Carper, C. E. Keller and J. S. Wilkes .92<br />

Structure and Dynamics of a Membrane Bound Polypeptide, T. A. Cross,<br />

R. R. Ketchem, W. Hu, K.-C. Lee, N. D. Lazo and C. L. North 96<br />

The Role of Metal Ions in Processes of Conformational Selection during Ligand-<br />

Macromolecule Interactions, E. Gaggelli, N. Gaggelli, G. Valensin and A. Maccotta .... 102<br />

Detection and Characterization of CFC, HCFC and HFC Gases in Foamed Insulation<br />

by High Field NMR Imaging, L. H. Randall, C. A. Fyfe, Z. Mei and S. Whitworth 108<br />

Mysterious Negative Peaks in the 1 H{ 1 H}NOE Difference Spectra of Some Thiopyran<br />

Compounds, C. Szantay, Jr 112<br />

H-l and C-13 NMR Spectra of the Carbanions Produced from Phenylpropene<br />

Derivatives, A. Yoshino, K. Aoki, M. Ushio and K. Takahashi 116<br />

Intracellular pH and Inorganic Phosphate Effects on Skeletal Muscle Force,<br />

E. R. Barton-Davis, R. W. Wiseman and M. J. Kushmerick 122<br />

A Simple Model for the Influence of Motion on the NMR Line Shape, M. Goldman,<br />

T. Tabti, C. Fermon, J. F. Jacquinot and G. Saux 126<br />

The Effect on Ti of Correlated Water Motions in the Polar Phase of Colemanite,<br />

J. Sun and A. Watton 131<br />

Measurement of Deuteron Spin Relaxation Times in Liquid Crystals by a Broadband<br />

Excitation Sequence, R. Y. Dong 134<br />

Carbon-13 Relaxation Mechanisms and Motional Studies in Selected Halomethane<br />

Molecules, A. A. Rodriguez, T. Davis and L. E. Nance 139<br />

An Efficient Large Sample Volume System for Solid State NMR, R. J. Pugmire,<br />

Y. J. Jiang, M. S. Solum and D. M. Grant . . . . 144<br />

Magnetic Resonance Spectroscopic Investigations of Poly(p-Phenylene Sulfide/Disulfide),<br />

PPS/DS, D. W. Lowman and D. R. Fagerburg 148<br />

Application of 2-D HETCOR NMR to Investigate Polymer Blend Heterogeneity,<br />

S. Kaplan 153<br />

New High Resolution NMR Studies in Polycrystalline Tetracyanoquinodimethane,<br />

M. T. Nunes, A. Vainrub, M. Ribet, F. Rachdi, P. Bernier, M. Almeida and G. Feio 156<br />

Use of NMR Relaxation Measurements to Derive the Binding Site of Plastocyanin<br />

in Complexes With Cytochrome-F and C, S. Modi, E. McLaughlin, D. S. Bendall, S. He<br />

and J. C. Gray 159<br />

Metal-peptide Interaction: Influence of the Aminoacid Sequence on the Binding<br />

of Co(II) to Glycyltryptophan and Tryptophylglycine Studied by X H NMR and<br />

Fluorescence, A. Spisni, G. Sartor, L. Franzoni, A. Orsolini, P. Cavatorta and M. Tabak . . . 165<br />

Assignments of the X H NMR Spectrum of a Consensus DN A-Binding Peptide from<br />

the HMG-I Protein, J. N. S. Evans, M. S. Nissen and R. Reeves 171<br />

Solution Structure of the DNA-binding Domain of GAL4 from Saccharomyces cerevisiae,<br />

J. D. Baleja, V. Thanabal, T. Mau and G. Wagner . 175


Vol. 14, No. 1-4 7<br />

Structural Investigation of Folic Acid by NMR Proton Relaxation and Molecular<br />

Mechanics Analysis, C. Rossi, A. Donati, S. Ulgiati, and M. R. Sansoni 181<br />

Characterization of Water-in-Bitumen Emulsions in Model Porous Media by NMR<br />

Microscopic Imaging Techniques, L. H. Randall and G. E. Sedgwick and C. A. Fyfe .... 186<br />

Computer Graphics for Pulse Sequence Analysis, J. Callahan, D. Mattiello and<br />

G. P. Drobny 191<br />

NMR Investigation of the Simultaneous Fermentation of Xylose and Glucose by<br />

a Selected Strain of Klebsiella Planticola (Gil), C. Rossi, A. Lepri, M. P. Picchi, S. Bastianoni,<br />

D. Medaglini, M. Vanassina and E. Cresta 197<br />

Interleukin-1 Receptor Antagonist Protein: Solution Secondary Structure from<br />

NOE's and l Ha and 13 Ca Chemical Shifts, B. J. Stockman, T. A. Scahill, A. Euvrard,<br />

N. A. Strakalaitis, D. P. Brunner, A. W. Yem and M. R. Deibel, Jr. 202<br />

Green's Function Calculation of Effective Nuclear Relaxation Times in Metals and<br />

Disordered Metals, M. Martin-Landrove and J. A. Moreno 208<br />

Stochastic Averaging Revisited, D. H. Jones, N. D. Kurur and D. P. Weitekamp 214<br />

Magnetic Resonance of Trapped Ions by Spin-Dependent Cyclotron Acceleration,<br />

P. J. Pizarro and D. P. Weitekamp . . 220<br />

Coordination Modes of Histidine Moiety in Copper (II) Dipeptide Complexes Detected<br />

by Multifrequency ESR, R. Basosi, R. Pogni, and G. D. Lunga 224<br />

An EPR and ab initio Study of a Phosphaalkene Radical Anion, and Comparison<br />

with other Phosphorus-Containing Radical Ions, M. Geoffroy, G. Terron, A. Jouaiti,<br />

P. Tordo and Y. Ellinger 229<br />

Conformational Substate Distribution in Myoglobin as Studied by EPR Spectroscopy,<br />

A. R. Bizzarri and S. Cannistraro . 234<br />

Effect of Paramagnetic Ions in Aqueous Solution for Precision Measurement of the<br />

Proton Gyromagnetic Ratio, A. R. Lim, C. S. Kim and S. H. Choh 240<br />

Magnetic Resonance of 23 Na and 14 N Nuclei in Single and Multi-Domain Crystals<br />

of Ferroelectric NaNO2, S. H. Choh and K. T. Han 246<br />

Knight Shifts and Spin Dynamics in Disordered Systems, M. J. R. Hoch and<br />

S. T. Stoddart .252<br />

Numerical Design and Evaluation of Broadband Pulse Sequences for 1=1 Spin<br />

Systems, D. L. Mattiello, J. Callahan, T. Alam and G. P. Drobny 255<br />

A BASIC Program to Calculate the Evolution of Cartesian Product Operators,<br />

S. Mammi 259<br />

Selective Long-Range Polarization Transfer via DEPT, T. Parella, F. Sanchez-Ferrando<br />

and A. Virgili . . 263<br />

Computer Simulations of High Resolution NMR Spectra, S. A. Smith, W, E. Palke<br />

and J.T. Gerig 267<br />

Variation of 13 C NMR Linewidths of Metallocenes as a Function of Magic Angle<br />

Sample Spinning Frequency, I. J. Shannon, K. D. M. Harris and S. Arumugam 273


Bulletin of Magnetic Resonance<br />

The Structural Role of Water in Silicate Glasses: X H and 29 Si NMR Evidence,<br />

J. Kiimmerlen, T. Schaller, A. Sebald and H. Keppler 278<br />

High-Resolution Solid-State NMR Study of Microstructures in Layered Aluminosilicate,<br />

S. Hayashi, T. Ueda, K. Hayamizu and E. Akiba 282<br />

Broadline NMR of Structural Ceramics, C. Connor 285<br />

Permeability of Liposomal Membranes to Molecules of Environmental Interest:<br />

Results from NMR Experiments Employing Shift Agents, F. G. Herring, W. R. Cullen,<br />

J. C. Nelson and P. S. Phillips 289<br />

Nuclear Magnetic Resonance Partitioning Studies of Solute Action in Lipid Membranes,<br />

L. Ma, T. F. Taraschi and N. Janes 293<br />

Weak Molecular Interactions: NMR Spectroscopy of Oriented Molecules,<br />

C. L. Khetrapal 299<br />

Structural Studies of Collagen by Solid State NMR, R. J. Wittebort and A. M. Clark . 303<br />

Calendar of Forthcoming Conferences 307<br />

Recent Magnetic Resonance Books 308<br />

Instructions for Authors 314


Vol. 14, No. 1-4<br />

Solid-State Polarization-Transfer Experiments<br />

Involving Quadrupolar Nuclei<br />

K. T. Mueller, C. A. Fyfe, H. Grondey, K. C. Wong-Moon and T. Markus<br />

1 Introduction<br />

Department of Chemistry, University of British Columbia<br />

2036 Main Mall, Vancouver, BC Canada V6T 1Z1<br />

Historically, cross-polarization experiments<br />

[1,2] have been used to transfer spin<br />

coherence from abundant spins to a dilute spin<br />

system. Protons (*H) have been used almost<br />

exclusively as the source of strong nuclear<br />

polarization for cross-polarization experiments,<br />

although there have been some examples where<br />

other abundant nuclei have been used. Coupled<br />

with magic-angle spinning (MAS) NMR [3],<br />

cross-polarization techniques have proven<br />

extremely powerful for the study of organic<br />

solids.<br />

Inorganic systems such as zeolites, gels,<br />

and ceramics are of great technological<br />

importance and interest and contain many<br />

quadrupolar nuclei but very few protons. For<br />

quadrupolar nuclei with non-integral spins such<br />

as n B, 17 O, and 27 A1, the second-order<br />

quadrupolar broadening of the readily observed<br />

central (+1/2 -1/2) transition is not<br />

completely averaged by MAS, and the NMR<br />

lines from quadrupolar spins are shifted and<br />

distorted in single-axis spinning experiments<br />

[4,5]. Very few examples exist of crosspolarization<br />

experiments involving quadrupolar<br />

nuclei, and they all involve magnetization<br />

transfer from protons to quadrupolar nuclei. We<br />

have undertaken a study to determine the<br />

feasibility of polarization transfer and dipolar<br />

dephasing experiments between spin pairs in<br />

these systems, particularly between 3l P (I = 1/2)<br />

and 27 A1 (I = 5/2). Our preliminary results show<br />

that these experiments are indeed possible [6],<br />

The attainment of cross-polarization from<br />

quadrupolar spin systems is particularly<br />

important in materials chemistry as these nuclei<br />

usually have very short Ti relaxation times.<br />

Spin-1/2 nuclei in dense inorganic systems may<br />

have Ti values ranging from many seconds to<br />

hours, effectively precluding their observation in<br />

many instances. By using cross-polarization<br />

from the quickly relaxing quadrupolar spins,<br />

spectra of the spin-1/2 nuclei could be obtained<br />

in a relatively short time. Additional<br />

information regarding the local structure and<br />

bonding in these systems might also be obtained<br />

through the distance dependence of the crosspolarization<br />

process.<br />

Similarly, dipolar-dephasing NMR<br />

experiments such as rotational-echo doubleresonance<br />

(REDOR) and transferred-echo doubleresonance<br />

(TEDOR) have been demonstrated to<br />

be useful for demonstrating connectivities and<br />

determining internuclear distances [7,8] in<br />

heteronuclear spin systems with dipolar<br />

couplings. Experimental verification of these<br />

experiments with the same heteronuclear spin<br />

pair ( 31 P and 27 A1) demonstrates the feasability<br />

of applying these techniques to systems<br />

containing quadrupolar nuclei.<br />

2 Experimental<br />

The sample chosen for study was the very<br />

large pore molecular sieve VPI-5, an<br />

aluminophosphate dihydrate containing 18membered-rings<br />

[9]. NMR experiments were<br />

performed under MAS conditions in a 9.4 T<br />

superconducting magnet where the resonance<br />

frequencies for 31 P and 27 A1 are 161.98 MHz<br />

and 104.26 MHz respectively. The rotational<br />

frequencies in all experiments were<br />

approximately 3.1 kHz, and 90° pulse times for<br />

the nuclei studied ranged from 9 to 12 (xsec.<br />

3 Results<br />

The spectra in Figure 1 demonstrate the<br />

transfer of magnetization using cross-polarization<br />

in both directions between the 27 A1 and 31 P<br />

spins in the Al-O-P bonding units in VPI-5.<br />

The cross-polarization is accomplished with an<br />

appropriate spin-locking pulse sequence [2] after<br />

a preparation pulse creates spin coherence for the<br />

nuclei used as the polarization source. With<br />

MAS the 31 P chemical shift anisotropies are<br />

averaged to their isotropic values for the three<br />

crystallographically inequivalent 31 P sites in the<br />

unit cell. For the 27 A1 nuclei, MAS partially<br />

averages the second-order quadrupolar interaction<br />

and two resonances are seen: One from the


10<br />

(b)<br />

0 -SO '<br />

ppm from 85% H3PO4<br />

(d)<br />

» 0 -JO<br />

ppm from A1(NO3>3<br />

Figure 1. Demonstration of 27 A1 —> 31 P and<br />

31 P —> 27 A1 cross-polarization in VPI-5<br />

(projection of unit cell shown at top). Spinning<br />

sidebands are marked with an (s).<br />

tetrahedrally coordinated aluminum sites (41<br />

ppm) and a second from the octahedrally<br />

coordinated aluminum (approximately -18 ppm).<br />

The observed signals are solely due to crosspolarization<br />

and not caused by direct irradiation<br />

during the spin-lock as proven by a series of<br />

cross-check experiments, of which spectra (b) and<br />

(d) of Figure 1 are representative.<br />

A two-dimensional heteronuclear<br />

correlation experiment [10] using crosspolarization<br />

can be performed by preparing the<br />

aluminum spins with a 90° pulse, and then<br />

encoding their evolution frequencies in an initial<br />

time period. The aluminum polarization is<br />

subsequently transferred to the phosphorus spins<br />

with a spin-lock, and a phosphorus free induction<br />

decay is accumulated after each of a set of<br />

aluminum evolution times. Two-dimensional<br />

Fourier transformation provides the correlation<br />

spectrum of Figure 2. From the twodimensional<br />

spectrum it is evident that each of<br />

the three 31 P resonances is connected to both<br />

tetrahedral and octahedral 27 A1 resonances, in<br />

agreement with the proposed crystal structure of<br />

VPI-5 [11].<br />

The REDOR experiment [7] was carried<br />

out in both directions between 27 A1 and 31 P<br />

nuclei in VPI-5, and the results are shown in<br />

Figure 3. In both directions, negative crosscheck<br />

experiments (not shown) were undertaken<br />

to ensure that the observed signal came from the<br />

dipolar dephasing phenomenon and not from<br />

timing missets or experimental artifacts.<br />

00<br />

3<br />

100 JO 0 - 5 0<br />

ppm from A1(NO3)3<br />

Bulletin of Magnetic Resonance<br />

Figure 2. Two-dimensional heteronuclear<br />

correlation spectrum of - 7 A1 and 31 P in VPI-5.<br />

The TEDOR experiment [8] was also<br />

accomplished with initial evolution of 27 A1<br />

spins and subsequent transfer to the 31 P after two<br />

rotor periods of preparative dephasing. The<br />

signal in Figure 3(c) was obtained after one<br />

additional period of dipolar evolution in order to<br />

create observable spin coherence from the<br />

antiphase signal which was transferred at the time<br />

of the 27 A1 spin echo. As before, negative<br />

cross-check experiments were performed and gave<br />

null signals.<br />

A two-dimensional TEDOR experiment.<br />

[12] was performed with 27 A1 spin frequency<br />

encoding before the initial dipolar dephasing<br />

period. After transfer of the coherence to the 31 P<br />

spins, an FID was accumulated and the ti value<br />

incremented. The two-dimensional correlation<br />

spectrum is shown in Figure 4, revealing crosspeaks<br />

between all three 31 P resonances and both<br />

the resonances from the tetrahedrally coordinated<br />

and octahedrally coordinated 27 A1 sites, in<br />

agreement with the proposed crystal structure and<br />

the results of the two-dimensional CP<br />

experiment discussed above. These results are<br />

taken as demonstrating the general success of the<br />

experiments. A more detailed interpretation of<br />

the results in terms of the individual T-sites from<br />

the present data alone is not attempted.


Vol. 14, No. 1-4 11<br />

oc<br />

"P REDOR<br />

IS C -B -SO -73<br />

ppm from 85% H3PO4<br />

Wrahedral<br />

r AI REDOR<br />

IX 75 X 25 0 -23 -SO<br />

ppm from A1(NO})3<br />

P TEDOR<br />

0 .J5 -JO -75<br />

ppm from 85% H3PO4<br />

Figure 3. One-dimensional REDOR and<br />

TEDOR spectra of 27 A1 and 3I P in VPI-5.<br />

SO 25 0 -25<br />

ppm from A1(NO3>3<br />

Figure 4. Two-dimensional TEDOR spectrum of<br />

27 Al and 31 P in VPI-5.<br />

4 Conclusions<br />

In summary, cross-polarization to and<br />

from quadrupolar nuclei has been experimentally<br />

verified using the 31 P and 27 A1 spin systems in<br />

an aluminophosphate molecular sieve. This<br />

bodes well for the use of heteronuclear<br />

correlations for further investigation of local<br />

micrdstructure in solids. Dipolar-dephasing<br />

experiments have also been accomplished, with<br />

both REDOR and TEDOR results confirming the<br />

connectivities detected by the cross-polarization<br />

studies. A two-dimensional TEDOR experiment<br />

has also been demonstrated that separates<br />

connectivities between distinct resonances.<br />

References<br />

1 S. R. Hartmann and E. Hahn Phys. Rev.,<br />

128, 2042, 1962.<br />

2 A. Pines, M. G. Gibby, and J. S. Waugh /.<br />

Chem. Phys., 59, 569, 1973.<br />

3 J. Schaefer and E. 0. Stejskal J. Am. Chem.<br />

Soc.,9%, 1031, 1976.<br />

4 H. J. Behrens and B. Schnabel Physic a,<br />

114B, 185. 1982.<br />

->A. Samoson. E. Kundla, and A. Lippmaa J.<br />

Magn. Reson.. 49, 350, 1982.<br />

6 C. A. Fyfe, H. Grondey, K. T. Mueller, K. C.<br />

Wong-Moon, and T. Markus /. Am. Chem.<br />

Soc, 114, 5876, 1992.<br />

7 T. Gullion and J. Schaefer J. Magn. Reson.,<br />

81, 196, 1989.<br />

8 Y. Pan and J. Schaefer J. Magn. Reson., 90,<br />

341, 1990.<br />

9 M. E. Davis, C. Sadarriaga, C. Montes, J.<br />

Garces, and C. Crowder Nature (London), 331,<br />

698, 1988.<br />

l0 P. Caravatti, G. Bodenhausen, and R. R.<br />

Ernst Chem. Phys. Lett., 89, 363, 1982.<br />

n L. B. McCusker, Ch. Baerlocher, E. Jahn,<br />

and M. Bulow Zeolites. 11, 308, 1991.<br />

12 C. A. Fyfe, K. T. Mueller. H. Grondey, and<br />

K. C. Wong-Moon, submitted to Chem. Phys.<br />

Lett.


12<br />

1. Introduction<br />

Magnetic Resonance<br />

Evidence for Superconductivity<br />

in a Semimetal<br />

I.P. Goudemond, G. J. HiU and M.J.R. Hoch<br />

Department of Physics and<br />

Condensed Matter Physics Research Unit,<br />

University of the Witwatersrand, Johannesburg<br />

The group V semimetals As, Sb and Bi have<br />

low carrier densities and may be viewed<br />

as rather poor metals. Their electronic<br />

properties have been studied using a variety<br />

of methods, including NQR [1], [2]. At<br />

temperatures below the Debye temperature<br />

the nuclear spin—lattice relaxation rate in As<br />

and Sb has been found to obey the Korringa<br />

relation [1], [2]. For As, this relation has<br />

been found to hold down to 150 mK [3].<br />

In the present work Ti measurements<br />

on As have been extended to still lower<br />

temperatures. Motivation has come from<br />

the interesting electrical conductivity<br />

behaviour found by Uher [4] for a single<br />

crystal sample in the vicinity of 100 mK.<br />

These results may be interpreted as evidence<br />

for a superconducting transition, although no<br />

further experiments appear to have been<br />

carried out to confirm this. Probing of the<br />

superconducting state in zero magnetic field<br />

using NQR methods offers interesting<br />

challenges and opportunities.<br />

2. Experimental Details<br />

The experiments were carried out in an<br />

Oxford dilution refrigerator using procedures<br />

that have been described previously [3].<br />

Pulsed NQR spin echo methods with signal<br />

averaging were used at 23.5 MHz on a<br />

powdered As sample. ( 75 As has I = 3/2 and<br />

100% abundance). The powdered material<br />

was prepared by crushing and sieving (25 \i<br />

mesh) high purity (99.9995%) arsenic,<br />

Bulletin of Magnetic Resonance<br />

followed by annealing in vacuo and further<br />

careful sieving. An oxide layer on the<br />

surface of the grains prevented metallic<br />

contact between neighbouring particles.<br />

Refrigerator<br />

Mixing Chamber<br />

Sintered Silver<br />

Heat Exchanger<br />

Vacuum Space<br />

Coaxial Cable<br />

Helium Fill<br />

Capillary<br />

Copper Block<br />

Cermanium<br />

Thermometer<br />

R F Coil<br />

Stycast Sample<br />

Holder<br />

Figure 1<br />

Sample holder and rf coil assembly for NQR<br />

measurements in the dilution refrigerator. The<br />

sample is immersed in liquid *Be, which is in<br />

contact with the sintered silver heat<br />

exchanger.<br />

Figure 1 depicts the sample<br />

arrangement used to ensure good thermal<br />

contact to the refrigerator mixing chamber.<br />

Liquid 4 He surrounds the sample and is<br />

in contact with a sintered silver heat<br />

exchanger. Further details may, be found<br />

in reference [3]. Fractional 'rf pulses<br />

were used to minimize heating effects.<br />

Temperatures were measured using a<br />

calibrated germanium thermometer mounted<br />

on the mixing chamber.


Vol. 14, No. 1-4 13<br />

3. Results and Discussion<br />

In an effort to establish that the sample<br />

was in good thermal contact with it's<br />

surroundings, careful measurements of the<br />

echo amplitude were made as a function<br />

of temperature down to the lowest<br />

temperatures reached in these experiments,<br />

40 mK. Down to 150 rnK Curie law type<br />

behaviour is observed. At lower<br />

temperatures the data depart from linear<br />

behaviour. We do not believe that this is<br />

due to heating effects or the loss of thermal<br />

contact. Changes in the pulse sequence<br />

repetition rate did not change the amplitude<br />

of the echo signal. It is likely that some<br />

mechanism, characteristic of the sample, is<br />

responsible for the departure from Curie law<br />

behaviour.<br />

At temperatures of 4 K and below, the<br />

skin depth 6 at 23.5 MHz is comparable to,<br />

or less than, the mean particle radius r. We<br />

estimate that 5 ~ .3.0 [im at 150 rnK, while<br />

r < 10 fim. It is clearly desirable that<br />

smaller particles should be used, although<br />

this is difficult because of rapid surface<br />

oxidation which occurs in air and the<br />

tendency of the arsenic particles to sinter<br />

during annealing.<br />

Taking into account the attenuation of<br />

rf pulses and the presence of a core of<br />

undisturbed spins in the particles, we have<br />

examined the situation in some detail. In<br />

order to see whether spin diffusion can<br />

operate between spins near the surface of a<br />

particle and those in the interior, we have<br />

calculated the spin diffusion coefficient<br />

D = V30 V^d 2 , where M2 is the second<br />

moment and d the spin spacing, and obtain<br />

D ~ 2 x 10" 13 cm 2 s"'. On the time—scale of<br />

our experiments (10 2 — 10 3 s) spin diffusion<br />

operates over a distance v/2Dt ~10~ l fim.<br />

It may be concluded that this mechanism<br />

should have negligible effects on our results.<br />

The change in the skin depth with<br />

temperatures below 1 K is of the order of 1<br />

to 2%, which our calculations show will not<br />

lead to detectable changes in the echo<br />

amplitude beyond the Curie law changes.<br />

We do not believe that the departure from<br />

Curie law behaviour is due to effects of this<br />

kind.<br />

The measured spin lattice relaxation<br />

rates are shown as a function of temperature<br />

in Figure 2. Note that, for magnetic<br />

relaxation in an I = 3 /2 system, a unique<br />

relaxation rate may be defined using<br />

1/T] = 6 Wm, where Wn, is the transition<br />

rate between the ±'/2 and ± 3 /2 spin states.<br />

At temperatures down to roughly<br />

120 mK the data obey the Korringa relation.<br />

Below this temperature the relaxation rate<br />

0.09<br />

8.64<br />

0.01 8.62<br />

30 50 70 90 110 130 150 170<br />

T (mK)<br />

Figure 2<br />

Plots of the ?!>As Spin lattice relaxation rate<br />

and of the resistivity of a single crystal<br />

arsenic sample [4] as a function of temperature<br />

down to 50 mK. The straight line which joins<br />

with the relaxation rate plot represents an<br />

extra— polation of the Korringa relation for<br />

arsenic found at higher temperatures.<br />

decreases less rapidly with T than expected<br />

from the Korringa relation. Figure 2 also<br />

shows the electrical conductivity data of<br />

Uher [4] in the same temperature range as<br />

the present measurements. It can be seen<br />

that the conductivity starts to decrease quite<br />

rapidly at ~100 mK. Inspection of the two<br />

plots in Figure 2 suggests a common<br />

underlying physical mechanism for the<br />

changes in behaviour observed in this<br />

temperature range.<br />

It appears likely that a fairly broad<br />

superconducting transition occurs with a Tc<br />

around 100 fiK. The T"i data suggest a<br />

slightly higher Tc value than the<br />

conductivity data.<br />

Cohen [5] has pointed out that the<br />

semimetals may be candidates for<br />

superconductivity through a BCS pairing<br />

mechanism. This is largely because of the<br />

multivalley character of these materials. On<br />

the basis of calculations given by Cohen for<br />

systems of this kind, a Tc of 100 mK is not<br />

unreasonable for As. Doped Bi has been<br />

found by Uher and Opsal [6] to have<br />

a Tc < 100 mK, depending on the<br />

concentration of the Sn or Te dopant.<br />

We have attempted to fit the observed<br />

relaxation data for As using the Hebel—<br />

Slichter expression based on BCS theory.<br />

Anisotropy of the gap is allowed for by<br />

introducing a parameter r = Ao(0)/A, where<br />

Ao is the BCS gap at 0 K and A is a<br />

measure of the gap anisotropy. However the<br />

theoretical curve does not fit the<br />

experimental data plotted in reduced form.<br />

Details will be given elsewhere. It is quite<br />

possible that the suggested transition in As<br />

is not of the standard BCS type. It appears,


14<br />

however, that other effects could be<br />

important.<br />

Cohen [5] suggests that the semimetals<br />

will become type II superconductors with<br />

rather low upper critical fields. In the NQR<br />

experiment rf fields of 30 or 40 G are used<br />

and it is possible that some remnant field<br />

effects may be produced in the sample which<br />

contribute to relaxation.<br />

Further experiments should be carried<br />

out to confirm the onset of<br />

superconductivity in As around 100 mK.<br />

Clearly, Meissner effect measurements<br />

should be attempted. Further NQR<br />

experiments involving CW methods with<br />

low rf fields may prove useful and complementary<br />

to the present measurements.<br />

4. Conclusion<br />

Evidence has been obtained of marked<br />

departures from Korringa relaxation<br />

behaviour in As below 120 mK. Taken<br />

together with previous electrical<br />

conductivity results, it appears likely that<br />

the anomalous behaviour is due to the onset<br />

of superconductivity.<br />

Bulletin of Magnetic Resonance<br />

The relaxation rate measurements are<br />

not in agreement with the Hebel—Slichter<br />

theory predictions. Further work is required<br />

to determine whether this is because of the<br />

non—BCS nature of the transition or some<br />

other cause, such as the magnitude of the rf<br />

fields used in our pulsed NQR. experiments.<br />

5. References<br />

1. J.M. Keartland, G.C.K. Folscher and<br />

M.J.R. Hoch, Phys. Rev. B 43, 8362<br />

(1991).<br />

2. J.M. Keartland, G.C.K. Folscher and<br />

M.J.R. Hoch, Phys. Rev. B. 45, 7882<br />

(1992).<br />

3. IP Goudemond, J M Keartland, and<br />

M J R Hoch, J. Low Temp. Physics 82,<br />

369 (1991).<br />

4. C Uher, J. de Physique, C6, 39, 1054<br />

(1978).<br />

5. ML Cohen, in "Superconductivity",<br />

edited by R D Parks (Marcel Dekker,<br />

New York, 1969), Vol.1.<br />

6. C Uher, J L Opsal, Phys. Rev. Letters<br />

40,1518(1978).


Vol. 14, No. 1-4 15<br />

NMR Spectroscopy in Cardiac Surgery<br />

Roxanne Deslauriers, Sylvain Lareau" 1 ", Rosalind S. Labow+, Wilbert J. Keon + , Gang-Hong Tian # , Anthony L. Panos*,<br />

Carlos A.M. Barrozo*, Imtiaz S. Ali *, Owayed Al-Nowaiser*, and Tomas A. Salerno*<br />

1 Introduction<br />

Institute for Biodiagnostics, National Research Council of Canada, Ottawa<br />

+ University of Ottawa Heart Institute, Ottawa<br />

# Department of Physiology, University of Ottawa, Ottawa, Ont.<br />

* Division of Cardiovascular Surgery, University of Toronto, Toronto, Ont., Canada<br />

Applications of magnetic resonance spectroscopy in<br />

medicine have been restricted mostly to the research<br />

laboratory. The technique is now entering the field of<br />

medical diagnosis and therapy. In the heart, levels of<br />

phosphorus metabolites are often correlated with function.<br />

Nuclear magnetic resonance spectroscopy has been used to<br />

monitor high energy phosphorus metabolite levels in the<br />

heart to evaluate the effect of work and ischemic stress. Our<br />

applications of magnetic resonance to the practice of cardiac<br />

surgical have been in three areas: a) preservation of tissue<br />

for transplantation b) optimization of myocardial protective<br />

techniques (cardioplegia) and c) monitoring of the heart<br />

during the aortic clamping period.<br />

2 Heart Preservation for Transplantation<br />

With increasing demand for a limited number of donor<br />

hearts, organ preservation during procurement is critically<br />

important. Controversy still exists over issues such as the<br />

optimal temperature, optimal solution, and maximum time<br />

limit for donor heart preservation. Numerous studies have<br />

been, and continue to be, conducted on various animal<br />

tissues. Not much data have been obtained in human<br />

myocardium primarily because of the difficulty in obtaining<br />

adequate quantities of viable tissue for laboratory<br />

investigation.<br />

a) Human Atrial Tissue<br />

Portions of human atrial appendages normally discarded<br />

during cannulation in the course of surgery requiring<br />

cardiopulmonary bypass have been used, with informed<br />

patient consent, for studies of heart preservation. We have<br />

used 31 P and ^H NMR spectroscopy to define the optimal<br />

temperature for long-term (up to 24 hours) preservation of<br />

high energy metabolite levels and contractile function, and<br />

to gain a fundamental understanding of the energy<br />

generating pathways in preserved human cardiac tissue [1,<br />

2]. The studies were carried out on a Bruker AM-360<br />

spectrometer. 31 P spectra were obtained using a 60° pulse<br />

and a 1 s recycle time. 1 H spectra were acquired using a<br />

spin-echo sequence based on the water-suppressing 1331<br />

pulse sequence. The acquisition of 31 P and *H NMR<br />

spectra were interleaved (Figure 1). 31 P spectra were used<br />

to measure ATP levels on a continuous basis, as an index of<br />

net energy preservation. *H spectra of lactate provided<br />

information on generation of ATP through the glycolytic<br />

pathway.<br />

10 0 -10 -20 1<br />

PPM<br />

PPM<br />

Lactate<br />

FIGURE 1 31 P (left) and l H (right) NMR spectra of an<br />

atrial appendage (ca. 0.5 g) preserved at 20°C in saline, as a<br />

function of time [1] {ref, reference capillary; PME,<br />

phosphomonoester; Pi, inorganic phosphate).<br />

Studies of atrial appendages preserved at 1°, 4°, 12° and<br />

20°C in physiological saline (0.9% NaCl) for up to 20 hours<br />

demonstrate that preservation of ATP is better at 1° and 4°<br />

than at 12° or 20°C. Based on measurements of lactate<br />

production, glycolysis is active at all the temperatures, its<br />

rate correlating positively with increasing temperature.<br />

However, the ATP generated by glycolysis falls short of


16<br />

demand at all temperatures, but the difference is small at 1°<br />

and 4°C (Table 1).<br />

TABLE 1<br />

Energy balance in human cardiac tissue preserved in NaCl<br />

0.9% [1].<br />

Temp.<br />

(°C)<br />

1<br />

4<br />

12<br />

20<br />

ATP* #<br />

loss<br />

7<br />

8<br />

12<br />

20<br />

Lactate*<br />

production<br />

43<br />

52<br />

106<br />

212<br />

ATP*+<br />

generated<br />

65<br />

78<br />

159<br />

318<br />

ATP* &<br />

utilization<br />

72<br />

86<br />

171<br />

338<br />

* nmoleg" 1 (wet myocyte mass) min' 1 .<br />

* From the rate of change of NMR-visible ATP.<br />

+ Assuming 1.5 mole of ATP produced per mole of lactate<br />

from glycolysis.<br />

& Calculated from (rate of generation of ATP by<br />

glycolysis) + (2 x rate of ATP loss). This takes into<br />

account the ATP generated by adenylate kinase.<br />

In a separate series of studies, we tested the possibility<br />

of improving the maintenance of high energy phosphates at<br />

12°C, one of the higher temperatures currently used in some<br />

institutions for the preservation of heart grafts. Our<br />

hypothesis was that the poor maintenance of high energy<br />

phosphates at 12° and 20°C results from the increased<br />

intracellular acidosis that occurs at higher temperatures [1].<br />

Ultimately, acidosis partially inhibits ATP production by<br />

glycolysis, the only metabolic pathway for generation of<br />

ATP in the anoxic heart. We postulated that the addition of<br />

a buffer to the solution used for heart preservation would<br />

increase the rate of transport of protons and lactate to the<br />

extracellular space, thereby maintaining better intracellular<br />

pH homeostasis. Our studies showed that at 12°C, the halftime<br />

for loss of ATP increased from 300 minutes in saline to<br />

over 900 minutes in a modified Krebs-Henseleit solution<br />

containing 100 mM buffer [2, 3]. This observation was<br />

confirmed independently using biochemical measurements<br />

of high energy phosphates [3]. The beneficial effects of<br />

high buffer concentration observed at 12°C did not occur at<br />

4°C (figure 2 [3]), which lead us to postulate that at that<br />

temperature, glycolysis was rate limited by the temperature<br />

rather than by the acidosis. These studies show the<br />

continued need for testing all the conditions to which a graft<br />

may be subjected, and the importance of avoiding broad<br />

Bulletin of Magnetic Resonance<br />

generalizations when dealing with complex multi-enzyme<br />

systems.<br />

4°C<br />

3-j<br />

Q 0.9% saline (3)<br />

3 • 20 mM PIPES (6)<br />

Ito<br />

o 60 mM PIPES (6)<br />

• 100 mM PIPES (6)<br />

2- —<br />

c/5<br />

•d<br />

1-<br />

0-<br />

0 5<br />

2-<br />

1 -<br />

12°C<br />

EP<br />

dm<br />

10 15<br />

Time<br />

><br />

20 25<br />

• 0.9% saline (3)<br />

• 20 mM PIPES (6)<br />

o 60 mM PIPES (6)<br />

• 100 mM PIPES (6)<br />

10 15 20 25<br />

Time (h)<br />

FIGURE 2. Effect of the concentration of PIPES buffer on<br />

the preservation of ATP in isolated atrial appendages<br />

preserved at 4° or 12 °C in modified Krebs-Henseleit<br />

solution [3].<br />

The importance of defining the relationship between<br />

high energy phosphate levels and contractile function in<br />

human cardiac tissue led to the development of a<br />

temperature-controlled NMR microprobe incorporating a<br />

perifusion system and a non-magnetic strain gauge (figure<br />

3). The system has been used to study human atrial<br />

trabeculae, which are small functional muscle fibers<br />

weighing 8-25 mg that can be isolated from atrial<br />

appendages. The perifusion system provides the tissue with<br />

the oxygen and nutrients required for its function. The strain<br />

gauge allows for measurement of developed force in


Vol. 14, No. 1-4<br />

electrically stimulated muscle fibers while- NMR<br />

simultaneously assesses the high energy phosphate<br />

compound levels [4]. In addition, the perifusion system can<br />

mimic preservation conditions by allowing muscles to be<br />

perifused at low temperature (down to 1°C) during<br />

acquisition of NMR data (figure 4).<br />

Stimulation wire<br />

I<br />

Muscle in the<br />

NMR tube<br />

Flexible tubing<br />

Reflector<br />

Optical fiber<br />

Fiber optic<br />

strain gauge Perfusate line<br />

Stimulation wire<br />

4<br />

Adjustable<br />

hook mount<br />

FIGURE 3. Schematic drawing of the microperifusion<br />

system used for NMR studies of human atrial trabeculae.<br />

The total length of the system is 6.2 cm.<br />

By studying atrial trabeculae in the presence of<br />

metabolic inhibitors under conditions simulating<br />

preservation, it is possible to assess the contribution of<br />

various cellular mechanisms of energy production to the<br />

total energy balance of the tissue. Our 31 P NMR studies of<br />

isolated human atrial trabeculae [5] preserved at 4° and 12°C<br />

in oxygenated St-Thomas II solution showed that the high<br />

energy phosphates (ATP and phosphocreatine (PCr)) are<br />

well maintained during 18 hours of preservation.<br />

Contractile studies performed under similar conditions<br />

showed high recovery of developed force. Glycolysis, the<br />

only pathway available for energy generation under the<br />

anoxic conditions existing in large preserved organs, is<br />

capable of maintaining ATP levels in hypothermically<br />

preserved tissue. Under anoxia, ATP levels are stable for 6 -<br />

10 hours at 12°C, and for a longer period at 4°C. In a<br />

resting heart, the major energy source is provide by the lipid<br />

catabolism. To test whether this pathway is active at low<br />

• temperatures, we have measured the ability of the oxydative<br />

pathway to maintains ATP levels. At 12°C, when glycolysis<br />

is inhibited by iodoacetate, the oxidative pathway can<br />

maintained ATP levels, but only if an external source of<br />

substrate (10 mM acetate) is present in the perfusate. Thus,<br />

the oxidative pathway is functional but depends on both,<br />

oxygen and glycolysis.<br />

In tissue preserved ischemically (no flow), metabolic<br />

waste products such as lactate cannot be eliminated. This<br />

results in considerable extracellular and intracellular acidosis<br />

[1] which has profound effects on energy production<br />

because glycolysis is inhibited by low pH. Using the atrial<br />

trabecula model, we simulated the conditions that exist in<br />

the ischemically preserved human heart. Such a large organ<br />

(300 - 450 gm) cannot obtain sufficient oxygen to maintain<br />

oxidative phosphorylation simply through diffusion from the<br />

surrounding medium as in the case with trabeculae.<br />

Trabeculae were subjected to acidosis by perifusing the<br />

trabeculae with modified St-Thomas preservation solution<br />

containing 10 mM lactate at pH 6.0. Anoxia was<br />

simultaneously induced with 1 mM cyanide, a potent<br />

inhibitor of oxidative phosphorylation, and nitrogen. In the<br />

trabeculae, these conditions reproduced those previously<br />

observed in the larger atrial appendages where an anoxic<br />

core probably existed [1]. Under acidosis and anoxia at<br />

12°C, ATP decreased linearly by 40 to 100% over a 12 h<br />

period. At 4°C, ATP decreased less over the same time<br />

period.<br />

40<br />

DMMP<br />

1 I •<br />

30<br />

1 I '<br />

20<br />

10<br />

a -ATP Y.ATP<br />

1 I '<br />

-10<br />

-20<br />

PPM<br />

FIGURE 4. 31 P NMR spectrum (147 MHz) of a 10.5 mg<br />

trabecula perifused at 12°C in modified St-Thomas II<br />

solution (60° pulse, 1 sec recycling delay, 2400 scans).<br />

These observations can be reconciled to the following<br />

model of energetics: the maintenance of cellular ATP<br />

depends on matching of supply and demand. At 4°C,<br />

glycolysis appears to be limited by temperature. ATP<br />

regeneration cannot be driven at an adequate rate until the<br />

feedback drive (ADP + Pj) is increased considerably above<br />

the normal level. ATP is then maintained at a low<br />

phosphorylation potential. At 12°C glycolysis is not limited<br />

by temperature but is limited by low intracellular pH.<br />

17


18<br />

Other than gaining a better understanding of energetics<br />

in atrial trabeculae, we can ask whether these studies<br />

provided the transplant surgeon with any information of<br />

practical value? In answer to our initial question on the<br />

optimal temperature for preservation of grafts, we have<br />

provided evidence that, for preservation times of 5 hours or<br />

less, ATP levels are better maintained at 12°C [3]. For<br />

longer preservation times, ATP levels are better preserved at<br />

the lower temperatures. In addition, increasing the buffer<br />

capacity of preservation solutions used at 12°C has a major<br />

impact on maintenance of high energy phosphates.<br />

b) Intact Hearts<br />

Most published NMR studies on heart preservation have<br />

used rodent hearts, with a particularly large number of<br />

studies being performed on the rat heart. We have<br />

developed the isolated perfused pig heart for preservation<br />

studies [6] because it is architecturally, biochemically, and<br />

in size most similar to the human heart. As we discussed<br />

above, provision of oxygen and removal of metabolic waste<br />

products are critically important for long term heart<br />

preservation. Perfusion preservation, which can enhance<br />

oxygen delivery and waste removal from the heart, has not<br />

yet achieved much clinical application and remains mostly<br />

in the realm of the research laboratory. Some of the reasons<br />

for this are related to the implementation difficulties in<br />

situations in which the heart must be transported over long<br />

distances under sterile conditions. In addition, hearts<br />

preserved ischemically for less than 5-6 hours generally<br />

show good recovery of mechanical function after<br />

transplantation.<br />

We have consequently focused our efforts on methods<br />

of improving long-term ischemic preservation of hearts.<br />

The goal is to optimize the conditions currently in use and to<br />

extend the safe preservation time between harvest and<br />

implantation of the donor organ. This should allow harvest<br />

of donor hearts to occur over a wider geographical range and<br />

provide for better immunological organ matching.<br />

The studies of isolated, Langendorff perfused pig hearts<br />

[6] are performed using a Bruker Biospec instrument<br />

equipped with a 4.7 T / 30 cm horizontal bore magnet. The<br />

heart is arrested and isolated using techniques similar to<br />

those used for human hearts. The isolated heart is then<br />

placed in an NMR probe to observe high energy phosphate<br />

levels and pH on a continuous basis, with a two minute time<br />

Bulletin of Magnetic Resonance<br />

resolution, during a hypothermic preservation period that<br />

usually lasts 8 hours. Following preservation, the heart is<br />

rewarmed to 37°C without removing it from the magnet and<br />

NMR spectra are then recorded in the beating heart. A<br />

balloon placed in the left ventricle measures the developed<br />

pressure and serves as an index of functional recovery of the<br />

heart [6]. In this manner, energy levels during and after<br />

preservation can be correlated with functional performance<br />

of the heart following preservation.<br />

A number of technical difficulties arise in trying to<br />

obtain quantitative results from large, isolated, perfused<br />

hearts because they change shape when beating and often<br />

swell when perfused with solutions other than whole blood.<br />

To alleviate the problems caused by the sample moving in a<br />

heterogeneous Bi field and the consequent uncertainty in<br />

received signal strength, a high Bi homogeneity probe was<br />

designed [7]. The prototype probe comprised four separate<br />

tuned rings on a spherical surface (Figure 5) giving a Bi<br />

field homogeneity of ± 5% over 60% of the radius of a 14<br />

cm sphere (Figure 6).<br />

tobalun<br />

coupling ring<br />

FIGURE 5. Geometry of the 4 coil system [7].<br />

The received signal was rendered less sensitive to the<br />

dielectric constant of the sample by distributing the<br />

capacitance around the rings. Inter-ring coupling and to a<br />

fifth ring used for matching was by induction. The coupling<br />

loop was tuned with its own capacitor to Larmor frequency,<br />

thereby ensuring that the probe was always on resonance,<br />

and rendering the tuning and matching independent. In<br />

addition, the use of a low input impedance preamplifier


Vol. 14, No. 1-4 19<br />

virtually eliminated the dependence of signal strength on<br />

coil loading [8].<br />

o V* °<br />

its<br />

/ £<br />

/ -<br />

0/ :g<br />

/ -2-<br />

/ *<br />

7 1<br />

/ I il<br />

0.8-<br />

0.6<br />

0.4'<br />

B1<br />

i<br />

\° \V\°\\V<br />

, . 1 1—-,—*-y<br />

-10 Distance in cm<br />

FIGURE 6 Plot down the Y axis of the Bi field of the<br />

probe (open circles, f = 7 cm at 81 MHz) and the form of the<br />

plot predicted by the theory.<br />

Strategies for improving hypothermic preservation have<br />

ranged from improving the buffer capacity of the<br />

preservation solution [6] to the use of secondary<br />

cardioplegia [6, 9] to maintain the heart in an arrested state<br />

during the entire rewarming phase prior to reperfusion. The<br />

purpose of secondary cardioplegia is to allow the energy of<br />

the heart to be directed towards the re-establishment of ionic<br />

balances that become disrupted by hypothermia and<br />

ischemia, rather than to expend energy in mechanical<br />

function. We have found that the use of secondary<br />

cardioplegia prior to reperfusion does not affect the net<br />

energy levels of the heart but rather eliminates ventricular<br />

fibrillation that is normally observed upon rewarming of<br />

hypotfaermically preserved hearts [6].<br />

As a result of the increasing requirement for donor<br />

organs, a number of organs are frequently harvested from a<br />

single donor. This has led to the need for a single<br />

preservation solution suitable for all thoracic and abdominal<br />

organs. The University of Wisconsin Cold Storage Solution<br />

(UW-CSS, DuPont Pharmaceuticals) is currently in use for<br />

the preservation of liver, kidney and pancreas. Its utility for<br />

heart preservation remains to be determined. We compared<br />

the efficacy of UW-CSS to St-Thomas II solution, which is<br />

in widespread use for heart preservation [10]. Pig hearts<br />

were preserved for 8 hours at 4° or 12°C and then tested<br />

functionally after rewarming to 37°C. These temperatures<br />

were chosen because they are both in use clinically. At 4°C,<br />

UW-CSS and St-Thomas II were equally effective for<br />

preservation of heart function. Figure 7 and 8 shows the<br />

results obtained with UW-CSS and St-Thomas at 12°C. The<br />

lack of functional recovery observed with UW-CSS shows<br />

that this solution is unsuitable for use at 12°C. The results<br />

also demonstrate the necessity for precise temperature<br />

regulation when the solution is used at 4°C. This is not<br />

necessary with St-Thomas II solution because recovery is<br />

not severely compromised by use at either 4° or 12°C.<br />

Ref.<br />

PCr<br />

20 0 -20<br />

PPM<br />

P-ATP<br />

20 0 -20<br />

PPM<br />

20 0 -20<br />

PPM<br />

St-Thomas UW-CSS<br />

4°C<br />

Reperfusion<br />

8 h ischemia<br />

4 h ischemia<br />

2 h ischemia<br />

Reperfusion<br />

8 h ischemia<br />

4 h ischemia<br />

2 h ischemia<br />

FIGURE 7. Typical time course of the 3 l P NMR spectra of<br />

4 hearts, two preserved at 4°C (top panels), two preserved at<br />

12°C (bottom panels) [10]. Spectra on the left were<br />

obtained from hearts stored with St-Thomas II and spectra<br />

on the right were from hearts preserved with UW-CSS. The<br />

ATP and PCr disappeared upon reperfusion in the hearts<br />

stored with UW-CSS at 12°C.<br />

One of the reasons for the failure of UW-CSS in heart<br />

preservation at 12°C could be the calcium paradox. This<br />

phenomenon occurs when a heart has been subjected to a<br />

calcium-free medium (UW-CSS contains no calcium) and<br />

then is reperfused with a solution containing calcium. The<br />

calcium paradox results in massive irreversible damage to<br />

cell membranes, as calcium from the reperfusion medium<br />

overloads the cells. This phenomenon is temperaturedependent<br />

and does not occur readily at low temperatures.<br />

In order to test the "calcium paradox" hypothesis, we added<br />

0.5 mM calcium (0.08 mM free calcium) to UW-CSS and


I<br />

III<br />

20 Bulletin of Magnetic Resonance<br />

repeated our studies at 12°C [11]. Figures 9 and 10 show<br />

the improvement observed in the high energy phosphates<br />

during reperfusion of a heart preserved with UW-CSS<br />

containing calcium.<br />

6000-<br />

•a 5000-<br />

4000-<br />

3000-<br />

Cu_ 2000-<br />

1000-<br />

0<br />

0<br />

6OOO-1<br />

^ 5000-<br />

.9<br />

-M 4000-<br />

1 3000-<br />

2000-<br />

looo H<br />

o-<br />

4°C A u f *<br />

10<br />

12 P C<br />

20 30 40<br />

Time (min)<br />

St-Thomas<br />

UW-CSS<br />

50 60<br />

* *<br />

UW-CSS<br />

0 10 20 30 40 50 60<br />

Tune (min)<br />

FIGURE 8 Time course of the rate pressure product (RPP:<br />

heart rate x developed pressure, a measure of heart function)<br />

during reperfusion (n = 7 in each group) [10]. The<br />

functional recovery was extremely poor in the hearts stored<br />

with UW-CSS at 12°C (* : p < 0.05)<br />

These studies show that NMR spectroscopy can be a<br />

valuable tool in the design and modification of solutions for<br />

protecting the myocardium prior to transplantation. Cardiac<br />

surgery is another area in which NMR spectroscopy is being<br />

used.<br />

3 Cardiac Surgery<br />

Cardiac surgery is being offered to higher and higher<br />

risk patients. This has led to the need for improved methods<br />

of myocardial protection during surgery. Traditionally the<br />

heart is arrested and kept cold (4°C) with one or more<br />

infusions of a crystalloid solution, such as the St-Thomas II<br />

solution described above. The hypothermia and ischemia<br />

associated with the use of cold crystalloid solutions can<br />

impose additional stress on the damaged heart. One of the<br />

most recent modifications to cardiac surgical practice has<br />

been the use of continuous normothennic blood cardioplegia<br />

(CNBC) [12]. The purpose of CNBC is to avoid ischemia.<br />

With CNBC, the heart is maintained at 37°C in an arrested<br />

state by increasing the potassium concentration of a blood<br />

solution that continuously flows through the coronary<br />

vessels. Many questions remain unanswered regarding<br />

CNBC, such as the route of administration (retrograde<br />

and/or antegrade); b) the optimal volume of cardioplegia; c)<br />

the flow distribution of cardioplegia, and d) metabolic<br />

monitoring of the heart during cardioplegic arrest. For these<br />

purposes, a blood-perfused porcine heart model was<br />

developed for NMR studies of CNBC. In this model, the<br />

heart is continuously perfused with blood while being<br />

isolated from the animal. The heart can then be placed in<br />

the NMR magnet and its initial function assessed before it is<br />

arrested in the magnet. NMR surveillance of the high<br />

energy phosphates during CNBC may allow optimization of<br />

flow rates and the route of delivery (antegrade and/or<br />

retrograde) for maintenance of the energy status of the<br />

myocardium. In this context, localized NMR spectroscopy<br />

using either spectroscopic imaging [13] or surface gradient<br />

coils [14], allows assessment of protective techniques in<br />

selected regions of the heart [13, 15] or at various depths<br />

across the heart wall [14], respectively.<br />

40 20 0 -20<br />

PPM<br />

-Calcium<br />

2 hour ischemia<br />

ji. Reperfusioa<br />

40 20 0 -20<br />

PPM<br />

+ Calcium<br />

FIGURE 9 Typical 3 *P NMR spectra of hearts preserved in<br />

unmodified UW-CSS (left panel), or with UW-CSS<br />

containing Ca 2+ (right panel) [11]. The peaks of ATP and<br />

PCr disappeared upon reperfusion in the heart stored with<br />

unmodified UW-CSS.<br />

The difference between the two curves is statistically<br />

significant (p < 0.001) [ll]The NMR technique


Vol. 14, No. 1-4<br />

may determine whether there is a safe time limit during<br />

which blood cardioplegia can be interrupted for surgical<br />

visualization. Initially, we evaluated the effect on cardiac<br />

energetics and function by interrupting the flow of CNBC,<br />

as occurs for instance during aorto-coronary bypass surgery.<br />

Our data show that the high energy phosphate profile<br />

deteriorates and PCr becomes unobservable within 14 ± 2<br />

minutes when flow is interrupted for 20 minutes in the<br />

middle of a 1 hour period of CNBC. This is associated with<br />

decreased left ventricular function when the heart is tested<br />

after reperfusion, in spite of the fact that both PCr and pH<br />

returned to normal within 3 minutes of resuming CNBC<br />

[16].<br />

600O,<br />

5000-1<br />

4000-<br />

3000-<br />

2000-<br />

Ca 2+<br />

1000- Control<br />

10 20 30 40 50 60<br />

Reperfusion Time<br />

(min)<br />

FIGURE 10. Time course of the rate pressure product of<br />

hearts reperfused with UW-CSS containing Ca 2+ (0.5 mM)<br />

and without Ca 2+ (mean ± SD, n=7 per group).<br />

It has been proposed that CNBC could resuscitate the<br />

damaged heart during surgery by providing continuous<br />

delivery of oxygen and nutrients to the heart. In order to test<br />

one aspect of resuscitation, we subjected isolated hearts that<br />

had been previously stressed by a 20 minute period of<br />

normothermic ischemia to two types of cardioplegia and<br />

measured functional recovery following reperfusion [17].<br />

Twenty minutes of normothermic ischemia reduced the ATP<br />

levels, measured by 31 P NMR, in CONTROL hearts (n=6)<br />

to 70 ± 7%. These hearts recovered 86 ± 18 % of theninitial<br />

function (systolic elastance) when reperfused with<br />

normal blood perfusate. The experimental hearts were<br />

subjected to either intermittent cold blood cardioplegia<br />

(ICBC, n=6) for 5 min at 14°C, every 20 minutes, and<br />

repeated 3 times, or CNBC (n=6) for 60 minutes at a flow<br />

rate of 0.5 mL mur 1 g" 1 heart wet weight. Both ICBC and<br />

CNBC prevented exacerbation of the initial ischemic injury.<br />

PCr recovered to initial levels following reperfusion in all<br />

three groups (CONTROL, ICBC and CNBC) indicating that<br />

the mitochondria still possessed sufficient phosphorylating<br />

capacity to maintain appropriate physiological activity. ATP<br />

levels did not recover to initial levels in any of the groups.<br />

This could be related to the loss of nucleotide precursors<br />

from the cells during the initial ischemic period. Functional<br />

recovery with CNBC was 115 ± 30% compared to ICBC<br />

which was 88 ± 9% but there were no significant difference<br />

among the three groups by ANOVA (p>0.05).<br />

Is there a safe period of normothermic ischemia? From<br />

a biochemical perspective, there may be a partial answer.<br />

Using 31 P NMR spectroscopy, we have seen that during an<br />

ischemic episode, PCr decreases before ATP. Theoretical<br />

calculations using the enzymatic equilibria of the creatine<br />

kinase and adenylate kinase reactions support this<br />

observation and have shown that there are two phases of<br />

energy depletion [18]: the buffering phase and the depletion<br />

phase. During the buffering phase, energy is derived from<br />

PCr and the adenine pool is stable. During the depletion<br />

phase, energy is primarily derived from adenine nucleotides.<br />

As ATP is consumed, AMP is produced which subsequently<br />

acts as a substrate for deamination and dephosphorylation<br />

reactions, whose products are lost from the cell. Upon<br />

reperfusion, although the PCr levels may return to normal,<br />

adenine nucleotides may not reach normal levels for a<br />

number of days. To avoid imposing a metabolic stress on<br />

the myocytes, PCr levels should not be allowed to become<br />

depleted to the point where adenine nucleotides will begin to<br />

be depleted. Although the role and critical level of ATP<br />

necessary for recovery of function in the ischemic heart are<br />

controversial, it seems logical to avoid any type of<br />

preventable metabolic stress to the heart during surgery.<br />

NMR spectroscopy is useful for monitoring the energy<br />

depletion and repletion processes in model systems such as<br />

the pig heart. This information can subsequently be used to<br />

verify the predictions of the theoretical calculations.<br />

NMR can monitor PCr and ATP levels directly and<br />

continuously in the isolated perfused heart and in vivo [19].<br />

In our studies, PCr levels reflect the balance of energy<br />

supply and demand; in the arrested heart they decrease and<br />

increase in concert with the availability of oxygen. However<br />

NMR techniques are not compatible with direct use in the<br />

surgical theatre. Recently developed fiber optic pO2 probes<br />

based on oxygen quenching of fluorescence have been used<br />

21


22 Bulletin of Magnetic Resonance<br />

to monitor the arrested heart during surgery. We have<br />

performed NMR measurements on isolated blood perfused<br />

pig hearts and correlated the data with simultaneously<br />

measured levels of pO2 using custom built (Innerspace,<br />

Irvine, Calif.) NMR-compatible probes. We have found a<br />

good correlation between tissue p(>2 (measured in mmHg)<br />

and PCr level in the normal heart. The data are illustrated in<br />

Figure 11. Although the data are preliminary, we see that<br />

PCr levels drop precipitously when the tissue pO2 decreases<br />

below 30 mmHg. By establishing similar relationships in<br />

metabolically damaged or physically abnormal<br />

(hypertrophied for instance) hearts it may be possible to<br />

provide the surgeon with insight into a) the acceptable limits<br />

of oxygen deprivation should delivery of cardioplegia be<br />

discontinued during surgery; b ) the optimal flow rates of<br />

cardioplegia; c) assessment of retrograde versus antegrade<br />

delivery; and d) cardioplegic flow distribution across the<br />

heart.<br />

150, -r. 120<br />

120<br />

0 20 40 60 80 100 120 140<br />

Time (min)<br />

20 40 60 80 100 120 140<br />

pO2 (mmHg)<br />

FIGURE 11. Correlation between the pC>2 and PCr levels<br />

in the normal heart. Top: pC


Vol. 14, No. 1-4<br />

7. D.I. Hoult and R. Deslauriers. A High-Sensitivity, High<br />

Bi Homogeneity Probe for Quantification of<br />

Metabolites, Magn. Reson. Med., 16,411-417 (1990).<br />

D.I. Hoult and R. Deslauriers. Elimination of Signal<br />

Strength Dependency upon Coil Loading - An Aid to<br />

Metabolite Quantitation when Sample Volume<br />

Changes, Magn. Reson. Med., 16,418-424 (1990).<br />

G.H. Tian, G.P. Biro, B. Xiang, K.W. Butler and R.<br />

Deslauriers. The Effect of Magnesium Added to<br />

Secondary Cardioplegia on Postischemic Myocardial<br />

Metabolism and Contractile Function - A<br />

10.<br />

31 P NMR<br />

Spectroscopy and Functional Study in the Isolated Pig<br />

Heart, Basic Res. Cardiol. (in press).<br />

G. H. Tian, K.E. Smith, G.P. Biro, K.W. Butler, N.<br />

Haas, J. Scott, R. Anderson and R. Deslauriers. A<br />

Comparison of University of Wisconsin Cold Storage<br />

Solution and St-Thomas Solution II for Hypothennic<br />

Cardiac Preservation: A 31 P NMR and Functional Study<br />

of Isolated Porcine Hearts, /. Heart Lung Transplant.,<br />

10, 975-985 (1991).<br />

11 G.H. Tian, G.P. Biro, K.W. Butler, B. Xiang, C. Vu and<br />

R. Deslauriers. The Effects of Ca<br />

13<br />

14.<br />

++ Ion on the<br />

Preservation of Myocardial Energy and Function with<br />

UW Solution. A 31 P NMR Study of Isolated Pig Hearts,<br />

/. Mol. Cell. Cardiol., 24, S. 190 (1992).<br />

12. A. Panos, S.J. Kingsley, A.P. Hong, T.A. Salerno and S.<br />

Lichtenstein. Continuous Warm Blood Cardioplegia,<br />

Surg. Forum, 61, 233-235 (1990).<br />

D. Bourgeois and R. Deslauriers. Phasing Spin-Echo<br />

Acquired 31 P Spectroscopic Images Using Complex<br />

Conjugate Data Reversal, Magn. Reson. Med. (in press).<br />

A. Jasinski, P. Kozlowski, A. Urbanski and J.K.<br />

Saunders. Hexagonal Surface Gradient Coil for<br />

Localized MRS of the Heart, Magn. Reson. Med., 21,<br />

296-301 (1991).<br />

15. R. Deslauriers, S. Lareau, G.H. Tian, A.L. Panos,<br />

C.A.M. Barrozo and T.A. Salerno. Surgical<br />

Technology - Applications of Magnetic Resonance<br />

Spectroscopy to Cardiac and Transplant Surgery,<br />

Current Surg., 49, 95-101 (1992).<br />

16. R. Deslauriers, A.L. Panos, C.A.M. Barrozo, O. Al-<br />

Nowaiser, K.W. Butler, N. Haas, K.H. Teoh and T.A.<br />

Salerno. Myocardial Protection: Energy Profile During<br />

Continuous Normothermic Blood Cardioplegia,<br />

Abstracts of the Works-in-Progress, Tenth Annual<br />

Meeting of the Society of Magnetic Resonance in<br />

Medicine, San Francisco, Aug. 10-16,1991, p. 1098.<br />

17. R. Deslauriers, K.W. Butler, N. Haas, C.A.M. Barrozo,<br />

A.L. Panos, I.S. Ali, O. Al-Nowaiser and T.A. Salerno.<br />

The Effects of Intermittent Cold, and Continuous<br />

Warm, Blood Cardioplegia on Isolated Pig Hearts: 31 P<br />

NMR and Functional Studies, Abstacts of the Eleventh<br />

Annual Meeting of the Society of Magnetic Resonance<br />

in Medicine, Berlin, Aug. 8-14,1992.<br />

18.<br />

RJ. Connett. Analysis of Metabolic Control: New<br />

Insights Using Scaled Creatine Kinase Model, Am. J.<br />

Physiol, 254, R949-R959 (1988).<br />

19. P.A. Bottomley, CJ. Hardy and P.B. Roemer.<br />

Phosphate Metabolite Imaging and Concentration<br />

Measurements in Human Heart by Nuclear Magnetic<br />

Resonance, Magn. Reson. Med., 14,425-434 (1990).<br />

23


24<br />

1. Introduction<br />

Topology and Spin Alignment in Organic<br />

High-Spin Molecules<br />

Yoshio Teki, Kazunobu Sato, Masayuki Okamoto, Atsuya Yamashita,<br />

Yoji Yamaguchi, Takeji Takui, Takamasa Kinoshita, and Koichi Itoh<br />

Department of Chemistry, Faculty of Science, Osaka City University,<br />

Sugimoto 3-3-138, Sumiyoshi-ku, Osaka 558, Japan<br />

Organic high-spin molecules are ideal model compounds<br />

for organic magnetic materials such as organic superparaand<br />

ferro-magnets which have recently been attracting<br />

increasing interest [1]. The related experimental [2] and<br />

theoretical works [3] have been done to serve their<br />

molecular design for the last two decades. High-spin<br />

polycarbenes as organic high-spin systems, in spite of their<br />

highly chemical reactivity, are very important from the<br />

view point .of organic magnetism as well as of spin<br />

ordering/spin control in chemistry from the following<br />

reasons, (i) One of the most prominent features of the highspin<br />

polycarbenes is multi-electron open shell systems in<br />

the ground or low-lying excited states which arise from<br />

degenerate delocalized 7t orbitals and from o dangling<br />

orbitals localized at divalent carbon atoms, the latter<br />

orbitals being nearly degenerate with the highest half-filled<br />

K orbitals. (ii) The degeneracy of their n orbitals is<br />

governed by a particular connectivity of n electron network,<br />

i.e. by the topology of the rc electron network.<br />

We have been studying a series of high-spin polycarbenes<br />

as well as their n topological isomers [2a,2c,4-19] which<br />

are designed by exploiting their % electron networks. We<br />

define a Ji-topological isomer as a molecule which differs<br />

from others only in the topology of its TC electron network,<br />

i.e. in the linking positions of its TC bonds. These high-spin<br />

carbenes have been successfully detected up to a tridecet<br />

polycarbene having twelve parallel spins (S=6) by means of<br />

ESR spectroscopy 118].<br />

In this work, we have studied the spin density distributions<br />

of high-spin polycarbenes and their topological<br />

isomers, biphenyl-n,n'-bis(phenylmethylene) (I: n,n'=3,3';<br />

II: n,n'=3,4') and m- and p-phenylenebis(phenylmethylene),<br />

in order to clarify the mechanism of the spin correlation<br />

and its role in the intramolecular spin alignment of organic<br />

systems.<br />

2. Topological Isomers and Their<br />

Spin States<br />

Figure 1 shows the ground states and the low-lying<br />

excited states of the high-spin polycarbenes and their<br />

topological isomers studied in this work. They were determined<br />

from our previous ESR experiments [2a,5,15,<br />

17]. Molecule I has a unique spin alignment in that it has<br />

low-lying high-spin states (S=l and S=2) above the low-<br />

Bulletin of Magnetic Resonance<br />

Figure 1. Topological Isomers and their spin states<br />

S=l<br />

S=0<br />

spin ground state (S=0) as shown in this figure. In<br />

contrast, its topological isomer II has the high-spin ground<br />

state. Both molecules have four unpaired electron spins in<br />

their nearly degenerate nonbonding n and o orbitals. On the<br />

other hand, molecule III is the first organic high-spin<br />

system with the quintet ground state (S=2) [2a], while its<br />

topological isomer IV has the low-spin ground state and the<br />

low-lying triplet excited state located ca. 200 cm"' above<br />

the ground state.<br />

3. Experimental<br />

The syntheses of the diazo precursors of III and IV were<br />

carried out according to the literatures [20,21]. The<br />

synthetic work of the carbon 13 labeled compounds of<br />

biphenyl-3,3'-bis(phenylmethylene) and other precursors<br />

will be published elsewhere. Each diazo precursor was<br />

diluted in a host single crystal of benzophenone-di()- Mixed<br />

single crystals were grown in the dark by slowly cooling an<br />

ethanol or an ether solution containing the diazo precursor.<br />

The polycabenes were generated at 2 - 4 K by the<br />

photolysis of the corresponding diazo precursors. The<br />

photolysis was carried out with an XBO 500W high pressure<br />

mercury lamp using a quartz rod which guided the 405<br />

nm light into an X-band TMon ENDOR cavity. All the<br />

EPR and ENDOR spectra were recorded with a Bruker ESP<br />

300/350 spectrometer equipped with an Oxford variable temperature<br />

controller ESR910. The spectra of the low-lying<br />

excited states were observed under the thermal excitation of<br />

these levels. Other spectra were measured at ca. 2 - 4 K.


Vol. 14, No. 1-4 25<br />

4. Theoretical<br />

In addition to the ENDOR work, we have calculated the<br />

spin density distributions of molecules I - VI using two<br />

model Hamiltonian approaches. The following unrestricted<br />

Hartree-Fock calculation based on a generalized Hubbard<br />

model [8] as well as the exact numerical solution of a<br />

valence-bond Heisenberg Hamiltonian [9] have provided<br />

rather satisfactory and complementary descriptions for spin<br />

structures of organic high-spin polycarbenes.<br />

The generalized Hubbard model Hamiltonian is given by<br />

[22]<br />

ft - -l Zj


26<br />

(a) 34 K<br />

QAt<br />

(b) 15K<br />

24.0<br />

(c) I5K<br />

—I<br />

0.2<br />

QB+<br />

T: Triplet<br />

Q: Quinlet<br />

QB-<br />

0.4 0.6<br />

MAGNETIC FIELD/T<br />

26.0 28.0 30.0<br />

FREQUENCY / MHz<br />

52.0 56.0 60.0 80.0<br />

FREQUENCY / MHz<br />

84.0 88.0<br />

Figure 2. Typical EPR, l H- and 13 C-ENDOR spectra of<br />

molecule I. (a) EPR spectrum, (b) ^-ENDOR spectrum<br />

(c) l^C-ENDOR spectrum. The external magnetic field is<br />

along the a axis. The microwave frequency v is 9413.0<br />

MHz for the spectra (a) and (b), and 9456.0MHz for the<br />

spectrum (c) of the carbon 13 labeled compound,<br />

respectively.<br />

the two end phenyl groups were deuterated, in order to<br />

facilitate the assignment of 1H-ENDOR transitions by<br />

the reduction of the spectral density. By a comparison of<br />

the ^H-ENDOR spectra of the deuterium labeled compound<br />

with those of the normal compound, the eight transitions<br />

labeled by the asterisk in figure 2(b) were assigned to the<br />

protons of the central biphenyl group and the remaining<br />

unlabeled transitions to those of the end phenyl groups.<br />

Thus, we have achieved reliable assignment for all the *H-<br />

ENDOR signals observed.<br />

On the basis of the assignment above, we have<br />

determined the K spin densities on the carbon sites<br />

adjacent to the hydrogen atoms and the n and a spin<br />

densities on each divalent carbon atom from the analysis of<br />

the angular dependence of the ENDOR frequencies as<br />

described in section 5. The experimentally obtained spin<br />

densities are given in figure 3(a). The spin densities on the<br />

six carbon atoms without circles could not be obtained,<br />

since they have no adjacent protons.<br />

We have also calculated the spin density distribution<br />

theoretically by two model Hamiltonian approaches<br />

[8,9,19] as described in section 4. The calculated spin<br />

Bulletin of Magnetic Resonance<br />

density distribution based on the generalized Hubbard model<br />

is shown in figure 3(b). In this calculation, we used the<br />

weakly interacting model in which molecule I, biphenyl-<br />

3,3'-bis(phenylmethyIene), is regarded as composed of two<br />

diphenylmethylene moieties, unit A and unit B, weakly<br />

interacting with each other. This model has shown to<br />

interpret well the observed particular relationship Dj = -<br />

3DQt as described in our previous work [5J. We have<br />

applied this model to account for the spin density<br />

distribution of molecule I. The spin densities p(S,Ms) in<br />

molecule I can be derived from p0(SA=l,Ms=l) of the<br />

isolated triplet diphenylmethylene moieties using the<br />

equation [15,19]<br />

p(S=l,Ms=l)i = (l/2)pO(SA=l,Ms=l)i (6)<br />

for the triplet state, and<br />

p(S=2,Ms=2)j = p°(SA=l,Ms=l)i (7)<br />

for the quintet state. Similar expressions hold also for unit<br />

B since A and B are equivalent. The spin densities of the<br />

isolated diphenylmethylene moieties were obtained from the<br />

UHF calculation based on the generalized Hubbard model.<br />

This calculation well interprets the ENDOR results as<br />

shown in figure 3. The observed and calculated spin density<br />

distributions show that the sign of the n spin density is<br />

alternatively distributed on the carbon sites within the<br />

diphenylmethylene moiety, thus forming the up-and-down<br />

(a) O : positive<br />

0.093<br />

0.105<br />

-0.019 n -086 0.092 -0.072<br />

0.096 0.098<br />

: negative<br />

0126 -0.075<br />

0.095<br />

0.105<br />

0.711 0.711<br />

Figure 3 Spin density distribution of the thermally excited<br />

triplet state of molecule I. (a) Experimental values, (b)<br />

Theoretical values obtained from the UHF generalized<br />

Hubbard calculation based on the weakly interacting model.<br />

(c) Theoretical values obtained by the valence-bond<br />

Heisenberg Hamiltonian approach.


Vol. 14, No. 1-4<br />

network of the K spin. We define the up-and-down network<br />

of the % spin as the pseudo spin density wave (pseudo-<br />

SDW), since in infinite systems the up-and-down spin<br />

network forms a spin density wave. This network resulting<br />

from spin correlation is most favorable in view of the total<br />

spin energy (the sum of the spin exchange-correlation<br />

energy). The observed spin density distribution is, in<br />

principle, symmetrical with respect to the center of the<br />

molecule as predicted by the weakly interacting model<br />

above. Thus, the pseudo 7t-SDW is formed within each<br />

diphenylmethylene moiety. As a result of these, the central<br />

two carbon sites of the biphenyl group (the bridged carbons)<br />

should have the same (negative) sign. This violates the upand-down<br />

pseudo JI-SDW in the whole molecule, resulting<br />

in a node of spin density distribution at the central bridged<br />

carbons. The existence of this node, which unstabilizes the<br />

spin-correlation energy, makes the observed triplet state to<br />

be an excited state above the spin-less ground state. A<br />

similar spin distribution is also expected for the quintet<br />

state from the comparison of eqs. (6) and (7).<br />

Figure 3(c) shows the spin density distribution calculated<br />

by the Heisenberg model Hamiltonian approach where we<br />

replaced the end phenyl groups of molecule I with the<br />

hydrogen atoms to reduce the dimensionality of the<br />

Hamiltonian matrix. In this calculation, we have obtained<br />

the results similar to figure 3(b) without using the weakly<br />

interacting model. In this approach, the spin correlation<br />

is exactly taken into account within the Heisenberg model,<br />

leading to the correct ordering of the ground state and the<br />

low-lying excited states [9,11].<br />

(B) ESR. ENDOR and Spin Density Distribution<br />

of Molecule III<br />

As mentioned above, the lowest spin ground state, i.e.<br />

the singlet ground state, is realized in molecule I, since the<br />

high-spin states are unstabilized because of the existence of<br />

the node at the central bridged carbons of the biphenyl<br />

group. However, if the position of the bridge is shifted by<br />

one carbon site as shown in figure 5, it is expected that the<br />

nodeless pseudo rc-SDW is formed in the whole molecule,<br />

leading to the high-spin ground state as a result of the<br />

stabilization of spin correlation energy. In order to confirm<br />

this, we have observed the ^H-ENDOR spectra of biphenyl-<br />

3,4'-bis(phenylmethylene) (molecule II). It was shown by<br />

our previous ESR experiments that the ground state of<br />

molecule II was a high-spin (S=2) state without low-lying<br />

excited states [17]. The fine structure parameters and the g<br />

value of the quintet ground state were determined as<br />

D=+0.1250 cm" 1 , E=-0.0065 cm" 1 , and g=2.003<br />

(isotropic). Figures 4(a) and (b) show a typical ESR and<br />

H-ENDOR spectra of the quintet ground state of molecule<br />

II observed at 2.7 K. The ENDOR spectrum was obtained<br />

by monitoring the B+ ESR transition (Ms=0*^+l). Three<br />

transitions by o in figure 4(b) correspond to the *H-<br />

ENDOR signals arising from the Ms = 0 spin sublevel and<br />

the remaining unlabeled signals to those from the Ms=+1<br />

sublevel. The eight transitions labeled by the asterisk<br />

are due to the protons of the central biphenyl group.<br />

(a) 2.7K<br />

A-<br />

B-<br />

0.20 0.30 0.40<br />

MAGNETIC FIELD/T<br />

(b)<br />

* *<br />

12.0 16.0 20.0 24.0<br />

FREQUENCY/MHz<br />

Figure 4. Typical ESR and 'H-ENDOR spectra of molecule<br />

II. (a) ESR spectrum, (b) I H-ENDOR spectrum. The<br />

external magnetic field is along the b axis. The microwave<br />

frequency vis 9378.7 MHz.<br />

(a)<br />

Q : positive<br />

0 : negative .o.os8 °- 224<br />

-0.064 0.218<br />

-0.138 0.198<br />

0.206<br />

(b) "205<br />

-0.154 0.230<br />

-0.180 "-I 54<br />

HIT = 2.0. JIT = 0.25<br />

-0.076<br />

0.151<br />

Figure 5 Spin density distribution of the quintet ground<br />

state of molecule II. (a) Experimental values, (b)<br />

Theoretical values obtained from the UHF generalized<br />

Hubbard calculation.<br />

The spin densities obtained from McConnell's equation<br />

- Qpi n /2S and the theoretical values calculated on the basis<br />

of the Hubbard model are given in figures 5(a) and 5(b),<br />

respectively. The observed and calculated spin density<br />

27


28 Bulletin of Magnetic Resonance<br />

distributions indicate that the nodeless pseudo rc-SDW is<br />

formed in the whole molecule as expected from the<br />

topology of the n electron network of molecule II.<br />

These findings give the following physical picture for<br />

the intramolecular spin alignment of molecule II: The unpaired<br />

n electrons are distributed over the carbon skeleton<br />

with alternating the sign of the spin density from carbon to<br />

carbon, thus forming the pseudo SDW in the K electron<br />

network. On the other hand, the two unpaired a spins in<br />

the localized o dangling orbitals become parallel to each<br />

other as a results of the ferromagnetic coupling to the<br />

unpaired n spins at each divalent carbon site, since the onecenter<br />

exchange integral / in eq. (3) between the nearly<br />

degenerate a and n orbitals on the same carbene site is<br />

usually ferromagnetic. Thus, this picture shows that the<br />

spin alignment in molecule II is predominantly determined<br />

by the formation of the pseudo rc-SDW.<br />

(O ESR. ENDOR and Spin Density Distribution<br />

of Molecule III and IV<br />

To demonstrate the role of spin correlation as determined<br />

by the topology of n electron networks, we have also<br />

investigated the spin density distribution in the quintet<br />

ground state (S=2) of the first organic, high-spin molecule,<br />

m-phenylenebis(phenylmethylene) III, and that in the lowlying<br />

excited triplet state (S=l) of its topological isomer, pphenylenebis(phenylmethylene)<br />

(molecule IV). In the<br />

previous work, we reported ESR studies for III and IV<br />

[2a,5] and ^-ENDOR experiments for III [2c]. Figures<br />

6(a)and6(b) show typical ESR and 'H-ENDOR spectra<br />

of III observed at 2 K. The signals labeled by the circle (o)<br />

in the ESR spectrum are those due to a minor byproducts.<br />

The primed pairs A'+ and B'±, and the unprimed pairs A+<br />

and B+ arise from the two magnetically nonequivalent sites<br />

occupied by the guest molecule III in the host single crystal.<br />

A typical ^C-ENDOR spectrum is shown in figure<br />

6(c). We have roughly estimated the n and a spin<br />

densities at each divalent carbon atom from the angular<br />

dependence of the * ^C-ENDOR frequencies in figure 6(c);<br />

the complete analysis by the numerical diagonalization of<br />

the spin Hamiltonian (4) is in progress. The anisotropy of<br />

the l^C hyperfine tensor is about one half of that of<br />

diphenylmethylene reported by Hutchison et al. 126]. This<br />

is due to the projection factor 1/(2S) in eq. (5). Our<br />

preliminary results indicate that the spin densities on both<br />

divalent carbons have values similar to those of<br />

diphenylmethylene. Thus, the sign of the spin densities on<br />

each divalent carbon site is determined to be positive as<br />

expected from the topology of the n electron network of<br />

III, and their n spin densities were estimated to be similar<br />

in magnitude. The K spin densities on the other carbon<br />

sites having adjacent protons were also determined from the<br />

isotropic term of each proton hyperfine tensor using<br />

McConnell's relation. The spin-density distribution<br />

experimentally determined showed that the pseudo 7t-SDW<br />

is formed in the n electron network in a manner similar<br />

to the ground state of II.<br />

A question arises as for the low-lying triplet excited state<br />

of IV whether there exists a node in the n spin distribution,<br />

since II and IV have similar spin structures, i.e. low-lying<br />

high-spin excited states above the low-spin (singlet) ground<br />

state (figure 1). To answer this question, we measured the<br />

^H-ENDOR spectra of the low-lying triplet excited state of<br />

p-phenylenebis(phenylmethylene) (molecule IV). Its typical<br />

ESR and *H-ENDOR spectra are shown in figures 7(a) and<br />

7(b), respectively. The spin density distribution obtained<br />

from the analysis of the angular dependence of the ENDOR<br />

signals is given in figure 7(c). This shows that the four<br />

carbon sites in the central phenyl ring have the same<br />

negative sign in the K spin density, leading to a node in the<br />

7t spin density distribution in the central ring. This finding<br />

confirms the physical picture for the formation of low-lying<br />

high-spin states above a low-spin/spin-less ground state in<br />

topological isomers of high-spin polycarbebes, as described<br />

in section 6(A): The node of the spin distribution<br />

violates the formation of the pseudo rc-SDW in the<br />

whole molecule. This unstabilizes the spin correlation<br />

energy, leading to the observed triplet state above the<br />

singlet ground state.<br />

0.1 0.2 0.3 0.4<br />

8.0 12.0 16.0<br />

FREQUENCY/MHz<br />

(c)<br />

6ab = 5 Temp. 3K<br />

20.0<br />

40.0 44.0 48.0 60.0 64.0 68.0<br />

FREQUENCY / MHz<br />

Figure 6. Typical ESR, 1 H- and l3 C-ENDOR spectra of<br />

molecule III. (a) ESR spectrum, (b) JH-ENDOR<br />

spectrum, (c) 13 C-ENDOR spectrum. The external<br />

magnetic field is along ©ab= 5° from the a axis. The<br />

microwave frequency v is 9430.4 MHz.


Vol. 14, No. 1-4 29<br />

1.2.0<br />

(c)<br />

0.28 0.32 036 ' O40<br />

MAGNETIC FIELD/T<br />

16.0 20.0 24.0<br />

FREQUENCY / MHz<br />

O Positive Spin (P = O.l9i - 0.293)<br />

9 Negative Spin (p = -o.oio~-o.08i)<br />

IF.NDOR measurement<br />

28.0<br />

Figure 7. Typical ESR and ' H- ENDOR spectra and spin<br />

density distribution of the low-lying triplet excited state of<br />

molecule IV. (a) ESR spectrum, (b) JH-ENDOR spectrum.<br />

The external magnetic field is along 0ab= 12° from the a<br />

axis. The microwave frequency v is 9457.0 MHz. (c)<br />

Experimentally determined spin densities.<br />

7. Conclusions<br />

We have investigated the spin alignment in the ground<br />

states and the low-lying excited states of the high-spin<br />

polycarbenes and their topological isomers shown in figure<br />

1. Their spin density distributions were determined by<br />

single-crystal 1 H- and ^G-ENDOR as well as by the<br />

theoretical calculations using the two model Hamiltonians<br />

(the generalized Hubbard model and the valence-bond<br />

Heisenberg model). The results obtained in this work can<br />

be summarized as follows: (1) It is shown that the spin<br />

alignment is highly dependent on the topological nature in<br />

the K electron network. (2) The pseudo JC-SDW governed<br />

by the topological nature plays an important role in the<br />

stabilization of the high-spin ground state. (3) The spin<br />

correlation also plays essential part for the formation of the<br />

low-lying spin states of molecule I and IV, as well as of<br />

the high-spin ground states of molecule II and HI. (4) The<br />

physical picture of the intramolecular spin alignment in<br />

polycarbenes has been clarified in view of spin correlation.<br />

8. References<br />

1. J.S. Miller and D.A. Dougherty, eds.. Proceedings of the<br />

symposium on Ferromagnetic and High-Spin Molecular<br />

Based Materials, 197th ACS Meeting, Dallas, USA (April<br />

9-12, 1989); Mol. Cryst. Liquid Cryst. 176, 1-562 (1989).<br />

2. (a) K. Itoh, Chem. Phys. Lett., 1, 235 (1967). (b) E.<br />

Wasserman et al., J. Am. Chem. Soc, 89, 5076 (1967).<br />

(c) T. Takui, S. Kita, S. Ichikawa, Y. Teki, T. Kinoshita,<br />

and K. Itoh, Mol. Cryst. Liquid Cryst., 176, 67 (1989),<br />

and references cited therein.<br />

3. (a) K. Itoh, Bussei, 12, 635 (1971). (b) N. Mataga,<br />

Theoret. Chim. Acta (Berl.), 10, 372 (1968). (c) N.<br />

Tyutyulkov, G. Olbvich, H. Brenzen, O, Polansky, Theoret.<br />

Chim. Acta, 73, 27 (1988), and references cited therein.<br />

4. T.Takui and K. Itoh, Chem. Phys. Lett.,19, 120 (1973).<br />

5. K. Itoh, Pure & Appl. Chem., 50, 1251 (1978).<br />

6. Y. Teki, T. Takui, K. Itoh, and H. Iwamura, J. Chem.<br />

Phys., 83, 539 (1985).<br />

7. Y. Teki, T. Takui, K. Itoh, H. Iwamura, and K.<br />

Kobayashi, J. Am. Chem. Soc, 108, 2147 (1986).<br />

8. Y. Teki, T. Takui, T. Kinoshita, S. Ichikawa, H. Yagi,<br />

and K. Itoh, Chem. Phys. Lett., 141, 201 (1987).<br />

9. Y. Teki, T. Takui, M. Kitano, and K. Itoh, Chem. Phys.<br />

Lett., 142, 181 (1987).<br />

10. Y. Teki, T. Takui, and K. Itoh, J. Chem. Phys., 88,<br />

6134(1988).<br />

11. K. Itoh, T. Takui, Y. Teki, and T. Kinoshita, J. Mol.<br />

Electronics, 4, 181 (1988).<br />

12. K. Itoh, T. Takui, Y. Teki, and T. Kinoshita, Mol.<br />

Cryst. Liquid Cryst, 176, 49 (1989).<br />

13. I. Fujita, Y. Teki, T. Takui, T. Kinoshita, K. Itoh, F.<br />

Miko, Y. Sawaki, H. Iwamura, A. Izuoka, and T.<br />

Sugawara, J. Am. Chem. Soc, 112, 4047 (1990).<br />

14. M. Matsushita, T. Momose, T. Sida, Y. Teki, T.<br />

Takui, and K. Itoh, J. Am. Chem..Soc, 112, 4701 (1990).<br />

15. M. Okamoto, Y. Teki, T. Takui, T. Kinoshita, and K.<br />

Itoh, Chem. Phys. Lett., 173, 265 (1990).<br />

16. M. Matsushita, T. Nakamura, T. Momose, T. Sida, Y.<br />

Teki, T. Takui, T. Kinoshita, and K. Itoh, J. Am. Chem.<br />

Soc, in press (1992).<br />

17. Y. Teki, I. Fujita, T. Takui, T. Kinoshita, and K. Itoh,<br />

J. Am. Chem. Soc, submitted (1992).<br />

18. K. Furukawa, T. Matsumura, Y. Teki, T. Kinoshita, T.<br />

Takui, and K. Itoh, J. Am. Chem. Soc, submitted (1992).<br />

19. Y. Teki, M. Okamoto, T. Takui, T. Kinoshita, and K.<br />

Itoh, J. Am. Chem. Soc, submitted (1992).<br />

20.R. W. Murray and A. M. Trozzolo, J. Org. Chem., 26,<br />

3109 (1961).<br />

21. S. I. Murahashi, Y. Yoshimura, Y. Yamamoto, and I.<br />

Moritani, Tetrahedron, 28, 1485 (1972).<br />

22. (a)R. M. White, Quantum Theory of Magnetism;<br />

Springer, p. 139 (1983). (b) K. Nasu, Phys. Rev., B33,<br />

330 (1986).<br />

23. S. A. Alexander and D. J. Klein, J. Am. Chem. Soc,<br />

110, 3401 (1988).<br />

24. N. Hirota, C. A. Hutchison Jr. and P. Palmer, J.<br />

Chem. Phys., 40, 3717 (1964).<br />

25. M. E. Rose, Elementary Theory of Angular<br />

Momentum, John Wiley & Sons: New York (1957).<br />

26. C. A. Hutchison Jr., B. E. Kohler, J. Chem. Phys.,<br />

51, 3327 (1979).


30 Bulletin of Magnetic Resonance<br />

New Developments in<br />

Pulsed Electron Paramagnetic Resonance:<br />

Relaxation Mechanisms of<br />

Nitroxide Spin Labels<br />

Colin Mailer, Bruce H. Robinson, and Duncan A. Haas<br />

Department of Chemistry<br />

University of Washington<br />

Seattle WA 98195<br />

Recent technical developments in this laboratory in the areas of pulsed Saturation Recovery Electron<br />

Paramagnetic Resonance (SR-EPR) and Saturation Recovery Electron-Electron Double Resonance (SR-ELDOR)<br />

have enabled experiments with these two techniques to be done with high sensitivity. We have studied the mechanisms<br />

of the relaxation rates of spin labels. The electron spin-lattice relaxation rate Tic' 1 and the nitrogen spinlattice<br />

relaxation rate (Tin" 1 ) of per-deuterated 15 N TEMPOL in glycerol-water solutions have been measured by<br />

SR-EPR and SR-ELDOR. The motional range covered is from a few picoseconds to hundreds of nanoseconds and<br />

the motion is characterized by simple Brownian dynamics. The dependence of Tie 1 upon the rotational<br />

correlation time is explained by a combination of spin rotation and electron -nuclear dipolar coupling mechanisms<br />

plus a proton spin diffusion process. Tin-1 is explained by the electron-nuclear dipolar mechanism and proton<br />

spin diffusion.<br />

1 Introduction<br />

Given the utility of nitroxide spin labels as a probe<br />

of molecular motion in the biological and physical<br />

sciences for the past 25 years it is somewhat surprising<br />

that a clear understanding of their relaxation<br />

mechanisms is still unknown. The quantities<br />

desired are the spin-lattice relaxation rates both<br />

electronic (Tie-i) and nuclear (TV 1 - 14 N or 15 N)<br />

of the nitroxide label. Pulsed EPR can be used to<br />

monitor such relaxation processes directly regardless<br />

of T/2e or line inhomogeneity. In a pulsed<br />

Saturation Recovery (SR) experiment a high rf<br />

field pump pulse is applied for a short time and the<br />

recovery of the magnetization is measured with a<br />

low power observer at the same frequency. The<br />

pulsed Electron-Electron Double Resonance (SR-<br />

ELDOR) experiment has the pump and observer<br />

frequencies quite different (commonly pump and<br />

observer are set to resonate on different spin manifolds).<br />

Huisjen and Hyde [1] pioneered the use<br />

of the SR-EPR technique in liquids and applied it<br />

to a number of systems [2], [3], [4], [5], [6]. The<br />

1 MHuisjen and J.S.Hyde. Rev. Sci. lost., 45, 669-675,<br />

1974.<br />

2 P.W. Percival and J.S. Hyde. /. Mag. Res., 23,249-<br />

257, 1976.<br />

3 T. Sarna and J.S. Hyde. J. Chem. Phys., 69, 1945-<br />

1948, 1978.<br />

Hyde group has used SR-EPR in conjunction with<br />

Continuous Wave (CW) - ELDOR to measure lateral<br />

diffusion of 14 N labelled lipids in bilayers [7],<br />

[8], [9]. Hyde et al. [10] have observed the transfer<br />

of energy from one line to another line in a<br />

CTPO 14 N spin label with pulsed SR-ELDOR.<br />

They clearly saw that cross relaxation took place,<br />

and measured the cross-relaxation rate. We discuss<br />

this experiment below.<br />

Freed and co-workers [11], [12], [13], [14] analyzed<br />

the EPR spectra of peroxylamine disulphon-<br />

10<br />

11<br />

C. Altenbach, W. Froncisz, J.S. Hyde, and W.L.<br />

Hubbell. Biophys. J., 56, 1183-1191, 1989.<br />

J-J. Yin and J.S. Hyde. Zeitschrift fur Physikalische<br />

Chemie. 153, 57-65, 1987.<br />

A. Kusumi, W.K. Subczynski and J.S. Hyde. Proc.<br />

Nat. Acad. Sci. USA, 79, 1854-1858, 1982.<br />

C.A. Popp and J.S. Hyde Proc. Nat. Acad. Sci. USA,<br />

79, 2559-2563, 1982.<br />

J.B. Feix, C.A. Popp, S.D. Venkataaramu, A.H. Beth,<br />

J.H. Park and J.S. Hyde. Biochemistry, 23, 2293-<br />

2299, 1984.<br />

J.-J. Yin, M. Pasenkiewicz-Gierula, and J. S. Hyde<br />

Proc. Nat. Acad. Sci. USA, 84, 964-968, 1987.<br />

.S. Hyde, W. Froncisz and C. Mottley Chem. Phys.<br />

Lett., 110, 621-625, 1984.<br />

J.H. Freed, in Spin Labelling: Theory and<br />

Applications vol I ed. LJ.Berliner. Academic Press<br />

1976. Chapter 3. pp.53-132.


Vol. 14, No. 1-4 31<br />

ate ( 14 N-PADS) and deuterated 2,2,6,6-tetramethyl-4-piperidinone-l-oxyl<br />

(pd-TEMPONE) in<br />

glycerol-water mixture solvents. The linewidths<br />

as a function of temperature were obtained,<br />

corrected for inhomogeneous broadening, and<br />

analyzed according to: l/T2e(M)= A+ B • M +<br />

C-M 2 where l/T2e(M) is the homogeneous<br />

line width of the Mth line (M= -1, 0, +1) in the<br />

three line spectrum. Spectral simulations and<br />

calculation of the A, B and C parameters were<br />

done using Redfield theory [15]. The best<br />

simulated EPR spectral fits to the experimental data<br />

for correlation times (TR) slower than 10" 8 sec<br />

used a non-Brownian spectral density function<br />

This extremely detailed and careful study took the<br />

linewidth analysis method to its limits, but this<br />

alone was not sufficient to completely determine<br />

the motion. The molecular dynamics was inferred<br />

from plots of the A, B, and C parameters versus<br />

each other or viscosity and temperature. Subtle<br />

changes in slopes indicated deviations from<br />

isotropic motion. The simulations to test motional<br />

models were not able to unequivocally determine<br />

the type of motion either. This work presents an<br />

independent test of the relaxation rates and the correlation<br />

times, as estimated by the CW-EPR lineshape<br />

analysis.<br />

2 Experimental<br />

Spectrometer The 9.3 GHz (X-band) pulsed<br />

EPR spectrometer for these studies follows published<br />

designs [16], [17]. There are three arms -<br />

pump, observer and detector bias. The observer<br />

and bias arms act as a conventional high sensitivity<br />

spectrometer for both linear EPR and ST-EPR experiments.<br />

The pump arm. klystron is phase<br />

locked to the observe klystron. For Free Induction<br />

decay (FID) measurements the frequency difference<br />

is zero. SR-EPR and SR-ELDOR experiments<br />

have the pump-observer frequency differ-<br />

12<br />

13<br />

S.A. Goldman.G.V. Bruno.CF. Polnaszek, and J.H.<br />

Freed, J. Chem. Phys., 56, 716-735, 1972.<br />

S.A. Goldman,G.V. Bruno, and J.H. Freed,. /. Chem.<br />

Phys., 59, 3071-3091, 1973<br />

J.S. Hwang,R. Mason,L.P. Hwang, and J.H. Freed, J.<br />

Phys. Chem., 79, 489-511, 1975.<br />

Spin labelling 1 chapter 2 Freed<br />

C Mailer, J.D.S. Danielson and B.H. Robinson. Rev.<br />

Sci. Inst., 56, 1917-1925, 1985.<br />

C Mailer, D.A. Haas, EJ. Hustedt, J.G. Gladden, and<br />

B.H. Robinson, J. Mag. Res., 91, 475-496, 1991.<br />

ence phase locked to a low frequency (MHz) oscillator.<br />

SR-ELDOR operation was simplified by the<br />

use of a Loop Gap Resonator (LGR) [18] instead<br />

of a bimodal EPR cavity. The Q of the LGR is<br />

approximately 300 which gives a 3 dB resonator<br />

bandwidth of 30 MHz. Figure 1 shows the measured<br />

value of rf field versus offset from the 9.3<br />

GHz resonant frequency of the LGR. It is clear<br />

that at the 60 - 70 MHz offset needed for SR-<br />

ELDOR the rf field in the resonator is about 50%<br />

of maximum<br />

20 40 60 80 100<br />

OFFSET FREQUENCY MHz<br />

Figure 1 Plot of 1 mm Loopgap Resonator Bandwidth.<br />

The data are the relative heights of the FID produced by<br />

a short pulse of rf power on the resonator at the frequency<br />

offset from LGR resonance as indicated on the<br />

abscissa. The solid line is the relative rf field amplitude<br />

produced by a resonator with a Q of 300.<br />

The small size of the LGR produces a high power<br />

density leading to high rf fields for moderate powers<br />

(less than 1 Watt).Typical experimental conditions<br />

were: pump power of 200 mW, observer<br />

power 100 nWatt, dead time 50 nanoseconds, acquisition<br />

time 2 nanoseconds/point (or longer) for<br />

1024 points. The pulse repetition rate was about 3<br />

kiloHertz. The decay curves took about 50<br />

seconds to obtain.<br />

CW-EPR We have carried out a CW-EPR and<br />

pulsed EPR study of per-deuterated 15 N TEMPOL<br />

in glycerol-water mixtures. The EPR linewidths in<br />

the fast motion region were measured and corrected<br />

18 Medical Advances, Milwaukee, Wis Loopgap<br />

Resonator Model #XP-0201


32 Bulletin of Magnetic Resonance<br />

Figure 2A. Pulsed ELDOR response of l5 N TEMPOL in 40 % glycerol- 60% water at a TR of 0.015 nanoseconds High field<br />

line pumped with a 250 mW 100 nanosecond duration pulse and the low field line observed with 100 microwatts in the LGR.<br />

Figure 2B. Pulsed. Saturation Recovery. Same experimental conditions as 2A except that pump and observer are both set to<br />

the low field line.<br />

for inhomogeneous broadening using Bales'<br />

method [19] to obtain the motionally dependent<br />

Lorentzian linewidths. These line widths when<br />

subtracted gave the B term of the A(M) = A + B<br />

M + C • M^ expression. (Taking the difference<br />

removed residual Lorentzian effects common to<br />

both lines - such as Heisenberg exchange [20]).<br />

To obtain longer correlation times the high glycerol<br />

percentage samples (> 85%) were cooled to subzero<br />

temperatures and the rotational correlation<br />

time (xR) was estimated from the Stokes-Einstein<br />

equation: tR = V • TJ/T where r\ and T are the<br />

19 B. Bales in Spin Labeling: Theory and Applications,<br />

Biological Magnetic Resonance vol 8 eds. L.J. Berliner<br />

and J. Reuben. Plenum Press NY 1989 Chapter 3.<br />

20 S. Lee and A. Shetty. J. Chem. Phys., 83, 499-505,<br />

1985.<br />

viscosity and temperature of the solvent, and V is<br />

the hydrodynamic volume of the spin probe. This<br />

information calibrates the motion and connects TR<br />

with the temperature and percent glycerol.<br />

Time Domain EPR An example of a pulsed<br />

ELDOR experiment is shown in Figure 2A. The<br />

magnetization from the pumped manifold arrives at<br />

the observer field position at rate Tin-l and then<br />

decays slowly to thermal equilibrium at rate Tie' 1 -<br />

A Saturation Recovery decay is shown in Figure<br />

2B. Some of the magnetization leaves for the<br />

other manifold (at the same rate as it arrived in the<br />

pulsed ELDOR experiment) and the remainder returns<br />

to thermal equilibrium by the Tie process.<br />

The opposite sign of the faster rate in the two experiments<br />

shows that it arises from cross-relax-


Vol. 14, No. 1-4 33<br />

ation between the nuclear manifolds. Superimposed<br />

on the data are the least-squares, best fits<br />

which assume two independent exponential<br />

relaxation components and a baseline. The relaxation<br />

rates of each component, the amplitudes<br />

and the baseline were adjusted using a Marquardt<br />

non-linear least squares algorithm [21].<br />

The electron and the nitrogen spin-lattice relaxation<br />

rates of per-deuterated 15 N TEMPOL in glycerolwater<br />

have been measured with pulsed ELDOR<br />

and SR. The TR range covered is from a few picoseconds<br />

to tens of nanoseconds as indicated on<br />

Figure 3. Also plotted are electron spin-lattice relaxation<br />

rates of *% TEMPOL obtained in the<br />

early pulsed experiments of Percival and Hyde [2].<br />

The Tje" 1 are similar for similar correlation times,<br />

despite the fact that glycerol-water mixtures between<br />

room temperature and -20°C were used in<br />

our experiments, and the Percival and Hyde results<br />

were obtained between -20°C and -90°C with the<br />

label in sec-buty\ benzene. This similarity suggests<br />

;<br />

_<br />

Id U)<br />

to z<<br />

I oUJ<br />

I<br />

UJ<br />

<<br />

,10 -12<br />

10*<br />

10r<br />

2<br />

10<br />

10" 10" 10" 10-8 10"<br />

"""I e is the spectrometer frequency.<br />

Spin diffusion represents the coupling of a paramagnetic<br />

center to the relaxation of distant spins in<br />

the solvent [24], most often protons, and in liquids<br />

the expression has the weak dependence on the<br />

rotational correlation time [25]:<br />

11 jSD = fi / x^ (4)<br />

where P is an adjustable constant The nuclear<br />

spin-relaxation rate is explained with just the electron-nuclear<br />

dipolar mechanism plus proton spin<br />

diffusion:<br />

1/TTOTAL=1/TEND l<br />

ln<br />

with the END term:<br />

((l/xR) 2 +(Ye.A/2) 2 )<br />

(5)<br />

(6)<br />

22 P.W. Atkins and D. Kivelson. J.Chem.Phys., 44,<br />

169, 1966.<br />

23 A. Abragam. The Principles of Nuclear Magnetism,<br />

Oxford University Press. London 1961<br />

24 A.Abragam. loc.cit. p379.<br />

25 B.I. Hunt and J.G. Powles. Proc. Phys. Soc, 88, 513-<br />

528, 1966.


34 Bulletin of Magnetic Resonance<br />

where A is the average of the A tensor. The spin<br />

diffusion term is identical in form to Tie" 1 w i*h a<br />

larger value for p.<br />

We believe that the present interpretation<br />

of the spin lattice relaxation rates, in<br />

large measure, solves the the long-standing<br />

mystery of the nature of Tie and Tin<br />

in spin labels. All the theoretical curves in<br />

Figure 3 assumed simple Brownian motion; there<br />

was no need for modified spectral density functions.<br />

It is important to note that virtually all the<br />

measured values of Tie-i and Tin 4 lie above the<br />

theoretical predictions. It is possible that residual<br />

oxygen in the glycerol-water solutions and spinlabel<br />

concentration effects may have slightly increased<br />

the observed rates.<br />

The maximum rate of T^" 1 (upper curve ) occurs<br />

when^ the motional frequency TR* 1 is equal to<br />

Ye • A / 2. Any deviation from Brownian motion<br />

would have reduced this maximum rate from the<br />

value found. The maximum value thereby becomes<br />

an extremely sensitive indicator of the type<br />

of motional process. The shape of the nuclear relaxation<br />

curve is quite different from that found by<br />

Hyde et al. [10] in their pulsed ELDOR work.<br />

The value of Tin" 1 found by them was about one<br />

MegaRadian/sec and was approximately independent<br />

of correlation time. Unfortunately these<br />

workers used a high Q EPR cavity which put an<br />

upper limit of a few MHz on the response bandwidth<br />

of their spectrometer, rendering the fast<br />

Tin 1 unobservable. Our pulsed ELDOR results<br />

show that the actual Tin' 1 is about a factor of ten<br />

shorter and clearly varies dramatically with the<br />

motion.<br />

In the study of Yin et al. [9] the lipids were<br />

moving sufficiently fast (TR about 50 picoseconds)<br />

that the Tin- 1 (~ 1.3 usec 1 ) was competitive with<br />

Tie 1 and with the 0.1-1 MHz exchange rate<br />

typical of lateral diffusion. The three rates differ<br />

by about factors of two and the data contains all<br />

three decays with the same sign. However,<br />

slower motion of the lipids (such as due to a phase<br />

transition) would lead to a rapid increase in Tin 1<br />

and make the faster exponentials difficult to<br />

distinguish. The only way to overcome this<br />

difficulty would be to perform pulsed ELDOR<br />

experiments, thereby changing the signs of one of<br />

the faster exponentials.<br />

A comment on using CW progressive saturation<br />

techniques for obtaining Tie 1 : CW-EPR assumes<br />

slow passage through the resonance line. The<br />

typical 100 kHz Zeeman modulation frequency<br />

used for detection must be much slower than<br />

Tie" 1 . The long Tie's found for motions slower<br />

than 10 nanoseconds violate this condition. This<br />

is one of the reasons why CW-EPR saturation experiments<br />

are difficult to perform: the modulation<br />

acts as a relaxation mechanism. Knowledge of the<br />

correct spin-lattice relaxation rate is important in<br />

order to avoid operating in saturation.<br />

4 Conclusions<br />

We have determined the electron and nuclear spinlattice<br />

relaxation rates in a nitroxide spin label.<br />

The rates found agree well with those predicted assuming<br />

isotropic Brownian motion. The mechanisms<br />

of spin rotation, proton spin diffusion and<br />

electron-nuclear dipolar coupling appear to explain<br />

the spin label spin-lattice relaxation mechanisms<br />

quite satisfactorily. Instrumental development is<br />

clearly at a stage where the full range of Tie" 1 and<br />

Tin" 1 rates can be studied with relative ease. We<br />

note that in this fast motional regime both spin lattice<br />

relaxation rates are approximately simply proportional<br />

to the correlation time but have opposite<br />

functional dependencies: T," 1 «= TR and T^" 1 «=<br />

TR" 1 . Pulsed SR-ELDOR is much more directly<br />

able (than CW-EPR) to connect the experimentally<br />

determinable parameters (spin lattice relaxation<br />

rates) to the general and powerful methods of<br />

interpretation in terms of motional correlation<br />

functions and the associated spectral density<br />

functions.


Vol. 14, No. 1-4 35<br />

New Developments in<br />

Pulsed Electron Paramagnetic Resonance:<br />

Direct Measurement of Rotational<br />

Correlation Times from<br />

Decay Curves<br />

Duncan A. Haas, Colin Mailer, Tetsukuni Sugano and Bruce H. Robinson<br />

Department of Chemistry<br />

University of Washington<br />

Seattle WA 98195<br />

Perdeuterated ^N TEMPOL in glycerol-water mixtures and spin-labeled hemoglobin are studied using the time<br />

domain technique of saturation recovery electron double resonance (SR-ELDOR), in the ultra-slow motional time<br />

regime. The electron spin lattice relation rate (Tig" 1 ) and the nitrogen spin-lattice relaxation rate (Tin* 1 ) are determined.<br />

Moreover, the characteristic rotational correlation times (TR) are measured directly. It is not possible,<br />

in general, to separate uniquely these rates from one another with Saturation Recovery Electron Paramagnetic<br />

Resonance (SR-EPR) experiments alone. The SR-ELDOR technique of pumping one spin manifold or orientation<br />

and observing at another changes the sign of the amplitudes of selected components contributing to the overall<br />

signal. Pooling the SR-EPR and various SR-ELDOR spectra enables all rates to be uniquely determined.<br />

1 Review of Isotropic motion<br />

Perdeuterated 15 N-Tanol [1] and spin-labeled<br />

hemoglobin (sl-Hb) [2] have been used as model<br />

systems undergoing Brownian isotropic rotational<br />

motion in the micro to millisecond time range as<br />

described by the characteristic rotational correlation<br />

time, (TR). The continuous wave technique of<br />

Saturation Transfer EPR (ST-EPR), which is very<br />

sensitive to motion on this time scale [3], relies on<br />

the competition among the different magnetization<br />

transfer mechanism of Tie, Tin, TR and the<br />

Zeeman modulation frequency, com. There are<br />

many applications of ST-EPR to slow motion,<br />

primarily using 14 N spin labels. However, the<br />

relaxation and rotational rates can only be indirectly<br />

inferred from ST-EPR [4]. We will show<br />

that these rates may be directly obtained from SR-<br />

ELDOR experiments.<br />

L.R. Dalton, B.H. Robinson, L.A. Dalton and P.<br />

Coffey, in Advances in Magnetic Resonance VIII,<br />

(J.S. Waugh, ed.) Academic Press, NY, 1976.<br />

D.D. Thomas, L.R. Dalton, J.S. Hyde, J. Chem.<br />

Phys., 65, 3006-3024, 1976.<br />

J.S. Hyde and L.R. Dalton, Chem. Phys. Lett., 16,<br />

568-572, 1972.<br />

A.H. Beth, and B.H. Robinson, in Biological Magnetic<br />

Resonance VIII Spin Labeling: Theory and<br />

Application (L J. Berliner and J. Reuben, eds.),<br />

Plenum Press, NY, 1989.<br />

Hyde and co-workers invented the technique of<br />

Continuous Wave Electron-Electron Double Resonance<br />

(CW-ELDOR) [5]. They demonstrated that,<br />

by pumping one point in a spectrum and observing<br />

at another point, energy transfer could be detected.<br />

However, CW-ELDOR is an extremely complex<br />

technique and the results can be influenced by instrumental<br />

artefacts [6]. The technique has been<br />

applied by Stetter et al. [7] to nitroxides moving<br />

with sub-nanosecond times and by Smigel et al.<br />

[8] to nitroxides with microsecond correlation<br />

times. From CW-ELDOR, one cannot estimate<br />

either Tie or Tin independently, only their ratio.<br />

The Stetter [7] work used different isotopes of nitrogen<br />

to separate exchange due to lateral diffusion<br />

from nitrogen relaxation. (There can be no nuclear<br />

spin lattice relaxation between different isotopes so<br />

any interaction between them must come from<br />

collisions.)<br />

J.S. Hyde, J.C.W. Chien and J.H. Freed, J. Chem.<br />

Phys.,4S, 4211-4226, 1968.<br />

L.A. Dalton and L.R. Dalton, in Multiple Electron<br />

Resonance Spectroscopy eds. M.M. Dorio and J.H.<br />

Freed. Plenum Press 1979 Chapter 5.<br />

E. Stetter, H.-M. Vieth, and K.H. Hauser, J. Mag.<br />

Res., 23, 493-504, 1976.<br />

M.D. Smigel, L.R. Dalton, J.S. Hyde, and L.A. Dalton,<br />

Proc. Nat.Acad. Sci. USA, 71, 1925-1929, 1974.


36 Bulletin o[ Magnetic Resonance<br />

Time Domain EPR: In a time domain EPR experiment,<br />

the spins are subjected to a short pulse<br />

of microwave power; the time dependence of the<br />

transient relaxation after the pulsing is monitored.<br />

The rates of relaxation are functions of the motional<br />

process. Two types of time domain experiments<br />

exist: one monitors spin echoes (a coherent<br />

detection of the x- and y-components of the magnetization)<br />

and the other monitors Saturation<br />

Recovery (as a polarization detection of the z-magnetization).<br />

Freed [9] has applied spin-echo methodology to<br />

study motion in liquids. A (90°-T-180°-i-Observe)<br />

sequence measures echo height as a function of the<br />

delay time t to give the phase memory decay time<br />

TM- Other workers have studied spin-labeled<br />

membranes and vesicles - showing changes in TM<br />

with temperature, but do not relate results to any<br />

theory [10]. Freed [9] has also developed the<br />

technique of Fourier Transform EPR using<br />

multiple pulse techniques, originally developed for<br />

NMR. The variation of TM across the spectrum<br />

can be directly obtained and related to motional<br />

models. Cross-relaxation from one manifold to<br />

another can also be measured. Rates are measured<br />

from the volumes of the main and cross-peaks in<br />

the 2-D display, and not from actual decay curves.<br />

Huisjen and Hyde [11] pioneered the use of the<br />

SR-EPR technique in liquids and applied it to a<br />

number of systems [12], [13]. The Hyde group<br />

[14] has used SR-EPR, in conjunction with CW-<br />

ELDOR, to measure lateral diffusion of 14 N labelled<br />

lipids in bilayers. These important papers on<br />

translational motion set the stage foT our work in<br />

rotational motion.<br />

Measurement of rotational diffusion directly in the<br />

time domain with SR-EPR alone has proved im-<br />

9 J. Gorcester, G.L. Millhauser, and J.H. Freed, in<br />

Modern Pulsed and Continuous-Wave Electron Spin<br />

Resonance, L. Kevan and M.K. Bowman(eds.) John<br />

Wiley 1990 Chapter 3.<br />

10 K. Madden, L. Kevan, P.D. Morse and R.N. Schwartz,<br />

/. Am. Chem. Soc, 104, 10, 1982.<br />

11 M. Huisjen and J.S. Hyde, Rev. Set Inst., 45, 669-<br />

675, 1974.<br />

12 P.W. Percival and J.S. Hyde, /. Mag. Res., 23, 249-<br />

257, 1976.<br />

13 A. Kusumi,W.K. Subczynski and J.S. Hyde, Proc.<br />

Nat. Acad. Sci. USA, 79, 1854-1858, 1982.<br />

14 J.-J. Yin, M. Pasenkiewicz-Gierula, and J.S. Hyde,<br />

Proc. Nat. Acad. Sci. USA, 84, 964-968, 1987.<br />

possible. Typical of an SR-EPR experiment is that<br />

of Fajer et al. on sl-Hb [15] which produced biexponential<br />

decays. The slower rate was<br />

(correctly) identified with Tie" 1 and the faster decay<br />

rate was associated with the rotational motion<br />

of the hemoglobin. The faster decay rate remained<br />

approximately constant despite a two order of<br />

magnitude change of correlation time. These authors<br />

then used a qualitative theory of spin diffusion<br />

to extract an effective correlation time. The<br />

data agreed only approximately with the known 1R<br />

values. The problem with this type of experiment<br />

is the strong influence of the nitrogen nuclear spinlattice<br />

relaxation, on the decays. As explained below,<br />

the observed decays were a mixture of all<br />

three relaxation processes. This makes the protocol<br />

for data collection and for analysis of the various<br />

decays much more complex than originally<br />

expected. Attempts in our laboratory using<br />

Saturation Recovery alone were similarly unsuccessful<br />

(unpublished experiments with Dr. P.<br />

Fajer). We believe we have now solved these<br />

fundamental problems using pulsed SR-ELDOR<br />

and SR-EPR together.<br />

We will demonstrate that Pulsed EPR techniques,<br />

such as SR-EPR and SR-ELDOR, can be used to<br />

obtain both the nuclear and electronic spin-lattice<br />

relaxation rates directly and monitor changes in exchange<br />

rates. Moreover, the correlation time connecting<br />

two spectral positions can be directly measured<br />

as well. Relaxation time measurements, especially<br />

nuclear spin-lattice relaxation (Tin). can<br />

also characterize the motion in the ultra-slow motional<br />

time range.<br />

2 Theory<br />

Figure 1 shows an absorption CW-EPR signal<br />

when the correlation time is longer than 0.1 microseconds.<br />

The figure illustrates the dependence of<br />

the signal on molecular orientation. The reorientation,<br />

characterized by a rotational correlation time,<br />

TR, moves the magnetization around within a given<br />

manifold. Nuclear spin flips, occurring at rate<br />

T^" 1 , move the magnetization from one manifold<br />

to the other without causing molecular reorientation.<br />

Regardless of orientation or manifold, the<br />

magnetization relaxes to the lattice at rate Tie'<br />

15 P. Fajer, D.D. Thomas, J.B. Feix, and J.S. Hyde,<br />

Biophys. J., 50, 1195-1202, 1986.


Vol. 14, No. 1-4 37<br />

Theory [4], [16] predicts for the SR-ELDOR experiment<br />

that the TR of a molecule should appear as<br />

an exponential decay with a rate (TR" 1 + T^" 1 ) and<br />

that Tin should appear as a decay of rate (T\a' 1 +<br />

T^" 1 ). These two processes have very similar<br />

rates, which is why motional rates have been so<br />

difficult to determine. However, suitable positions<br />

of pump and observer can be chosen so that the<br />

signs and the amplitudes of both Tin and the motional<br />

decay curves can be individually changed,<br />

enabling analysis to disentangle the rates. The SR-<br />

ELDOR experiment measures the decay of the<br />

transient component of the z magnetization,<br />

(Mz(t;v,8)), where v is the nuclear manifold<br />

(which is ±1/2 for 15 N), and 6 is the orientation,<br />

as shown in Figure 1. The evolution of the magnetization<br />

can be understood, qualitatively at least,<br />

by a simple population analysis treatment<br />

Lattice!<br />

^Figure 1: Linear Absorption EPR spectrum of sl-Hb<br />

illustrating the positions where SR-ELDOR experiments<br />

are performed and the dependence of the resonance<br />

'positions of the magnetization on spin-label orienta-<br />

°" The figure further illustrates how TR, and Tin<br />

! spectral diffusion, whereas Tie induces true spin-<br />

"J relation.<br />

^•Sugano, C. Mailer, and B.H. Robinson, /. Chem.<br />

zfkys. 87, 2478-2488, 1987.<br />

Therefore the master equation is:<br />

{(M2(t;v,e))-(Mz(t;-v,0))}<br />

-DV2 (Mz(t;v,e))<br />

(1)<br />

where D is the Einstein Rotational Diffusion<br />

Coefficient (D = 1/6 tR) and V 2 , is the angular<br />

Laplacian operator. The system is pumped with a<br />

weak selective pulse so that the spin system is<br />

excited in manifold vp and at orientation 8p. The<br />

the observer frequency is set to manifold v0 and at<br />

orientation 80. The recovery signal then has the<br />

approximate form:<br />

(Mz(t;vo,eo vp,ep)) =<br />

* (l + P2(cos(9o)) • P2(cos(8p))e- t/x R)<br />

where is the magnetization at<br />

the observer position subject to the condition that<br />

all of the magnetization was at the pump position at<br />

time zero; and (Mz(O;vp,0p)) is the initial magnetization<br />

at time zero (when the pumping is completed).<br />

P2 (cos(8)) = (3 cos 2 (8) - l)/2 is the I =<br />

2 component of the Legendre polynomials P;(x),<br />

and fvy = +1 if v = v' and fvv- = -1 if v = -v'<br />

(which is the case of going from one manifold to<br />

the other). This result assumes a short duration<br />

pump time, a low observer amplitude and that the<br />

higher rotational functions are not significant [17].<br />

Equation 2 predicts that one will observe 4 exponentially<br />

decaying components, wherein the rate of<br />

each component is a linear combination of T^" 1 ,<br />

Ti,," 1 , and TR" 1 . The amplitude of the components<br />

containing Tin" 1 , may be switched in sign by setting<br />

the pump and observer positions to different<br />

spin manifolds. The amplitude of the components<br />

containing TR" 1 may be switched in sign by setting<br />

the pump and the observer to different ends of the<br />

same manifold.<br />

17 T. Sugano, "A Study of Very Slow Rotational<br />

Diffusion by SR-EPR", Ph.D. Thesis, University of<br />

Washington, 1987.


38 Bulletin of Magnetic Resonance<br />

3 Experiment<br />

The details of the spectrometer and its performance<br />

are discussed more fully in another paper in these<br />

Proceedings [18]. SR-EPR and SR-ELDOR experiments<br />

are both polarization experiments which<br />

measure the recovery of the system to equilibrium.<br />

A selective pulse is applied to the spins which alters<br />

the polarization and burns a partial hole in the<br />

line. Under conditions of SR-EPR the frequency<br />

of observation is the same as the pump and therefore<br />

the signal is that of a pure recovery as the polarization<br />

spreads throughout the spin system and<br />

out to the lattice. Under conditions of SR-ELDOR,<br />

the frequency of the observer is different from that<br />

of the pump and the arrival of magnetization, as<br />

well as relaxation to equilibrium, are both detected.<br />

It is neither necessary, nor desirable, for the pump<br />

to be coherent with the observer. The experiments<br />

obtain decay curves taken with various pump times<br />

and pump-observer frequency differences; the rotational<br />

correlation time is determined by the solvent<br />

viscosity and temperature. The pump time<br />

controls the relative amplitudes of the various components<br />

which contribute to the overall relaxation<br />

of the magnetization; and the pump-observer frequency<br />

difference controls the signs of the amplitudes<br />

of the exponentially decaying components in<br />

ways predicted by equation 2.<br />

The experiment proceeds as follows: for one set of<br />

experimental conditions, a number of decay curves<br />

over different time scales are obtained. These decays<br />

are linked by the fact that the exponentials<br />

which comprise them are all identical and only the<br />

time range of data collection is altered. For a different<br />

set of conditions e.g. in pulse length or in<br />

observer field position, the relative amplitudes<br />

and/or signs of the individual decays are altered<br />

and another set of linked spectra is collected Both<br />

sets of linked data are then pooled with the constraint<br />

that the rates of the exponentials which<br />

make up the spectra are common to all and the<br />

amplitudes of each component within a linked set<br />

are the same. This technique [19], as an<br />

18 C. Mailer, B.H. Robinson and D.A. Haas, "New<br />

Developments in Pulsed EPR: Relaxation Mechanisms<br />

of Nitroxide Spin Labels", [these proceedings] 1992.<br />

19 J.M. Beecham, E. Gratton, M. Ameloot, J.R.<br />

Knutson, and L. Brand, in Fluorescence<br />

Spectroscopy: Principles J.R. Lakowicz (ed.) Plenum<br />

Press NY Volume 2 Chapter 5,1991.<br />

application of Global Analysis, is well recognized<br />

in optical spectroscopy as being extremely effective<br />

in determining relaxation rates.<br />

Once one has obtained a unique set of decay rates<br />

that minimizes the global x 2 one must identify the<br />

processes which gives rise to each individual decay<br />

rate i.e.whether it is due to Tie" 1 , Tin" 1 o f rotation.<br />

Suitable choice of experimental conditions<br />

for acquisition of the decays enable the rates to be<br />

distinguished. The strategies we have found useful<br />

are:<br />

(i) Pumping and observing at the magic angle<br />

will reduce the amplitude of the motional rate,<br />

leaving just the Tie and Tin" 1 .<br />

(ii) Pumping in one spin manifold and observing<br />

in the other will always invert the amplitude of<br />

a Tin containing term.<br />

(iii) Pumping at one extreme turning point of a<br />

spin manifold and observing the other will invert<br />

the amplitudes of the motionally dependent components.<br />

(iv) Suppression of free induction decay (FID)<br />

is always important. In the nanosecond range of<br />

motion the FED has approximately the same time as<br />

Tie and Tin.<br />

(v) A pump time that is long compared to the<br />

relaxation time of a particular component will tend<br />

to suppress the amplitude of that component.<br />

(vi) Pooling the Saturation Recovery EPR and<br />

pulsed ELDOR decay curves for analysis is essential<br />

for all rates to be found - no single experiment<br />

alone is sufficient.<br />

4 Results<br />

We have obtained correlation times from perdeuterated<br />

15 N TEMPOL in glycerol-water mixtures<br />

over the TR range from 0.1 to 100 microsec- .<br />

onds: Bi-expOnential recovery curves were ob-1<br />

tained when using the SR-ELDOR protocol de- ^<br />

scribed above, when the "magic angle" (8 = 54°);l<br />

was chosen for the pumping and observing posi-Jj<br />

tions (as shown in Figure 1). In Figure 2 are<br />

plotted the Tie' 1 and Tin" 1 rates versus correlation<br />

time in the slow motion range. The very weak;<br />

dependence of T^' 1 on correlation time suggest|<br />

that the more traditional relaxation mechanisms"


Vol. 14, No. 1-4<br />

such as spin rotation and Electron-Nuclear-Dipolar<br />

coupling (END) are not the dominant ones. The<br />

power law dependence of Tie' 1 is on the order of<br />

1/TR 1/8 . The NMR literature suggests that a model<br />

of spin diffusion in liquids will have the same<br />

power law dependence as that experimentally<br />

observed [20]. This same power law dependence<br />

was also observed in similar measurements using<br />

SR-EPR [15]. The solid line superimposed on the<br />

Tie" 1 data is defined as:<br />

(coexR)<br />

(3)<br />

where fie is an adjustable parameter, and the second<br />

term (r.h.s.) is the END term and the third term<br />

u<br />

q><br />


40<br />

SR-EPR: «B<br />

Bulletin of Magnetic Resonance<br />

Figure 3: Three different SR-ELDOR spectra, taken at different pump and observation positions (see Figure 1). The top<br />

curve is the SR-EPR spectrum at position B (in Figure 1), the middle curve is the SR-ELDOR spectrum pumping at position<br />

A and observing at position B, and the bottom spectrum is the SR-ELDOR spectrum pumping at position B and observing<br />

at position C. Superimposed on each spectra is the least-squares best fit consisting of an adjustable baseline and three<br />

components each of which is a single exponential decay (see equation 2), the rates are given in Figure 2.<br />

Figure 3 shows an example of the SR-EPR and<br />

SR-ELDOR data for sl-Hb tumbling with xR in the<br />

1 to 5 microsecond time range. The top curve,<br />

(SR-EPR at B) is the SR-EPR data acquired at<br />

point B (defined in Figure 1). This curve is a<br />

simple recovery and all components have the same<br />

sign. The bottom curve (SR-ELDOR from B to C)<br />

is the data when the pump is set on B and the observer<br />

frequency is set to C. This represents a<br />

jump from one manifold to the other and the amplitude<br />

of the components, containing a Tin" 1 in the<br />

rate, change sign. The middle trace of Figure 3<br />

(SR-ELDOR from A to B) is the data for the pump<br />

on A and the observer on B. This corresponds to<br />

the pump-observe case within a manifold.<br />

According to equation 2, the amplitude of the<br />

components containing TR in the rate will change<br />

sign. Clearly there is a peculiar dip in the shape of<br />

this SR-ELDOR curve. The dip is characteristic of<br />

the component that depends directly on the correlation<br />

time. Superimposed on each of the data sets<br />

is a least-squares best fit simulation composed of<br />

three exponentials and a baseline. The three rates<br />

of the exponentials are the same in all three data<br />

sets. The rates arc interpreted, according to equation<br />

2, as Tie* 1 , Tin" 1 and TR' 1 as 0.0937, 1.983<br />

and 0.329 MRad/sec respectively with around a|<br />

10% error. Notice that the motional rate is in be-.|<br />

tween the spin-lattice relaxation rate and the nu-|<br />

clear relaxation rate. These data are also shown inf<br />

Figure 2. The nominal correlation time, d cte f1<br />

mined from solvent viscosity and hydrodynami


Vol. 14, No. 1-4 41<br />

radius of sl-Hb, is 1.5 |j.sec. From the L"/L ratio<br />

calibration curve of the ST-EPR [4] one may estimate<br />

the correlation time to be 4.0 p.sec and the<br />

rotational correlation time measured by SR-<br />

ELDOR is 3.3 jisec, which is in excellent agreement<br />

with the ST-EPR calibration data. This<br />

work, demonstrates that the different components<br />

of the experimental curve can be uniquely identified<br />

and their rates quantitatively measure, and<br />

hence Tie, Tin and TR may be directly measured.<br />

This is the first time that xR has been directly<br />

measured in EPR and represents a<br />

fundamental advance in time domain<br />

methodology.<br />

5 Conclusions<br />

We have measured the values of the electron and<br />

nuclear spin lattice relaxation times in the ultraslow<br />

motion region using SR-ELDOR and SR-<br />

EPR and find that the measured values of Tie' 1<br />

depend on 1/TR 1 / 8 , or a power law dependence,<br />

which is consistent with a model of spin diffusion<br />

in liquids. The measured values of Tin are well<br />

described by the electron-nuclear dipolar relaxation,<br />

(which has no adjustable parameters in it)<br />

and only the rates at motional times longer than<br />

100 fisec suggest the need for a spin diffusion<br />

mechanism for Tin as well as Tie. Notice that for<br />

the case of spin labels the nuclear relaxation is<br />

much faster than the electron relaxation (an unusual<br />

situation). This is primarily a result of the<br />

ability of the electron, at this particular rotational<br />

rate, to efficiently relax the nucleus and the fact<br />

that the electron has very few other spins to relax<br />

it. By using SR-ELDOR and by pooling the data<br />

at different orientations the rotational correlation<br />

time can be measured directly, and is seen experimentally<br />

as a single exponential relaxation. This<br />

is, we believe, the first time rotational motion has<br />

been quantitatively measured as a single exponential<br />

decay in this type of experiment. These exper-<br />

- imental results suggest that it may be possible to<br />

. directly measure rotational reorientation by SR-<br />

, ELDOR even for systems characterized by aniso-<br />

> tropic motion.<br />

M-<br />

|r. These type of time domain experiments are a necunderpinning<br />

and improvement on tradi-<br />

CW techniques. Quantitative simulation of<br />

f-EPR spectra requires knowledge of the relax-<br />

ation times of the spin system. When performing<br />

progressive saturation studies using CW-EPR, the<br />

experiments can only be used to detect relative<br />

changes in relaxation times when motion is longer<br />

than a nanosecond [22]. (Absolute values are difficult<br />

to obtain accurately because the effective relaxation<br />

time for the CW experiment is a complex<br />

mixture of competing relaxation processes.) Direct<br />

measurement of Tie" 1 an d Tin w ^h pulse techniques<br />

is clearly superior.<br />

Prior to the time domain experiments it was found<br />

in this laboratory that the Tin* 1 value in calculations<br />

of ST-EPR spectra in the microsecond motional<br />

range had to be set artificially high for good<br />

agreement between simulation and experiment<br />

[23]. The simulation assumes that the electronnuclear<br />

dipolar (END) interaction is the sole mechanism<br />

for nuclear spin-lattice relaxation. The data<br />

in Figure 2 clearly show that other mechanisms<br />

add to the END rate (e.g. from proton spin diffusion)<br />

justifying the ad hoc addition of an extra rate<br />

into the calculations.<br />

22 B.H. Robinson, C. Mailer and D.A. Haas, Biophys. J.<br />

61, A167, Abstract # 960, 1992.<br />

23 D. Haas, R. St Denis, C. Mailer, B.H. Robinson, 13th<br />

Intl. EPR Conf., Denver, CO, 1990.


42 Bulletin of Magnetic Resonance<br />

Non-Linear effects in Standard 2D NOE<br />

experiments in Coupled spin systems<br />

R.Christy Rani Grace* and Anil Kumar*t<br />

''Department of Physics and ^Sophisticated Instruments Facility<br />

Indian Institute of Science, Bangalore - 560 012, INDIA.<br />

INTRODUCTION<br />

The nuclear Overhauser effect (NOE)<br />

which monitors the transfer of magneti-<br />

zation from one spin to another, is criti-<br />

cally dependent on the internuclear dis-<br />

tance and has therefore become a pow-<br />

erful tool for elucidation of the struc-<br />

tures of Biomolecules. Experimental<br />

methods for monitoring these effects of-<br />

ten use radio frequency pulses which si-<br />

multaneously excite and/or detect sev-<br />

eral spins at a time. If the spins are<br />

not coherently coupled (no J coupling),<br />

there are no non-linear effects of the<br />

pulses, except for a scaling factor. The<br />

non-linear effects in the presence of J-<br />

coupling for one-dimensional NOE ex-<br />

periments are well known(l,2). In this<br />

paper the non-linear effects in the 2D<br />

NOE (NOESY) experiment are anal-<br />

ysed in detail.<br />

The standard NOESY experiment<br />

uses the sequence 90° — ii — 90° — rm —<br />

90° —12, in which relaxation takes place<br />

during the mixing interval rm . The<br />

rate equations governing relaxation are<br />

exactly identical to the transient NOE<br />

experiment(3-5). It has been known<br />

that for uncoupled spins each cross-<br />

section in the NOESY experiment is<br />

equivalent to a ID transient NOE exper-<br />

iment in which the peak corresponding<br />

to the diagonal peak is selectively in-<br />

verted^). When there are J-couplings<br />

present in the spin system, selective in-<br />

version has to be carefully defined. Re-<br />

cently, it has been shown that for small<br />

values of the second pulse (90° — t\ —<br />

a — Tm — 90° — t2 at frequency u>i = ua is<br />

equivalent to a ID difference transient<br />

NOE experiment in which the transition<br />

at frequency ua is selectively inverted.<br />

This is true irrespective of the strength<br />

of the coupling(6,7). It has also been<br />

shown for weakly coupled spins that in


Vol. 14, No. 1-4 43<br />

the standard NOESY experiment, any<br />

cross-section parallel to CJ2 at u\ = u>a, is<br />

equivalent to a ID transient experiment<br />

in which, the whole multiplet of which<br />

ua is a part is non-selectively inverted.<br />

When the spins are strongly coupled the<br />

90° pulse distributes the perturbation<br />

over all the transitions of the strongly<br />

coupled network and the 2D NOE ex-<br />

periment is not equivalent to any stan-<br />

dard transient ID experiment. In ad-<br />

dition, the third pulse in the NOESY<br />

experiment (the measuring pulse) mea-<br />

sures the state of the spin system in a<br />

non-linear manner for finite angles. As<br />

a result it is shown here that in strongly<br />

coupled spin systems one can obtain<br />

'cross-peaks' in the standard NOESY<br />

experiment without relaxation. The ori-<br />

gin of these cross-peaks in terms of the<br />

non-linearity of the second and/or the<br />

third pulse is also discussed with the<br />

help of an ABX spin system.<br />

Cross-correlations between pairwise<br />

dipolar relaxation and between dipolar<br />

and other mechanism of relaxation such<br />

as chemical shift anisotropy(CSA) are<br />

known to yield a multiplet effect in J-<br />

coupled spectra(7-ll). A measurement<br />

of this effect in one and two dimensional<br />

spectra is carried out using small angle<br />

pulses. Recently Osckinat et al. have<br />

used small angles for the second and<br />

the third pulses in the NOESY experi-<br />

ment and have shown that in the initial<br />

rate approximation the effect of cross-<br />

correlations is present in all the mul-<br />

tiplets of an AMX spin system(7). In<br />

their experiment the direct pumping ef-<br />

fects and cross-correlation effects both<br />

give rise to multiplet effects. We pro-<br />

pose here simple modifications which al-<br />

lows the direct pumping effects to be<br />

absent, with the cross-correlations ex-<br />

clusively exhibiting multiplet effects in<br />

weakly coupled spins.<br />

A. STRONG COUPLING<br />

INDUCED CROSS-PEAKS IN<br />

NOESY<br />

The signal in a NOESY experiment<br />

utilizing 90° — ti — a — rm — j3 — t2 se-<br />

quence in which only longitudinal mag-<br />

netization is retained during rm period<br />

can be expressed as,<br />

S(h,t2) =<br />

Tr{{Fx)exp(-iHt2)exp(-if5Fx)<br />

[exp(—iaFx)exp(—iHti)exp(—i'^Fy)<br />

exp(WTm)exp(i(3Fx)exp(iHt2)}<br />

(1)<br />

where the prime indicates retention of<br />

only the diagonal elements of the den-<br />

sity matrix after the a pulse, W is the<br />

matrix governing relaxation during rTO<br />

period and cr0 is the initial density ma-<br />

trix. If cr0 is an equilibrium density ma-


44 Bulletin of Magnetic Resonance<br />

trix, then only single quantum coher-<br />

ences are created during ii period and<br />

since during period t2 only single quan-<br />

tum coherences are detected, the above<br />

equation can be written as(7)<br />

exp(WNxNrm)<br />

(2)<br />

X) represents a matrix which<br />

transforms the N populations into M<br />

single quantum coherences by a pulse<br />

of angle 7X . The N populations are<br />

arranged in descending order of energy<br />

while the M coherences represented by<br />

vectors


Vol. 14, No. 1-4 45<br />

PN*M{I) =<br />

-s 2<br />

-C 2 +<br />

S 2 (l - v 2 )<br />

v 2 S 2<br />

c 2<br />

-C 2<br />

-u' 2 S 2<br />

C 2 +<br />

S 2 (u 2 - 1)<br />

s 2<br />

—S 2<br />

u 2 S 2<br />

—C 2 —<br />

S 2 (u 2 - 1)<br />

c 2<br />

-C 2<br />

c 2 -<br />

S 2 (l - v 2 )<br />

-v' z S 2<br />

s 2<br />

u 0 0 0<br />

0 v 0 0<br />

0 0 v 0<br />

0 0 0 u<br />

(6) '<br />

where S = Sin(7 / 2) ; C = Cosfr / 2); u = Cos 9 + Sin 9 ;v = Cos 9 - Sin 9 and<br />

tan(29) — JAB/($A — &B) defines the strength ofthe coupling.<br />

Frequencies<br />

U)2<br />

(1) Diagonal peaks<br />

1-3<br />

3-4<br />

1-3<br />

3-4<br />

1-2 1-2<br />

2-4 2-4<br />

(2) Auto-peaks<br />

1-3 2-4<br />

3-4 1-2<br />

2-4 1-3<br />

1-2 3-4<br />

(3) Cross-peaks<br />

1-3 3-4<br />

3-4 1-3<br />

2-4<br />

1-2<br />

1-3<br />

2-4<br />

1-2<br />

3-4<br />

1-2<br />

2-4<br />

1-2<br />

3-4<br />

1-3<br />

2-4<br />

V<br />

Table. 1.<br />

u\4GIC}<br />

+ 4S 2 aS 2 {v 2 + (1 - ^ 2 ) 2 } — (1 — C2aC2/3)(l — V 2 )]<br />

u 2 v '[-4^(1-«V) +2(1<br />

-c2ac<br />

U 2 V<br />

Intensities *(—5*2<br />

%CIC} + 4S 2 aS 2 p{u 2 + {v 2<br />

2 M^(i-^) + 2(i<br />

v 4 [-2C 2 aC 2 -2S 2 aS 2 {uU<br />

- m<br />

— {u 2 \o — zu<br />

2 — v)(C2a<br />

-<br />

- (u 2 - 1)<br />

2 } + (1 - C2aC2P)v 2 }<br />

- (1 - v'<br />

A)<br />

— (1 — ^2cc^- / 2/3)(l — u )\<br />

iff) + (1 - V 2 )(C2c - CV)]<br />

w) - (1 - v 2 )(C2a - C2P)}<br />

\2\ i^ /i /"i /"i \ 2*1<br />

/ J ' v ^- / 2a^- / 2/5/^ J<br />

u r — '^\ -*- — ^- J '2f"y^-^2/?/<br />

Cos(i) ; ^i = Sin(i) ; d = Cos(^) ; Si = Sin(^) where i = a, f3 and<br />

u = Cos 9 + Sin 9 ; v = Cos 9 - Sin 9.


46<br />

The origin of these cross-peaks lies in<br />

the creation of a initial state in which<br />

the initial perturbation is distributed<br />

over all the transitions of a strongly cou-<br />

pled spin system as well as due to the<br />

non-linear measurement of the strongly<br />

coupled spin system by the third 90°<br />

pulse. The initial state in this experi-<br />

ment can also be described using mag-<br />

netization modes(12,13). For an AB<br />

system the initial state in terms of the<br />

magnetization modes at various cross-<br />

sections parallel to UJ^ is given in Ta-<br />

ble.2. From these it is seen that the<br />

single spin modes of both the spins are<br />

created in each cross-section. This is the<br />

origin of the cross-peaks in strongly cou-<br />

pled spins. In the limit of weak coupling<br />

(u = v = 1) each cross-section contains<br />

only one single spin mode belonging to<br />

the inverted spin and the cross-peaks<br />

are absent.<br />

ABX spin system<br />

If there are two groups of spins which<br />

are strongly coupled among themselves<br />

but weakly coupled to others then it<br />

is not a priory clear that there will<br />

be cross-peaks between the two groups.<br />

This is investigated here with the help<br />

of an ABX spin system. Fig.2 shows<br />

the standard NOESY (a = j8 = 90° )<br />

spectrum calculated using eqn.[5] for an<br />

Bulletin of Magnetic Resonance<br />

ABX spin system with zero mixing time.<br />

Actual intensities of the peaks in the<br />

2D spectrum is obtained by multiply-<br />

ing the expressions given in Table.3 with<br />

the corresponding ID intensities in both<br />

u>i and ui2 dimensions of that particu-<br />

lar peak. From this it is seen that ev-<br />

ery peak has a cross-peak to every other<br />

peak. The cross-peaks in this spectrum<br />

including those between A and B spins<br />

and between AB spins and X spin arise<br />

due to the strong coupling among the A<br />

and B spins, and disappear under weak<br />

coupling approximation. The appear-<br />

ance of these cross-peaks needs further<br />

investigation in terms of whether they<br />

are due to the non-linearity of the sec-<br />

ond or the third pulse. To investigate<br />

this, calculations have been carried out<br />

for the cases when the excitation pulse is<br />

small(in the linear regime) or the detec-<br />

tion pulse is small(in the linear regime).<br />

The following results were noted from<br />

these experiments:<br />

(i) 90° - a - 90° Experiment<br />

The ABX spin system has eight AB-<br />

transitions and six X-transitions two of<br />

which are between states which are un-<br />

perturbed by strong coupling (the so ><br />

called pure states 1,2,7 and 8)(14). In<br />

this system in the 90° — a — 90° experi-<br />

ment there are no cross-peaks from the


Vol. 14, No. 1-4 47<br />

K.-<br />

At ui<br />

< AA2 >Tm=0<br />

< ABZ >Tm=0<br />

< AAZBZ >Tm=0<br />

u 2<br />

u 2<br />

V 2<br />

V2<br />

-u 2 (l - v 2 )<br />

-u 2 (l+v 2 )<br />

0<br />

Table. 2.<br />

~V 2 (1 + U 2 )<br />

v 2 (l - v 2 )<br />

0<br />

2 u 2<br />

• • • 1i<br />

• • • 4<br />

• • • *<br />

to,<br />

I<br />

•<br />

v 2<br />

-u\l + v 2 )<br />

-u 2 (l - v 2 )<br />

0<br />

V 2 {l-V 2 )<br />

-v 2 (l + u 2 )<br />

0<br />

= -(1+v 4 )<br />

• - (1-v 4 )<br />

- (1-u 4 )<br />

- (1-U 2 V 2 )<br />

Figure 1. Schematic spectrum of an AB spin system calculated for the 90°—90°—90°<br />

2D NOESY experiment with zero mixing time. The symbols represent -PMxiv(90°) x<br />

-fWxM(90°), the |FX| 2 are given along the ID spectra and the final intensities are<br />

obtained using eq [5].


48<br />

B<br />

A<br />

C_<br />

2<br />

V_<br />

U<br />

V<br />

u<br />

V.<br />

AiB3<br />

A1B1<br />

A,<br />

An<br />

A12<br />

37 13<br />

A2X2<br />

A2B3<br />

A2B2<br />

A3X4<br />

^3X3<br />

A3X2<br />

A3X1<br />

A3B3<br />

A3B1<br />

-43<br />

u 2<br />

26 47<br />

A4X2<br />

AtB3<br />

B1X4<br />

fijA'3<br />

BiX2<br />

Bu<br />

B13<br />

B12<br />

B,<br />

u 2<br />

B2X4<br />

4"<br />

- Bt for i = 1 io 4<br />

B2X3<br />

B2X2<br />

B2X1<br />

B24<br />

B23<br />

B2<br />

B3X6<br />

B3X5<br />

B3X4<br />

B3X3<br />

B3X2<br />

BsXj<br />

B34<br />

B3<br />

68<br />

B4X4<br />

B4X3<br />

B4X2<br />

B4X1<br />

B4<br />

36,<br />

A'l4<br />

X13<br />

X12<br />

X!<br />

12<br />

Bulletin of Magnetic Resonance<br />

A'26<br />

X25<br />

x24<br />

x2<br />

46 35<br />

X35<br />

X34<br />

X3<br />

X4<br />

x4<br />

78<br />

45 1<br />

Xe<br />

.1 SI 1 Cl Cl I Si<br />

o>2<br />

* V — V — V<br />

A2 — A5 — A25<br />

D{j A12 — A15<br />

AjBi for i ^ j and j > i A"23 = X35<br />

— B{Xj for i = 1 to 4 (md A24 = A45<br />

A{X2 for j = 1 io 6 A26 = A56<br />

5.A-.2 " AXX, = A4A6<br />

A3JB4 A^X'3 = A3X4<br />

A34 A2A4 = A4A3<br />

A3X1 = A2A6<br />

Figure 2.


Vol. 14, No. 1-4<br />

a3<br />

a4<br />

Peaks<br />

Al<br />

A2<br />

A3<br />

A4<br />

Au<br />

Tl^ —<br />

A13<br />

A14<br />

A23<br />

A24<br />

Xi<br />

x2<br />

x3<br />

x4<br />

X6<br />

XX2<br />

Xi3<br />

X\4<br />

Xm<br />

x23<br />

X24<br />

X26<br />

^34<br />

-^36<br />

X4Q<br />

Strong<br />

Coupling<br />

-(1 + 5/6^)<br />

-(1 + 56?,)<br />

-(1 + *%)<br />

-(1 + s/b 2 )<br />

-(1 +5)<br />

-(1 +564/63)<br />

-(1 + 5/6x62)<br />

-(1 + 56x62)<br />

-(1 + 363/64)<br />

-(a 2 4 + ka 2 3)<br />

-(l + k)<br />

-(al + kaf)<br />

—(al + ka\)<br />

-(a 2 + kal)<br />

sa3 — a4<br />

ka2a4 — axa.3<br />

kaxa4 - a2a3<br />

-a3a4(l + k)<br />

—(ax + 502)<br />

—(a2 + 5ax)<br />

504 — a3<br />

-axa2(l + k)<br />

Table 3. Intensities of the peaks<br />

Weak<br />

Coupling<br />

-2<br />

-2<br />

-2<br />

-2<br />

-2<br />

-2<br />

-2<br />

-2<br />

_2<br />

0<br />

-2<br />

-2<br />

-2<br />

0<br />

0<br />

0<br />

0<br />

0<br />

-2<br />

_2<br />

0<br />

0<br />

0<br />

0<br />

61 = V+/U+<br />

62 = v_/u_<br />

h =<br />

b4<br />

=<br />

— ,,2<br />

U +~ V +<br />

u_v+v_/u+;<br />

u+u_v+/v_;<br />

m2<br />

k<br />

= u<br />

s<br />

-<br />

=<br />

f,,2 ,,2<br />

Peaks<br />

AlBl<br />

A2B2<br />

A3B3<br />

A4B4<br />

AiB2<br />

AiB3<br />

AXB4<br />

A2B3<br />

A2B4<br />

AxXx<br />

AiX2<br />

AiX3<br />

A1X4<br />

AiX6<br />

A2XX<br />

A2X2<br />

A2X3<br />

A2X4<br />

A2X&<br />

A3X2<br />

A3X3<br />

A3X6<br />

A4X1<br />

A4X2<br />

A4X4<br />

Strong<br />

Coupling<br />

s/b 2 -l<br />

562 -1<br />

sb 2 -l<br />

s/bl - 1<br />

5-1<br />

564/63 - 1<br />

563/64 - 1<br />

56162 - 1<br />

5/6i62-l<br />

2 2<br />

a4miif+ — a3m2U_<br />

[n3mx — ra2ui]<br />

a2m2u 2 _ - axmiv\<br />

ax[n3mx — rn2u 2 _]<br />

-a4[n3mx — m2u 2 _]<br />

—a3[nxmx — m2f 2 _]<br />

—[nxmx — m2v 2 _]<br />

r 2 1<br />

—a2[ni?Tix ~ Tn2v_\<br />

axm2v 2 _ - a2mxu%<br />

a3mxv\ — a4m2v 2 __<br />

—[n2m2 — mx^]<br />

a2JVl2 m 2 ~~ ^•1^4.]<br />

— a4\jl2V(l2 — T^lU-iJ<br />

—a3[?i47TT.2 — mxu , ]<br />

-[n4m2 - mxu\]<br />

ax\n4m2 — mxU^.1<br />

u%v\ ulv 2 _)/2<br />

J- •= Cos(6+-9_); u+ - Cos(0+) + Sin(6+); U_<br />

>- = Sin(8+-6_); v+ = Cos(0+) - Sin(0+); v_<br />

Weak<br />

Coupling<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

= Cos{BJ) + Sin{0J)<br />

= Cos{dJ) - Sin(9_)<br />

49


50<br />

X transitions between pure states to<br />

all AB transitions, while there are cross-<br />

peaks between the other X transitions<br />

to all AB transitions and also cross-<br />

peaks between all AB transitions to all<br />

X transitions. The selective inversion of<br />

X^ or X^ by a small angle a pulse<br />

does not cause any perturbation of the<br />

strongly coupled states and hence there<br />

are no cross-peaks from these transi-<br />

tions to all AB transitions. The 90°<br />

third pulse mixes the X-magnetisation<br />

unequally between all the X transitions<br />

giving rise to the auto-peaks. On the<br />

other hand, the selective inversion of<br />

an AB transition perturbs the strongly<br />

coupled states leading to cross-peaks<br />

to X transitions between mixed states.<br />

The non-linear detection pulse in turn<br />

mixes the intensities of all the X transi-<br />

tions giving rise to cross-peaks to even<br />

the X transitions between pure states.<br />

The spectrum is not symmetrical 15).<br />

(ii) 90° - 90° - a Experiment<br />

In this experiment there are no cross-<br />

peaks between all AB transitions to X<br />

transitions between pure states, while<br />

there are cross-peaks to all X transitions<br />

between the mixed states and also cross-<br />

peaks between all X transitions to all<br />

AB transitions. This is due to the fact<br />

Bulletin of Magnetic Resonance<br />

that the second 90° pulse perturbs un-<br />

equally all the transitions of AB as well<br />

as the X spin. The mixed states of the<br />

AB spins do not give directly any cross-<br />

peak to X-pure transitions. Since the<br />

detection pulse is a small angle pulse<br />

it does not mix the X transitions be-<br />

tween pure and mixed states and there-<br />

fore there are no cross-peaks from AB<br />

to X-pure transitions. The appearance<br />

of cross-peaks between X transitions be-<br />

tween pure states and the AB transi-<br />

tions is due to the mixing produced be-<br />

tween the various transitions of the X<br />

spin by the second 90° pulse. This state<br />

of the system is faithfully measured by<br />

the detection pulse. Here also the spec-<br />

trum is not symmetrical 15).<br />

The results of 90° - a - 90° and<br />

90° — 90° — a experiments are trans-<br />

pose of each other. This is due to the<br />

fact that PJVXMIOJX) = -PMXJVC 0 ^)-<br />

The conversion of populations into co-<br />

herences and vice versa are described by<br />

mirror operations(15).<br />

Experimental<br />

Experimental observation of these •<br />

cross-peaks was carried out in acetone<br />

oriented in liquid crystal ZLI 1167.<br />

Acetone oriented in liquid crystal is a<br />

strongly coupled spin system of the type<br />

(A3A3) with C3u C3v symmetry. The


Vol. 14, No. 1-4 51<br />

spectra is shown in Fig.3. From this<br />

spectrum it is clear that there are cross-<br />

peaks from every peak to all others<br />

within the same irreducible representa-<br />

tion. Theoretical simulations of these<br />

cross-peaks show a very good match<br />

with the experimental results(6), con-<br />

firming the existence of strong coupling<br />

induced cross-peaks in the 2D NOE ex-<br />

periments even in the absence of relax-<br />

ation.<br />

B. CROSS-CORRELATIONS IN<br />

2D NOE<br />

If a spin has more than one pathway<br />

for relaxation, then there can be cross-<br />

terms between these pathways that may<br />

contribute to the relaxation of the spin.<br />

For example, if there is another spin<br />

nearby, and the mutual dipolar interac-<br />

tion contributes to the relaxation of the<br />

spin and if in addition the first spin has<br />

a partial relaxation by CSA, there can<br />

be cross-terms between the dipolar re-<br />

laxation and CSA(16). If on the other<br />

hand there is a third spin contributing<br />

to the relaxation of the first two through<br />

dipolar relaxation then there can be<br />

cross-terms between various dipolar in-<br />

teractions and between the dipolar and<br />

CSA interactions contributing to the re-<br />

laxation of the various spins. These<br />

cross-terms known as cross-correlations<br />

are often neglected in the relaxation<br />

analysis such as those using generalized<br />

Solomons equations(17). It turns out<br />

that while the cross-terms may be sig-<br />

nificant in magnitude their manifesta-<br />

tion in a particular experiment may be<br />

small. For example the dominant ef-<br />

fect of the cross-terms is to make the re-<br />

laxation of various transitions of a spin<br />

unequal. In a given spin system or in<br />

an experiment if these transitions are<br />

not resolved then this dominant effect<br />

of cross-terms is absent. This can hap-<br />

pen for example when the spins are not<br />

J-coupled or if one uses a 90° pulse for<br />

measuring the intensities of the multi-<br />

plet. In the later case the non-linearity<br />

of the pulse yields an average intensity<br />

over all the transitions of a spin oblit-<br />

erating the multiplet effect and largely<br />

the cross-correlation effects. The use<br />

of a small flip angle for the measuring<br />

pulse is a necessary requirement for the<br />

observation of the multiplet effect and<br />

in turn the cross-correlation effects in<br />

the ID and the standard NOESY ex-<br />

periments^).<br />

In two-dimensional NOE experiment<br />

the most significant attempts to ob-<br />

serve the effect of cross-correlations<br />

have been made by Bodenhausen and<br />

his group(7,16,18-20). One of the ex-


52<br />

periments they have used is a small<br />

flip angle NOESY experiment namely<br />

90°—1\—a—rm—a—£2 , where a is small.<br />

Each cross-section of the small flip angle<br />

NOESY (NOESY 90° - a - a) is then<br />

equivalent to a ID difference transient<br />

NOE experiment in which the peak cor-<br />

responding to the diagonal peak is se-<br />

lectively inverted. This experiment has<br />

both the direct pumping effects and the<br />

cross-correlation induced multiplet ef-<br />

fects present which are measured by the<br />

small angle third pulse. For example the<br />

intensities of the X diagonal and the AX<br />

cross-peak multiplet in a weakly coupled<br />

three spin (AMX) system, in the initial<br />

rate approximation are given by(7),<br />

Xi<br />

x2<br />

x3<br />

x4<br />

Ax<br />

A2<br />

A3<br />

A4<br />

dn<br />

l[A<br />

KM<br />

_ hAM<br />

r rw<br />

r3 0)<br />

pf<br />

X2 X3 X4<br />

1\A<br />

'lM<br />

r2 x)<br />

(o)<br />

P3<br />

r4 0)<br />

«A<br />

r(0)<br />

P2 0)<br />

r3 1}<br />

P ( 4 1}<br />

^2AM<br />

L<br />

dTT<br />

_<br />

pf 1<br />

r(o)<br />

r 2<br />

P3 X)<br />

r 4<br />

(7)<br />

where X1; X2, X3, X4 are the four<br />

X transitions and Ax, A2, A3, A4 are<br />

the four A transitions. The expressions<br />

for the various intensities of the peaks<br />

are given in ref (7) except that when<br />

the cross-correlations due to CSA and<br />

Bulletin of Magnetic Resonance<br />

dipole-dipole interaction are included<br />

W11 ^ wu ? WU £ w^ and l\i ^ l[i<br />

where i = A, M or X. The r and p<br />

terms signify regressive and progressive<br />

peaks respectively. From eqn[7] it is<br />

seen that while the cross-correlation in-<br />

formation is contained in the small flip<br />

angle NOESY experiment, it is coupled<br />

with the direct pumping effects.<br />

We propose here simple modifications<br />

to the small flip angle NOESY (NOESY<br />

90° — a — a). If the second or the third<br />

pulse is made 90° then the intensities<br />

in the initial rate approximation are ob-<br />

tained as averages of the multiplets in<br />

either u>x or o>2 direction respectively.<br />

This removes the direct pumping effects<br />

from the 2D spectra. The following re-<br />

sults are obtained.<br />

90° - a - 90° NOESY<br />

The intensities of the various peaks in<br />

the initial rate approximation are given<br />

by<br />

x2<br />

x3<br />

x4<br />

Ax<br />

A2<br />

A3<br />

A4<br />

Xx X2 X3<br />

Rz<br />

2 ^2<br />

C3 Cz<br />

Ri<br />

R2<br />

Rz R2<br />

Rz<br />

Cz<br />

x4<br />

R2<br />

Rz<br />

Ci<br />

C2<br />

(8)


Vol. 14, No. 1-4<br />

where<br />

i2i =<br />

R2<br />

=<br />

Rs =<br />

R4 —<br />

Co =<br />

c4 =<br />

l0AM<br />

l\M<br />

XM +<br />

rx ^AX-<br />

Sx)]rm<br />

+ hAM + l\M +<br />

- (Px -<br />

A A<br />

(!) , (1) 1 (0) , (0)<br />

r x + Px + r x + Pi<br />

MX<br />

(9)<br />

(10)<br />

Here p^ is the rate of self relaxation of<br />

spin X, a AX 1S the cross-relaxation rate<br />

between spins A and X, 6X = SAXMXI<br />

which gives the cross-correlation rate<br />

between the dipolar vectors AX and<br />

MX and A^ gives the cross-correlation<br />

raten between the dipolar vector AX<br />

and the CSA of spin X. The expres-<br />

sions for the spectral density functions<br />

for the various relaxation rates (p, cr,<br />

A, 6) are given in ref(ll). The inten-<br />

sities of the various peaks in each mul-<br />

tiplet are identical in u>2 dimension and<br />

differ in o?j dimension, the differences<br />

directly yielding the cross-correlations.<br />

If the multiplet is resolved in the wi<br />

dimension the difference in the intensi-<br />

ties of the inner or the outer lines gives<br />

the dipole-CSA cross-correlations(A s)<br />

and the difference between the inner and<br />

outer lines gives the dipole-dipole cross-<br />

correlations (6 s). The diagonal multi-<br />

plet result is identical to the differences<br />

in the initial rates of recovery of the<br />

outer and inner multiplets in inversion-<br />

recovery Ti measurements (21). How-<br />

ever many analyses of inversion recov-<br />

ery measurements including (21) ignore<br />

CSA-dipole cross-correlations, while re-<br />

taining dipole-dipole cross-correlations.<br />

90° - 90° - a NOESY<br />

The intensities of the diagonal and<br />

the cross-peak multiplets in the initial<br />

rate approximation in this case are given<br />

by<br />

x2<br />

X,<br />

At<br />

where<br />

c2 =<br />

c3 =<br />

c: =<br />

X4<br />

R\ R2 Rs R4<br />

R\ R2 Rs R4<br />

R\ R2 R3 R4<br />

r> r> r? r><br />

ti\ Ii2 Kz it.4<br />

' ri' ri' p' ri'<br />

1 2 o 4<br />

ri' ri' rt' ri 1<br />

^1 ^2 U 3 W<br />

ri' r" r>' r>'<br />

O-i uo Lyo Ly A<br />

X Jr O 4<br />

ri' r" si' ri'<br />

Oi Wo L/o Ly^<br />

2{aAX + A; lx + 6A)Tm<br />

2(VAX - A; AX ~~ "A) T m<br />

2{


54<br />

a b cdefg h<br />

* •<br />

4 %<br />

* »<br />

500<br />

i j k<br />

#<br />

• *<br />

: • ' .<br />

I I<br />

pmm'l<br />

\> til,<br />

0<br />

CO2<br />

1e' d c b o<br />

,, d c b<br />

Ml./<br />

* f • »•- * •<br />

• « #<br />

•t: : :::.<br />

* •<br />

0 9 • • *.<br />

• • •<br />

• • • • • •<br />

Bulletin of Magnetic Resonance<br />

Figure 3. 2D NOESY spectrum of oriented acetone recorded at 400 MHz with<br />

rm = 20 /isec. The cross-peaks are mainly due to strong coupling. Zero- quantum<br />

interference during rm was shifted out in another experiment and the residual strong<br />

coupling peaks showed satisfactory correlation with the calculated intensities(6).<br />

(a)<br />

(b)<br />

-500<br />

_Jj\A<br />

Figure 4. Cross-sections taken from (a) 90° - 90° - 15° (b) 90° - 90° - 90° 2D<br />

NOESY experiment with rm = r0 + k using a 400 MHz spectrometer. r0 was 400<br />

msec and k was randomly varied between 10 and 1000 msec.<br />

-500<br />

500


Vol. 14, No. 1-4<br />

From these expressions it is seen that<br />

the intensities differ in u>2 and an aver-<br />

aging takes place along u>i . The sec-<br />

ond 90° pulse excites the multiplet as<br />

a whole, the correct state being moni-<br />

tored by the small angle a pulse. The<br />

differences in the intensities again yield<br />

the cross-correlations except that in this<br />

case the AX multiplet yields 8% and<br />

A^. The diagonal multiplet has in-<br />

tensities identical to 90° — a — 90° ex-<br />

periment. Since in the 90° — 90° — a<br />

experiment the intensities differ in UJ2<br />

domain which is easier to resolve, this<br />

experiment may be preferred over the<br />

90° - a - 90° experiment. In addi-<br />

tion, since all the lines of a multiplet<br />

along uj\ have equal intensities, a u)\ -<br />

decoupled (90° - (A + h)/2 - 180° -<br />

(A - *i)/2 - 90° -rm-a-t2) NOESY<br />

experiment can replace the undecoupled<br />

(90° - tx - 90° - rm - a - t2) NOESY<br />

experiment without loss of information.<br />

Experimental<br />

Two-dimensional NOESY experiment<br />

was carried out in 2,3-dibromo propi-<br />

onic acid using the 90° - 90° - a se-<br />

quence with small a(~ 15°) and a mix-<br />

ing time of 400msec plus a random vari-<br />

ation from 10 to 1000msec. Some of<br />

the cross-sections are shown in Fig.4.<br />

( The differences between the intensities<br />

of various transitions of a multiplet in-<br />

dicate the presence of cross-correlations.<br />

Lack of any particular symmetry in<br />

these multiplets indicates the presence<br />

of both dipole-dipole and CSA-dipole<br />

cross-correlations.<br />

CONCLUSIONS<br />

The use of 90° angle for the ex-<br />

citation or detection pulses allows an<br />

easier method for studying the cross-<br />

correlations in 2D NOE experiment.<br />

One can use either the second or the<br />

third pulse as small angle pulse to high-<br />

light the cross-correlation effects. In<br />

strongly coupled spins the non-linearity<br />

of the pulses can give rise to cross-peaks<br />

even in the absence of relaxation. The<br />

origin of these cross-peaks arising due<br />

to the non-linearity of the second or the<br />

third pulse are discussed with the help<br />

of an ABX spin system.<br />

Acknowledgement<br />

AK wishes to acknowledge discussion<br />

with Ms. Irene Burghardt of University<br />

of Lausanne regarding Fig.l.<br />

References<br />

*R. R. Ernst, G. Bodenhausen and<br />

A. Wokaun,'Principles of Nuclear Ma g-<br />

netic Resonance in One and Two Di-<br />

mensions', Clarendon, Oxford, 1987.<br />

55


56<br />

2 J. Keeler, D. Neuhaus and M. P.<br />

Williamson, J. Magn. Reson. 73, 45-<br />

68 (1987).<br />

3 I. Solomon, Phys. Rev. 99, 559-565<br />

(1955).<br />

4 I. Solomon and N. Bloembergen, J.<br />

Chem. Phys. 25, 261-266 (1956).<br />

5 S. Macura and R. R. Ernst, Mol.<br />

Phys. 41, 95-117 (1980).<br />

6 R. C. R. Grace and Anil Kumar, J.<br />

Magn. Reson. 97, 184-191 (1992).<br />

7 H. Oschkinat, D. Limat, L. Emsley<br />

and G. Bodenhausen, J. Magn. Reson.<br />

81, 13-42 (1989).<br />

8 J. Keeler and F. S. Fernando, J.<br />

Magn. Reson. 75, 96-109 (1987).<br />

9 T. E. Bull, J. Magn. Reson. 72,<br />

397-413 (1987).<br />

10 V. V. Krishnan and Anil Kumar, J.<br />

Magn. Reson. 92, 293-311 (1991).<br />

11 C. Dalvit and G. Bodenhausen,<br />

Adv. Magn. Reson. 14, 1-32 (1990).<br />

12 L. G. Werbelow and D. M. Grant,<br />

Adv. Magn. Reson. 9, 189-299 (1977).<br />

13 D. Canet, Prog. NMR. Spectrosc.<br />

21, 237-291 (1989).<br />

14 J. A. Pople, W. G. Schneider<br />

and H. J. Bernstein, 'High Resolu-<br />

tion Nuclear Magnetic Resonance Spec-<br />

troscopy', McGraw-Hill, New York,<br />

1959.<br />

15 R. C. R. Grace and Anil Kumar,<br />

(unpublished results)<br />

Bulletin of Magnetic Resonance<br />

16 I. Burghardt, R. Konrat and G. Bo-<br />

denhausen, Mol. Phys. 75, 467-486<br />

(1992).<br />

17 J. H. Noggle and R. E. Schirmer,<br />

'The Nuclear Overhauser Ef-<br />

fect : Chemical applications', Academic<br />

Press, New York, 1971.<br />

18 H. Oschkinat, A. Pas-tore and G.<br />

Bodenhausen, J. Am. Chem. Soc. 109,<br />

4110-4111 (1987).<br />

19 C. Dalvit and G. Bodenhausen, J.<br />

Am. Chem. Soc. 110, 7924-7926<br />

(1988).<br />

20 J. M. Bohlen, S. Wimperis and G.<br />

Bodenhausen, J. Magn. Reson. 77,<br />

599-605 (1988).<br />

21 E. Ilyina and Daragan (in press).


Vol. 14, No. 1-4 57<br />

Deriving Structures from 2D NMR. A Method for Defining the<br />

Conformation of a Protein Adsorbed to Surfaces<br />

With the development of multi-dimensional<br />

NMR methods for the specific assignment of many of<br />

the *H NMR signals of small proteins, it is now possible<br />

to determine three-dimensional solution structures<br />

and dynamics for these proteins. Unfortunately<br />

these methods fail to provide detailed structural<br />

information on proteins bound to macroscopic<br />

surfaces because of the very slow overall rotational<br />

correlation time of the particle-bound protein. Indeed<br />

other spectroscopic and structural methods<br />

have provided few details of the structure of proteins<br />

when adsorbed to surfaces (1). Commercially,<br />

surface-bound, immobilized enzymes provide an important<br />

method for efficient utilization of these catalysts<br />

in bioreactors. The development of biosensors<br />

and novel biomaterials such as self-assembled monolayers<br />

(2) requires a better understanding of the<br />

structure of proteins bound to surfaces (1). Identification<br />

of specific residues of proteins involved in<br />

interactions with stationary phase surfaces is critical<br />

in understanding their chromatographic behavior<br />

(3). Finally, the association of proteins with<br />

hydrophobic or glass surfaces has been shown to<br />

cause denaturation or partial unfolding of the pro-<br />

. tein at the surface, with often irreversible loss of<br />

the strongly adsorbed protein. It is believed that<br />

;.these irreversibly bound proteins undergo a confor-<br />

|mational change to expose a portion of the interior<br />

sidues to the surface for effective adsorption.<br />

J this report we describe an NMR methodology<br />

[it for the first time allows us to probe the detailed<br />

^formation of a protein bound to a macroscopic<br />

This was accomplished by 2D amide hy-<br />

«i exchange NMR spectroscopy (4, 5, 6). In a<br />

?i, the NH exchange rates for different residues<br />

over a factor of greater than 10 8 (4, 5).<br />

fate of amide hydrogen/deuterium exchange<br />

"I on the pH (being both acid and base cat-<br />

! the accessibility of the hydrogen to solvent,<br />

Mlization by hydrogen-bonded secondary as<br />

tertiary structure and finally, local fluctu-<br />

•jf the protein (7). Importantly amide hy-<br />

David A. Keire and David G. Gorenstein*<br />

Department of Chemistry<br />

Purdue University<br />

West Lafayette, Indiana 47907<br />

drogen exchange can be studied under conditions<br />

where high resolution resolvable proton NMR signals<br />

would otherwise not be observed - this would<br />

be true for a protein immobilized at a polymeric or<br />

glass surface. However, analysis of the degree of NH<br />

exchange is feasible if the protein can be desorbed<br />

from the surface and the 2D NMR study of the free<br />

protein carried out in solution.<br />

As demonstration of the feasibility of the<br />

method, we have used amide hydrogen exchange<br />

NMR spectroscopy of lysozyme bound to a hydrophobic<br />

chromatographic stationary phase support.<br />

Hen egg white lysozyme (E.C.3.2.1.17) was<br />

chosen for this initial study because the ^-NMR<br />

assignments have been made (8), the crystal structure<br />

was known to 2 A resolution (9) and much was<br />

known about its adsorption onto hydrophobic substrates<br />

(10). Lysozyme is one of the major constituents<br />

of protein deposits on contact lenses (11)<br />

and is commonly used as a standard in the chromatographic<br />

separation of proteins on reverse-phase<br />

columns.<br />

In our protocol we first adsorb lysozyme to the<br />

hydrophobic surface (solid 5 \i diameter polystyrene<br />

divinylbenzene chromatographic stationary phase<br />

support; Polymer Labs Inc., Amherst, Mass.). This<br />

forms a tightly bound single monolayer on the<br />

surface (10). We then expose the surface bound<br />

lysozyme to D2O under fast amide hydrogen exchange<br />

conditions (high pH), desorb the protein under<br />

slow NH exchange conditions (low pH) using<br />

a detergent, and after removal of the detergent, run<br />

the high resolution 2D spectra in D2O solution. The<br />

intensities of the 2D NHCHa TOCSY crosspeaks<br />

then reflect the degree of exchange of the specific<br />

residue amide hydrogens when adsorbed to the hydrophobic<br />

surface.<br />

The stock solution of lysozyme (Sigma) was purified<br />

using a home built hollow fiber bundle dialysis<br />

device versus 10 mM NaH2PC>4 pH = 7.4 buffer. An<br />

aliquot of the stock solution containing 125 mg of


58<br />

lysozyme was diluted to 10 mL in buffer and mixed<br />

with 10 g of hydrated stationary phase support for<br />

30 min. with stirring. The solution was then removed<br />

by the use of a filter funnel with a medium<br />

grain glass frit.<br />

Eighty milliliters of buffer were used to wash the<br />

stationary phase support in 10 to 20 mL aliquots.<br />

For each aliquot a sample of the filtrate was taken<br />

for UV determination of protein concentration (e280<br />

= 2.313 O.D. mL/mg) to ensure that only an "irreversibly"<br />

bound monolayer coverage remained.<br />

Typically, out of 125 mg of lysozyme initially added<br />

~25 mg would remain on the support after extensive<br />

washing. The stationary phase support has a<br />

surface area of ~3 A/gm and a rough calculation<br />

using the molecular dimensions of lysozyme (4.5 x 3<br />

x 3 nra 3 ) shows that ~50 mg of protein could adsorb<br />

to 10 gm of support with monolayer coverage. The<br />

25 mg of lysozyme adsorbed represents 50% coverage<br />

which is in good agreement with the 65% coverage<br />

by the irreversibly bound layer determined by<br />

Schmidt et al. (10) to a similar hydrophobic surface.<br />

At this point a 10 mM NaH^PCU D2O solution<br />

pH* = 7.0 (uncorrected pH meter reading) was prepared<br />

and mixed with the lysozyme covered support.<br />

The adsorbed protein was then stirred for 45<br />

min. during which time the amide hydrogens that<br />

are exposed to solvent can exchange with deuterons.<br />

After 45 min. the solution was aspirated away<br />

and another 10 mL of the buffered D2O solution at<br />

pH*=2.5 was added to quench the amide exchange.<br />

Two aliquots of 0.1% Triton X-114-RS (Sigma)<br />

detergent in 10 mL D2O, pH* = 2.5 were used to<br />

remove the tightly adsorbed protein. The reduced<br />

form of Triton was used to allow determination of<br />

desorbed protein concentration by UV. The ~30<br />

mL of desorbed lysozyme/Triton solution was then<br />

concentrated by pressure dialysis (Amicon, Beverly,<br />

Mass.) to ~4 mL using a 5000 molecular weight cutoff<br />

membrane. The 4 mL was then passed through<br />

an Extracti-Gel column (Pierce) equilibrated with<br />

pH* = 2.5 D2O buffer to remove the Triton. The<br />

column eluent was further concentrated to ~750 fib<br />

and placed in an NMR tube. Approximately 5 mg<br />

(0.5 mM) of desorbed lysozyme was obtained by this<br />

procedure for the NMR study. Much of the protein<br />

loss occurred during the repurification scheme since<br />

the adsorbed protein was nearly quantitatively removed<br />

from the surface by the detergent treatment.<br />

Bulletin of Magnetic Resonance<br />

In the control experiment the identical procedure<br />

was used except that only 5 mg of lysozyme was<br />

used and no stationary phase support was present.<br />

An additional control was run without stationary<br />

phase support and without Triton to ensure that<br />

any residual Triton not detectable in the ID NMR<br />

spectrum did not interfer with the amide exchange<br />

results. This control gave essentially the same integrated<br />

area for the TOCSY cross peaks of the NH-<br />

CaH region as the control involving Triton.<br />

The 600 MHz X H-NMR NH-CaH region of the<br />

30 ms TOCSY spectra of the control and surface<br />

adsorbed/desorbed lysozyme is shown in Figure 1.<br />

Assignment of the signals was based upon complete<br />

analysis of the TOCSY spectra, taking advantage<br />

of the reported 1 H assignments of lysozyme at 500<br />

MHz (8). Normally, a series of TOCSY spectra at<br />

various time points are taken to measure the rate<br />

of NH exchange. Since this has been shown to be<br />

first order, comparing the spectra at a fixed time<br />

of exchange under conditions where the protein is<br />

either surface-bound or not will also provide a measure<br />

of the relative degree of NH protection. We<br />

define protons as being "exposed" when their crosspeaks<br />

present in the control spectrum are decreased<br />

in intensity by >40% relative to the surface exposed<br />

protein spectrum. Protons are "protected" if the<br />

crosspeaks are either absent in the control spectra or<br />

increase in intensity by >40% in the surface exposed<br />

sample spectra. Three control experiments (without<br />

the support present) were used to calculate the percent<br />

deviation from the mean of the integrated areas<br />

of 39 of the slow exchanging amide proton NH-CaH<br />

TOCSY cross peaks. The mean of the 39 deviations<br />

from the means of the integrated areas was 20±13%.<br />

Thus, only those changes in integrated area >40%;<br />

were considered exposed or protected after averag|<br />

ing the cross peak areas from two separate adsorbaj<br />

protein and control experiments. These results ar|<br />

tabulated for the slow exchanging amides [as definw<br />

by Redfield & Dobson (8) at pH = 3.8, 35°C, n/2.<br />

the half life for exchange > 1-5 h] in Table I. Also<br />

included in Table I are the integrated intensities oj<br />

2 intermediate exchange (1 h < r1/2 < 5 h) and on«<br />

fast exchange amide hydrogens (T1/2 < 1-5 h) whicg<br />

are protected relative to the control.<br />

All of the amides which are protected (Table.*<br />

with the exception of the 2 arginines are neutr<br />

or hydrophobic residues. A CPK model {Figuiej


Vol. 14, No. 1-4 59<br />

B<br />

Fl (ppm)<br />

5.5 3.5 3.0 2.5<br />

Fl (ppm)<br />

1: Identical 600 MHz ^-NMR TOCSY experiments at 35°C and pH*=2.5 were run on both the<br />

>1 and adsorbed lysozyme samples. The data was collected in phase sensitive mode with a mixing time<br />

using a MLEV17 spin lock and 2 ms trim pulses. 2K points were collected in F2 and 312 in Fl<br />

^sweep width of 7500 Hz and a repetition time of 3 s. 32 transients were coadded at each ti increment.<br />

£JD signal was suppressed through low power decoupling at the HOD frequency. The spectra were<br />

with zero filling to 2K points in the Fl dimension and a 45° phase shifted sine bell apodization in<br />

tensions.


60 Bulletin of Magnetic Resonance<br />

based upon the crystal structure of lysozyme (9)]<br />

shows that all of the protected amide residues (labeled<br />

by shading) are found on the surface. Four<br />

of the protected amide residues (T40, Q41, A82<br />

and L84) are located close in space at the hinge region<br />

between the a-helical and /3-sheet domains of<br />

lysozyme opposite the active site cleft.<br />

All of the exposed amides - those that show enhanced<br />

amide exchange upon binding to the surface<br />

- with the exception of W63 are involved in ahelices<br />

or /3-sheet structures and are mostly buried<br />

in the interior of lysozyme (Table I). Of the exposed<br />

amides only D52 and W63 show significantly exposed<br />

side chains as shown in the CPK model (Figure<br />

2; exposed residues shown by stripes).<br />

The protein amide hydrogens that are observable<br />

by 2D amide exchange spectroscopy are mostly the<br />

slow exchanging amides (58/129) which are either<br />

involved in secondary or tertiary structure hydrogen<br />

bonding or buried in the interior of the protein<br />

and inaccessible to solvent (12). Surprisingly, one<br />

fast exchanging amide hydrogen (III 14), and two intermediately<br />

exchanging hydrogens (T40 and Q41)<br />

CaH-NH cross peaks were observable in the spectrum<br />

of lysozyme exposed to the hydrophobic surface<br />

which were not present in the control spectrum<br />

(Figure 1).<br />

Figure 2A shows that the protected (stippled)<br />

amide hydrogen side chains lie on one face of the<br />

globular protein in a narrow ridge from R125 to<br />

L17. In contrast, in the back-side view (Figure 2B),<br />

only a few of the largely buried side chains of the<br />

exposed residues (striped) are visible. None of the<br />

protected residues are visible in this view. These results<br />

suggest that lysozyme is not randomly oriented<br />

with respect to the surface, but that it is oriented<br />

with a relatively hydrophobic ridge facing towards<br />

the hydrophobic surface. This is consistent with the<br />

observed retardation of amide hydrogen exchange in<br />

binding an amphiphilic helix to a micelle (13), believed<br />

to result from burial of the hydrophobic face<br />

of an amphiphilic helix into the hydrophobic interior<br />

of a detergent micelle.<br />

In addition a number of residues exchange more<br />

rapidly when the protein is surface bound than when<br />

it is in solution. All of these are not exposed to<br />

the solvent in the native structure but are located<br />

on elements of structure that are likely involved in<br />

segmental motion of the two domains that form the<br />

active site cleft (14). Binding of the hydrophobic<br />

ridge of the protein to the surface thus appears to<br />

induce a conformational change, exposing the active<br />

site residues and residues adjacent to the active site<br />

to solvent.<br />

These perturbations in the amide exchange rates<br />

do not simply reflect proximity of the protein to<br />

the surface because enhancement and protection towards<br />

exchange are observed. In addition most<br />

of the amide hydrogen exchange rates which are<br />

not exposed or protected are the same for surface<br />

bound or free lysozyme. In contrast only decreases<br />

in amide exchange rates are observed in binding an<br />

amphiphilic helix to a micelle (13) and in crystalline<br />

lysozyme (15).<br />

Taken together these data indicate that<br />

lysozyme adsorbs to the surface on the side opposite<br />

the active site cleft. This protects this narrow ridge<br />

of amides from exchange by either blocking solvent<br />

access to these residues at the hydrophobic surface<br />

or by reducing the rate of local fluctuations in the<br />

surface oriented residues. In addition upon binding<br />

to the hydrophobic surface a significant disruption<br />

of several of the buried a-helices and the active site<br />

cleft occurs as the protein partially unfolds at the<br />

surface presumably by opening of the "hinge" at<br />

the active site (16), exposing a number of buried<br />

residues that now show enhanced rate of hydrogen<br />

amide exchange. This conformational change alters<br />

the amide solvent exposure and/or local fluctuations<br />

that allow access of solvent to these interior residues.<br />

This partial unfolding exposes the catalytically<br />

important D52 residue and other important residues<br />

in the active site cleft. This model is supported by<br />

the fact that the enzyme is inactive on adsorption<br />

to the hydrophobic surface of alkylated silica (10).<br />

These conclusions are also in agreement with the<br />

results of Fausnaugh and Regnier (3) based upon<br />

analysis of chromatographic behavior of various bird<br />

lysozymes that also suggests the protein adsorbs on<br />

a side opposite the active site with a contact surface j<br />

that extends from residues 41 to 102 and 75 to 89.1<br />

Computer modeling has also revealed a relatively^<br />

hydrophobic patch identified as a possible binding<br />

site in this region and total internal reflection i j<br />

sic fluorescence (TIRIF) shows a decreased quantun^<br />

yield (and hence altered conformation) of the<br />

sorbed hen lysozyme (17).<br />

Importantly, the surface desorbed lysozj


Vol. 14, No. 1-4<br />

A<br />

B<br />

wire 2: Three dimensional structure of hen lysozyme. A) Front-side view of a CPK model [MIDAS modeling<br />

", UCSF (21)] showing amide NH's which are relatively protected (shaded and stippled) or exposed<br />

and striped) when the protein is bound to the polystyrene surface. The active site cleft is oriented<br />

ne upper left. B) No residues that are protected from exchange (shaded and stippled) can be seen<br />

^.ackside view of the protein. Several of the side chains of residues that are more exposed (shaded and<br />

,


62<br />

Lysozyme CH-NH TOCSY Cross Peaks<br />

A9<br />

All<br />

L17<br />

W28<br />

C30<br />

A31<br />

N37<br />

T40<br />

Q41<br />

D52<br />

W63<br />

A82<br />

L84<br />

C94<br />

198<br />

R114<br />

W123<br />

R125<br />

1.19<br />

4.26<br />

4.53<br />

1.06<br />

0.68<br />

0.27<br />

4.71<br />

4.36<br />

0.99<br />

1.15<br />

2.55<br />

2.40<br />

2.73<br />

0.59<br />

0.32<br />

1.26<br />

0.71<br />

1.30<br />

2.84<br />

2.47<br />

1.03<br />

1.49<br />

1.28<br />

0.24<br />

0.87<br />

1.91<br />

1.45<br />

Slow Exposed<br />

Slow Exposed<br />

Bulletin of Magnetic Resonance<br />

Slow Protected<br />

Slow Exposed<br />

Slow Exposed<br />

Slow Exposed<br />

Slow Protected<br />

Inter. Protected<br />

Inter. Protected<br />

Slow Exposed<br />

Slow Exposed<br />

Slow Protected<br />

Slow Protected<br />

Slow Protected<br />

Slow Exposed<br />

Fast Protected<br />

Slow Protected<br />

Slow Protected<br />

a.) Blanks indicate no observable cross peak.<br />

b.) The integrated area of the adsorbed protein CH-NH TOCSY cross peaks were normalized via the integraM<br />

area of the non-exchangeable Trp 108 H4-H5 cross peak in the control and adsorbed protein spectra. Wt<br />

integrated areas in the table are the average of two control and two adsorbed protein experiments. Jjj<br />

c.) Slow, intermediate and fast exchange amide protons are classified as per Redfield and Dobson (4). |ip<br />

d.) Exposed amides are present in the control and diminished by >40% in integrated area or absent inj|<br />

adsorbed protein spectra. Protected amides are present in the adsorbed protein spectra and diminish^<br />

integrated area by >40% or absent in the control spectra. |i


Vol. 14, No. 1-4 63<br />

spectra have identical chemical shifts as the native<br />

form spectra. This suggests that while the protein<br />

partially unfolds at the surface, it refolds by<br />

the same kinetically or thermodynamically favorable<br />

pathway upon desorption. Thus, the surface<br />

unfolded state may represent a protein folding intermediate<br />

of native hen lysozyme. Only a transient<br />

folding intermediate has been previously identified<br />

in lysozyme (12). By binding the protein to a surface<br />

it may be possible to trap a partially unfolded<br />

state of lysozyme that bears some resemblance to<br />

the transient folding intermediate observed by Miranker<br />

et al. (12). This would be complementary to<br />

the rapid quench NMR amide hydrogen exchange<br />

spectroscopy methods (12, 18, 19, 20) which provides<br />

structural details on the initial refolding kinetic<br />

intermediates. Our method would presumably<br />

provide information on the initial unfolding intermediate.<br />

Surface 2D amide exchange spectroscopy offers<br />

a new method by which protein adsorption can be<br />

monitored at the level of individual residues. This<br />

work demonstrates for the first time the feasibility<br />

of the method which should be applicable to a number<br />

of other enzymes. The method offers a more detailed<br />

picture of protein adsorption than provided by<br />

currently used techniques (e.g. TIRF, ATR-FTIR<br />

and Raman spectroscopy (1)). Future work will<br />

include the extension of the methodology to other<br />

surfaces and proteins. Indeed at the XVth International<br />

Conference on Magnetic Resonance in Bi-<br />

. ological Systems (Jerusalem, Israel, August 16-21,<br />

,., 1992, abstracts) K. Kawano et al., described simii<br />

-lar application of the NH exchange experiment to<br />

\j the binding of lysozyme to hydroxyapatite. They<br />

'also find that certain residues are protected from<br />

nange when the protein is adsorbed to the sur-<br />

However unlike our results no enhancement of<br />

•exchange is observed.<br />

!' Acknowledgments<br />

rSupported by the Office of Naval Research<br />

jp00l4-91-J-1686) and the Purdue University Bio-<br />

4cal Magnetic Resonance Laboratory which is<br />

ported by the NSF Biological Facilities Center on<br />

Secular NMR, Structure and Design at Pur-<br />

[grants BBS 8614177 and DIR-9000360 from<br />

jjyision of Biological Instrumentation) and NIH<br />

•ferences<br />

1. J. D. Andrade and V. Hlady, Adv. Polym.<br />

Sci. 79, 1-63 (1986).<br />

2. K. L. Prime and G. M. Whitesides, Science<br />

252, 1164-1166 (1991).<br />

3. J. L. Fausnaugh and F. E. Regnier, J. Chromatogr.<br />

359, 131- 146 (1986).<br />

4. H. Roder, "Methods in Enzymology" (N. J.<br />

Oppenheimer and T. L. James, eds. ), 446-473, Academic<br />

Press, Inc., 1989.<br />

5. S. W. Englander and N. R. Kallenbach, Q.<br />

Rev. Biophys. 19, 521-655 (1984).<br />

6. H. Roder, G. Wagner, and Wuthrich, Biochemistry<br />

24, 7396-7407 (1985).<br />

7. Y. Paterson, S. W. Englander, and H. Roder,<br />

Science 249, 755-759 (1990).<br />

8. C. Redfield and C. M. Dobson, Biochemistry<br />

21, 122-136 (1988).<br />

9. C. C. F. Blake, G. A. Koenig, A. C. Mair, C.<br />

T. North, D. C. Philips, and V. R. Sarma, Nature<br />

206, 757-761 (1965).<br />

10. C. F. Schmidt, R. M. Zimmermann, and H.<br />

E. Gaub, Biophys. J. 57, 577-588 (1990).<br />

11. E. J. Castillo, J. L. Koenig, and J. M. Anderson,<br />

Biomaterials 6, 338-344 (1985).<br />

12. A. Miranker, S. E. Radford, M. Karplus, and<br />

C. M. Dobson, Nature 349, 633-636 (1991).<br />

13. C. Karslake, M. E. Piotto, Y. M. Pak, H.<br />

Weiner, and D. G. Gorenstein, Biochemistry 29,<br />

9872-9878 (1990).<br />

14. C. C. F. Blake, G. A. Mair, A. C. T. North,<br />

D. C. Phillips, and V. R. Sarma, Proc. R. Soc.<br />

Lond. Ser. B 167, 365-377 (1967).<br />

15. T. G. Pedersen, B. W. Sigurskjold, K. V.<br />

Andersen, M. Kjaer, F. M. Poulsen, C. M. Dobson,<br />

and C. Redfield, J. Mol. Biol. 218, -413-426<br />

(1991).<br />

16. J. A. McCammon, B. R. Gelin, M. Karplus,<br />

and P. G. Wolynes, Nature 262, 325-326 (1976).<br />

17. D. Horsley, J. Herron, V. Hlady, and J. D.<br />

Andrade, "Proteins at Interfaces", 290-305, 1987.<br />

18. H. Roder, G. A. Elove, and S. W. Englander,<br />

Nature 335, 700 (1988).<br />

19. J. B. Udgaonkar and R. L. Baldwin, Nature<br />

335, 694-699 (1988).<br />

20. C. M. Dobson and P. A. Evans, Nature 335,<br />

666 (1988).<br />

21. T. E. Ferrin, C. C. Huang, L. C. Jarvis, and<br />

R. Langridge, J. Mol. Graphics 6, 13-27 (1988).


64<br />

Flavoridin, a protein with 70<br />

amino acids from the venom of<br />

Trimeresurus gramineus, is a very<br />

potent inhibitor of blood platelet<br />

aggregation. The protein contains<br />

a local Arg-Gly-Asp (RGD)<br />

sequence at position 49 to 51. This<br />

primary sequence element is<br />

known to inhibit fibrinogen<br />

binding by a specific interaction<br />

with the integrin-type platelet<br />

receptor GPIIb/IIIa. By now, a<br />

rather large family of homologous<br />

RGD containing snake toxins have<br />

been sequenced. Most of these<br />

proteins contain n=12 cysteines<br />

which are all linked by disulfide<br />

bridges. Previous biochemical<br />

studies, however, have so far not<br />

revealed the native pattern of the<br />

individual cysteine pairings.<br />

^ l , Werner Klaus and Pau, Gerber<br />

The ! H NMR spectrum of<br />

Flavoridin was almost fully<br />

assigned in aqueous solution by<br />

conventional 2D ^H NMR methods<br />

(2QF-COSY, clean TOCSY, 2Qspectroscopy,<br />

NOESY). The 3Dstructure<br />

calculation with<br />

distance geometry methods<br />

(DIANA) proceeded in several<br />

rounds:<br />

(1) The global fold of Flavoridin<br />

was calculated from the collected<br />

set of NOE distance constraints<br />

(#666) and dihedral angle<br />

constraints obtained from vicinal<br />

coupling constants (#88) but<br />

without the use of any disulfide<br />

bridge constraints.<br />

(2) The interatomic cP-C-P distances<br />

between all possible pairs [n(nl)/2]<br />

of cysteines were measured<br />

in a set of 20 converged distance<br />

geometry structures. A<br />

probability weight wjj (0 - wij - 1)<br />

i« assigned to each individual<br />

Bulletin of Magnetic Resonance<br />

cysteine pair according to a<br />

gaussian shaped distribution<br />

function. When the average<br />

crystallographic CP-CP distance of<br />

a cystine disulfide bridge is<br />

matched, the weight wy = 1.<br />

(3) All combinatorial patterns of 6<br />

disulfide-bridges involving the 12<br />

cysteines in Flavoridin were<br />

calculated and the 6 individual<br />

weights, w^j, for each pattern<br />

were multiplied. Only patterns<br />

with individual wjj — 0.3 were<br />

considered. On this objective basis,<br />

a single Cys-Cys pairing pattern<br />

could unambiguously be<br />

determined. The method was<br />

validated with known protein<br />

crystal structures of Cys-rich<br />

proteins. Also the experimentally<br />

observed NOE's between Ci«H/CjPH<br />

and/or CiP/CjP of Cys residues,<br />

which commonly are taken as<br />

direct evidence for Cys-Cys<br />

disulfide links, do agree' with the<br />

computationally evaluated Cys-Cys<br />

pattern.<br />

(4) In a second step of the<br />

structure calculation, the Cys-Cys<br />

pairing was used as additional<br />

input for the distance geometry<br />

program. Finally, the 50 best<br />

structures were refined by<br />

energy minimization.<br />

Our structural results show that<br />

the polypeptide backbone is folded<br />

in two domain-like structures -<br />

composed of 8 turns and stabilised<br />

by 6 cystine cross-links. The<br />

conformation of the Arg-Gly-Asp<br />

(RGD) sequence is located in an<br />

extended loop structure exposed at<br />

the tip of a so called hairpin,<br />

which is rather flexible.


Vol. 14, No. 1-4 65<br />

1. Introduction<br />

NMR Approaches to Large Proteins:<br />

trp Repressor and Chloramphenicol Acetyltransferase<br />

L.-Y. Lian, J. P. Derrick, V. Ramesh, R. O. Frederick,<br />

The current upper limit for protein structure<br />

determination by *H nmr [1] is in the region<br />

of Mr 12-15,000. The use of stable isotope<br />

labelling, with 2 H, 13 C or 15 N, can<br />

substantially extend this limit, perhaps to Mr<br />

30,000 (for methodological reviews, see [2]-<br />

[4]). However, there remain many interesting<br />

and important proteins which are much larger<br />

than this, and we have been trying to<br />

establish what information nmr can provide in<br />

such cases. Two systems we have been<br />

studying in this context are the E. coli trp<br />

repressor and the enzyme chloramphenicol<br />

acetyltransferase.<br />

The trp repressor is a dimer of total Mr<br />

•; 25,000, and is a member of the "helix-turnpielix"<br />

family of DNA binding proteins. A<br />

umber of high resolution crystal structures<br />

the protein are available, but the<br />

hanisms of its activation and DNAinding<br />

specificity remain poorly understood,<br />

"lain priority is therefore to determine<br />

solution structure of the proteinion<br />

ucleotide complex, and the effects on<br />

of changes in protein and oligonucleotide<br />

S ure; our work to date has lar s el y been<br />

* to developing the necessary methods<br />

S. E. H. Syed and G. C. K. Roberts<br />

Biological NMR Centre, University of Leicester,<br />

PO Box 13 8, Medical Sciences Bldg.,<br />

University Road, Leicester LEI 9HN, U.K.<br />

for this. We have, in collaboration with the<br />

group of Jardetzky at Stanford, achieved a<br />

virtually complete sequence-specific assignment<br />

of the *H nmr spectrum of the protein<br />

[5]-[7], and Jardetzky's group have used this<br />

data to determine the solution structure of<br />

the repressor-tryptophan complex [8]. Nmr<br />

studies of the binding of corepressors such as<br />

L-tryptophan and the inducer, indole-3propionic<br />

acid, show that the environments of<br />

the two classes of ligand in the protein differ,<br />

and strongly suggest that this arises from a<br />

difference in the orientation of their indole<br />

rings [9]. Even the repressor alone is of such<br />

a size that these studies required the<br />

extensive use of isotope labelling, and we<br />

have developed appropriate expression<br />

systems for efficient biosynthetic labelling of<br />

the protein. The overall molecular mass of<br />

the repressor-tryptophan-operator oligonucleotide<br />

complex is ca. 38,000, and for this<br />

complex rather little useful information can<br />

be obtained by nmr without labelling.<br />

Chloramphenicol acetyltransferase is<br />

responsible for resistance to the antibiotic<br />

chloramphenicol in bacteria. Its crystal<br />

structure is known to high resolution, and it<br />

is the subject of a substantial programme of<br />

protein engineering in the laboratory of Prof.


66 Bulletin of Magnetic Resonance<br />

W.V. Shaw in Leicester [10]. Since it is a<br />

trimer of total molecular mass 75,000, it is<br />

very large for study by nmr, and we are using<br />

it as a test system for the development of nmr<br />

methods suitable for large proteins, focussing<br />

particularly on complexes, such as that with<br />

the product, diacetyl-chloramphenicol, which<br />

have not proved amenable to crystallographic<br />

study.<br />

2. Deuteration<br />

Selective deuteration has been used to<br />

simplify the *H nmr spectra of proteins for<br />

many years [11], and more recently has been<br />

combined with 2D nmr spectroscopy (e.g.,<br />

[12]). Selective deuteration of carefully<br />

chosen combinations of residues has been<br />

combined with the sequential assignment<br />

method of Wiithrich (see [1]) to yield a<br />

considerable number of resonance<br />

assignments in the trp repressor, both alone<br />

and in its complex with an operator<br />

oligonucleotide [6], [7]. With the improved<br />

resolution of the 3D spectra, we have<br />

recently assigned all the intermolecular NOEs<br />

between protons of the protein and those of<br />

bound tryptophan, permitting the kind of<br />

'ligand-docking' experiments which we have<br />

already carried out on phospholipase A2 [13].<br />

In a species as large as the repressor-operator<br />

complex, the substitution of most of the<br />

protons in the protein by deuterons leads to a<br />

notable decrease in the resonance linewidth<br />

of the remaning protons; similarly, we have<br />

employed random fractional deuteration [14]<br />

to good effect in these large systems. In<br />

particular, deuteration of chloramphenicol<br />

acetyl-transferase to the level of about 85%<br />

had three beneficial effects: (a) it allowed the<br />

unambiguous observation and assignment of<br />

resonances of the bound substrate, (b) it<br />

allowed the observation of intra-molecular<br />

NOEs in the bound substrate, thus defining<br />

its conformation, and (c) it allowed the<br />

observation of inter-molecular NOEs<br />

between protons of the (isotopically normal)<br />

substrate and nearby residues of the protein<br />

[15].<br />

3. 13 C and 15 N labelling<br />

Notwithstanding the usefulness of selective<br />

deuteration, complete assignments of the<br />

backbone resonances of the repressor have<br />

required the use of 15 N-labelled protein, in<br />

combination with 3D nmr spectroscopy [7].<br />

The same approach has led to a substantial<br />

number of resonance assignments in the<br />

repressor-operator complex, and to the clear<br />

observation of inter-molecular NOEs [7],<br />

[16]. In addition to the standard NOESY-<br />

HMQC and TOCSY-HMQC experiments,<br />

we have found the HMQC-NOESY-HMQC<br />

experiment [17] valuable in this a-helical<br />

protein in which there is significant overlap of<br />

both *H and 15 N chemical shifts of the<br />

backbone amides. In some cases, there is a<br />

considerable benefit in selective labelling; in<br />

the trp repressor we have used this<br />

successfully for 15 N-leucine and for [amide-<br />

15 N]-asparagine and glutamine [16], and in<br />

chloramphenicol acetyltransferase for<br />

[imidazole 2- 13 C]-histidine [18]. The latter<br />

experiment permitted the detection of the<br />

histidine C2- 1 H signals, and the use of 2D<br />

*H- 13 C correlation spectra allowed<br />

overlapping signals to be resolved, even in a<br />

protein of Mr 75,000. We are also exploring<br />

the usefulness of 50-65% perdeuteration in<br />

combination with 15 N-labelling as a means of<br />

improving the quality of the 3D spectra.<br />

In chloramphenicol acetyltransferase we<br />

have used nmr to study the binding of<br />

diacetyl[ 13 C]-chloramphenicol by means of<br />

13 C-edited l U- l B. NOESY experiments. A<br />

number of clear intermolecular NOEs were<br />

observed, in particular to aromatic protons of<br />

the protein. Candidate aromatic residues<br />

were identified by model-building on the<br />

basis of the crystal structure, and were<br />

replaced in turn by isoleucine residues. 13 Credited<br />

NOESY spectra of the complexes with<br />

these mutants allowed the two aromatic<br />

residues with which the acetyl groups of the<br />

ligand made contact to be identified<br />

unambiguously, thus allowing the orientation<br />

of the product in the binding site to be


Vol. 14, No. 1-4 67<br />

defined [15], and assisting in the modelling of<br />

the transition-state complex.<br />

4. References<br />

[1]<br />

[2]<br />

[3]<br />

[4]<br />

[5]<br />

[6]<br />

[7]<br />

[8]<br />

[9]<br />

[10]<br />

Wiithrich, K., Science 243 (1989) 45.<br />

Bax, A. (1989) Ann. Rev. Biochem. 58,<br />

223.<br />

Oppenheimer, N.J., and James, T.L.<br />

(1989) Methods in Enzymology, 176,<br />

Academic Press, New York.<br />

Oppenheimer, N.J., and James, T.L.<br />

(1989) Methods in Enzymology, 177,<br />

Academic Press, New York.<br />

Hyde, E.I., Ramesh. V., Roberts.<br />

G.C.K., Arrowsmith, C.H., Treat-<br />

Clemons, L., Klaic, B., and Jardetzky,<br />

O. (1989) Eur. J. Biochem. 183,545.<br />

Arrowsmith, C.H., Pachter, R.,<br />

Altman, R.B., Iyer, S.B., and<br />

Jardetzky, O. (1990) Biochemistry 29,<br />

6332.<br />

Ramesh, V., Frederick, R.O., Syed,<br />

S.E.H., and Roberts, G.C.K., unpublished<br />

work; Arrowsmith, C.H., and<br />

Jardetzky, O., unpublished work.<br />

Arrowsmith, C.H., Pachter, R.,<br />

Altman, R.B., and Jardetzky, O. (1991)<br />

Eur. J. Biochem., 202, 53.<br />

Hyde, E.I., Ramesh, V., Frederick, R.,<br />

and Roberts, G.C.K. (1991) Eur. J.<br />

Biochem., 201, 569.<br />

Shaw, W.V., and Leslie, A.G.W.<br />

(1991) Ann. Rev. Biophys. Biophys.<br />

Chem., 20, 363.<br />

[11] Jardetzky, O. & Roberts, G.C.K.<br />

(1981) NMR in Molecular Biology,<br />

Academic Press, New York.<br />

[12] Feeney, J., Birdsall, B., Akiboye, J.,<br />

Tendler, S.J.B., Barbero, J.J., Ostler,<br />

G., Arnold, J.R.P., Roberts, G.C.K.,<br />

Kuhn, A., & Roth, K. (1989) FEBS<br />

Lett., 248, 57.<br />

Bennion, C, Connolly, S., Gensmantel,<br />

N.P., Hallam, C, Jackson, C.G.,<br />

Primrose, W.U., Roberts, G.C.K.,<br />

Robinson, D.H., and Slaich, P.K.<br />

0992) J. Med. Chem., in press.<br />

[14] LeMaster, D.M. (1990) Quart. Rev.<br />

Biophys., 23, 133.<br />

[15] Derrick, J.P., Lian, L.-Y., Roberts,<br />

G.C.K., and Shaw, W.V. (1992)<br />

Biochemistry, in press.<br />

[16] Ramesh, V., Frederick, R.O., Syed,<br />

S.E.H., and Roberts, G.C.K. (1992)<br />

manuscript in preparation.<br />

[17] Frenkiel, T., Bauer, C, Carr, M.D.,<br />

Birdsall, B., and Feeney, J. (1990) J.<br />

Magn. Reson., 90, 420.<br />

[18] Derrick, IP., Lian, L.-Y., Roberts,<br />

G.C.K., and Shaw, W.V. (1991) FEBS<br />

Lett., 280, 125.


68<br />

Bulletin of Magnetic Resonance<br />

2D NMR STUDY OF DRUG-PROTEIN INTERACTIONS : ETHIDIUM BROMIDE -<br />

NEOCARZINOSTATIN COMPLEX<br />

Introduction<br />

Smita Mohanty*> Larry C. Sieker"*" and Gary P. Drobny*<br />

Department of Chemistry* and Biological Structure 1 "<br />

University of Washington<br />

Seattle, WA 98195, USA<br />

N eocarzinostatin (NCS) is a small<br />

acidic holo-protein isolated from the<br />

culture broth of Streptomyces<br />

Carzinostaticus [1]. It has a protein<br />

component (apo-NCS) of 113 amino acid<br />

residues and a non-covalently bound<br />

heat and light sensitive chromophore<br />

(NCS-chr) (Fig.l). This protein possesses<br />

antibiotic activity against organisms<br />

such as Sarcina Lutea and antitumor<br />

activity against the experimental tumors<br />

Ascitic Sarcoma 180, Ascitic<br />

Leukemia SN-36, Leukemia L-210<br />

[1-3]. It is known that the chromophore<br />

is responsible for all the biological<br />

activities and the apo-protein stabilizes<br />

and acts as a carrier for this UV<br />

sensitive component of the antitumor<br />

drug [4].<br />

Though the secondary and tertiary<br />

structure of the apo-protein is well<br />

understood from X-ray and NMR studies,<br />

little is known about the binding region<br />

and the amino acid residues involved in<br />

the drug - protein interactions in the<br />

holo-protein.<br />

Both the crystal structure at 2.8 A<br />

[5] and the 2-D NMR work done on apo-<br />

NCS [6-8] indicate that a major part of<br />

the protein is composed of a seven<br />

strand antiparallel B-sandwich formed<br />

by a three strand B-sheet and a four<br />

strand B-sheet. The rest of the protein is<br />

composed of two loops oriented<br />

somewhat perpendicularly to the B<br />

sandwich, thus forming a distinct Ushaped<br />

cleft between the four strand<br />

face of the sandwich and one of the<br />

loops of the molecule (Fig. 2). The<br />

crystal structure of holo-NCS at 2.0 A<br />

resolution indicates that the<br />

chromophore is located in this cleft. But<br />

a detailed knowledge in the region of<br />

the chromophore could not be obtained<br />

from the 2.0A map. Preliminary NMR<br />

studies also indicate that the<br />

chromophore binds within the cleft and<br />

interacts with amino acid residues in the<br />

region C37-D48 [9, 10]. Although NCS<br />

does not bind the chromophore of other<br />

streptomyces derived anti-tumoi<br />

proteins (e.g. Auromomycin) [11] it is<br />

known from X-ray studies [12] tc<br />

strongly bind a number of drugs<br />

including ethidium bromide (Fig. 3) anc<br />

daunomycin. In order to furthe<br />

elucidate the nature of drug-NC!<br />

interactions, we have initiated a study o<br />

both holo-NCS and the complex betwee<br />

ethidium bromide and NCS.<br />

Materials and Methods<br />

Apo-NCS solution used in o 1<br />

experiments was prepared from hoi<br />

NCS. The chromophore was extracted<br />

i


Vol. 14, No. 1-4<br />

Fig. 1: NCS-Chromophore<br />

60<br />

Fig. 2: Ribbon picture of NCS by X-ray<br />

Fig. 3: Eihidium Bromide<br />

69


I:<br />

I<br />

70<br />

the procedure of Napier et.al.[13].<br />

Purified lyophilized holo-NCS was a gift<br />

from Kayaku Co. Ltd. The 1:1 Ethidium<br />

Bromide-NCS complex was prepared by<br />

adding the protein solution in to vials<br />

containing the drug. The final solution<br />

was purified by passing through<br />

sephadex G-25 column followed by<br />

lyophilization. The lyophilized complex<br />

was brought up in lOmM acetate buffer<br />

(pH 5, 90% H2O/ 10% D2O for non<br />

exchanged protein sample and 99.98%<br />

D2O for exchanged protein sample) and<br />

lOmM EDTA. The concentration was<br />

adjusted between 2.0 mM to 2.5 mM for<br />

500|il sample.<br />

All NMR experiments were<br />

performed on a Bruker AM-500<br />

Spectrometer at 313 K. DQF- and TQF-<br />

COSY [14], RELAY [15], TOCSY [16] and<br />

NOESY [17] were acquired in TPPI mode<br />

with standard phase cycling schemes.<br />

The water resonance was presaturated<br />

by selective irradiation between 1.5 s to<br />

2 s. RELAY spectra were performed with<br />

mixing times of 30 ms (90%H2O) and 25<br />

ms (for 99.98%D2O). TOCSY spectra were<br />

performed with a variety of mixing<br />

times ranging between 40 to 80 ms.<br />

NOESY spectra were recorded with 150<br />

ms mixing time, randomly varied by<br />

10%. The data were processed with<br />

FTNMR software of Dr. Dennis Hare [18].<br />

Results and Discussion<br />

Our H-NMR assignments of<br />

ethidium bromide - NCS complex in<br />

solution indicates the drug to be located<br />

in the cleft region. There are two lines of<br />

evidence which support this conclusion.<br />

First, the chemical shifts of numerous<br />

Bulletin of Magnetic Resonance<br />

SI nn — horn ethidium<br />

to the ring ^ aromatic system,<br />

bromide's extensive aromatic y<br />

£> tp"t e -45 Gln Cy;-47 y , Asp-**<br />

Cy's-9 Gln-94, Leu-97 in the<br />

fingerprint region (Fig. 5 and F.g. 6).<br />

0.0<br />

ppm<br />

Fig 4: DQF-COSY spectrum in D2O<br />

' snowing uprtcldshid or Uu_4<br />

chemical shifts (a): apo-NCS. (b).<br />

P ,


Vol. 14, No. 1-4<br />

O36<br />

3 '•<br />

Q94" 5<br />

C47 'C37<br />

XL -<br />

L97<br />

8.8 8.0<br />

ppm<br />

9<br />

7!2<br />

Fig. 5: DQF-COSY fingerprint region of apo-NCS recorded at 40'C .<br />

so<br />

10 03<br />

C93<br />

DO<br />

OK)<br />

00<br />

i< 00<br />

00<br />


72<br />

8.0<br />

ppm<br />

7. 2<br />

Bulletin of Magnetic Resonance<br />

Fig.7: DQF-COSY and NOESY spectra of the complex in D2O<br />

showing some of the intermolecular NOE observed between<br />

the aromatic protons of ethidium bromide (shown by dotted<br />

lines) and B-protons of certain residues within the cleft.


Vol. 14, No. 1-4 73<br />

While residues Cys-93, Gln-94,<br />

Leu-97 are in the four strand face of the<br />

6-sandwich, which forms one side of the<br />

cleft, residues Gln-36, Cys-37, Ala-38,<br />

Trp-39, Leu-45, Cys-47, Asp-48 are in<br />

one of the loops that forms the other<br />

side of the cleft. Second, a number of<br />

NOEs have been observed to occur<br />

between protons on ethidium bromide<br />

and residue protons within the cleft. The<br />

Leu-45 methyl group is ring current<br />

shifted and shows NOEs to the<br />

methylene protons and aromatic protons<br />

of ethidium bromide. Intermolecular<br />

NOEs are also observed between the<br />

aromatic protons of ethidium bromide to<br />

the aromatic proton of Trp-39 and to the<br />

8 protons of Ser-98, Cys-37 and Gly-96<br />

(Fig. 7). Additional intermolecular NOEs<br />

are observed but have not been<br />

unambiguously assigned.<br />

Conclusion<br />

NMR assignment based on<br />

i coherence transfer experiments and<br />

y


74<br />

[11] L. S. Kappen, M. A. Napier, I. H.<br />

Goldberg, and T. S. A. Samy,<br />

Biochemistry., 19, 4780.(1980).<br />

[12] L. C. Sieker, Personal communication(<br />

1991).<br />

[13] M. A. Napier et.al., Biochem.<br />

Biophys. Res. Commun., 89,<br />

635. (1979).<br />

[14] U. Piantini, O. W. S0rensen, and<br />

R. R. Ernst, J. Am. Chem. Soc,<br />

104, 6800. (1982).<br />

[15] G. Eich, G. Bodenhausen, and R.<br />

R. Ernst, J. Am. Chem. Soc, 104,<br />

3732. (1982).<br />

[16] L. Braunschweiler and R.R.<br />

Ernst, /. Magn. Reson., 53, 521.<br />

(1983).<br />

[17] Anil Kumar, R. R. Ernst, and K.<br />

Wuthrich, Biochem. Biophys.<br />

Res. Commun., 95, 1. (1980).<br />

[18] D. Hare, Hare Research:<br />

Woodinville, WA.<br />

BulJetin of Magnetic Resonance<br />

'•A


Vol. 14, No. 1-4 75<br />

1 Introduction<br />

Transferred NOE experiments are being utilized to<br />

determine the structure of small peptide ligands bound to<br />

proteins of molecular weight up to 10 6 daltons (1-4). For<br />

larger peptides, 2D phase-sensitive NOESY is the optimal<br />

method for obtaining transferred NOE's between all the<br />

protons in the free ligands (1-2). In case of further<br />

spectral overlap, transferred NOE's between protons of<br />

selected amino acid residues can be obtained by use of the<br />

homonuclear TOCSY-editted 2D NOESY technique (5).<br />

Ultimately, NOE's between all the spin systems may be<br />

resolved by use of the 3D NOESY-TOCSY experiment<br />

(6). The resolved NOE's can be input into a procedure for<br />

the refinement of the dynamic structures of bound ligands<br />

(7).<br />

In practice, the interpretation of NOESY data is<br />

complicated by the presence of the large solvent resonance<br />

and baseplane problems, especially for samples with very<br />

low concentration of the material under study. In addition,<br />

magnetization transfers from the binding protein often<br />

result in further spectral distortions. We have been<br />

, developing methods to improve the quantitative accuracy<br />

|of transferred NOE's (7-9). In this paper, we summarize<br />

|Procedures developed for the optimized acquisition and<br />

ssing of multi-dimensional transferred NOE spectra<br />

• that more accurate quantitative results can be obtained<br />

i experimental data.<br />

Quantitative Analysis<br />

in Multi-Dimensional Transferred NOE Experiments:<br />

Improved Spectral Acquisition and Processing<br />

jlimination of Baseline Distortions in<br />

volution Dimensions<br />

evel of tj ridges in 2D phase-sensitive NOESY<br />

~ can be minimized if the FID matrix is recorded<br />

proportional phase incrementation (TPPI) with<br />

lulation along n (10). Since the tj interferograms<br />

zero at the zero time for a sine-modulated<br />

to our implementation is to start the ex-<br />

Feng Ni<br />

Biotechnology Research Institute<br />

National Research Council Canada<br />

6100 Royalmount Avenue<br />

Montreal, Quebec, Canada H4P 2R2<br />

periment at the second FID with a compensated initial<br />

delay of,<br />

where IN = At/2 is the increment time between successive<br />

FID's and x9o is the width of the 90° pulse. This<br />

implementation of the sine modulated NOESY experiment<br />

completely removes Fj baseline distortions (Figure 1A)<br />

with no need for any further baseline correction along this<br />

dimension (8). Furthermore, the resulting NOESY matrix<br />

can be phased to absorption along the F i direction with a<br />

Oth order phase of exactly 90° (in effect, a sine transform).<br />

This eliminates possible baseline distortions associated<br />

with phase correction after a real Fourier transformation<br />

(11). The procedure of delay-compensated sine modulation<br />

has also been applied successfully along the evolution<br />

dimensions of both homonuclear and heteronuclear 3D<br />

experiements (9,12).<br />

3 Ridge Suppression Along Detection<br />

Dimension<br />

The cause for ridges (baseline offsets and curvatures) in the<br />

detection dimension is more complicated than that for<br />

ridges along the evolution dimensions. There is often<br />

need for further correction in the frequency domain. Figure<br />

2a shows the baseline points recognized based on the first<br />

derivative (13) of a row slice of the NOESY spectrum in<br />

Figure 1A. The regions with sharp peaks are filled with<br />

interpolations (linear or polynomial) from adjacent<br />

baseline points. It is seen that the recognized baseline<br />

includes all the broad signals in the original spectrum.<br />

One can then construct a smooth curve through the<br />

available baseline points with some sort of curve fitting<br />

(Figures 2b and 2c). We adopted a simple and fast method<br />

[1]


76<br />

Bulletin of Magnetic Resonance<br />

Figure 1: Enhanced processing of 2D transferred NOE spectra. The FID matrix was acquired using sine-modulation along the<br />

ti direction. All spectra were processed by use of cosine-square windows in both directions and were plotted with the same<br />

parameters except that post-acquisition water suppression (8) was applied in (A)-(D); linear-prediction baseline correction was<br />

applied in (B); polynomial baseline correction was applied in (C); and baseline Fourier filtering was applied in (D).<br />

0.0


Vol. 14, No. 1-4<br />

10.0 8.0 0.0<br />

ppm<br />

Figure 2: Baseline fitting from incomplete data, (a) baseline points from one row of Figure 1 A. The missing points were<br />

filled with linear interpolations, (b) one possible baseline reconstructed by Fourier smoothing, (c) an approximate baseline<br />

calculated by a fit to a polynomial of fifth degree.<br />

for data smoothing based on Fourier filtering (8). Figure<br />

2b is the result of a 30-point Fourier smoothing of<br />

Figure 2a. Figure ID shows the NOESY spectrum<br />

(Figure 1A) after subtraction of the Fourier-filterred<br />

baseline points along the F2 direction. In comparison to<br />

-••• other methods (Figures IB and 1C), there is a dramatic<br />

... improvement in the clarity of the spectrum with complete<br />

^ elimination of broad signals from the original spectrum.<br />

Optimized Spectral Processing<br />

frequency domain spectrum is usually generated via a<br />

i Fourier transform of the interferogram if the data are<br />

"pled using the TPPI procedure (14). The computed<br />

"jtrum contains both real and imaginary parts which are<br />

|;cosine or sine transforms of the original real data,<br />

ctively. Theoretically, an absorption spectrum is<br />

ly the cosine transform of a cosine modulated FID or<br />

i;ne transform of a sine modulated FID (10). In<br />

however, both the real and imaginary parts have<br />

Iculated so that they can be suitably combined<br />

»g) to correct for possible phase distortions. For<br />

|«mensional NMR experiments, the evolution<br />

not ~ns can usually be sampled using the procedure of<br />

[jpensated sine (or cosine) modulation (see section<br />

2) to obtain well-phased interferograms (8,9,15). Thus,<br />

the processing procedure can be greatly simplified if only<br />

the sine or the cosine transform of the data is retained. We<br />

thus implemented the procedures for fast sine (or cosine)<br />

transformation (16). Compared to a real Fourier<br />

transform, there is in principle a factor of two increase in<br />

computational efficiency (16).<br />

To minimize truncation artifacts, special attention<br />

must be paid to the selection of window functions and/or<br />

to data extension by use of linear prediction. Attenuation<br />

of truncation is usually accompanied by broadening of<br />

spectral peaks (14). Linear prediction data extension, on<br />

the other hand, tends to increase the noise level of the<br />

spectrum as a result of errors accumulated with predicted<br />

data points (17, 18). We found that linear prediction<br />

followed by data windowing usually gives a good compromise<br />

between spectral resolution and noise suppression<br />

(9). In this case, truncation is removed by linear<br />

prediction while linear prediction errors are attenuated by<br />

window functions (18). In multi-dimensional NMR,<br />

linear prediction should be used only with the last spectral<br />

dimension after Fourier transformations along all other<br />

directions. Otherwise, linear prediction parameters must<br />

be carefully optimized to minimize error accumulation in<br />

the intermediate stages of multi-dimensional Fourier<br />

transformation.<br />

77


78<br />

5 Sensitivity Enhanced 3D NOESY-<br />

TOCSY<br />

In the usual pulse sequence for 3D NOESY-TOCSY,<br />

spin-locked coherence transfers are achieved by use of the<br />

trimmed MLEV-17 sequence (Figure 3A). We replaced<br />

the MLEV-17 sequence by a z-filtered WALTZ-16 pulse<br />

sequence (Figure 3B) to obtain sensitivity enhancement.<br />

This method involves post-acquisition combinations of<br />

the FIDs acquired from sub-groups of the full phase cycle<br />

utilized for the pulse sequence (9). The two detected FIDs<br />

B<br />

MLEV-17<br />

WALTZ-16 ($,,<br />

acq(t3)<br />

Figure 3: Pulse sequences for 3D NOESY-TOCSY. (A).<br />

The original 3D sequence incorporates a MLEV-17 pulse<br />

sandwiched by two short cw spin-lock pulses (trimmed<br />

MLEV-17). (B). The new 3D NOESY-TOCSY features<br />

a z-filtered WALTZ-16 sequence for spin-locking. The rf<br />

phases are cycled as 4>i=x, -x, y, -y; 2=-x, -x, -y -y;<br />

fa=-x, -x, -y, -y; 4=x, x, y, y; sl=y, y, -x, -x; 5=x, x,<br />

y, y; acq=x, -x, y, -y. For the same values of l\ and t2,<br />

5 is incremented by 180 degrees and another FID is<br />

acquired and stored in a different memory location. These<br />

seperated FIDs can be suitably combined to obtain a<br />

sensitivity-enhanced 3D NOESY-TOCSY spectrum<br />

(Figure 4).<br />

are stored in separate memory blocks for subsequent<br />

processing. Adequate combinations of these FIDs would<br />

restore the two orthogonal components, Iiy and l[x. If<br />

both the Iiy and 1^ FIDs are processed to yield absorptive<br />

spectra, they can be combined to produce a spectra with<br />

peaks doubled in size compared to each individual<br />

spectrum. The key to sensitivity enhancement lies in the<br />

fact that noise components in the two spectra are<br />

statistically independent and the post-processing combinations<br />

are then equivalent to signal averaging, reducing<br />

noise in the process (19).<br />

Bulletin of Magnetic Resonance<br />

With the sensitivity-enhanced NOESY-TOCSY, we<br />

utilized delay-compensated sine modulation along both ti<br />

and t2 to eliminate baseline distortions. Baseline<br />

adjustments along F3 were achieved in the frequency<br />

domain by use of Fourier baseline reconstruction (8). An<br />

absorptive spectrum can be obtained from Ijy if a sine<br />

transform is applied along both tj and t2- With the Iix<br />

components, an absorptive spectrum can be obtained only<br />

if a sine transform is used along t\ and a cosine transform<br />

used along t2. However, the points of the 1^ matrix for t2<br />

= 0 can not be sampled due to the finite widths of the 90<br />

degree pulses. Therefore, the missing first point was<br />

estimated via a linear prediction algorithm (17).<br />

Alternatively, the first data point can be left as zero during<br />

cosine transformation. Baseline offsets as a result can be<br />

corrected afterwards in the frequency domain.<br />

Sensitivity-enhanced NOESY-TOCSY was applied to<br />

an anticoagulant pep tide (2). The concentration of the<br />

peptide was 6 mM and the thrombin concentration was 0.5<br />

mM in an aqueous solvent of 90% H2O and 10% D2O at<br />

pH 5.5. The experiment composed of four scans for each<br />

of the two FIDs with each pair of fixed values for ti and<br />

t2- The data were acquired on a Briiker AMX-500 MHz<br />

NMR spectrometer and the 3D FID matrix (I(y or 1^ components)<br />

was of the sizes 512(t3) x 160(ti) x 62(t2).<br />

Residual solvent signals were suppressed by linear<br />

prediction time-domain convolution (8). The data along ti<br />

was extended to 200 and along t2 to 80 by linear prediction<br />

(17). Kaiser windows were used in all dimensions to<br />

reduce the effects of error propagation in linear prediction.<br />

Frequency-domain baseline adjustments were applied only<br />

along the F3 spectral dimension. The final sizes of the<br />

spectra were 512(F3) x 256(Fj) x 128(F2). Figure 4A is<br />

the F1-F2 plane sliced through the frequency of one of the<br />

well-resolved 6CH2 protons of Pro along the F3<br />

dimension of the Iiy 3D spectral matrix. The corresponding<br />

IiX spectrum (Figure 4B) also contains similar<br />

information but with somewhat reduced intensities for<br />

some of the crosspeaks. This is probably due to the fact<br />

that the 1^ components travel through different transfer<br />

pathways and are much more sensitive to prolonged delays<br />

and/or phase offsets both before and after the WALTZ-16<br />

spin-lock pulse. Nonetheless, combination of the two<br />

spectra still produced a sensitivity-enhanced spectrum<br />

(Figure 4C). This is evident if one compares the NOE<br />

crosspeaks between the aCH and PCH protons of ProfiO<br />

and the NH proton of Glufi! (60A/61N and 60B/61N of<br />

Figure 4) and if one inspects the selected slices (Figure 5)<br />

through the sensitivity-enhanced spectrum compared to the.,<br />

original spectrum. |


Vol. 14, No. 1-4<br />

A<br />

t<br />

608/6 IN<br />

.' i |,<br />

608/63E<br />

600/630 1<br />

.6 ft* fom<br />

\ ft<br />

60A/6IN AX<br />

0 /<br />

6.0 4.0<br />

F1( ppm )<br />

a /^ 59C/600<br />

608/6 IN<br />

2.0 8.0 6.0 4.0<br />

F"K ppm)<br />

2.0<br />

Figure 4. 3D transferred NOES Y-TOCSY spectra of an anticoagulant<br />

peptide, DSS-F-E-E-I-P-E-E-Y-L-QGS. (A) the<br />

plane was extracted from the conventional 3D NOESY--<br />

TOCSY Iiy spectrum. Only positive levels above 0.008<br />

are plotted. (B) the same plane as in (4A), but extracted<br />

from the 1^ spectrum. Only negative levels below -0.008<br />

are plotted. (C) the same plane as in (4A) and (4B), but<br />

from the combined spectra of both the and the Ijy components.<br />

The contour levels are above 0.012.<br />

79


80<br />

10.0<br />

o.o<br />

Figure 5. Selected slices from the sensitivity-enhanced 3D<br />

spectrum (top trace in each box) compared to those from<br />

the conventional spectrum (bottom trace in each box).<br />

Top trace of Box A, ID spectrum sliced through the PCH2<br />

chemical shift of Pro GO (labelled as 60B in Figure 4C);<br />

bottom trace of Box A, the corresponding slice from<br />

Figure 4A. Box B, slices corresponding to the aCH<br />

chemical shift of Pro^o (Labelled as 60A in Figure 4).<br />

6 Summary<br />

We have optimized procedures for both acquisition and<br />

processing of homonuclear 2D and 3D spectra in aqueous<br />

solutions. These include a new 3D NOESY-TOCSY<br />

pulse sequence that can be used with sine modulation to<br />

simplify spectral processing and to improve spectral<br />

baselines. It is also demostrated that orthogonal<br />

components of the spin-locked magnetizations can be<br />

suitably combined during processing to achieve sensitivity<br />

enhancements for 3D NOESY-TOCSY. These improved<br />

schemes are not limited to transferred NOE experiments.<br />

They should be of general applicability for resonance<br />

assignments and structure determination of dilute proteins.<br />

7 References<br />

Bulletin of Magnetic Resonance<br />

!R Ni, Y. Konishi, R. B. Frazier, H. A. Scheraga, and S.<br />

T. Lord, Biochemistry 28,3082 (1989).<br />

2 F. Ni, Y. Konishi, and H. A. Scheraga, Biochemistry<br />

29, 4479 (1990).<br />

3A. P. Campbell, and B. D. Sykes, J. Mol. Biol. 222,<br />

405 (1991).<br />

4 S. J. Landry, R. Jordan, R. McMacken, and L. M.<br />

Gierasch, Nature 355,455 (1992).<br />

5 V. Sklenar, and J. Feigon, J. Am. Chem. Soc. 112,<br />

5644 (1990).<br />

6G. W. Vuister, R. Boelens, and R. Kaptein, /. Magn.<br />

Reson. 80,176 (1988).<br />

TF. Ni, J. Magn. Reson. 96,651 (1992).<br />

8F. Ni, /. Magn. Reson. (1992a), in press.<br />

9F. Ni,7. Magn. Reson. (1992b), in press.<br />

10<br />

G. Otting, H. Widmer, G. Wagner, and K.<br />

Wuthrich, J. Magn. Reson. 66,187 (1986).<br />

11D. Marion, and A. Bax, 7. Magn. Reson. 79, 352<br />

(1988).<br />

12K. A. Carpenter, and F. Ni, J. Magn. Reson.<br />

(1992), in press.<br />

13 W. Dietrich, C. H. Riidel, and M. Neumann, J.<br />

Magn. Reson. 97,1(1991).<br />

14<br />

R. R. Ernst, G. Bodenhausen, and A. Wokaun, "<br />

Principles of Nuclear Magnetic Resonance in One and<br />

Two Dimensions", Clarendon Press, Oxford, 1987.<br />

15<br />

D. Marion, and A. Bax, J. Magn. Reson. 83, 205<br />

(1989).<br />

i&W. H. Press, B. P. Flannery, S. A. Teukolsky, and<br />

W. T. Vetterling, "Numerical Recipes'. The Art of<br />

Scientific Computing," Chap.13 and 14, Cambridge Univ.<br />

Press, Cambridge, 1986.<br />

«F. Ni, and H. A. Scheraga, /. Magn. Reson. 70,<br />

506 (1986).<br />

l«E, T. Olejniczak, and H. L. Eaton, /. Magn. Reson.<br />

87,628 (1990).<br />

19 J. Canavagh, and M. Ranee, J. Magn. Reson. 88,<br />

72 (1990).


Vol. 14, No. 1-4 81<br />

TIME-RESOLVED SOLID-STATE NMR: SMALL MOLECULES AND ENZYMES<br />

IN RAPIDLY FROZEN SOLUTION<br />

INTRODUCTION<br />

Jeremy N. S. Evans§**, Richard J. Appleyard§ and Wendy Shuttleworth§<br />

Departments of Biochemistry/Biophysics§ and Chemistry^,<br />

Washington State University, Pullman, WA 99164-4660. U. S. A.<br />

An important aspect of understanding enzymatic reaction<br />

mechanisms is the determination of the molecular structures<br />

of enzyme-bound substrates and products. NMR spectroscopy<br />

is in a unique position in that it is the only relatively<br />

high resolution structural technique which can focus<br />

on specific parts of the molecule, such as enzyme-bound<br />

intermediates. It has been widely applied to the study of enzyme<br />

mechanisms in both the solution and solid states [1-3]<br />

at ambient temperatures and at sub-zero temperatures [4].<br />

However, the current molecular weight limit in solution is<br />

around 50 kDa [2]. In contrast, with solid-state NMR spectroscopy<br />

[5] anisotropic and dipolar broadening can be<br />

reduced by cross-polarization magic angle sample spinning<br />

(CP-MASS), and there is no known molecular weight limit.<br />

We have therefore sought to develop a new technique called<br />

time-resolved solid-state NMR spectroscopy [6] which involves<br />

applying CP-MASS NMR to studying rapidly<br />

freeze-quenched enzyme-substrate mixtures.<br />

The rapid freeze-quench method involves rapidly mixing<br />

enzyme and substrate together and freezing by spraying the<br />

mixture directly into a secondary cryogen such as liquid<br />

propane cooled to ~ 85 K. CP-MASS NMR of the enzymesubstrate<br />

mixture is carried out as a function of mixing<br />

time, and the transient enzyme-bound species may be<br />

detected. Since it is largely unknown how the NMR spectrum<br />

of solutes are affected by the structure of water in<br />

frozen solution, we have studied this with a small molecule,<br />

glycine [7], and report the results in preliminary form here.<br />

We have subsequently applied these methods to the direct<br />

observation of an enzyme-intermediate complex by timeresolved<br />

solid-state NMR spectroscopy [6] and report our<br />

preliminary results here.<br />

MATERIALS AND METHODS<br />

NMR spectroscopy was carried out using a wide-bore,<br />

T Chemagnetics CMX-400 spectrometer, operating at<br />

iPO.l MHz for *H and 100.6 MHz for 13 C. The small<br />

nolecule studies were carried out using zirconia rotors (7<br />

m) in a double resonance, variable temperature (VT),<br />

ic angle sample spinning (MASS) Chemagnetics proto-<br />

Pencil-rotor probe. The enzyme studies were carried out<br />

*? Dotv °SI-368 double resonance probe using a 5mm<br />

jpnire rotor. Stable spinning was achieved with a<br />

"lagnetics spinning speed controller, which uses a<br />

'"-cessor controlled valve on the drive gas line to<br />

the speed ±5 Hz. The MAS probe was of a triple<br />

mel (dnve, bearing and VT) design. Boil-off N2 gas<br />

was cooled by an exchange dewar filled with liquid nitrogen<br />

(280 kPa, -110K). The sample cooling was performed by<br />

cooling the VT line, with the drive and bearing lines at approximately<br />

room temperature (-293K). Stable VT operation<br />

could be achieved for over 24h with this apparatus.<br />

The nuclear relaxation constants were determined by<br />

established techniques [8-10]. The proton Zeeman spinlattice<br />

relaxation, Ti H , was determined using the inversion<br />

recovery technique [11], with 13 C detection via cross<br />

polarization. The carbon Zeeman spin-lattice relaxation,<br />

Ti c , was determined using a modified cross polarization and<br />

inversion recovery technique [12]. The rotating-frame spinlattice<br />

relaxations, Tip H and Tjp C , were determined using<br />

an appropriate spin lock (50kHz) of varied length, with 13 C<br />

detection via cross polarization [9,13]. Unlabelled glycine<br />

was purchased from Sigma (St Louis, MO). 99% [1-<br />

13 C] Glycine was purchased from Cambridge Isotope<br />

Laboratories (Cambridge, MA), and 99% [2- 13 C]Glycine<br />

was purchased from MSD Isotopes (Canada). Propane gas<br />

was purchased from Bemzomatic (Medina, NY). The crystalline<br />

glycine sample was re-crystallized from a saturated<br />

aqueous solution, using ethanol as the triturant. The dimensions<br />

of the crystals were consistent with the a-form. A 1M<br />

1:1 solution (pH 7.5) of [l- 13 Ci]glycine and [2-<br />

13 Ci]glycine was used for the frozen samples. The slow<br />

frozen samples were prepared in situ. A 200 pL sample was<br />

loaded into the rotor and spun slowly (


82 Bulletin of Magnetic Resonance<br />

TABLE 1 Activation Energies for motions in different states of glycine from Ti and T\p data.<br />

Ea (kj mol* 1 )<br />

Crystalline<br />

Slow Frozen<br />

Fast Frozen<br />

(10"<br />

Crystalline<br />

Slow Frozen<br />

Fast Frozen<br />

1 Ks" 1 )<br />

(10 5 Ks 4 )<br />

(10" 1 Ks" 1 )<br />

(10 5 Ks" 1 from Ti<br />

)<br />

H data<br />

/romTj<br />

23.5<br />

21.1<br />

19.4<br />

from Tip** data<br />

28.2<br />

24.6<br />

23.1<br />

c Cl<br />

data<br />

23.5<br />

16.5<br />

19.5<br />

from Tip c data<br />

Cl<br />

13.6<br />

4.4<br />

5.2<br />

BL21(XDE3)(pLysS)(pWS230) and purified by literature<br />

methods [14].<br />

RESULTS<br />

The structures of frozen water have been extensively studied<br />

in electron microscopy [15] and materials science [16],<br />

and has been shown to be greatly affected by the freezing<br />

rate used. The main problem with water is that due to its<br />

strong propensity for hydrogen bonding, it is incredibly hard<br />

not to form crystalline hexagonal ice (I^). Two other forms<br />

of frozen water known are cubic ice (Ic) and vitreous water<br />

(Iv), also referred to as amorphous solid water (ASW). ^ ice<br />

is a metastable form of ice (with respect to 1^ ice) produced<br />

when Iv is heated. In Iv, the water molecules are randomly<br />

distributed throughout the solid phase having been unable to<br />

form an Ih lattice during the freezing process.<br />

The freeze-quench apparatus used in these studies is a<br />

modified version of that used in rapid freeze quench ESR<br />

spectroscopy studies [17,18]. It involves firing a solution<br />

at a pre-determined flow-rate through a fine nozzle into a<br />

receptacle containing a secondary cryogen (e.g. liquid<br />

propane) cooled by immersion in a primary cryogen (usually<br />

liquid nitrogen). The cooling rate is dependent on the size of<br />

the droplets created by the exit nozzle and can be determined<br />

by the study of a reaction with a known rate. ESR studies<br />

[17,18] have estimated the freezing time by this technique to<br />

be 2-6 ms and the cooling rate to be -10 5 Ks' 1 .<br />

We have investigated how nuclear relaxation rates of a<br />

solute in frozen solution are affected by the freezing rate,<br />

since they have been shown to be a sensitive probe of<br />

molecular motion in the solid state [19,20]. We have chosen<br />

to study glycine because of its simple mode of motion that<br />

has been characterized previously using NMR relaxation<br />

techniques [21]. Since the relaxation properties of pure ice<br />

have also been studied [22,23], this provides a good model<br />

for characterization of the molecular motions of solute<br />

molecules in frozen solution.<br />

Previous studies [21,24] of polycrystalline glycine have<br />

shown that the main source of Tj H relaxation is the random<br />

modulation of the proton magnetic dipolar interaction by the<br />

C2<br />

30.2<br />

17.6<br />

19.4<br />

C2<br />

0.26<br />

3.3<br />

1.6<br />

reorienting ammonium group (-NH3+). The activation<br />

energy of the -NH3 + rotation in glycine was determined to be<br />

28.6 kJ mol" 1 which agrees well with our T^ and Ti c Ea<br />

values for crystalline glycine shown in Table 1. The Ea values<br />

from the Tip C data (Table 1) display a deviation from<br />

this value, which is even more evident for the slow and fast<br />

frozen states. However, since the latter fits are less welldefined,<br />

these Ea values should be regarded as less reliable.<br />

The crystal structures of a-glycine has been solved to a<br />

high degree of accuracy by X-ray [25] and neutron diffraction<br />

[26] techniques. The glycine molecule is in the zwitterion<br />

form. The dipolar glycine molecules are linked by two short<br />

-N-H--O- hydrogen bonds to form layers connected in an<br />

antiparallel manner by weaker -N-H •••0- hydrogen bonds.<br />

The close packing nature of the crystalline lattice, coupled<br />

with the hydrogen bond interactions, will produce large barriers<br />

to rotation for both the -NH3"*" and -CO2' groups as<br />

observed.<br />

The relaxation process in the frozen solution is more<br />

complicated due to the presence of an additional relaxation<br />

pathway provided by the ice lattice. Previous studies [22,23]<br />

have shown the measured activation energy for the Tj and<br />

Tip processes in pure crystalline ice to be 59.8 kJ mor 1 .<br />

This gives a xc (-10°C) = 7.5 (is, and it has been concluded<br />

[23] that the dominant magnetic relaxation mechanism must<br />

be the diffusion of Schottky defects through the ice lattice.<br />

Although there are insufficient data points for the number of<br />

variables to fit two correlation times, it is clear that the diffusion<br />

due to Schottky defects is not detected since no activation<br />

energy of ~60 kJ mol" 1 is evident from the Ti dau<br />

for the frozen solutions given in Table 1. The field strengu<br />

used is much higher than those in previous studies<br />

Therefore the (Ti)min occurs at a temperature above tin<br />

melting point of the ice and cannot be detected.<br />

Table 1 clearly shows a drop in Ea for -NH3 group rot*<br />

tion of glycine calculated from the Ti H and Tip H data in tr<br />

order, crystalline; slow-frozen solution; fast-frozen soluua<br />

The structure of the frozen solution samples is less cle|<br />

Many studies [16,27] have been made of water frozen unfl


Vol. 14, No. 1-4 83<br />

a variety of different freezing rates. Much effort has been<br />

made to develop techniques of freezing with rates that<br />

approach 10 5 - 10 10 Ks- 1 , the estimated [28-31] rates<br />

necessary to form vitreous ice, Iv (amorphous solid water).<br />

It is now generally accepted that freezing rates >10 5 Ks" 1<br />

are required before Iv will form, however there are other<br />

factors to consider, such as sample size, freezing medium<br />

and so on. For solutions frozen at rates well below the 10 5 -<br />

10 10 Ks -1 range, the frozen sample is thought to be made<br />

up of bulk hexagonal ice and aggregated solute. For a slowfrozen<br />

1M glycine solution, aggregates of crystalline<br />

glycine will form along with the Ih, i.e. the freezing rate is<br />

slow enough that the crystalline states have time to form.<br />

The crystalline glycine is distributed throughout the Ih<br />

lattice, resulting in a large surface to volume ratio. The<br />

glycine molecules exposed to the ice front have fewer lattice<br />

interactions, giving the -NH3 + group a lower energy barrier<br />

to rotation. This explains the drop in Ea observed in Table I<br />

between crystalline and slow frozen glycine.<br />

Even freezing rates of -10 5 Ks' 1 have been shown [32]<br />

not to prevent completely the segregation of solute from<br />

solvent, nor to form Iv, in both pure water and dilute aqueous<br />

solutions. However, the inclusion of a solute increases<br />

the propensity for vitrification [16], and is the basis for the<br />

use of cryoprotectants. In sufficiently large quantities, a<br />

solute will depress the free energy of the liquid water relative<br />

to the Ih lattice and reduce the driving force for crystallization.<br />

It is therefore unclear what the Iv / Ih ratio is under the<br />

conditions used here. However, once the secondary cryogen<br />

(liquid propane) is removed under vacuum at 223 K, any Iv<br />

that might have been formed at 85 K would readily undergo<br />

a phase transition to Ih at -160 K [33]. Therefore, at the<br />

temperatures used in this NMR study, it is unlikely that Iv<br />

is present. Rapid cooling has been shown [34] to cause a<br />

highly dispersed Ih phase regardless of the degree of vitrification.<br />

A rapidly frozen 1M glycine solution will consist of<br />

.' highly dispersed ice and glycine phases and the very high<br />

^ surface: volume ratio achieved renders the system metastable.<br />

'Under these conditions, maturation can take place with the<br />

^crystal size distribution broadening and shifting to larger<br />

fSfy?^ 1 dimensions. However, the rate is dependent on the<br />

'" ' temperature and is not significant within the timeof<br />

the relaxation studies. It can be rationalized that<br />

very high surface: volume ratio contributes to the<br />

r drop in Ea observed between the slow frozen and fast<br />

solutions. It is also possible that the glycine is<br />

*nt predominantly in the amorphous form in the rapidly<br />

*en sample, since the freezing rate is fast enough to<br />

Jtnt the glycine from forming a microcrystalline<br />

" ""The barriers to the -NH3+ group rotation in the<br />

"i state will be lower since there will not be such a<br />

» of ordered hydrogen bonding, and this will<br />

'- further to the drop in Ea.<br />

applied these methods to a well-characterized<br />

g-€nolpyravylshikimate-3-phosphate (EPSP) syn-<br />

thase (EC 2.5.1.19), which catalyzes the penultimate step in<br />

the aromatic amino acid biosynthetic pathway in higher<br />

plants and bacteria. EPSP (4) is formed from shikimate-3phosphate<br />

(S3P, 1) and phosphoenolpyruvate (PEP, 2) (see<br />

Scheme 1). The enzyme is a monomer with molecular<br />

weight Mx = 46,000 and the cloned E.coli gene has been<br />

used to generate a hyperexpressing strain [14], so that the<br />

bacterial enzyme is available in gram quantities.<br />

Furthermore, EPSP synthase is the primary site of action of<br />

the herbicide glyphosate [35], or N-phosphonomethylglycine.<br />

This is a broad spectrum post-emergence herbicide<br />

with worldwide applications in agriculture and horticulture.<br />

This enzyme has been extensively studied by kinetic and<br />

biophysical methods in the last 5 years. The direct observation<br />

of the enzyme-intermediate (E«I) complex was first reported<br />

by our laboratory [1,2], later confirmed by another<br />

laboratory [36]. There are only a handful of enzymes for<br />

which the full kinetic and thermodynamic profile has been<br />

determined, and EPSP synthase is one of this select group<br />

[37].<br />

Concerns about the fate of the protein under these conditions<br />

of freezing have been addressed [38], and at the protein<br />

concentrations and freezing rates (10 5 K s 1 , or millisecond<br />

time regime) employed here, there is significant dispersal of<br />

the solute in the frozen water [7]. Furthermore the frozen<br />

water is probably largely amorphous [39], with the protein<br />

itself acting in a similar manner to a cryoprotectant [39]<br />

thereby reducing the formation of hexagonal ice that is<br />

detrimental to the protein. We have found that the specific<br />

activity of EPSP synthase employed in the experiments<br />

reported here was found to be unchanged before and after<br />

rapid freezing at these high protein concentrations.<br />

Figure 1 shows 13 C CP-MASS solid-state NMR spectra<br />

of EPSP synthase«S3P mixed with [2- 13 C]PEP under<br />

steady-state conditions (in the presence of the product, inorganic<br />

phosphate) and under pre-steady state conditions where<br />

/ indicates the time elapsed from the start of the reaction.<br />

The intermediate (E«I) is clearly visible at 104 ppm [1,2]<br />

under steady-state conditions and its build-up demonstrated as<br />

the reaction proceeds under pre-steady state conditions. It is<br />

worth noting that the intensities of the E«I resonance correlate<br />

well with the concentrations observed by chemical<br />

quench methods [37]. On allowing the pre-steady state reaction<br />

to proceed for a few minutes, the turnover of intermediate<br />

(E«I) to product (E«EPSP) is evident. In addition to the<br />

resonance due to the E«I complex, the resonance due to the<br />

E'EPSP product complex builds up at 155 ppm, and one<br />

tentatively assigned to the E-PEP substrate complex appears<br />

transiently at 151 ppm. Note that under the conditions<br />

which these spectra were obtained, the free small molecules<br />

(substrate and product) are not detected due to their relative<br />

isotropic motion in frozen solution. Although the spin<br />

locking fields used provide excellent cross-polarization for<br />

rapidly-frozen solutions of EPSP synthase and enzymebound<br />

species also provide very poor cross-polarization for<br />

PEP and EPSP.


84 Bulletin of Magnetic Resonance<br />

240.0 160.0 80.0 0.0<br />

ppm<br />

Figure 1 The 9.4 T 13 C CP-MASS solid-state NMR<br />

spectra of EPSP synthase at 233 K under conditions indicated:<br />

steady-state (EPSP synthase (4 mM) in 20 mM phosphate<br />

buffer, pH 7.8,15 % D2O in the presence of S3P (9.1<br />

mM) and [2- 13 C]PEP (7.6 mM) slow frozen in the NMR<br />

rotor over 90 s; rapidly mixed and freeze-quenched after time<br />

1 (EPSP synthase (4 mM) in 50 mM tris buffer containing<br />

5 mM P-mercaptoethanol, pH 7.8, in the presence of S3P<br />

(40 mM) and rapidly mixed with [2- 13 C]PEP (40 mM) and<br />

sprayed into liquid propane at -85 K.<br />

DISCUSSION<br />

These relaxation studies have provided evidence that the<br />

distribution of the Ih and the solute becomes more<br />

dispersive as the freezing rate is increased. This is<br />

manifested as a lowering of the activation energy for the<br />

-NH3 + group rotation as the amount of glycine free from<br />

crystalline packing forces increases. This is rationalized by<br />

two factors: the increase in surface: volume ratio in the<br />

dispersed phase; and the increase in the percentage of<br />

amorphous glycine present This demonstrates the difference<br />

between fast and slow freezing and suggests that at 10 5 Ks"<br />

1 , either some Iv or highly dispersed Ih is forming. This is<br />

encouraging from the point of view of studying rapidlyfrozen<br />

proteins, since this suggests minimal possible<br />

damage through Ih formation. The characterization of the<br />

freezing rate also indicates that it will be sufficient for<br />

trapping transient intermediates in an enzymatic reaction.<br />

Time-resolved solid-state NMR spectroscopy provides a<br />

major technological advance in the study of enzymatic reaction<br />

mechanisms. One important consideration in any<br />

attempt to detect transient intermediates in addition to their<br />

lifetimes, is their pre-steady state concentrations. This is<br />

dependent upon the kinetics and thermodynamic stability of<br />

intermediates of each individual enzyme. Some enzymes,<br />

like EPSP synthase, have unusually stable intermediates.<br />

However, we would expect that the majority of enzymes are<br />

evolving towards "perfection" [40], and stabilize highly unstable<br />

intermediates. Furthermore, when coupled with the<br />

elegant solid-state NMR distance measurements that have<br />

been introduced recently [41] this technique will be uniquely<br />

able to "map out" molecular conformations of intermediates<br />

and enzyme active site-intermediate distances as an enzymatic<br />

reaction proceeds. This will provide the crucial missing<br />

structural details which Laue X-ray diffraction and allied<br />

techniques cannot provide, and enable the complete definition<br />

of the molecular events of enzyme catalysis.<br />

ACKNOWLEDGEMENTS<br />

We should like to thank Prof. R.C.Bray (University of<br />

Sussex) for helpful discussions in the design of the freezequench<br />

apparatus, and Dr. Yves Dupont of Biologic Co. for<br />

implementing some of these designs. We are also grateful to<br />

Drs. Allan Palmer and Jim Frye of Chemagnetics/Otsuka<br />

Electronics Ltd. for help with the custom design of the lowtemperature<br />

equipment for the NMR spectrometer, member!<br />

of WSU technical services instrument shop and Fret<br />

Schuetze of WSU electronics shop for numerous customiza<br />

tions, and Dr. Dave Cleary for help with the ESR spec<br />

troscopy. This work was supported in the early stages by th;<br />

donors of the Petroleum Research Fund (American ChemicJ<br />

Society) and National Institutes of Health grant RR O60Q|<br />

and mainly by National Institutes of Health graf<br />

GM43215. The WSU NMR Center is supported by NJ<br />

grant RR 0631401, NSF grant CHE-9115282 and Battd<br />

Pacific Northwest Laboratories Contract No.12- 097718-/<br />

L2.


Vol. 14, No. 1-4 85<br />

REFERENCES<br />

11] P.N.Barlow,R.J.Appleyard, BJ.O.Wilson and<br />

J.N.S .Evans, Biochemistry, 28, 7985 and p. 10093<br />

[2]<br />

[ 3]<br />

[4]<br />

[5]<br />

[6]<br />

[7]<br />

[8]<br />

[9]<br />

[10]<br />

[11]<br />

[12]<br />

[13]<br />

[14]<br />

[15]<br />

U6]<br />

[17]<br />

B US]<br />

[19]<br />

:[20]<br />

(1989).<br />

J. N. S. Evans, "NMR and Enzymes", in "Pulsed<br />

Magnetic Resonance, Optics and Imaging (Honoring<br />

E. L. Hahn)", (Ed. D. Bagguley), Oxford University<br />

Press, in press (1992).<br />

J.N.S.Evans, CBurton, P.E.Fagerness,<br />

N.E.Mackenzie & A.I.Scott, Biochemistry 25, 905<br />

(1986); JP.G.Malthouse, Prog, in NMR Spectroscopy<br />

18, pp 1-59 (1985); PHosch, Ibid. 18, 123 (1986);<br />

K.Kanamori & J.D.Roberts, Acc.ChemJRes. 152, 35<br />

(1983); W.W.Bachovchin, Biochemistry 25, 7751<br />

(1986).<br />

J.P.G.Malthouse, M.P.Gamcsik, A.S.F.Boyd,<br />

KE.Mackenzie, and Ai.Scott, JAmer.Chem.Soc.<br />

104, 6811 (1982); N.E.Mackenzie, J.P.G.Malthouse<br />

and A.I.Scott, BiochemJ. 219, 437 (1984).<br />

S.O.Smith, LPalings, V.Copie, D.P.Raleigh,<br />

J.Courtin, J.A.Pardoen, Jlugtenburg, R.A.Mathies<br />

and R.G.Griffin, Biochemistry 26, 1606 (1987).<br />

J.N.S.Evans, RJ.Appleyard & W.A,Shuttleworth,<br />

submitted (1992).<br />

RXAppleyard and JJsf.S.Evans, submitted (1992).<br />

J. S. Frye, Concepts Magn. Res. 1, 27 (1989).<br />

C. A. Fyfe, "Solid State NMR for Chemists," p. 48,<br />

CFC Press, Canada, 1983.<br />

D. W. McCall, Ace. Chem. Res. 4, 223 (1971).<br />

R. L. Void, J. S. Waugh, M. P. Klein and D. E.<br />

Phelps, J. Chem. Phys. 48, 3831 (1968).<br />

D. A. Torchia, /. Magn. Res. 30, 613 (1978).<br />

E. 0. Stejskal, J. Schaefer, M. D. Sefcik and R. A.<br />

McKay, Macromolecules 14,275 (1981).<br />

W.A.Shuttleworth, CD.Hough, KP.Bertrand and<br />

J.N.S .Evans, Protein Engineering 5, in press (1992).<br />

H. Plattner and L. Bachmann, Int. Rev. Cytol. 79,<br />

237 (1982).<br />

C. A. Angell and Y. Choi, /. Microsc. 141, 251<br />

(1986).<br />

R. C. Bray, D. J. Lowe, C. Capeillere-Blandin and E.<br />

M. Fielden, Biochem. Soc. Trans. 1, 1067 (1973).<br />

D. P. Ballou and G. A. Palmer, Anal. Chem. 46,<br />

1248 (1974).<br />

A. M. P. Goncalves, Prog. Solid St. Chem. 13, 1<br />

(1980).<br />

A. Weiss, Angew. Chem. Internal. Edit. 11, 607<br />

(1972).<br />

E. R. Andrew, W. S. Hinshaw, M. G. Hutehins and<br />

R. O. I. Sjoblom, Mol. Phys. 31, 1479 (1976a).<br />

[22] D. E. Barnaal and I. J. Lowe, /. Chem. Phys. 48,<br />

4614 (1968).<br />

[23] M. Weithase, F. Noack and J. von Schiitz, Z. Phys.<br />

246, 91 (1971).<br />

[24] E. R. Andrew, W. S. Hinshaw, M. G. Hutehins and<br />

R. O. I. Sjoblom, Mol. Phys. 34, 1695 (1977).<br />

[25] R. E. Marsh and J. Donohue, Advanc. Protein Chem.<br />

22, 235 (1967).<br />

[26] P.-G. Jonsson and A. Kvick, Ada Cryst. B28, 1827<br />

(1972).<br />

[27] E. Mayer, /. Appl. Phys. 58, 663 (1985).<br />

[28] P. T. Sargeant and R. Roy, Mater. Res. Bull. 333,<br />

265 (1968).<br />

[29] D. Tumbull, Contemp. Phys. 10, 473 (1969).<br />

[30] N. H. Fletcher, Rep. Prog. Phys. 34, 913 (1971).<br />

[31] D. Uhlmann, /. Non-Cryst. Solids 7, 337 (1972).<br />

[32] P. Briiggeller and E. Mayer, Nature 288, 569 (1980).<br />

[33] D. R. MacFarlane and C. A. Angell, J. Phys. Chem.<br />

88, 759 (1984).<br />

[34] A. Calvelo, "Development in Meat Science," p. 125,<br />

Applied Science Publishers, London, 1981.<br />

[35] H.C.Steinrucken and KAmrhein, EurJ.Biochem.<br />

143, 351 (1984).<br />

[36] K.S.Anderson et a!., Biochemistry 29, 1460 (1990).<br />

[37] K.S.Anderson, J.A.Sikorski & K.A.Johnson,<br />

Biochemistry 27, 7395 (1988).<br />

[38] See for example, Proteins at Low Temperatures (Ed.<br />

O. Fennema), Am. Chem. Soc. Adv. in Chem. Series<br />

No. 180 (1979); RFranks, Biophysics and<br />

Biochemistry at Low Temperatures, Cambridge<br />

University Press (1985).<br />

[39] L. Bachmann & E. Mayer, in Cryotechniques in<br />

Biological Electron Microscopy (Ed. R.A. Steinbrecht<br />

and K. Zierold), Springer-Verlag, Berlin p.3. (1987);<br />

P.Douzou, Cryobiochemistry: An Introduction,<br />

Academic Press (1977).<br />

[40] WJ.Albery and JJl.Knowles, Biochemistry 15, 5631<br />

(1976).<br />

[41] D.P.Raleigh, M.H.Levitt & R.G.Griffin,<br />

Chem.PhysJLett. 146, 71 (1988); D.P.Raleigh,<br />

F.Creuzet, S.K.Das Gupta, M.H.Levitt &<br />

R.G.Griffin, JAm.Chem.Soc. Ill, 4502 (1989);<br />

T.Gullion & J.Schaefer, JMagnReson. 81, 1%<br />

(1989); G.R.Marshall et aL.JAm.Chem.Soc. 112,<br />

963 (1990); S. M. Holl, R. A. McKay, T. Gullion, J.<br />

Schaefer, J. Magn. Reson. 89, 620 (1990); V.Copi6 et<br />

al.. Biochemistry 29, 9176 (1990).


86<br />

Abstract<br />

Coupled methyl groups in dimethyl sulphide<br />

M.R. Johnson, S. Clough, A.J. Horsewill and I.B.I. Tomsah<br />

Bulletin of Magnetic Resonance<br />

Department of Physics, University of Nottingham, Nottingham. NG7 2RD.<br />

Low field methyl tunnelling spectra of dimethyl sulphide<br />

have been measured using field cycling NMR and indicate<br />

the existence of two distinct methyl groups with tunnel.<br />

frequencies of 100kHz and 750kHz. Spectra show a significant<br />

broadening of the Larmor peak at those fields at<br />

which the Larmor and tunnel frequencies are equal. These<br />

are resonances between the rotational and spin dynamics<br />

of the methyl groups. An associated" change in the intensity<br />

and position of the 100kHz sidebands suggests that<br />

methyl group coupling may be responsible for these effects.<br />

Narrow 100kHz sidebands are restored by irradiating<br />

with an external RF field of this frequency which we<br />

suggest is due to decoupling of the rotor pairs. Measurements<br />

on partially deuterated dimethyl suphide show that<br />

the coupling cannot be intra-molecular.<br />

1 Introduction<br />

The methyl group tunnel frequency wt depends on the<br />

height of the potential barrier hindering rotation due to<br />

the molecular environment. Three stationary states are<br />

distributed equally in three potential wells and form an<br />

energy singlet and doublet split by ftwt. Excitations consist<br />

of a superposition of a pair of eigenfunctions to form<br />

a partially localised wavepacket which rotates in a definite<br />

sense at a frequency given by the energy splitting,<br />

ie ±wt or 0. With the aid of external fields the rotation<br />

frequency of a wavepacket can be changed by exerting a<br />

rotational impulse her. The rotation frequency is then determined<br />

by the difference between two of the three values<br />

(2w where i, y and z are the spin states, a or /?, at ,,<br />

proton sites which are defined by the hindering P° ten .*|<br />

[1]. The protons themselves are each equally ditbufr<br />

of


Vol. 14, No. 1-4<br />

between the three sites satisfying proton indistinguishabilitj.<br />

In this represenoation the matrix elements cf the<br />

tunnelling interaction reflect the balance of clockwise and<br />

anti-clockwise rotation of an undriven methyl group.<br />

< xyz\HR\zxy >=< zxy\HR\xyz >= A (2)<br />

A is the overlap integral between states localised in neighbouring<br />

wells of the three-fold hindering potential.<br />

Away from the anti-crossings the methyl group eigenstates<br />

are linear combinations of the localised states and<br />

are the delocalised representations of the Cz symmetry<br />

group,<br />

1 [\xyz> + exp(t2irn/3)\zxy (3)<br />

where n = 0 for an A-state and n = ±1 for Ea and £&<br />

states. Temporarily ignoring the dipole-dipole interaction,<br />

these symmetrised states have energies<br />

(4)<br />

E{\, m) = u>Lm - 2A cos(2;rn/3) (5)<br />

where OIL is the Larmor frequency and m is the spin component<br />

of \xyz > in the direction of the magnetic field.<br />

Mixtures of these eigenstates are partially localised, rotating<br />

wavepackets which have energies resulting from the<br />

differences between the rotational and magnetic energies<br />

of the component states E(Xi,rm) — E(Xj ,mj). Magnetic<br />

and rotational wavepacket energies add when the Larmor<br />

precession and the rotation have the same sense and they<br />

subtract when these senses are opposite.<br />

The exact eigenstates and eigenvalues of the methyl<br />

' group in the magnetic field are calculated numerically by<br />

;:"'diagonalising H. The field dependence of the energy levels<br />

Mis shown in figure 1. The level anti-crossings can be seen<br />

^ . 1.2 and 2.4mT for the 100kHz methl group and 8.8 and<br />

|17.6mT for the .750kHz rotor, where the A and E symmetry<br />

states reflect off each other and exchange symmedue<br />

to the finite non-secular parts of the intra-methyl<br />

1 dipole-dipole interaction.<br />

a energy levels diagram of figure 1 can be transformed<br />

1 diagram of energy differences, or wavepacket ener-<br />

|..as shown in figure 2 from which frequency spectra at<br />

magnetic fields can be predicted by taking horsections.<br />

In this way the level anti-crossings are<br />

manifest themselves as resonant motional broad-<br />

M the Larmor peak. Symmetry mixing at the<br />

sings results in stationary wavepackets with en-<br />

Pjroportional to u>£ evolving into rotating wavepackets<br />

~" T gy Proportional to wfc (x to y in figure 2). This<br />

he ability of the non-secular parts of the dipoleraction<br />

for converting spin angular momentum<br />

•tional angular momentum of the methyl group.<br />

18<br />

A(3/2)<br />

^(-3/2)<br />

A(-3/2)<br />

B(mT)<br />

Figure 1: Energy levels of dimethyl sulphide in a magnetic<br />

field showing resonances between tunnelling motion and<br />

spin dynamics as level anti-crossings at 1.2, 2.4 ,8.8 and<br />

17.6mT<br />

15-; '1:-.<br />

10-<br />

Reld(mT)<br />

5 - • .?• .:•?" ,v*<br />

0 S~=r<br />

200 400 600 800<br />

Frequency(kHz)<br />

1000<br />

Figure 2: Field/frequency plane of dimethyl sulphide<br />

showing resonant broadening of the Larmor peak at the<br />

level anti-crossings, notably at 9mT and 18mT<br />

87


88<br />

3 Low field, field cycling experiments<br />

Measuring magnetisation at low field using a Faraday law<br />

detector results in poor signal to noise. Thus in order to<br />

measure a low field frequency spectrum either a different<br />

type of detector [5] or a different technique such as field<br />

cycling must be employed. Field cycling has been used<br />

in this investigation. Each cycle of the experiment begins<br />

with the preparation of a standard magnetisation by<br />

destroying all magnetisation with a train of 90° pulses followed<br />

by the preparation period of 20 seconds in a field<br />

of 0.6T, The field is then switched rapidly in about 2 seconds<br />

to the chosen magnetic field where the initial magnetisation<br />

is destroyed by spin-lattice relaxation by an<br />

amount proportional to the relaxation, rate T{~ and the<br />

time at low field, 15 seconds. Anomalously rapid relaxation<br />

can be stimulated by an external RF field of scanning<br />

frequency / if this frequency matches the characteristic<br />

wavepacket frequencies in figure 2 and can therefore<br />

excite these wavepackets. The remnant magnetisation is<br />

measured by a single 90° pulse after a rapid field switch<br />

to 0.6T. The cycle is repeated increasing / each time, producing<br />

a frequency spectrum which is a flat plateau in<br />

which holes are drilled where f is equal to the frequencies<br />

predicted in figure 2. These spectra are inverted to give<br />

peaks for presentation.<br />

A slightly more complicated experiment entails the application<br />

of a second RF field of fixed frequency fsU at<br />

low field , which is applied on the same coils as the scanning<br />

field in an alternating sequence of short bursts of 0.5<br />

seconds of each field. This is a stirring experiment which<br />

may affect the intensity of spectral peaks by depopulating<br />

energy levels and enhancing transitions which would otherwise<br />

become suppressed as the populations of the initial<br />

and final states become equal [6].<br />

4 Sample preparation<br />

Dimethyl sulphide was obtained from the Aldrich Chemical<br />

Company and was used in an unsealed sample tube.<br />

The deuterated sample was prepared in the Chemistry Department<br />

by mixing aqueous solutions of CHzS~ No. and<br />

CD2I as outlined in [7]. It was also used in an unsealed<br />

sample tube.<br />

5 Low field frequency spectra of<br />

dimethyl sulphide<br />

A typical low field spectrum of dimethyl sulphide, measured<br />

at 3.5rnT, is shown in figure 3 and it is in excellent<br />

agreement with the corresponding slice through figure 2,<br />

indicated by the broken line. The Larmor peak occurs at<br />

175kHz indicating a true magnetic field of 4.1mT and the<br />

Am = 2 version of this peak occurs at 350kHz. These are<br />

Bulletin of Magnetic Resonance<br />

250 BOO 750 1000<br />

Frequency [kHz]<br />

1250<br />

Figure 3: Frequency spectrum of dimethyl sulphide<br />

mesaured at 3.5mT<br />

labelled 1 and 2 respectively. They are each flanked by a<br />

pair of 100kHz sidebands (labelled 1± and 2±) and also<br />

by a pair of 750kHz sidebands (labelled 1± and 2±). Since<br />

the Larmor frequency is much smaller than the larger tunnel<br />

frequency, the reflected low frequency sidebands give<br />

rise to a 750kHz spectrum which appears as a Am = 0<br />

peak split into five peaks, each separated by the Larmor<br />

frequency.<br />

Frequency spectra of dimethyl sulphide have been measured<br />

at many magnetic fields up to 35mT and a selection<br />

of these are shown in figure 4. The spectra are aligned by<br />

their Larmor peaks and only the frequency range incorporating<br />

the 100kHz sidebands is covered. Away from the<br />

level anti-crossings of the 750kHz rotor, that is below 7mT,<br />

between 10 and 14mT, and above 20mT, the Larmor peak<br />

and sidebands are well defined and reasonably symmetrical.<br />

However in the vicinity of the level anti-crossings,<br />

between 7 and lOmT and between 14 and 20mT, this discrete<br />

structure is lost as the Larmor peak broadens, the<br />

low frequency sideband virtually disappears and the high<br />

frequency sideband is poorly defined. Where a frequency<br />

separation of the Larmor peak and a sideband can be determined<br />

it is apparent that rotation frequency of these<br />

wavepackets is less 100kHz and is sometimes as small a?<br />

70kHz. . 1<br />

Although these effects occur at the level anti-crossing<br />

of the 750kHz rotor, the field range over which they occ^<br />

and the extent of the frequency broadening are too gre^<br />

for them to be explained by dipolar broadening alone. Ful<br />

thermore the the dramatic change in the sideband mtei<br />

sity and frequency is not predicted by the dipole-dipole"<br />

teraction. That it is the 100kHz spectrum which is so pr'<br />

foundly affected by the level anti-crossings of the 750K<br />

rotor suggests that these spectra may be evidence of sp


Vol. 14, No. 1-4 89<br />

Z<br />

o<br />

m zo<<br />

:<br />

111<br />

DC<br />

Q<br />

LU<br />

h-<br />

DC<br />

LU<br />

• \<br />

:<br />

v"V<br />

• ".}'<br />

• ' : I'.<br />

I<br />

I<br />

• I<br />

i '."<br />

I<br />

I<br />

I<br />

I V<br />

coz-1OO co'z ©2+100<br />

22.5<br />

20<br />

• 4'~.~*~ 15<br />

• • !•<br />

'.. ' i.<br />

• r-^ '12.5<br />

Si W9<br />

A 5<br />

• •"?• y '<br />

E<br />

Q<br />

UJ<br />

U_<br />

o<br />

LU<br />

z<br />

CD<br />

<<br />

FREQUENCY/kHz<br />

4: Field dependence of the Larmor peak and the<br />

coupling between the 100kHz and 750kHz rotors.<br />

The 100kHz sidebands can be restored in a stirring experiment<br />

with an external RF field with a stirring frequency<br />

of 100kHz, as shown in the spectra in figure 5. It<br />

appears that stirring at 100kHz drives rotation of methyl<br />

wavepackets at this frequency and therefore prevents coupling<br />

which has been seen to modify this frequency, as<br />

shown in figure 4.<br />

In order to probe the methyl group coupling a partially<br />

deuterated sample of dimethyl sulphide, in which<br />

one methyl group per molecule is fully deuterated, was<br />

investigated. Figure 6 shows a low field spectrum of the<br />

deuterated sample measured at 4mT which is very similar<br />

to the spectrum of the fully protonated sample shown in<br />

figure 3, one difference being the increase in the large tunnel<br />

frequency from 750kHz to 780kHz. Field dependent<br />

spectra of the deuterated sample covering the frequency<br />

range which incorporates the Larmor peak and the 100kHz<br />

sidebands are shown in figure 7 and they too are almost<br />

identical to the spectra from the fully protonated sample<br />

of figure 4. It therefore appears that the coupling of<br />

protonated rotors persists in the deuterated sample, suggesting<br />

that the packing of the molecules in the solid regarding<br />

the relative positions of protonated and deuterated<br />

methyl groups is random. Deuteration reduces the<br />

number of coupled protonated rotors and slightly distorts<br />

the packing of the molecules, decreasing the magnitude<br />

of the hindering potential of the 750kHz rotor which becomes<br />

780kHz rotor. These spectra indicate clearly that<br />

the anti-crossing effects cannot arise from intra-molecular<br />

methyl group coupling since each molecule has a protonated<br />

and a deuterated methyl group. One new feature of<br />

these spectra which may arise from coupling of protonated<br />

and deuterated rotors is a small, sharp, field independent<br />

peak at 200kHz which is indicated by a V in figure 7.<br />

6 Discussion - Coupled rotating<br />

wavepackets<br />

Tunnelling of small molecular rotors like CH3, CH4 and<br />

NHf is generally single particle in character. Few examples<br />

exist,of coupled tunnelling of rotors (see [8] and references<br />

therein) perhaps the most notable example being of<br />

coupled methyl groups in lithium acetate [9]. In both of<br />

these papers the coupling of methyl groups is mechanical,<br />

being propagated by the modulation of the hindering potential<br />

which has the three-fold symmetry of the individual<br />

methyl groups. The coupled states of the system are a<br />

direct product of the symmetrised eigenstates of each rotor.<br />

Lithium acetate has weakly hindered methyl groups,<br />

the tunnelling spectrum of the coupled system has been<br />

observed using neutron scattering, and it is thought that<br />

such a coupling is unlikely to be observable in strongly hindered<br />

methyl groups, although the computational method<br />

in [8] has been extended to consider such systems.<br />

This kind of coupling is in stark contrast to that seen


90<br />

G)z-100<br />

"FREQUENCY/kHz<br />

Figure 5: Stirred spectra of dimethyl sulphide showing '<br />

restoration of the 100kHz sidebands<br />

Bulletin of Magnetic Resonance<br />

O 250 BOO 7S0 1000 1250<br />

Frequency (kHz]<br />

Figure 6: Frequency spectrum of partially deuterated<br />

dimethyl sulphide mesaured at 4mT<br />

in dimethyl sulphide. An analysis in terms of uncoupled<br />

rotors was pursued in previous sections because first, the<br />

coupling only occurs at those magnetic fields corresponding<br />

to level anti-crossings and secondly, the form of the<br />

spectra indicate that the coupled states are not simple<br />

products, with three-fold symmetry, of the eigenstates of<br />

the individual rotors. Furthermore, these methyl groups<br />

are very strongly hindered in comparison with the methyl<br />

groups in lithium acetate.<br />

It is noted from figure 2 that at the level anti-crossings<br />

there is a matching of the frequencies associated with<br />

the rotational and magnetic evolution of the states of<br />

the 100kHz and 750kHz rotors. If the stable states of<br />

the methyl groups at these magnetic fields are rotating<br />

wavepackets then the frequency of evolution of the spins<br />

on each proton site is modulated by the rotation in a way<br />

that depends upon the composition bf the wavepacket.<br />

For example a coherent mixture of A and Ea states with<br />

the same spin quantum number is a partially localised<br />

wavepacket, which precesses clockwise at the tunnel frequency<br />

and consequently modulates the longitudinal component<br />

of the spins at this frequency. A similar mixed<br />

symmetry state, but composed of states with spin quanne<br />

turn numbers differing by unity modulates the transverse<br />

component of the spin states at each proton site at a frequency<br />

equal to the Larmor frequency plus or minus the<br />

tunnel frequency, depending on whether the rotation and<br />

Larmor precession have the same or opposite senses. Thus,<br />

providing the proton sites of the two methyl groups arft<br />

close enough together, the rotational motions may be<br />

pled by the spin dynamics and angular momentum may<br />

transferred between the groups. This results in an imb<br />

ance of clockwise and anti-clockwise rotation, the unic l u


Vol. 14, No. 1-4 91<br />

ide<br />

O<br />

<<br />

CO<br />

LU<br />

<<br />

Hi<br />

DC<br />

Q<br />

LU<br />

DC<br />

LU<br />

,v<br />

J •<br />

J<br />

v/AV \<br />

V<br />

00<br />

C0z-100<br />

/ ; •<br />

i<br />

ll<br />

V /<br />

\ '<br />

1/<br />

v \<br />

v 21<br />

\ 20<br />

v<br />

\<br />

19<br />

18<br />

16<br />

u<br />

10<br />

8<br />

h-<br />

E<br />

Q<br />

_1<br />

UJ<br />

LL<br />

g<br />

FREQUENCY/kHz<br />

7: Field dependence of the Larmor peak and the<br />

z sidebands in partially deuterated dimethyl sul-<br />

tunnel frequency being replaced by a pair of discrete rotation<br />

frequencies. A reduced rotation frequency is clearly<br />

observed in figure 4.<br />

Taken as evidence not only of coupled methyl groups<br />

but also for rotating wavepackets at 4K, these results are<br />

very significant. The wavepackets do not have three-fold<br />

symmetry which throughout the history of methyl dynamics<br />

has wrongly been regarded as an essential prerequisite<br />

for satisfying the indistinguishabilty of the methyl protons<br />

(e.g. [10]). Each basis function \xyz > accomodates proton<br />

indistinguishability since each proton has an equal amplitude<br />

at each proton site and therefore any combinations<br />

of these functions, ranging from delocalised states which<br />

ressemble spin symmetry species to completely localised<br />

functions, satisfy proton indistinguishability requirements.<br />

7 Acknowledgements<br />

The authors are grateful to the B.P. Venture Research<br />

Unit for supporting this work. I.B.I.T. would like to express<br />

his gratitude to the Sudanese government for his<br />

fellowship. We would also like to thank Dr D.K. Knight<br />

of the Chemistry Department for preparing the deuterated<br />

sample.<br />

References<br />

[1] S. Clough, A.J. Horsewill, M.R.Johnson and<br />

I.B.I.Tomsah (1992) submitted to Molec. Phys.<br />

[2] P.J. McDonald, G.J. Barker, S. Clough, R.M. Green<br />

and A.J. Horsewill (1986) Molec. Phys. 57,901<br />

[3] S. Clough, A. Heidemann, A.J. Horsewill, A.J. Lewis<br />

and M.N.J. Paley (1982)J. Phys. C 15,2495<br />

[4] S. Clough, A.J. Horsewill, P.J. McDonald and F.O.<br />

Zelaya (1985)Phys.Rev. Lett. 55,1794<br />

[5] C. Connor, A. Chang and A. Pines (1986)Rev. Sci.<br />

Instrum. 61,1059<br />

[6] M.J. Barlow, S. Clough, P.A. Debenham and A.J.<br />

Horsewill (1992)J. Phys. C 4,4165<br />

[7] L.Pierce and M. Hayashi (1960)J. Chem. Phys. 35,479<br />

[8] W. Hausler and A. Huller (1985) Z. Phys. B 59,177<br />

[9] S. Clough, A. Heidemann, A.J. Horsewill and M.N.J.<br />

Paley (1984) Z. Phys. B 55,1<br />

[10] J.H. Freed (1965)J. Chem. Phys. 43,1710


92 Bulletin of Magnetic Resonance<br />

INTRODUCTION<br />

NMR RELAXATION STUDIES OF MICRODYNAMICS IN<br />

CHLOROALUMINATE MELTS<br />

Pamela A. Shaw, W. Robert Carper, Charles E. Keller<br />

Room temperature molten salts consisting of mixtures of<br />

A1C13 and l-ethyl-3-methylimidazolium chloride (MEIC1),<br />

are of interest as aprotic solvents for studying a wide range<br />

of both organic and inorganic compounds [1-7]. These<br />

chloroaluminate molten salts possess considerable potential<br />

as battery electrolytes and various types of electrochemical<br />

agents [8-10].<br />

The composition of a chloroaluminate melt has a<br />

considerable effect on its physical properties. The<br />

variations in physical properties of the melt are due to a<br />

combination of factors including ion-ion interactions [4],<br />

and Lewis acid-base properties. Chloroaluminate melts<br />

with A1C13 present in excess (mole fraction, N, of A1C13 ><br />

0.5) are termed acidic with A1C14' and A12C17" the<br />

predominant anions.<br />

The use of NMR relaxation methods provides useful<br />

information about the dynamics and structure of various<br />

chemical systems and chloroaluminate systems in<br />

particular. In a previous work[ll], 13 C NMR relaxation<br />

measurements were used to investigate the motion and<br />

interactions of the MEI cation. The results indicate that<br />

A1C14' in a Na + 0.22MEI + 0,7gAlCl4" melt forms a complex by<br />

interacting with the C-2, C-4 and C-5 hydrogens on the<br />

MEI + ring. This investigation was followed by studies<br />

[12,13] in which the Dual Spin Probe method [14]<br />

supported the existence of MEI(AlCl4)n


Vol. 14, No. 1-4 93<br />

The quadrupole coupling constant, QCC, is given by:<br />

QCC = [e 2 Qq/h] (4)<br />

The DSP method has been applied to chloroaluminate<br />

melts[12,13] and has provided evidence that the ring<br />

hydrogens of MEI + interact with the tetrachloroaluminate<br />

anion. The existence of these complexes has been<br />

supported by linear plots of 13 C dipolar relaxation<br />

rates(R, dd ) vs. quadrupolar 27 A1 relaxation rates(R,) that<br />

pass through the origin as predicted by equation (5):<br />

where a = [3TT 2 /10][(2I + 3)/I 2 (2I -<br />

QCC = x-<br />

= R,( 27 Al)/aX 2<br />

(5)<br />

+ (z 2 /3)], and<br />

In this study, the DSP method is applied to melts<br />

containing MEIC1, A1C13 and EtAlCl2. The inclusion of<br />

EtAlCl2 provides a "baseline" as there is a covalent bond<br />

between the ethyl group and aluminum in EtAlCl2. The<br />

existence of covalent bonding(or complexation) between<br />

quadrupolar and dipolar nuclei in a molecule results in a<br />

linear plot of eqn. (5) that passes through the origin. In<br />

the MEICl-EtAlCl2 melts reported herein, we observe a<br />

linear plot of eqn (5) that passes through the origin when<br />

applied to the terminal CH3 carbon in EtAlCl2 and one of<br />

the peaks in the 27 A1 NMR of the melts.<br />

EXPERIMENTAL<br />

Materials<br />

The l-ethyl-3-methylimidazolium chloride (MEIC1) and<br />

. chloroaluminate molten salts were prepared as described<br />

. previously [1]. Ethylaluminum dichloride (EtAlCy was<br />

•^.obtained from Aldrich. All materials were stored under<br />

ft^hydrous helium gas atmosphere in a dry box. All<br />

molten salt preparations and manipulations were performed<br />

• the dry box. Samples were loaded into 5 mm sample<br />

s > capped in the dry box, removed, and sealed<br />

Jiately with a torch.<br />

Measurements<br />

and 27 A1 NMR spectra were recorded on a Varian<br />

300- spectrometer at 75.43 or 78.15 MHz.<br />

measurements were calibrated against<br />

or ethylene glycol and are accurate to within<br />

§£Pulse widths(90°) were typically 8.6 (75.43 MHz)<br />

1^(78.15 MHz) jts. Longitudinal relaxation times<br />

ured by the the inversion-recovery method<br />

(180°-r-90°-T) with T> 10T,. At least 12 delay times(r)<br />

were used and the results fitted to a three parameter<br />

exponential. NOE measurements were made using the<br />

gated decoupler method[18]. It is likely that the error in<br />

the NOE measurements is in the 5-10% range[18].<br />

RESULTS AND DISCUSSION<br />

The ability of both A1C13 and EtAlCl2 to form C2H<br />

dimers[19,20] led us to examine the 27 A1 spectra of: (1)<br />

neat EtAlCl2, (2) mixtures of MEICl-EtAlCl2 and (3)<br />

ternary melts (N = AlCl3/MEICl/EtAlCl2)[21]. The neat<br />

EtAlCl2 27 A1 NMR spectrum contains two peaks [21].<br />

Peak 1 is a broad downfield peak that domi-nates the<br />

spectrum. The second peak (upfield) overlaps peak 1 and<br />

is only a fraction of peak 1 in total peak area. Peak 2<br />

collapses into peak 1 as the temperature is lowered from<br />

60 to 25°C. These two aluminum sites are consistent with<br />

the extent of monomer-dimer formation in liquid<br />

EtAlCl2[21].<br />

The MEICI-EtAlCl2 (N = 0.5/0.5) melt 27 A1 NMR<br />

spectrum also has two peaks. In this case, peak<br />

1 (downfield) is very broad while peak 2 is very sharp, and<br />

has a low peak area. Peak 2 increases slightly in area and<br />

peak 1 broadens as the temperature is lowered from 70 to<br />

0°C. We have previously[21] made the tentative assignments<br />

of EtAlCl3- for peak l(downfield) and EtjAljClj" for<br />

peak 2.<br />

0.25<br />

0.20<br />

0.15<br />

O<br />

a.<br />

O o.io<br />

O CO<br />

0.05<br />

0.00<br />

EtAia<br />

10 20 30 40<br />

27AI R1<br />

Fig. 1. °C Dipolar Rl's vs "AI Rl's(25 to 70°C) for AI<br />

peak 1 (127-131 ppm from A1(H2O)6 3+ ).


94<br />

In this study, we first apply the DSP method to the CH3<br />

carbon in EtAlCl2 and 27 A1 NMR peaks 1 and 2 from<br />

several melt combinations and neat EtAlCl2. Fig. 1<br />

contains the results for 27 A1 peak 1 (downfield) and Fig. 2<br />

contains the results for 27 A1 peak 2. The fact that both<br />

plots are linear and pass through the origin, indicate that:<br />

(1) the DSP method is appropriate for these systems and<br />

(2) the species associated with each peak contains EtAlCl2.<br />

Furthermore, the slopes of these lines can be used to<br />

cn<br />

DIPOL<br />

13C<br />

. u<br />

0 .20<br />

0 .15<br />

0 .10<br />

0 .05<br />

n nn<br />

| EtAlClj /<br />

r<br />

I J .25/.40/.35 /<br />

/<br />

y<br />

• I * /<br />

t 1 A35/.40/.25<br />

i ; /<br />

i/<br />

0 240 480 720 960 1200<br />

27 Al R1<br />

Fig. 2. "C Dipolar Rl's vs 27 A1 Rl's(25 to 70°C) for Al<br />

peak 2 (102.5-103.0 ppm from Al(H2O)6 3+ ).<br />

calculate the relative quadrupole coupling constants for the<br />

EtAlCl2 -containing species in solution. The QCC values<br />

obtained from Fig. l(Al peak 1) are 171, 119, 106 and 93<br />

MHz for the (.5/.5), (.35/.40/.25), (.25/.40/.35) melts and<br />

neat EtAlCl2, respectively. The QCC values obtained from<br />

Fig. 2(A1 peak 2) are 6.9, 20, 11 and 93 MHz for the<br />

(.5/.5), (.35/.40/.25), (.25/.40/.35) melts and neat<br />

EtAlCl2(repeated).<br />

Results of the Dual Spin Probe method (eqn. [5]) applied<br />

to the (.5/.5), (.35/.40/.25) and (.25/.40/.35) melts<br />

indicate interactions between the Al-containing species in<br />

peak 2(102.5-103.0 ppm relative to A1(H2O)6 3+ ) and both<br />

the NCH3 and ethyl terminal CH3 groups of MEI + . Fig.<br />

3 contains the plots for the NCH3 group in each melt and<br />

Fig. 4 contains data for the terminal CH3 on the MEI ethyl<br />

group.<br />

cc<br />

0.50<br />

0.40<br />

o.oo<br />

.5/.5<br />

Bulletin of Magnetic Resonance<br />

• / .35/.40/.<br />

40/.25<br />

0 64 128 192 255 320<br />

27AI R1<br />

Fig. 3. 13 C Dipolar Rl's vs. 27 A1 Rl's(25 - 70 C) for<br />

NCH3 carbon vs Al peak 2(25 - 70°C).<br />

0.55<br />

0.44<br />

T .5/.5<br />

» •<br />

I / .35/.40/.: 40/.25<br />

64 128 192 256 320<br />

27AI R1<br />

Fig. 4. l3 C Dipolar Rl's for ethyl CH3 carbon vs 27 A1<br />

Rl's(25 - 70°C) for Al peak 2.<br />

The QCC's obtained from the slopes in Fig. 3(MEI<br />

NCH3) are 1.7, 2.3 and 4.4 MHz for the (.5/.5),<br />

(.35/.40/.25) and (.25/.40/.35) melts. The QCC's froffl|


Vol. 14, No. 1-4 95<br />

Fig. 4(terminal CH3 on the MEI ethyl group) are 1.6, 6.9<br />

and 1.3 MHz for the (.5/.5), (.35/.40/.25) and<br />

(.25/.40/.35) melts.<br />

Finally, there is no correlation between the ring hydrogen<br />

dipolar Rl's and any of the 27 A1 peak Rl's. This result is<br />

directly opposite to that found in MEIC1-A1C13 systems<br />

[11,12].<br />

CONCLUSIONS<br />

Application of the DSP probe method to these mixed<br />

melt systems indicates a lack of complexation between the<br />

ring hydrogens of MEI + and any of these aluminum<br />

containing species. These and previous results[21] suggest<br />

that the formation of various charged dimers containing<br />

EtAlCl2 takes precedence over the formation of complexes<br />

between EtAlCl3" and the MEI + ring hydrogens. However,<br />

there is evidence of interaction between the various Alcontaining<br />

species and the CH3 groups(NCH3 and terminal<br />

CH3 in the ethyl group) of MEI + in these melts.<br />

ACKNOWLEDGMENTS<br />

This work was partially supported by a National<br />

Research Council and Summer Faculty Research<br />

Associateship to W. R. C. and a summer Graduate<br />

Fellowship to P. A. S.<br />

REFERENCES<br />

[1] J. S. Wilkes, J. A. Levisky, R. A. Wilson and C. L.<br />

, Hussey, Inorg. Chem., 21 1263 (1982).<br />

(2| J. S. Wilkes, J. S. Frye and G. F. Reynolds, Inorg.<br />

.•Chem., 22(1983)3870.<br />

jfI3| A. A. Fannin, L. A. King, J. A. Levisky and J. S.<br />

'"Vilkes.J. Phys. Chem., 55(1984)2609.<br />

| 4 1 A. A. Fannin, D. A. Floreani, L. A. King, J. S.<br />

toders, B. J. Piersma, D. J. Stech, R. L. Vaughn, J. S.<br />

jfllkes and J. L. Williams, J. Phys. Chem., 88 (1984)<br />

514.<br />

M. Dieter, C. J. Dymek, N. E. Heimer, J. W.<br />

and J. s. Wilkes, J. Amer. Chem. Soc, 110<br />

2722.<br />

J - D ymek and J. J. P. Stewart, Inorg. Chem., 28<br />

,1472.<br />

[7] J. A. Boon, J. A. Levisky, J. L. Pflug and J. S.<br />

Wilkes, J. Org. Chem., 51 (1986) 480.<br />

[8] C. J. Dymek, J. L. Williams, D. J. Groeger and J. J.<br />

Auborn, J. Electrochem. Soc, 131 (1989) 2887.<br />

[9] C. J. Dymek and L. A. King, J. Electrochem. Soc.,<br />

132 (1985) 1375.<br />

[10] C. L. Hussey, T. B. Scheffler, J. S. Wilkes and A.<br />

A. Fannin, J. Electrochem. Soc, 133 (1986) 1389.<br />

[11] W. R. Carper, J. L. Pflug, A. M. Elias and J. S.<br />

Wilkes, J. Phys. Chem. 96 (1992) 3828.<br />

[12] W. R. Carper, J. L. Pflug and J. S. Wilkes,<br />

Inorganica Chimica Acta 193 (1992) 201.<br />

[13] W. R. Carper, J. L. Pflug and J. S. Wilkes,<br />

Inorganica Chimica Acta (in press).<br />

[14] J. J. Dechter and U. Henriksson, J. Magn. Res., 48<br />

(1982) 503.<br />

[15] A. Abragam, "Principles of Nuclear Magnetism",<br />

Oxford University Press, Oxford (1961).<br />

[16] K. F. Kuhlmann and D. M. Grant, J. Amer. Chem.<br />

Soc, 90 (1968) 7355.<br />

[17] B. Lindman and S. Forsen, in "NMR Basic Principles<br />

and Progress," P. Diehl, E. Fluck and R. Kosfeld,<br />

Editors, Vol. 12, p. 22, Springer-Verlag, New York<br />

(1976).<br />

[18] D. Neuhaus and M. Williamson, "The Nuclear<br />

Overhauser Effect in Structural and Conformational<br />

Analysis", VCH Publishers, New York (1989).<br />

[19] J. Weidlein, J. Organomet. Chem., 27(1969)213.<br />

[20] B. Gilbert, Y. Chauvin and I. Guibard, Vib.<br />

Spectros., 1 (1991)299.<br />

[21] W. R. Carper, C. E. Keller, P. A. Shaw, M. P. and<br />

J. S. Wilkes, in "Eighth International Symposium on<br />

Molten Salts", Electrochem. Soc, New York (in press).


96<br />

Structure and Dynamics of a<br />

Membrane Bound<br />

Polypeptide<br />

T. A. Cross, R.R.Ketchem, W. Hu, K.-C. Lee,<br />

N.D.Lazo & C.L. North<br />

Department of Chemistry &<br />

Institute of Molecular Biophysics<br />

Florida State University, Tallahassee, FL 32306-3006<br />

INTRODUCTION:<br />

Orientational constraints can be used to build-up<br />

three dimensional structures of biological macromolecules<br />

in much the same way that distance constraints are used<br />

today. In an anisotropic environment where molecular<br />

motions do not average NMR signals to their isotropic<br />

average the resonant frequencies become orientation<br />

dependent. Consequently, chemical shift frequencies, dipolar<br />

couplings, and quadrupolar interactions are all dependent on<br />

the specific orientation of the interaction tensor with respect<br />

to the magnetic field. In unoriented samples this results in<br />

broad spectral lineshapes, which arc useful in their own<br />

right, but if the samples are aligned so that all molecules<br />

have the same orientation with respect to the magnetic field,<br />

then sharp line spectra can be obtained that reflect the<br />

orientation dependence of the spin interactions.<br />

to interpret these frequencies for orientational<br />

constraints two pieces of information are critical. The first<br />

is a refined knowledge of the static tensor element<br />

magnitudes and orientation with respect to the molecular<br />

frame. The tensor element magnitudes represent the unique<br />

frequencies of the orthogonal axes of the interaction<br />

ellipsoid. In other words, the magnitudes represent the<br />

resonant frequency when the individual axes are aligned with<br />

the magnetic field. These values can, in many cases, be<br />

readily obtained from spectra of unoriented samples. The<br />

second critical characterization is a knowledge of the<br />

molecular motions that result in averaging both the tensor<br />

element magnitudes and orientation. Consequently, a<br />

detailed picture of dynamics is needed to characterize the axis<br />

about which motions are occurring, the amplitude of the<br />

motions and whether the motions are continuous or<br />

discontinuous, such as a flip of an aromatic ring. In fact, it<br />

Bulletin of Magnetic Resonance<br />

is this separation of dynamics and structure that represents<br />

the most difficult challenge for the biological solid state<br />

NMR spectroscopist. By using low temperature (120 K)<br />

experiments, torsional motions are essentially eliminated<br />

except in specific cases such as methyl group three site<br />

jumps. Once both the static tensors and motional averaging<br />

of the tensors are characterized it is possible to achieve very<br />

high resolution quantitative constraints. Here we present<br />

both an overview of the structural constraints and the<br />

dynamic characterizations achieved for the channel forming<br />

polypeptide, gramicidin A in a fully hydrated lipid bilayer.<br />

Gramicidin A is a hydrophobic polypeptide of 15<br />

amino acid residues with both end groups blocked so that<br />

there are no formal charges. As a dimer it forms a helical<br />

channel with a 4A pore that accommodates a single file of<br />

water molecules and cations that are stripped of all but two<br />

waters in the primary hydration sphere. The peptide linkages<br />

that line the channel are thought to help solvate the cations<br />

during transport by rotating the carbonyl oxygens toward<br />

the channel axis. While crystal structures of gramicidin A<br />

have been achieved in organic solvents (Langs, 1988;<br />

Wallace & Ravikumar, 1988; Langs et al., 1991) no crystal<br />

structure of the channel conformation has been achieved.<br />

However, much is known about the channel conformation.<br />

The backbone folding motif is a (5-sheet type of structure<br />

that has been wound into a helix (Urry, 1971). This is<br />

possible because the amino acids of gramicidin alternate in<br />

stereochemistry between D and L configurations. The helix<br />

sense has been determined from the orientational constraints<br />

of the 15 N amide sites (Nicholson & Cross, 1989).<br />

Recently, the first backbone torsion angles for the channel<br />

conformation have been determined from orientational<br />

constraints (Teng et al., 1991). Complementing this work


Vol. 14, No. 1-4 97<br />

f<br />

13 C Chemical Shift<br />

15 N Chemical Shift<br />

15 N- 13 C Dipolar<br />

^NMH Dipolar<br />

14 N- 13 C Dipolar<br />

15 N- 15 N Dipolar<br />

13 C- 13 C Dipolar<br />

2 H Quadrupolar<br />

Figure 1: A model of the gramicidin channel dimer showing the interaction tensors that have been studied. The model<br />

structure is that of Lomize et al., 1992. Each tensor is represented by a orthogonal set of three unit vectors that have<br />

been oriented correctly for the specific site of interest. All backbone and tryptophan indole I5 N chemical shift and I5 N-<br />

! H dipolar tensors have been studied with the exception of the ethanolamine blocking group. About half of the 15 N-<br />

13 Cj dipolar interactions in the backbone have been studied, four of the backbone carbonyl l ^C chemical shift tensors<br />

and in so doing the 14 N electric field gradient tensor has been characterized. Four sidechains have been deuterated and the<br />

quadrupole splittings obtained. 15 N spin diffusion has been observed between selectively labeled sites. To achieve this<br />

data approximately 50 separate syntheses of gramicidin A have been performed (Fields et al., 1989).


98<br />

are solution NMR studies of gramicidin in SDS micelles<br />

that have also shown the same backbone folding motif<br />

(Arseniev & Barsukov, 1986; Lomize et al., 1992).<br />

RESULTS & DISCUSSION:<br />

Fig. 1 illustrates the large number of nuclear spin<br />

interactions that have been studied in gramicidin. For each<br />

interaction the tensor represented by an orthogonal set of<br />

unit vectors is placed on each atomic site that has been<br />

studied. More than 40 different isotopically labeled (Fields<br />

et al., 1989) gramicidins have been synthesized and more<br />

than 100 different structural constraints have been<br />

developed. The structural task is one of determining the<br />

torsion angles. If it is assumed that the peptide linkages are<br />

planar then there are two backbone torsion angles for each<br />

amino acid residue, one for valine sidechains and two each<br />

for leucine and tryptophan sidechains. There are two<br />

additional torsion angles in the ethanolamine blocking<br />

Val<br />

Ala<br />

Leu,<br />

End View<br />

Side View<br />

Figure 2: A partial structure for the gramicidin channel<br />

dimer, experimentally derived. The N-terminal four peptide<br />

planes have been oriented with respect to the bilayer normal<br />

through a combination of ^N-'H, 15 N- 13 Ci dipolar and<br />

15 N chemical shift interactions in uniformly aligned bilayer<br />

samples. The experimentally verified symmetry of the two<br />

monomers has permitted the docking of these two partial<br />

structures to form a turn of the helix that supports previous<br />

models with 6.3 residues per turn of the channel helix. The<br />

amino acid residues have been identified for one of the<br />

monomers and the amide protons and oxygens labeled.<br />

Bulletin of Magnetic Resonance<br />

group of the carboxy terminus. Therefore, the structural<br />

problem comes down to finding the solution for 52 torsion<br />

angles.<br />

Fig. 2 shows the N-terminal / N-terminal<br />

backbone junction for the channel as determined by<br />

orientational constraints. For each peptide linkage plane the<br />

* S N - !Hand 15 N - 13 Ci dipolar interactions, which have<br />

their unique static tensor elements directed along the<br />

internuclear vectors, were obtained. These two vectors define<br />

the orientation of the plane with some ambiguity that is<br />

minimized by an interpretation of the 15 N chemical shift for<br />

each plane. The orientation of the chemical shift tensor<br />

elements with respect to the molecular frame has been<br />

determined for each of these i 5 N sites (Teng & Cross, 1989;<br />

W. Mai, W. Hu, C. Wang and T.A. Cross- unpublished<br />

results).<br />

This structure clearly shows the right-handed<br />

helical sense and the ^-helical class of torsion angles, in<br />

that the orientation of adjacent peptide planes alternates<br />

between parallel and antiparallel with respect to the channel<br />

axis. Furthermore, like many of the computationally refined<br />

structures (e.g. Roux and Karplus, 1988; Chiu et al., 1991)<br />

and unlike the original structural model (Urry, 1971) many<br />

of the carbonyl oxygens are rotated in toward the channel<br />

axis. The prime exception to this observation is the formyl<br />

oxygen of the amino terminal blocking group. Here we<br />

suspect that this is an artifact of the assumption that this<br />

peptide linkage is planar. Preliminary evidence suggests that<br />

the © torsion angle for this plane is far from 180°, the<br />

value for normal trans-planar linkages.<br />

Fig. 3 shows that we have also studied the<br />

sidechains of gramicidin A. In (A) the chemical shift powder<br />

pattern of a dry powder sample of 15 N indole labeled<br />

tryptophan is shown. The chemical shift tensor elements<br />

agree fairly well with those from the dry powder sample of<br />

gramicidin shown in (B). However the tensor elements for<br />

the fast frozen hydrated lipid bilayer preparation of<br />

gramicidin is very different. Both o"22 and a33 tensor<br />

elements differ by 10 ppm from the dry samples. This<br />

sample was frozen in liquid propane to avoid distortions in<br />

the lipid bilayer which occur when a bilayer preparation is<br />

slowly cooled through its phase transition temperature. The<br />

sample was then transferred to liquid nitrogen and from there<br />

into our low temperature NMR probe. It is unlikely that<br />

the former two samples have the nitrogen bound hydrogen<br />

involved in a hydrogen bond. However, there is some<br />

electrophysiological evidence that these hydrogens for each<br />

of the tryptophan rings in the gramicidin channel state are<br />

hydrogen bonded to the carbonyl oxygens of the ester<br />

linked lipids. Consequently, it may be that the substantial<br />

changes seen in the chemical shift tensor element<br />

magnitudes reflect hydrogen bonding of the indole group.<br />

Further indirect evidence, given below, for hydrogen<br />

bonding comes from the orientation of the indole N-H with


Vol. 14, No. 1-4<br />

respect to the bilayer surface.<br />

200 100<br />

ppm<br />

Figure 3: Powder pattern spectra of 1 5 N labeled indole. The<br />

spectra were obtained at 40 MHz for 15 Nona spectrometer<br />

that has been home built around a Ghemagnetics data<br />

acquisition system and an Oxford 400/89 superconducting<br />

magnet. A] Dry powder sample of the amino acid,<br />

tryptophan obtained at room temperature. Spectral<br />

simulation yields tensor elements: (?i i = 35, o~22 = 104, and<br />

C33 = 158 ppm. B] Dry powder of 15 N-Trp9 gramicidin A:<br />

On = 36, 022 = 106, and a33 = 161 ppm. C] Fast frozen<br />

sample of 1S N-Trp9 gramicidin A in fully hydrated lipid<br />

bilayers: a^ = 36, G22 = 116, and a33 = 171 ppm.<br />

The 2 H quadrupolar spectrum (Fig. 4) of d5-Trpj r<br />

gA provides an example of the spectroscopic data that has<br />

been used for structural constraints. The arrows point to a<br />

linewidth of 1.4 kHz that represents an uncertainty for the<br />

0<br />

orientation of ±0.2°. Not only does this spectrum describe a<br />

single conformation for this tryptophan ring, but it also<br />

dictates that the orientation for all of the gramicidin<br />

molecules in the sample is remarkably uniform (Moll &<br />

Cross, 1990). For this sample there are five deuterons on<br />

the tryptophan ring and five quadrupole splittings are<br />

observed. The analysis of this data combined with the 15 N<br />

chemical shift and 15 N-iH dipolar interaction for the indole<br />

nitrogen (W. Hu, K.-C. Lee and T.A. Cross - unpublished<br />

results) yields two possible orientations for this ring. Each<br />

of these structures has the same orientation with respect to<br />

the channel axis and the orientation is similar to that shown<br />

in Fig. 1 for Trpn, where the N-H points towards the<br />

bilayer surface. In recent computational (Meulendijks et al.,<br />

1989) and electrophysiological (O'Connell et al., 1990)<br />

studies it has been suggested that the indoles of gramicidin<br />

A are hydrogen bonded to the carbonyl oxygens of the esterlinked<br />

lipids. These results indicate that such hydrogen<br />

bonding may be present and this may be one of the prime<br />

reasons why this conformation is present in lipid bilayers<br />

rather than the double-helical structures that dominate<br />

organic solutions (Veatch and Blout, 1974; Zhang et al.,<br />

1992). Furthermore, the tryptophan dipole moment is<br />

oriented primarily along the channel axis, rather than radial<br />

to this axis. This has significant implications for cation<br />

transport by gramicidin A.<br />

Figure 4: 2 H NMR spectrum of ds-Trpj ] gramicidin A in an<br />

oriented lipid bilayer preparation. The arrows indicate a<br />

resonance linewidth of 1.4 kHz. The assignment of two of<br />

these quadrupole splittings is clear from a knowledge of<br />

other spin interactions in this ring system: £3 = 19 2; and<br />

T|2 = 9 9 kHz.<br />

Detailed dynamics have come from a combination<br />

of lineshape simulation and relaxation studies. Below the<br />

gel to liquid crystalline phase transition temperature of the<br />

bilayers all large amplitude motions cease. Both the global<br />

motion about the bilayer normal and the local backbone<br />

-80<br />

99


100<br />

motions become slower than the chemical shift frequency<br />

scale (Nicholson et al., 1989; 1991). This is because the<br />

backbone motions are dependent upon motion of the C a-Cp<br />

axis and hence, if the sidechain conformations are frozen in<br />

the gel phase the backbone motions will be greatly<br />

impeded. When oriented samples are lowered through the<br />

phase transition temperature the broadened resonance<br />

represents a considerable orientational dispersion. Instead of<br />

the backbone having a single conformation in this static<br />

environment a range of conformational substates has been<br />

trapped (Frauenfelder et al., 1988; Nicholson et al., 1989).<br />

This set of substates represents the range of orientations<br />

over which local motions occur above the phase transition<br />

temperature. From the observed lineshape it has been<br />

possible to determine the orientation of the axis about<br />

which the local backbone motions occur with respect to the<br />

magnetic field (Nicholson et al., 1991). This axis has been<br />

shown to be coincident with the Ca-Ca axis for each<br />

peptide linkage. Since the number of conformational<br />

substates is greater than three, the local motion can be<br />

modeled as a diffusion within a gaussian well. Furthermore,<br />

the amplitude of the motion could be estimated.<br />

Once such a motional model has been<br />

experimentally developed for a specific backbone site in the<br />

gramicidin channel conformation it is possible to interpret<br />

relaxation data for these same i*N sites. Tj relaxation times<br />

were obtained from oriented samples of fully hydrated lipid<br />

bilayers. To fix the frequency of the local motions it has<br />

been necessary to measure the relaxation times at two<br />

different field strengths. In Fig. 5 an analysis of such data is<br />

shown. The global correlation time (tp) and the local<br />

motion correlation time (tj) are variables as well as the<br />

amplitude of the local motion. However, we have a<br />

previous estimate of the motional amplitude (±15°,<br />

Nicholson et al., 1991) as well as the global correlation<br />

time (200 ns, Seelig and Macdonald, 1987). Consequently,<br />

it is possible to achieve a unique solution for the local<br />

motion correlation time (10 ns). A similar correlation time<br />

has been reported for an even numbered site (Leu4) in the<br />

gramicidin channel (North and Cross, 1992)<br />

This is a remarkably slow correlation time for a<br />

motion of a molecular group with nominally such a small<br />

molecular weight. Roux and Karplus (1988) have analyzed<br />

the normal modes in the gramicidin channel and concluded<br />

that fluctuations occur with frequencies of 4.6 to 20 cm'<br />

corresponding to harmonic oscillator periods of 2 to 8 ps,<br />

approximately, a factor of lO^-lO 4 slower than the<br />

determination reported here. Similarly, estimates from<br />

molecular dynamics indicate frequencies for these local<br />

motions that are in a similar range to those of the normal<br />

mode analysis.<br />

The experimental evidence presented here is not<br />

A.<br />

-7.0<br />

Log %. -9.0<br />

B.<br />

Log<br />

c.<br />

-11.0<br />

-7.U<br />

T i " 9 -°<br />

-11.0<br />

-7.0<br />

LogTj -9.0<br />

-11.0<br />

Bulletin of Magnetic Resonance<br />

i!! 1<br />

j||<br />

iii<br />

•>t<br />

ii!<br />

•IJJX,<br />

-7.0 -6.0<br />

.f |||1|lW<br />

iii<br />

HI<br />

Hi<br />

Hi i<br />

IP' i<br />

lit} I<br />

a<br />

° 4.7 T<br />

• 9.4 T<br />

' lll|ltI| ! ll "3J l lJJ;;<br />

° 4.7 T<br />

• 9.4 T<br />

• 1 ' • • r— 1 1 1 , . 1<br />

-7.0 -6.0<br />

iii*<br />

ill<br />

f iii<br />

111<br />

Hi<br />

ill<br />

-7.0 -6.0<br />

i<br />

° 4.7 T<br />

• 9.4 T<br />

Figure 5: T, relaxation times that were obtained from 15 N-<br />

Gly2 gramicidin A in oriented lipid bilayers at 20 and 40<br />

MHz have been analyzed to achieve a solution for local and<br />

global correlation times as well as the local motional<br />

amplitude. A] calculated for 12° rms deviation; B] 15°; and<br />

C] 18°. Because of previous estimates for the amplitude and<br />

global correlation time, it is possible to determine the local<br />

correlation time as 10 ns with an rms amplitude of 15°.<br />

unique for slow motions in polypeptide backbones, in fact<br />

most of the experimental evidence suggests much slower<br />

motions than the computational techniques. Analysis of Tj<br />

and NOEs from the backbone of the filamentous virus fd<br />

suggested correlation times of 1 ns (Cross and Opella,<br />

1982). A similar timescale has been reported for collagen


Vol. 14, No. 1-4 101<br />

(Sarkar et al., 1985). More recently, Cole and Torchia<br />

(1991) have reported local motional frequencies in the<br />

backbone of crystalline staphyloccal nuclease in the ns or<br />

near-ns timescale.<br />

For the gramicidin channel there appear to be two<br />

primary reasons for the discrepancy between the<br />

computational and experimental frequencies. First, it maybe<br />

as suggested by Venkatachalam and Urry (1984) that the<br />

local motions are correlated along the polypeptide backbone.<br />

Computational studies have argued that the extent of such<br />

correlations are limited to nearest neighbors. The second<br />

reason is that the lipid environment may damp the backbone<br />

motions severely. Evidence has already been presented that<br />

the lipid environment can reduce the motional frequencies<br />

below the kHz range when the lipids are in the gel phase.<br />

Above the phase transition temperature it is well known<br />

that the lipid environment damps global motional<br />

frequencies. For instance, the 200 ns global correlation time<br />

for gramicidin (1880 daltons) would be consistent with a<br />

protein of 50,000 daltons or greater in aqueous solution.<br />

Furthermore, evidence of specific gramicidin - lipid<br />

interactions described here suggests an additional mechanism<br />

for damping by the lipid environment.<br />

The transit time for cations to move between<br />

dipeptide carbonyl sites in the polypeptide backbone can be<br />

estimated from kinetic measurements (Anderson, 1983) and<br />

analyses of the energetics (Roux and Karplus, 1991) to be<br />

in the range of 10 ns. Consequently, it is now likely that<br />

there is a correlation between kinetics and the local<br />

dynamics of the polypeptide backbone. This unique<br />

correlation between structure, dynamics and function<br />

emphasizes the importance of pursuing such detailed studies<br />

of polypeptides and proteins in functional native-like<br />

environments.<br />

ACKNOWLEDGEMENTS:<br />

We are indebted to the staff of the FSU NMR<br />

facility, Joseph Vaughn, Richard Rosanske, and Thomas<br />

Gedris for their skillful maintenance, modification and<br />

service of the NMR spectrometers. This effort has been<br />

supported in large part through grant #AI-23007 from the<br />

National Institutes of Health. TAC also gratefully<br />

acknowledges support of the Alfred P. Sloan Foundation for<br />

a Research Fellowship.<br />

REFERENCES:<br />

Anderson, O. S. (1983) Biophys. J. 41:119.<br />

Arseniev, A.S. and Barsukov, V.F. (1986) in Chemistry of<br />

Peptides and Proteins, Vol. 3; W. Voelter.E. Bayer,<br />

Y.A. Ovchinnikov, V.T. Ivanov, Eds. Walter de Gryter &<br />

Co. Berlin, pgs. 127-158.<br />

Chiu, S.W., Nicholson, L.K., Brenneman, M.T., Teng, Q.,<br />

Subramaniam, S., McCammon, J.A., Cross, T.A.,<br />

Jakobsson, E. (1991) Biophys. J. 60:974-978.<br />

Cole, H.B.R. and Torchia, D.A. (1991) Chem. Phys.<br />

158:271.<br />

Cross, T.A. and Opella, S.J. (1982) J. Mol. Biol. 159:543-<br />

549.<br />

Fields, C.G., Fields, G.B., Noble, R.L. and Cross, T.A.<br />

(1989) Int. J. Peptide Protein Res. 33:298-303.<br />

Frauenfelder, H., Parak, F. and Young, R.D. (1988) Annu.<br />

Rev. Biophys. Biophys. Chem. 17:451-479.<br />

Langs, D.A. (1988) Science 241:188-191.<br />

Langs, D.A., Smith, G.D., Courseille, C, Precigoux, G.<br />

and Hospital, M. (1991) Proc. Natl. Acad. Sci. USA<br />

88:5345-5349.<br />

Lomize, A.L., Orechov, V. Yu. and Arseniev, A.S. (1992)<br />

Bioorgan. Khimia 18:182-200.<br />

Meulendijks, G.H.W.M., Sonderkamp, T, Dubois, J.E.,<br />

Nielen, R.J., Kremers, J.A. and Buck, H.M. (1989)<br />

Biochim. Biophys. Acta 979:321-330.<br />

Moll, F. Ill, and Cross, T.A. (1990) Biophys. J. 57:351-<br />

362.<br />

Nicholson, L.K. and Cross, T.A. (1989) Biochemistry<br />

28:9379-9385.<br />

Nicholson, L.K. LoGrasso, P.V. and Cross, T.A. (1989) J.<br />

Am. Chem. Soc. 111:400-401.<br />

Nicholson, L.K., Teng, Q. and Cross, T.A. (1991) J. Mol.<br />

Biol. 218:621-637.<br />

North, C.,L. and Cross, T.A. (1992) J. Magn. Res. - in<br />

press.<br />

O'Connell, A. M., Koeppe, R. E. II and Anderson, O. S.<br />

(1990) Science 250:1256-1258.<br />

Roux, B. and Karplus, M. (1988) Biophys. J. 53:297-309.<br />

Roux, B. and Karplus, M. (1991) J. Phys. Chem. 95:4856.<br />

Sarkar, S.K., Sullivan, C.E. and Torchia, D.A. (1985)<br />

Biochemistry 24:2348-2354.<br />

Seelig, J. and Macdonald, P.M. (1987) Ace. Chem. Res.<br />

20:221-228.<br />

Teng, Q. and Cross, T.A. (1989) J. Magn. Res. 85:439-<br />

447.<br />

Teng, Q., Nicholson, L.K. and Cross, T.A. (1991) J. Mol.<br />

Biol. 218:607-619.<br />

Urry, D.W. (1971) Proc. Natl. Acad. Sci. USA 68:672-676.<br />

Veatch, W.R. and Blout, E.R. (1974) Biochemistry<br />

13:5257-5264.<br />

Venkatachalam, CM. and Urry, D.W. (1984) J. Comp.<br />

Chem. 5:64.<br />

Wallace, B.A. and Ravikumar, K. (1988) Science 241:182-<br />

187.<br />

Zhang, Z., Pascal, S. and Cross, T.A. (1992) Biochemistry<br />

- in press.


102 Bulletin of Magnetic Resonance<br />

The Role of Metal Ions in<br />

Processes of Conformational<br />

Selection during Ligand-<br />

Macromolecule Interactions<br />

E.Gaggelli, N.Gaggelli, G.Valensin<br />

Department of Chemistry, University of Siena<br />

Pian dei Mantellini 44<br />

Siena 53100, Italy<br />

and<br />

A.Maccotta<br />

Department of Chemistry, University of Basilicata<br />

Via N.Sauro 85<br />

Potenza 85100, Italy<br />

1 Introduction<br />

The interaction with macromolecules<br />

plays a major role in<br />

eliciting the biochemical activity<br />

of relatively small flexible<br />

ligands. Several processes are<br />

involved in such interaction<br />

such that the key and hole<br />

assumption may often fail. One<br />

of the processes not carefully<br />

considered so far is that of<br />

conformational selection, in<br />

which the macromolecule<br />

stabilizes the conformation of<br />

the ligand at the bound site<br />

after a selection among several<br />

conformational arrangements.<br />

The conformation at the bound<br />

state may or may be not connected<br />

to some low-energy<br />

conformation or to the conformation<br />

stabilized at the solid<br />

state. The comprehension of<br />

this process is expected to<br />

provide a valuable aid in<br />

rationale drug design where<br />

synthesized molecules are<br />

sought yielding the same or<br />

even more specific responses<br />

than the natural ligands.<br />

Investigation of this pro-


Vol. 14, No. 1-4 103<br />

cess requires to delineate the<br />

change in conformation when<br />

going from the free to the<br />

bound state and, from this<br />

point of view, NMR is the<br />

technique of choice.<br />

Here we present evidence<br />

that NMR allows to<br />

detect and delineate the<br />

change in conformation experienced<br />

by a flexible ligand,<br />

the dipeptide carnosine (p -<br />

alanyl-L-histidine), when it<br />

binds to the serum protein<br />

albumin. We show also that<br />

the process of conformational<br />

selection is favoured by the<br />

presence of divalent metal<br />

ions, such as Ca(II) and Cu(II),<br />

that stabilize, in the metal<br />

complex, a conformation of the<br />

ligand very close to that assumed<br />

in the bound state.<br />

2 NMR Parameters<br />

The preferred conformation<br />

assumed by the ligand in its<br />

free state in solution can be<br />

easily delineated by measuring<br />

dipolar interaction energies<br />

between pairs of homo- or<br />

hetero-nuclear spin. Since such<br />

interaction terms are functions<br />

of tc/r 6 , distances can be calculated<br />

if the reorientational<br />

dynamics can be characterized,<br />

even in some approximate<br />

way. This last purpose can be<br />

accomplished by measuring<br />

and interpreting the 13 C-NMR<br />

spin-lattice relaxation rates,<br />

that are, in general, determined<br />

by the one bond (r = 1.09<br />

A [1]) 13C-1H dipole-dipole interaction.<br />

Once the motional correlation<br />

time(s) is (are) determined,<br />

relevant geometric<br />

features can be obtained by<br />

one or more of the following<br />

experiments:<br />

a) evaluation of the iH-pH}<br />

n.O.e. if spectral resolution<br />

is not limited and crosscorrelation<br />

effects can be<br />

neglected;<br />

b) measurement of singleand<br />

double-selective l H-<br />

NMR spin-lattice relaxation<br />

rates of selected proton<br />

pairs; these provide a<br />

means of calculating absolute<br />

values of pairwise<br />

dipolar cross-relaxation<br />

terms, Ojj [2-4];<br />

c) measurement of the * 3 C -<br />

pH} n.O.e. upon selective<br />

presaturation of resolved<br />

proton resonances [5];<br />

d) evaluation of relative<br />

values of cross-relaxation<br />

terms from intensities of<br />

cross peaks in 2D NOESY<br />

maps [6].<br />

All these methods are<br />

very efficient in providing the<br />

desired information on the<br />

preferred conformation in<br />

solution of any 'NMR visible'<br />

ligand and, eventually, the<br />

change in conformation caused<br />

by the presence of metal ions.<br />

In this last case, if the metal is<br />

paramagnetic, a great piece of<br />

structural information is<br />

gained by investigating the<br />

paramagnetic effects on nu-


11 ;<br />

104 Bulletin of Magnetic Resonance<br />

clear relaxation rates and chemical<br />

shifts [7,8] or on 2D<br />

spectra [9,10].<br />

In order to investigate<br />

the eventual change in the<br />

conformation of the ligand<br />

when it binds to a macromolecule<br />

either in the presence or<br />

in the absence of metal ions,<br />

the previously outlined NMR<br />

methods, at least not all of<br />

them, are not as efficient as in<br />

the free state. The concentration<br />

of the macromolecule is a<br />

limiting factor, since it must be<br />

kept quite small if spectral<br />

distortion is to be avoided. As<br />

a consequence exchange of the<br />

ligand between the free (bulk)<br />

and the bound state must be<br />

taken into consideration, yielding:<br />

Pobs = xfPf + xbPb<br />

where P is any observed NMR<br />

parameter, f and b refer to the<br />

free and bound states and the<br />

x's are molar fractions. It<br />

follows that the change in P<br />

from the free to the bound<br />

state is expressed by:<br />

AP = xbPb<br />

where Xb is usually of the<br />

order 0.01-0.1. The consequence<br />

is that spin-lattice<br />

relaxation rates and chemical<br />

shifts are no longer suitable<br />

parameters for delineation of<br />

geometric features of the<br />

bound state, unless a paramagnetic<br />

centre, either intrinsic<br />

or extrinsic, is present. One is<br />

then left with measurements<br />

of ID or 2D transferred n.O.e.<br />

or, which we prefer, of singleand<br />

double-selective proton<br />

spin-lattice relaxation rates. In<br />

fact, in absence of spectral<br />

distortions, the same measurements<br />

can be easily accomplished<br />

for the free ligand as<br />

well as in the system where<br />

the ligand is exchanging<br />

between the two states, and<br />

absolute values of the crossrelaxation<br />

rate can be separately<br />

obtained for the free and<br />

the bound ligand.<br />

3 Free Carnosine<br />

The relevant features of<br />

carnosine in water solution can<br />

be summarized as follows:<br />

a) reorientational dynamics<br />

can be interpreted in<br />

terms of a principal correlation<br />

time (Tc = 58 ps)<br />

describing reorientation<br />

around a molecular axis<br />

passing through the imidazole<br />

ring, coupled with<br />

segmental motion of the<br />

amino-terminal moiety<br />

and librational motion of<br />

the ring (Tg = 10 ps);<br />

b) predominance of the g~<br />

rotamer around the C6-C7<br />

bond (CH2-CH segment of<br />

the histidyl residue);<br />

c) folding of the fi-alanyl<br />

moiety towards the imidazole<br />

ring.


Vol. 14, No. 1-4 105<br />

4 Calcium Complex<br />

Calcium forms two complex<br />

speies with carnosine in solution:<br />

a 1:1 complex where the<br />

carbonyl and carboxyl oxygens<br />

are the metal binding atoms<br />

and a dimeric complex where<br />

the two carbonyl and one carboxyl<br />

oxygens and the imidazole<br />

nitrogen are the four<br />

coordinated atoms. The overall<br />

dissociaton constant of the<br />

complexes is Kd = 0.04 mol<br />

dm"3. In the monomer complex<br />

the dipeptide retains the<br />

conformation detected by NMR<br />

as the 'preferred' one in the<br />

free state in solution. In the<br />

dimer species extensive intermolecular<br />

interactions are<br />

favoured and the conformation<br />

of the dipeptide is less folded<br />

than in the free state or in the<br />

1:1 complex.<br />

It is concluded that calcium<br />

ions stabilize a particular<br />

geometric arrangement that is<br />

itself representing the 'preferred'<br />

conformation in solution,<br />

as it raises from motional averaging<br />

of all the particular<br />

conformations assumed by the<br />

flexible peptide.<br />

5 Interaction with HSA<br />

The interaction of carnosine<br />

with human serum albumin<br />

(HSA) can be detected and<br />

delineated by measuring<br />

selective and double-selective<br />

proton spin-lattice relaxation<br />

rates of carnosine protons in<br />

the presence of low molar<br />

fractions of the protein [11,12].<br />

Detection of binding is allowed<br />

by appreciable relaxation rate<br />

enhancements of selective<br />

relaxation rates of the imidazole<br />

protons, as well as of the<br />

His Ha:<br />

AR sel = pbRb sel<br />

Even at very low fractions of<br />

bound carnosine the selective<br />

relaxation rate at the bound<br />

site is so fast that the observed<br />

rate undergoes enhancements<br />

as high as 50-100 %. The effect<br />

allows also a titration of the<br />

binding process, yielding an<br />

apparent dissociation constant<br />

of 2.5xlO- 4 mol dm- 3 .<br />

The observed enhancements<br />

are consistent not only<br />

with a very tight binding of<br />

the whole peptide molecule to<br />

the protein but also with<br />

occurrence of dipolar interaction<br />

between ligand* and protein<br />

protons, although there is<br />

no possibility of obtaining<br />

quantitative estimations of<br />

such interactions.<br />

More information can be<br />

obtained by measuring the<br />

double-selective proton spinlattice<br />

relaxation rates within<br />

the His Ha-Hpi-H(32 moiety. As<br />

in the free state in solution,<br />

such measurements yield the<br />

dipolar cross-relaxation rate<br />

between the excited protons,<br />

e.g.:


106<br />

a,Bl sel<br />

aa.pl = R« - Ra<br />

where the first term on the<br />

right hand defines the doubleselective<br />

spin-lattice relaxation<br />

rate measured on Ha when<br />

both Ha and Hpi are excited.<br />

The cross-relaxation rate<br />

in the bound ligand is obtained<br />

by:<br />

Gobs - Gf<br />

Ob =<br />

As a consequence of binding,<br />

the cross-relaxation rate changes<br />

from positive to negative<br />

values and allows to gain geometric<br />

information on the<br />

bound molecule.<br />

In fact, substitution of<br />

the motional correlation time<br />

of the protein in the equation:<br />

ob = -0.1<br />

provides a means of evaluating<br />

proton-proton distances in the<br />

bound peptide. It comes out<br />

that at least the investigated<br />

moiety retains the conformation<br />

that was shown to be<br />

stabilized by calcium ions with<br />

exclusive occurrence of the g~<br />

rotamer.<br />

6 Effect of Calcium<br />

The same experiments used to<br />

detect and delineate binding of<br />

carnosine to HSA can be repea-<br />

h<br />

Bulletin of Magnetic Resonance<br />

ted in the presence of Ca(II) at<br />

equimolar ratios with the<br />

ligand. No substantial change is<br />

observed, as far as the selective<br />

relaxation rate enhancement<br />

and the change in the<br />

double-selective relaxation<br />

rates are concerned. An<br />

appreciable change can be<br />

however observed in the<br />

ligand-protein dissociation<br />

constant that is now measured<br />

at 2.0xl0- 5 mol dm" 3 .<br />

It is therefore possible to<br />

conclude that the effect of the<br />

metal ion is to stabilize the<br />

conformation that occurs at the<br />

bound site. In absence of the<br />

ion, such conformation has to<br />

be selected among all the<br />

several conformations that are<br />

possibly assumed by the<br />

flexible ligand in solution. This<br />

process leads to an appreciable<br />

reduction of the binding constant.<br />

7 Effect of Copper<br />

It is important to underline<br />

that the same experiments<br />

cannot be carried on when<br />

using a paramagnetic ion such<br />

as copper. It is still possible to<br />

delineate the geometry of the<br />

metal complex, but, only of<br />

that having the maximum<br />

number of ligands in the<br />

coordination sphere. One is in<br />

fact forced to work at very low<br />

[metal]/[ligand] ratios.<br />

By the same way, it is<br />

possible to detect and delineate<br />

the ternary complexes


Vol. 14, No. 1-4 107<br />

formed in the presence of the<br />

protein but there is no way of<br />

shedding light on the geometrical<br />

and conformational features<br />

of the ligand bound to<br />

the macromolecule.<br />

References<br />

[1] Dill,K. and Allerhand,A.<br />

J.Am.Chem.Soc.lOl, 4376<br />

(1979).<br />

[2] Hall,L.D. and Hill, H.D.W.<br />

J.Am.Chem.Soc. 98, 1269<br />

(1976).<br />

[3] Gaggelli, E., Kushnir, T.,<br />

Navon, G. and Valensin, G.<br />

Magn.Reson.Chem., in the<br />

press.<br />

[4] Marchettini, N. and<br />

Valensin, G. J.Phys.Chem.<br />

94, 4508 (1990)<br />

[5] Niccolai, N., Rossi, C,<br />

Mascagni, P., Neri, P. and<br />

Gibbons, W.A. Biochem.<br />

Biophys. Res. Commun.<br />

124, 739 (1984).<br />

[6] Jeener, J., Meier, B.H.,<br />

Bachmann, P. and Ernst,<br />

R.R. J.Chem.Phys. 71,<br />

4546 (1979).<br />

[7] Niccolai, N., Tiezzi, E. and<br />

Valensin,G. Chem.Rev. 82,<br />

359 (1982)<br />

[8] Bertini, I. and Luchinat, C.<br />

"NMR of paramagnetic<br />

species in biological<br />

systems", Benjamin Cummings,<br />

Menlo Park, 1986.<br />

[9] Gaggelli, E., Tiezzi, E. and<br />

Valensin, G. J.Chem.Soc,<br />

Faraday Trans. II 84,<br />

141 (1988).<br />

[10] Gaggelli, E., Gaggelli, N.,<br />

Maccotta,A. and Valensin,<br />

G. Inorg.Chem., submitted.<br />

[11] Valensin, G., Valensin, P.E.<br />

and Gaggelli, E. in "NMR<br />

spectroscopy in drug research"<br />

(Jaroszewski, J.W.,<br />

Schaumburg,K. and Kofod,<br />

H. eds.), Munksgaard,<br />

Copenhagen, 1988, p.409.<br />

[12] Gaggelli, E., Di Perri, T.,<br />

Orrico, A., Capecchi, P.L.,<br />

Laghi Pasini, F. and<br />

Valensin, G. Biophys.<br />

Chem. 36, 209 (1990).


108 Bulletin of Magnetic Resonance<br />

1. Introduction<br />

Detection and Characterization Of CFC, HCFC AND HFC<br />

Gases in Foamed Insulation by High Field NMR Imaging<br />

Leslie H. Randall<br />

Alberta Research Council, PO Box 8330, Station F,<br />

Edmonton, Alberta, T6H 5X2.<br />

In recent years, there has been considerable concern<br />

regarding the environmental impact of<br />

chlorofluorocarbons (CFC's). CFC's have been<br />

widely used as blowing agents for both cellular<br />

polyurethane and polystyrene insulation applications.<br />

[1] In many non-critical applications, GFC's are<br />

being replaced by non-fluorochemical blowing agents,<br />

but in insulation based applications where the final<br />

performance of the product is dependent on the<br />

superior insulating characteristics of a closed cell<br />

network which retains the fluorochemical blowing<br />

agent, the use of similar blowing agents will probably<br />

need to continue. Hydrochlorofluorocarbons<br />

(HCFC's) have been identified as intermediate<br />

replacements for the CFC's, but hydrofluorocarbons<br />

(HFC's) are likely to become the blowing agents of<br />

choice in insulation based products. The eventual<br />

environmental fate of the blowing agent combined<br />

with the dependency of product performance on the<br />

fluorocarbon distribution within the cellular structure<br />

requires an accurate knowledge of its spatial<br />

distribution over a period of time in order to optimize<br />

performance. Currendy, there is no reliable and<br />

readily accessible technique with which the<br />

Colin A. Fyfe, Zhiming Mei<br />

University of British Columbia,<br />

Dept. of Chemistry and Pathology,<br />

Vancouver, B.C., V6T 1Y6.<br />

and<br />

Steve Whitworth<br />

Du Pont Canada Research Centre,<br />

Kingston, Ontario, K7L 5A5.<br />

distribution of these gases can be detected or monitored.<br />

Microscopic imaging has recently emerged as an excellent<br />

technique by which the distribution of mobile fluids in<br />

polymeric materials can be monitored. [2] - [8] In principal, it<br />

should be possible to perform similar experiments on samples<br />

which contain gaseous materials, the limiting factor being the<br />

signal to noise. In the present study, we demonstrate that 19 F<br />

NMR microscopic imaging is ideally suited for measuring the<br />

changes that occur in the spatial distribution as a function of<br />

time and yields quantitatively reliable information which will<br />

be critical to the fabrication of optimized insulating materials.<br />

2. Experimental<br />

Aged foam samples were provided by Dupont Canada and<br />

contained either a single fluorinated gas or a mixture of gases.<br />

NMR measurements were made on a Bruker MSL 400<br />

spectrometer equipped with a microimaging system. All<br />

experiments were performed using the microimaging probe<br />

supplied except that the probehead was modified by replacing<br />

the vertical saddle proton rf coil by a 16 mm horizontal<br />

solenoid coil that was tuned to fluorine (376.13 MHz). The<br />

nonselective 90° and 180° rf pulses were 12.5 us and 25 ps<br />

respectively. Quadrature phase cycling was used in all<br />

spectroscopic measurements.


Vol. 14, No. 1-4 109<br />

One dimensional 'H NMR spectra were obtained<br />

by the standard one pulse method and by the Carr-<br />

Purcell spin-echo experiment The Carr-Purcell NMR<br />

sequence was used to determine the T2 spin-spin<br />

relaxation times. [9] The inversion-recovery pulse<br />

sequence was used to determine the Tj spin-lattice<br />

times. [10] ID quantitative spectroscopic data was<br />

obtained by using a short 1 us ring-down delay.<br />

The spin-echo imaging sequence [11] was used<br />

for all samples. For samples which contained a single<br />

blowing agent, non-selective 90° and 180° rf pulses<br />

were used. This reduced the echo time to 2 ms which<br />

was advantageous, since spin-spin relaxation time<br />

constant, T2 for the fluorinated gases absorbed into<br />

the foam matrix was less than 5 ms. Images were<br />

composed of 128 phase encoding steps. The number<br />

of transients per experiment was typically 160. The<br />

in-plane resolution was typically 270 pm using a<br />

frequency encode gradient on the order of 7 G/cm.<br />

Due to the short Tj relaxation time of the fluorinated<br />

gases, the recycle delay was 100 ms which allowed<br />

the acquisition of an image in under 30 minutes.<br />

Concentration profiles were obtained using a<br />

frequency selective gradient in a time period on the<br />

order of 20 seconds.<br />

Samples which contained a mixture of gases were<br />

examined by a spin-echo imaging sequence in which<br />

the initial excitation pulse was frequency selective<br />

(Gaussian shaped, 300 ps duration). The echo time<br />

for this experiment was typically 3 ms.<br />

3. Results and Discussion<br />

To compare the aging characteristics, rectangular<br />

samples 1 cm x 1 cm x 2 cm in size were cut from<br />

foam boards which had been aged for several months.<br />

A variety of fluorinated gases were examined for both<br />

polyurethane and polystyrene (high and low density)<br />

foams. To devise the most appropriate imaging<br />

protocol, it is important to know the relaxation<br />

parameters of the 19 F nuclei. Tj was typically 10 on<br />

the order of 10 ms while T2 was on the order of 4<br />

ms. (Table 1). Thus, 19 F imaging will be very<br />

efficient in that the experiment can be repeated very<br />

quickly due to the short Tj values. The data must also<br />

be acquired with a short time between excitation and<br />

data acquisition due to the short T2 value. Using a<br />

spin-echo sequence comprised of hard 90° and 180°<br />

pulses, the echo time was reduced to 2 ms.<br />

Figure 1 shows the 19 F image of a sample of<br />

polystyrene foam of rectangular cross-section which<br />

has been cut from the outside edge of a sheet of<br />

foamed insulation. The sheet has been aged at room<br />

temperature for 14 months. The concentration of gas<br />

(CH3CF2C1) decreases from the outside edge inwards. The<br />

quantitative distribution of cell gas is also obtained from an<br />

examination of the image projection (Figure IB). This shows<br />

the actual distribution as a function of distance. Although<br />

there is only minute quantity of fluorinated gas present in the<br />

sample, an image is easily obtained at high magnetic fields due<br />

to the high sensitivity of the 19 F nucleus and its short spinlattice<br />

relaxation time (10 ms).<br />

Figure 1. Concentration of CH3CF2C1 in a sample of insulating<br />

foam which has been aged 14 months,<br />

(a) Projection, (b) Spin echo image.<br />

Edge of foam<br />

Table 1: I9 F NMR Relaxation Behaviour of<br />

CFC, HFC and HCFC gases in Insulating Foams<br />

Foam Cell Gas<br />

Polyurethane CFC13<br />

T2<br />

5<br />

3.6 5.4 -5.0<br />

Polyurethane CF3CHC12 13.3 3.7 -85.9<br />

Polystyrene CF2C12 4.4 4.5 -13.5<br />

Polystyrene CH3CF2C1 4.9 4.2 -53.1<br />

Polystyrene CF3CH2F 7.4 3.0 -85.0<br />

-246.4<br />

Tj and T2 are in ms.


110 Bulletin of Magnetic Resonance<br />

Polystyrene samples which had been formed<br />

using a mixture of blowing agents (CF2C12 and<br />

CH3CF2C1) was also examined. The distribution of<br />

the two gases was compared in samples which had<br />

been aged two months and 14 months (Figure 2).<br />

The concentration and distribution of CF2C12 gas in<br />

the two samples is similar. However, the<br />

concentration of CH3CF2C1 as a function of distance<br />

from the foam edge has substantially decreased. This<br />

implies that the loss of CF2C12 is somehow slowed by<br />

the presence of the CH3CF2C1. These observations<br />

have been verified by one- dimensional spectroscopic<br />

measurements in which a calibrated standard has been<br />

used.<br />

Figure 2. (a) Concentration of CF2C12 gas<br />

(i) after 2 months and (ii) after 14 months<br />

Figure 2. (b) Concentration of CH3CF2C1 gas<br />

(i) after 2 months and (ii) after 14 months<br />

In a separate study, a cylindrical piece of insulating foam<br />

which had been formed using CH3CF2C1 was then treated to<br />

CF3CH2F in an oven at 80° C. Chemical shift selective<br />

imaging was performed on the sample and revealed that the<br />

post-treatment gas has penetrated the outside edge of the foam<br />

and has formed a ring, approximately 2-3 mm in thickness.<br />

(Figure 3a) The initial blowing gas, CH3CF2C1 has an almost<br />

uniform distribution in the centre of the foam, which decreases<br />

smoothly to the outside edge and includes the volume occupied<br />

by the CF3CH2F gas (Figure 3b). The loss of CH3CF2C1 is<br />

substantially less in these samples as compared with samples<br />

that had been placed in the oven (in air). Thus, in the<br />

presence of a fluorinated gas, the blowing agent is lost at a<br />

much slower rate.<br />

Figure 3. (a) Concentration of CH3CF2C1 gas.<br />

Figure 3. (b) Concentration of CF3CH2F gas.


Vol. 14, No. 1-4 11!<br />

4. Conclusions<br />

These data are typical of those we have obtained on<br />

a variety of fluorocarbon gas/foam matrices and<br />

clearly indicate that 19 F NMR microscopic imaging is<br />

ideally suited for measuring the distribution of<br />

fluorinated hydrocarbons in polystyrene and<br />

polyurethane foams. Presently, a protocol is under<br />

development which will ensure that quantitatively<br />

reliable information is being obtained. This will<br />

allow us to monitor the changes that occur in the<br />

spatial distribution as a function of time (accelerated<br />

aging tests) or as a function of blowing agent. This<br />

information can then be used to optimize the<br />

fabrication of these insulating materials.<br />

5. References<br />

1. Proceedings of the Polyurethane World Congress,<br />

1987.<br />

2. Rothwell, W. P., Holeck P. R. & Kershaw, J. A.<br />

/. Polym. Sri. Polym. Lett. Ed., 22, 241 (1984).<br />

3. Blackband, S. & Mansfield, P. J. Phys. C: Solid<br />

State Phys., 19, L49 (1986).<br />

4. Marcei, T. H., Donstrup, S. & Rigamonti, A. J.<br />

Mol. Liquids, 38, 185, (1988).<br />

5. Weisenberger, L. A. & Koenig, J. L. Appl.<br />

Spectrosc, 42, 1117(1989).<br />

6. Weisenberger, L. A. & Koenig, J. L.<br />

Macromolecules, 23, 2445 (1990).<br />

7. Webb A. G. & Hall, L. D. Polym. Commun. 11,<br />

422 (1990).<br />

8. Webb A. G. & Hall, L. D. Polym. Commun. 11,<br />

425 (1990).<br />

9. H. Y. Carr and E. M. Purcell, Phys. Rev., 94, 630<br />

(1954).<br />

10. R. L. Void, J. S. Waugh, M. P. Klein and D. E.<br />

Phelps, J. Chem. Phys., 48, 383 (1968).<br />

11. Edelstein, W.A.; Hutchinson, J.M.S.; Johnson G.;<br />

Redpath, T. Phys. Med. BioL, 25, 751 (1980).


112<br />

MYSTERIOUS NEGATIVE PEAKS<br />

IN THE 1 H{ 1 H} NOE DIFFERENCE SPECTRA<br />

OF SOME THIOPYRAN COMPOUNDS<br />

Thioyrans 1-4 were formed in the course of<br />

regiospecific and stereoselective Diels-Alder<br />

cycloadditions involving 2-(N-Acylamino)-lthia-l,3-dienes.<br />

The constitution and the main<br />

conformational features of thiopyrans 1-4 were<br />

established using classical high field NMR<br />

methods: 2D ^ H and 13 C- X H shift correlation,<br />

selective 13 C{ l U} and ^H} NOE difference<br />

experiments [1], [2].<br />

Conformational aspects:<br />

a) ring pseudorotation:<br />

In the case of compounds 3 and 4 the<br />

thiopyran ring exhibits a slow ring flip between<br />

the relevant half-chair forms. This gives rise to<br />

saturation transfer effects in the 1 H{ 1 H} NOE<br />

difference experiments, as illustrated for 4 in<br />

Fig. 1.<br />

Csaba Szantay, Jr.<br />

Chemical Works of Gedeon Richter Ltd.<br />

H-1475, Budapest, POB 27, Hungary.<br />

Bulletin of Magnetic Resonance<br />

In 1 and 2 the ring conformational equilibrium<br />

is completely shifted towards the form<br />

in which H-2 and H-3 are antiperiplanarly arranged<br />

(Fig. 2).<br />

b) The C(3) sidechain:<br />

The sidechain shows free rotameric mobility<br />

at room temperature. The conformational features<br />

were characterized in terms of the<br />

measured NOEs and MM calculations [1]. The<br />

relative contributions of the major sidechain<br />

conformations are depicted in Fig. 2.<br />

Ac2N-<br />

AcN-<br />

anti syn<br />

Ph<br />

H<br />

s cis—trans<br />

H P \ h f<br />

. trans—cis<br />

\h<br />

Ph<br />

H<br />

0-<br />

i %<br />

1<br />

A-<br />

~100%<br />

2<br />

Ac2N—(/<br />

AcN—V<br />

Et S<br />

Figure 2. The main rotameric forms of the<br />

C(3) sidechain in compounds 1-4.<br />

Q


Vol. 14, No. 1-4 113<br />

minor<br />

NHMe<br />

major<br />

H-5<br />

U.I<br />

NH<br />

HMe<br />

H-3<br />

NMe<br />

NHMe<br />

•» ifr<br />

COMe<br />

Figure 1. One of the several NOE difference spectra of compound 4 showing the presence of<br />

saturation transfer as a result of a slow ring interconversion.<br />

H-5<br />

JL<br />

G.00 5.S0<br />

Ph<br />

-14X<br />

H-2<br />

PPH S.00<br />

4.SO<br />

+9X<br />

H-4<br />

JL<br />

Figure 3. A segment of the NOE difference<br />

spectra showing negative peaks in<br />

the case of 2. (CDC13,25 °C, 400 MHz).<br />

Ph<br />

2 PPH<br />

-30°C<br />

+7X<br />

' I ' • ' ' I ' ' • • I ' • « " T " t ' • • ! • ! •<br />

6.0 5.S 5.0 4.5 4.0<br />

PPH<br />

Figure 4. A segment of the NOE difference<br />

spectra showing negative peaks in<br />

the case of 1. (CDC13, -30 °C, 400 MHz).


t<br />

,!<br />

11<br />

114<br />

The mystery:<br />

During these invetigations a highly unusual<br />

phenomenon was observed in the ^^H} NOE<br />

difference spectra of compounds 1 and 2, in that<br />

the H-3 and H-2 protons are connected by massive<br />

negative peaks while all other NOEs are<br />

positive as illustrated in Figs. 3 and 4.<br />

The facts:<br />

a) The observed effect involves H-2 and H-3<br />

only.<br />

b) The phenomenon appears to be specifically<br />

linked to the presence of the C = O unit in<br />

the sidechain, and is compeletely absent in 3<br />

and 4.<br />

c) As measured at 400 and 300 MHz, respectively,<br />

the negative peaks showed no field dependence<br />

beyond that attributable to<br />

experimental errors.<br />

d) The intensity of the negative enhancements<br />

increases markedly with decreasing<br />

temperature and viscosity, and tends towards<br />

zero at elevated temperatures in accordance<br />

with the antiperiplanar arrangement of H-2 and<br />

H-3:<br />

Negative enhancements at 400 MHz,<br />

measured at various temperatures using<br />

otherwise identical experimental<br />

parameters:<br />

1<br />

-30°C<br />

-20%<br />

CDCb<br />

+ 25°C<br />

-3%<br />

+25°C<br />

-14%<br />

+50°C<br />

0%<br />

+ 50°C<br />

-4%<br />

DMSO<br />

+25°C<br />

-14%<br />

e) The nonselective Tx relaxation times of all<br />

ring protons in 1 (CDC13, +25°C, 400 MHz)<br />

were measured to be ca. 1 s. This accords with<br />

the size and expectedly fast tumbling of the<br />

molecule.<br />

Bulletin of Magnetic Resonance<br />

Conceivable (but rejectable) explanations:<br />

The effect seems to be inexplicable in terms<br />

of the well understood cross-relaxation or conceivable<br />

saturation transfer processes:<br />

a) The trivial case of irradiation spillover can<br />

be discounted because: 1) This would be incompatible<br />

with the observed temperature dependence<br />

of the negative peaks; 2) Fig. 4. illustrates<br />

a situation in which the signal due to H-3 is<br />

"halfway" in between those of H-5 and H-2, of<br />

which only the latter shows the negative effect.<br />

(The irradiation power levels used in all experiments<br />

were: 54L (Bruker AM 400) and<br />

DLP = 30 (Varian VXR-300).<br />

b) A three-spin effect is geometrically unjustified,<br />

and is incommensurate with the magnitude<br />

of these negative enhancements.<br />

c) The possibility of saturation transfer may<br />

be considered as being the result of some form<br />

of chemical exchange in 1 and 2. Possible options<br />

are: 1) direct proton transfer between H-2<br />

and H-3. However, *H and 2 H NMR studies on<br />

the C(3)- 2 H labeled isotopomer of 1 have conclusively<br />

shown that any possibility of a<br />

stereoselective H-2—H-3 proton transfer can be<br />

discounted. 2) A possible chemical exchange<br />

may also involve restricted ring pseudorotation<br />

or restricted C(3) sidechain mobility, both<br />

being slow on the chemical shift time-scale. Assuming<br />

that in such a hypothetical situation the<br />

H-3 signal of the minor conformer lies directly<br />

underneath the H-2 signal of the major conformer<br />

(and vice versa !), this could lead to<br />

negative peaks that are (vaguely) similar to<br />

those observed in the NOE difference spectra.<br />

However, such a situation can be ruled out on<br />

account of several quite obvious considerations,<br />

the most trivial being that 1) the multiplet patterns<br />

of the observed negative peaks do not<br />

conform to those expected for either of the<br />

above possibilities for slowly interconverting<br />

conformational species; 2) the phenomenon is<br />

observed only in relation to H-2 and H-3.


Vol. 14, No. 1-4<br />

Moreover, the observed temperature and<br />

viscosity dependence of the effect is contrary to<br />

that normally expected for saturation transfer:<br />

1) The longer correlation times associated with<br />

higher solution viscosity (in our case in DMSO)<br />

or lower temperatures are, in the extreme narrowing<br />

limit, related to more efficient Ti relaxation<br />

which works against saturation transfer. 2)<br />

Decreased temperatures usually diminish the<br />

efficiency of the exchange process.<br />

d) A major contribution from scalar relaxation<br />

[3] which could be perceived as being<br />

brought about by the somewhat labile character<br />

of H-3 can be ruled out since the J(H-2,H-3)<br />

couplings show no sign of the rapid modulation<br />

that would result in the collapse of the relevant<br />

rmiltiplet patterns.<br />

e) Strong coupling effects [4] are not present,<br />

and would not give negative peaks of this size.<br />

f) Anisotropic reorientation in the intermediate<br />

region between fast and slow tumbling<br />

can be considered. This carries the possibility of<br />

exhibiting small positive and small negative enhancements<br />

simultaneously. However, all the<br />

115<br />

positive enhancements measured between spatially<br />

analogously related protons in 1-4 were of<br />

similar magnitude, and in line with that expected<br />

for a fastly tumbling molecule.<br />

The phenomenon therefore still awaits adequate<br />

rationalization. The question of how the<br />

effect may be related to the geometrical and<br />

constitutional specifics of these molecules is<br />

still being pursued.<br />

1. I.T. Barnish, C.W.G. Fishwick, D.R. Hill,<br />

Cs. Szantay Jr., Tetrahedron, 45 (1989) 6771.<br />

2. Cs. Szantay Jr., I. Moldvai, C.W.G. Fishwick,<br />

D.R. Hill, Tetrahedron Lett, 32 (1991)<br />

2529.<br />

3. D. Neuhaus, M.P. Williamson, The<br />

Nuclear Overhauser Effect in Structural and<br />

Conformational Analysis, VCH Publishers, New<br />

York, 1989, pp. 203-207.<br />

4. J. Keeler, D. Neuhaus, M.P. Williamson, /.<br />

Magn. Reson., 73 (1987) 45.


116 Bulletin of Magnetic Resonance<br />

H-1 and C-13 NMR Spectra of the Carbanions<br />

Produced from Phenylpropene Derivatives<br />

Akihiro Yoshino, Kensuke Aoki, Masahiro Ushio,<br />

and Kensuke Takahashi<br />

Department of Applied Chemistry, Nagoya Institute of Technology,<br />

Gokiso-cho, Showa-ku, Nagoya 466, Japan<br />

Abstract l,l-Diphenyl-2-methyl-l-propene produces an equimolar mixture<br />

of two carbanions in contact with excess potassium-sodium alloy in<br />

tetrahydrofuran. The carbanions can be interpreted as the products of disproportionation<br />

reaction between two radical anions produced by one-electron<br />

reduction of the phenylpropene with alkali metal.<br />

1 Introduction<br />

Various carbanions can be prepared<br />

from substituted ethylenic<br />

compounds (1) in contact with alkali<br />

metal in tetrahydrofuran(THF).<br />

In these reactions a radical anion<br />

(2) is formed as shown in Scheme 1<br />

Scheme 1<br />

and its reactivity or stability will<br />

depend on the nature of the substituents.<br />

The three routes of reactions<br />

can be considered as follows.<br />

(1) If the ethylene has bulky<br />

substituents on both Ci and C2, the<br />

corresponding ethylene dianion is<br />

produced (Scheme l).[l-4] (2) If<br />

the ethylene has no or sterically<br />

small substituent such as only one<br />

methyl group on either Ci or C2, the<br />

corresponding dimer dianion is produced<br />

(Scheme 2).[5-7] (3) In the<br />

present paper, three title phenyl<br />

R, © • _R<br />

2 X >C,-C2 '>C,-CH< 3 + >C,-C2< '<br />

^ R4THF R/ V R4 R/ 2K R4<br />

2 5 6<br />

Scheme 3<br />

carbanions, 5 and 6 respectively,<br />

whose structures are confirmed by<br />

!H and 13 C NMR spectra. One of<br />

these anions(5) is a phenylalkyl<br />

carbanion and the other is a phenylallyl<br />

carbanion (6). Thus the result<br />

may be interpreted as a dispro-


Vol. 14, No. 1-4 117<br />

portionation reaction between two<br />

anion radicals generated first by<br />

one-electron reduction of the<br />

starting phenylpropene with alkali<br />

metal. As far as we know, this type<br />

of simultaneous formation of two<br />

different carbanions has not yet<br />

been reported. The conditions of<br />

these reactions will be discussed in<br />

terms of steric bulkiness of the<br />

substituents.<br />

2 Experimental<br />

The starting materials were prepared<br />

from dehydration of the corresponding<br />

phenylpropanols which<br />

were prepared by Grignard reaction<br />

and followed by dehydration with<br />

anhydrous acetic acid. The starting<br />

materials dissolved in THF or THFd8<br />

were kept in contact with<br />

potassium-sodium alloy in vacuum<br />

at room temperature for about 24 h.<br />

The resulting dark red solutions<br />

were filtered, concentrated if necessary,<br />

and then sealed into a 5-mm<br />

NMR sample tube. Their concentrations<br />

were about 1 M. The *H and<br />

13 C NMR measurements were carried<br />

out at 22°C using Varian XL-<br />

200 or Unity-400 spectrometer.<br />

The chemical shifts were evaluated<br />

from the upfield peak of THF or<br />

THF-dg, used as an internal reference.<br />

This peak was taken as 1.79<br />

or 1.75 for l H and 26.4 or 26.0 ppm<br />

for 13 C resonances, respectively,<br />

from TMS.<br />

3 Results and Discussion<br />

1 H NMR Spectra of the Carbanions.<br />

A typical l H NMR spectrum and<br />

chemical shifts are shown in Fig. 1<br />

and Table 1, respectively. In Fig. 1,<br />

there are three signals in the region<br />

from 0 to 3 ppm, which are assigned<br />

to one methyl and one isopropyl<br />

groups. There are two<br />

ethylenic signals in the region from<br />

3 to 4 ppm. One problem in the<br />

spectrum is the origin of hydrogen<br />

in the isopropyl group. To clarify<br />

the origin, the experiments using<br />

two different solvents, THF and<br />

THF-d8 were carried out. Since<br />

these two spectra are similar each<br />

Table 1. *H Chemical Shifts of Related Carbanions and Their Precursors.<br />

No.<br />

1 a<br />

1 b<br />

1 c<br />

5 a<br />

5 •b<br />

5<br />

6<br />

6<br />

6<br />

c<br />

a<br />

b<br />

c<br />

R?<br />

Ph<br />

Ph<br />

CH3<br />

Ph<br />

Ph<br />

CH3<br />

Ph<br />

Ph<br />

CH3<br />

R 4<br />

CH3<br />

C2H5<br />

CH3<br />

CH3<br />

C2H5<br />

CH3<br />

CH3<br />

C2H5<br />

CH3<br />

H2<br />

2 .904<br />

2 .591<br />

2 .540<br />

t rans ci s<br />

1<br />

H3<br />

1.862<br />

1.107<br />

.580 a) 1.<br />

1.115<br />

1.114<br />

0.875<br />

780<br />

4 .035<br />

4<br />

3<br />

b> 4.<br />

.375 b) 4.<br />

.141 b> 695<br />

811<br />

3. 343<br />

a )<br />

b)<br />

b)<br />

b)<br />

Ho<br />

( H o )<br />

1<br />

1<br />

1<br />

6.741<br />

6.756<br />

5.084<br />

5.368<br />

7.119<br />

7.091<br />

6.584<br />

H m<br />

(H»- )<br />

7.24<br />

7.25<br />

7.14<br />

6.593<br />

6.563<br />

6.076<br />

6.044<br />

6.711<br />

6.661<br />

6.656<br />

Hp<br />

1<br />

1<br />

5.760<br />

5.726<br />

4.364<br />

6.121<br />

5.966<br />

5.735<br />

a) Trans(R4) and cis(Ra) configurations are defined for<br />

b) They are defined for Ci.<br />

Hothe<br />

CH3<br />

1.842<br />

1.950<br />

0.872<br />

1.287<br />

1.870<br />

1.068<br />

1.858<br />

1.971<br />

r s<br />

CH2<br />

2.802<br />

1.467<br />

1.628<br />

2.252


118<br />

6Ho<br />

5Ho<br />

l6Hm 6HMe 5H3<br />

Ph CH3 Ph CHs<br />

\ - / \ - /<br />

C-C H + C-C<br />

/ 1 2 \ / 1 2 ^<br />

P h 3 C H i P h 3 C-H<br />

- /<br />

H<br />

5a 6a<br />

8 4<br />

6 I ppm<br />

Fig.l !H NMR spectrum of an equimolar mixture<br />

of 5a and 6a in THF-d8 at 200MHz.<br />

5Cm<br />

6C2<br />

6Ci 5Ci<br />

1 f<br />

160<br />

6Cm<br />

6Co P h e n , P h C H ,<br />

\- / \- /<br />

C-CH + C-C<br />

5CO / 1 2 \ /12V<br />

Ph 3CHj Ph H 3 C-H<br />

5a 6a<br />

6Cp<br />

Bulletin of Magnetic Resonance<br />

5Cp 6CMe<br />

5C3<br />

5C2<br />

6C3<br />

5C1<br />

Jt<br />

l l l 1 1 i i<br />

80<br />

Fig .2. 13 C NMR spectrum of a mixture of 5a<br />

and 6a in THF-d8 at 50.3 MHz.<br />

l i i i i i i<br />

61 ppm<br />

I<br />

0<br />

0


Vol. 14, No. 1-4<br />

other, it must be concluded that the<br />

CH hydrogen in the isopropyl group<br />

of 5a is coming from another radical<br />

anion (2), but not from the solvent.<br />

This is also supported by an<br />

experimental fact that the product<br />

ratio of 5 and 6 is 1 to 1. Another<br />

point of interest is that two<br />

methylene hydrogens of 5b give<br />

different chemical shifts, 1.467<br />

and 1.628 ppm, respectively. This<br />

is ascribed to the presence of an<br />

asymmetric center on C2.<br />

13 C NMR Spectra of the Carbanions.<br />

A typical 13 C NMR spectrum and<br />

chemical shifts are shown in Fig. 2<br />

and Table 2, respectively. In Fig. 2,<br />

there are 15 signals for a mixture<br />

of 5a and 6a except for two solvent<br />

peaks. One methyl signal is overlapped<br />

with a more shielded solvent<br />

peak. The carbon species can be<br />

differentiated by the DEPT technique.<br />

The assignment of each peak<br />

is shown in Fig. 2.<br />

IH and 13 C NMR Chemical Shifts.<br />

All the *H and 13 C except for C2 and<br />

Cj of the carbanions (5 and 6) show<br />

upfield shifts as compared with<br />

those of the starting materials.(1)<br />

These upfield shifts are mainly<br />

Table 2. 13 C Chemical Shifts of Re<br />

No. R2 R.4 C<br />

(C<br />

la Ph CH3 138.48 131.24<br />

1 b Ph C2H5 138 .47 136.74<br />

1 c CH3 CH3 146.11 127.47<br />

5 a<br />

5 b<br />

5 c<br />

6 a Ph<br />

6 b Ph<br />

6 c CH3<br />

Ph CH3<br />

Ph C2H5<br />

CH3 CH3<br />

90.26<br />

89.56<br />

78.19<br />

30.30<br />

38.24<br />

29.10<br />

CH3 89.61 147.54<br />

C2H5 88.65 156.93<br />

CH3 80.24 144.30<br />

ascribed to the extra negative<br />

charges on carbons of the carbanions.<br />

Since the carbanions have their<br />

stability due to partial delocalization<br />

of the extra charge into the<br />

phenyl rings, phenyl rings are<br />

necessary for formation of a stable<br />

carbanion. In fact, tetramethylethylene<br />

does not react with a<br />

potassium-sodium alloy in a<br />

similar condition used for the<br />

present study.<br />

The extent of charge delocalization<br />

can be thus estimated by comparison<br />

of the *H or 13 C NMR chemical<br />

shifts. The *H and 13 C chemical<br />

shifts of Hp and Cp of 6a are less<br />

shielded than those of 5a, by about<br />

0.361 and 5.49 ppm, respectively.<br />

These shift differences divided<br />

by 10.7 and 160 ppm, respectively,<br />

gave about 0.034 unit of extra<br />

charge difference on Cp. [8]<br />

Therefore, the chemical shift<br />

difference between Hp or Cp of 6a<br />

and 5a is explained in terms of<br />

their localized excess charges. The<br />

same is true for 6b and 6c in<br />

comparison with 5b and 5c, respectively.<br />

The especially large<br />

lated Carbanions and Their Precursors.<br />

C<br />

) (c<br />

V-» o t h e r s<br />

CH3<br />

22.71 144.22 130.58 128.61 126.85 ---<br />

19.26 144.25 130.17 128.64 126.83 13.64<br />

20<br />

21<br />

22<br />

19<br />

21<br />

64<br />

07<br />

11<br />

46<br />

76<br />

130.40 128.71 126.89<br />

131.14 129.08 128.71 126.55 22.33<br />

CH2<br />

119<br />

29.20<br />

144.28 118.75 128.90 107.91 —<br />

145.02 118.77 128.81 107.80 14.19 29.45<br />

135.54 105.30 130.75 88.50 12.09<br />

105.47 130.25<br />

93.16 148.38 124.86 128.61 113.40 24.76<br />

97.66 148.25 123.21 128.61 112.11 15.87 29.94<br />

76.36 146.24 118.64 128.59 107.46 19.39<br />

27.07


120<br />

difference between 6c and 5c, is<br />

ascribed to the delocalized capacity<br />

for excess charge consisting of only<br />

one phenyl ring. Therefore, Ci of 5c<br />

is more shielded by about 10 ppm<br />

than that of 5a or 5b. The same is<br />

true for Ci and C3 in 6c; they are<br />

more shielded than those in 6a or<br />

6b. Thus the p-orbital of Ci can<br />

conjugate with those in both phenyl<br />

and allyl groups in 6., while that of<br />

5 can only conjugate with phenyl<br />

rings.<br />

For the allyl anions the average<br />

shift of the outer carbons (Ci and<br />

C3) is shielded by about 60 ppm<br />

than that of the center carbon (C2),<br />

as shown in Table 2. This is a characteristic<br />

nature of the allyl anions.[12]<br />

Dunkelbium and Brenner reported<br />

on the same carbanion (6a) with<br />

lithium as a counterion.[13]<br />

Although their chemical shifts of<br />

6a were deshielded by about 0.1<br />

ppm, their *H NMR data are consistent<br />

with ours in consideration of<br />

different counterion.<br />

Internal Rotations. Three internal<br />

rotations can be considered around<br />

the bonds of Ci to Ci, Ci to C2, and<br />

C2 to C3 for 5 and 6 (these will<br />

hereafter be referred to as rotations<br />

A, B, and C). In the *H and 13 C<br />

NMR spectra of 5c, two signals for<br />

Ho, Co, Cm, Hm were observed and<br />

their chemical shifts are given in<br />

Tables 1 and 2, respectively. This<br />

is interpreted as a restricted rotation<br />

A for the phenyl ring of 5c<br />

even at room temperature. A similar<br />

observation was presented earlier<br />

in a methylbenzyl anion.[14] A<br />

chemical shift difference of 0.284<br />

ppm for Ho of 5c is smaller than<br />

that of 0.43 ppm for Ho of methyl-<br />

Bulletin of Magnetic Resonance<br />

benzyl anion. The cause of the<br />

shifts cannot be clarified yet; the<br />

shift differences for Ho, Co, Cm,<br />

and Hm changed to 0.284, 0.17,<br />

0.50, and 0.032 ppm. These four<br />

sites are adjacent to each other.<br />

But the values change alternatively.<br />

Therefore the origin of the shift<br />

may not be explained simply. On the<br />

other hand, the internal rotations A<br />

of the phenyl groups of 5a and 5b<br />

were not restricted at room temperature.<br />

At lower temperature,<br />

however, the rotation B was inhibited<br />

at -60°C for 5b. For 5a the<br />

aromatic carbon signals were significantly<br />

broadened at -80°C but<br />

were not split. In comparison with<br />

this behavior, however, those aromatic<br />

carbon signals of 6a and 6b<br />

were observed sharply even at<br />

-80°C. Further study is necessary.<br />

Two Ho signals of 5c at 5.084 and<br />

5.368 ppm were correlated with<br />

two Co signals at 105.30 and<br />

105.47 ppm, respectively, in a 2D<br />

CH COSY spectrum for their assignment<br />

purpose. [15] Similarly, two<br />

Hm signals (6.044 and 6.076 ppm)<br />

are correlated with two Cm signals<br />

(130.25 and 130.75 ppm, respectively).<br />

Which one of the two Ho's<br />

locates near to methyl or isopropyl<br />

group? NOE of the methyl signal<br />

(1.287 ppm) was observed on the<br />

more shielded Ho signal (ca. 20 %).<br />

Therefore the methyl group on Ci is<br />

nearer to Ho than Ho'. NOE of another<br />

methyl signal (0.875 ppm)<br />

was observed on the less shielded<br />

Ho" signal (ca. 10 %).<br />

Condition of Disproportionation<br />

Reaction. The condition can be<br />

discussed in terms of bulkiness of<br />

the substituents. Four groups are


Vol. 14, No. 1-4 121<br />

used as stubstituents on Ci and C2;<br />

namely H, CH3, C2H5, and CeHs. They<br />

are differentiated to have approximately<br />

1, 2, 3, and 5 A in their diameters<br />

as spherical models. For<br />

simplified discussion, we consider<br />

that 2 has bulkier substituents on<br />

Ci than on C2. Thus the C2 site will<br />

be more reactive than the Ci site<br />

for dimerization. In this occasion<br />

discussion will be rather limited to<br />

the bulkiness of substituents on C2.<br />

In cases where substituents are<br />

two C6H5 or one CgHs and one H,<br />

monomer dianions are formed<br />

(Scheme 1). Dimerization of the<br />

radical anion occurs in cases where<br />

the substituents are two hydrogens<br />

or one H and one CH3 (Scheme 2). In<br />

cases where the substituents are<br />

two CH3, disproportionate occurs<br />

(Scheme 3). Therefore bulkiness of<br />

the substituents may control the<br />

progress of the reaction. Two alkyl<br />

groups are large in size for dimerization,<br />

and small for dianion formation,<br />

and may be suitable for<br />

disproportionation. For disproportionation<br />

reaction, the substituent<br />

must have a hydrogen to be abstracted.<br />

One CH3 on C2 of la is<br />

substituted by C2H5 in order to investigate<br />

which hydrogen in two<br />

substituents, either CH3 or C2H5, is<br />

easily abstracted. From analyses of<br />

a mixture of the products, it is<br />

concluded that the hydrogen-leaving<br />

power is about ten times stronger<br />

in CH3 than in C2H5.<br />

Acknowledgement<br />

The authors wish to thank Mr. K.<br />

Kushida and A. Ono (Varian<br />

Instruments, Ltd.) for their kindness<br />

in measuring several 2D CH<br />

COSY spectra at 100.6 MHz.<br />

4 References<br />

1) K.Takahashi,Y.Inoue, and R.Asami,<br />

Org. Magn. Reson., 3, 349 (1971).<br />

2) M.Morikawa, H.Matsui,A.Yoshino,<br />

and K.Takahashi,£«//. Chem.Soc.Jpn.,<br />

57,3327(1984).<br />

3) H.Fujiwara,A.Yoshino,<br />

Y.Yokoyama, and K.Takahashi,<br />

Bull.Chem.Soc.Jpn., 65, No.8, in<br />

press (1992).<br />

4) Y. Yokoyama, T.Koizumi, and<br />

O.Kikuchi, Chem. Lett., 1991, 2205.<br />

5) K.Takahashi and R.Asami,<br />

Bull.Chem.Soc.Jpn., 41, 231 (1968).<br />

6) K.Takahashi,M.Takaki, and<br />

R.Asami, J.Phys.Chem.,<br />

75,1062(1971).<br />

7) Y.Yokoyama, et.,<br />

Bull.Chem.Soc.Jpn., 61, 1557(1988).<br />

8) The charge calculation was followed<br />

by using an empirical equation<br />

proposed by Fraenkel et al.[9]<br />

and proportional factors presented<br />

by Schaefer et al.[10,ll]<br />

9) G.Fraenkel, et al.,<br />

J.Am.Chem.Soc.,82, 5846(1960).<br />

10) T.Schaefer and W.G.Schneider,<br />

Can. J. Chem., 41, 966 (1963).<br />

11) H.Spiesecke and W.G.Schneider,<br />

Tetrahedron Letters, 1961, 468.<br />

12) D.H.O'Brien/'Comprehensive<br />

Carbanion Chemistry," ed by<br />

E.Buncel and T.Durst, Elsvier, New<br />

York (1980), Vol 5A, p.273.<br />

13) E.Dunkelblum and S.Brenner,<br />

Tetrahedron Letters, 1973, 669.<br />

14) K.Takahashi, et al.,<br />

Org.Magn.Reson., 3, 539 (1971).<br />

15) The 2D CH COSY spectra were<br />

measured at 75.4 MHz with a<br />

Hitachi R-3000 spectrometer. The<br />

authors wish to thank Mr. M. Tamura<br />

(Instrument Division Hitachi, Ltd.)<br />

for his kindness in measuring the<br />

spectra.


122 Bulletin of Magnetic Resonance<br />

Intracellular pH and Inorganic Phosphate<br />

Effects on<br />

Skeletal Muscle Force<br />

E. R. Barton-Davis 1 , R. W. Wiseman 2 ,<br />

and<br />

M. J. Kushmerick 1 ' 2<br />

Depts. of Physiology and Biophysics* and of Radiology 2 ,<br />

University of Washington,<br />

Seattle, WA 98195, USA<br />

INTRODUCTION<br />

The molecular mechanisms of skeletal muscle<br />

fatigue are not fully understood. Intracellular pH<br />

(pHi) and inorganic phosphate (Pi) have been<br />

implicated as causes of peripheral fatigue. Both a<br />

rise in [H + ] (decreased pHi) and elevated Pi<br />

result in the inhibition of force generation in<br />

skinned fibers (2) and sometimes in whole<br />

muscle preparations (1, 3). During exercise, an<br />

accumulation of H + and Pi occur simultaneously<br />

over time, which confounds the interpretation of<br />

the relative role of each with regard to skeletal<br />

muscle fatigue. The specific protonation state of<br />

Pi might also inhibit force generation, because<br />

the relative amount of Pi species (H2PO4" and<br />

HPO4 2 ") is dependent on pHi, and these amounts<br />

change over the physiologic range of pHi (pKa =<br />

6.79) (3). In order to evaluate the role these<br />

metabolites have in muscle force, the effects of<br />

pHi and Pi on force have been dissociated by<br />

independently manipulating each.Changes in the<br />

amount of CO2 in the perfusate (PCO2)<br />

manipulates pHi (1), and exogenous pyruvate<br />

lowers Pi (6). Performing simultaneous NMR<br />

spectroscopic and mechanical measurements<br />

can assess the effects of these treatments. We<br />

have compared the relationship between force<br />

and pHi in the presence and absence of inorganic<br />

phosphate and interpreted the relative<br />

contribution of Pi and pHi on muscle force<br />

generation.<br />

METHODS<br />

Isolated soleus from male Swiss Webster mice<br />

were attached to a Harvard Apparatus isometric<br />

force transducer and placed in a bath containing<br />

MOPS Ringers (llmM glucose, pH=7.4, 95%<br />

O2, 5% CO2) at 23°C. Tetanic force, as well as<br />

rise and relaxation time, were measured at<br />

optimum length every 10 minutes. Stimulation<br />

was at supramaximal voltage and fusion<br />

frequency for 1 second. The collateral muscle<br />

was placed in a custom designed NMR probe (5)<br />

and superfused from the same source of Ringers<br />

as the mechanics bath for 31 P-NMR spectra were<br />

acquired on a General Electric 7 Tesla GN300 at<br />

121 MHz. Intracellular pH measurements were<br />

made using the chemical shift of Pi relative to<br />

PCr and the equation<br />

pHi = 6.79 + log{(.89 - 8)/(8 - 3.19)}.<br />

Mechanical and metabolic measurements were<br />

repeated with control (11 mM glucose) and<br />

substrate (20mM pyruvate) solution (PYR),<br />

varying pHi by equilibration with 5%, 25%, and<br />

50% CO2, which resulted in PCO2 values of 22,<br />

174, and 266 torr respectively. All isometric<br />

contractions were bracketted by llmM Glucose<br />

5% CO2 treatment to normalize data. Statistical<br />

analysis was performed using 2-way ANOVAS<br />

with relative force as the dependent variable, and<br />

pHi and Pi as independent variables. Paired<br />

comparisons of means were performed between<br />

each PCO2 l eve * aa ^ substrate treatment. Values<br />

of p


Vol. 14, No. 1-4<br />

RESULTS<br />

PYR lowered Pi levels from ~5 mM to


124 Bulletin of Magnetic Resonance<br />

TABLE II: Total, Diprotonated, and<br />

Monoprotonated Phosphate Levels in<br />

Soleus<br />

Glue<br />

5%<br />

CO2<br />

Glue<br />

25%<br />

CO2<br />

Glue<br />

50%<br />

CO2<br />

Pyr<br />

5%<br />

CO2<br />

Pyr<br />

25%<br />

CO2<br />

Pyr<br />

50%<br />

CO2<br />

pHi<br />

7.19<br />

+.015<br />

6.81<br />

+.018<br />

6.54<br />

+.013<br />

7.22<br />

+.04<br />

6.74<br />

+.04<br />

6.48<br />

+0.0<br />

DISCUSSION<br />

Total<br />

Pi<br />

(fi«/g)<br />

6.76<br />

+.73<br />

(n=22)<br />

7.74<br />

+2.10<br />

(n=4)<br />

6.74<br />

+.05<br />

(n=2)<br />

2.64<br />

+1.01<br />

(n=2)<br />

1.69<br />

(n=l)<br />

N.D.<br />

H2PO4<br />

(M«/g)<br />

1.91<br />

+.22<br />

3.68<br />

+.95<br />

4.25<br />

+.02<br />

0.69<br />

+.22<br />

0.88<br />

N.D.<br />

HPO4-<br />

(uste)<br />

4.85<br />

+.52<br />

4.06<br />

+1.16<br />

2.48<br />

+.07<br />

1.96<br />

+.78<br />

0.81<br />

N.D.<br />

Data presented in mean ± SE<br />

Decreased [Pi] increased force at all PCO2<br />

levels, and did enhance the dependence of force<br />

generation on pHi. The effect of pHi on force<br />

concurs with skinned fiber data (2), but does not<br />

agree with results from hypercapnic<br />

manipulation in perfused muscle preparations<br />

(1). There are several possible explanations for<br />

this discrepency. One possibility is that there is<br />

inadequate O2 delivery to the isolated muscle<br />

preparation. However, since the mouse soleus is<br />

small (~1 mm in diameter), there is little chance<br />

of O2 limitation during hypercapnia because the<br />

diffusional distances are less than 500p.m. It is<br />

also possible that at low pHi, a 1 second tetanic<br />

stimulation was not sufficiently long enough for<br />

the muscle to attain maximum force . However,<br />

longer stimulations did not appreciably increase<br />

force (data not shown). Force measured at 1<br />

second tetani was not significantly different from<br />

that of longer tetani. Hence, the results observed<br />

give an accurate presentation of the effects pHi<br />

has on tetanic force in isolated whole muscle.<br />

It has been suggested that a specific protonation<br />

state of phosphate is correlated to muscle force<br />

(3). The diprotonated species (H2PO4") is<br />

thought to lower force by binding more tightly to<br />

actomyosin crossbridges. Because force<br />

production is correlated with the release of Pi<br />

from the crossbridge, the higher binding affinity<br />

of H2PO4" would result in a decrease in force. As<br />

pHi decreases, there is a greater proportion of<br />

H2PO4" present, indicating that this<br />

phenomenon is an indirect effect of pHi on<br />

skeletal muscle force. We examined the<br />

dependence of force on pHi, HPO4 2 ", and<br />

H2PO/. The slopes of each regression line show<br />

a higher correlation between force and pHi<br />

(R^.966 in normal [Pi]) (Figure 1) than between<br />

force and either phosphate species (H2PO4",<br />

R^.138; HPO4 2 ", R 2 =.O35) alone (Table II). In<br />

these experiments, the primary effect of pHi on<br />

force does not appear to be mediated by a<br />

specific species of phosphate.<br />

Table I shows that pHi and Pi influence other<br />

properties during tetanic stimulation. Proton<br />

accumulation could alter kinetics of force<br />

development (rise and relaxation times) in<br />

several ways. Enzymes, including Ca 2+ -ATPase,<br />

actomyosin ATPase, and those in glycolysis,<br />

such as phosphofructokinase, are inhibited when<br />

pHi falls. The decline in enzymatic activity<br />

inhibits the speed at which a muscle can develop<br />

force or relax after stimulation. Low Pi causes<br />

faster rise yet slower relaxation times by<br />

different mechanisms which are not fully<br />

understood. Faster kinetics are expected since the<br />

phosphate release step in the crossbridge cycle<br />

would be more favorable due to lower Pi.<br />

However, this hypothesis can not explain why<br />

relaxation rates are significantly slower than in<br />

control conditions. It would seem plausible that<br />

Ca 2+ uptake at the sarcoplasmic Ca 2+ -ATPase<br />

would be enhanced by the increase in AG for<br />

ATP breakdown caused by low free Pi. Yet this<br />

is in direct opposition to our observations. If<br />

phosphorylation of the Ca 2+ -ATPase is<br />

necessary to stimulate Ca 2+ transport, this type<br />

of upregulation is not as effective as in normal<br />

conditions because free Pi has been depleted.


Vol. 14, No. 1-4 125<br />

Ca 2+ would remain in the sarcoplasm for a<br />

longer period of time, and relaxation would<br />

slow. It seems that there are two independent<br />

mechanisms responsible for the effect of low [Pi]<br />

on rise and relaxation.<br />

By maintaining muscles in a resting state, our<br />

experimental design avoids introducing factors<br />

arising from constant stimulation, specifically, a<br />

concomitant rise of proton and phosphate levels.<br />

The independent diminution of Pi in this<br />

preparation allows the examination of the relative<br />

contribution of pHi on force generation. In so<br />

doing, we conclude that the relation between<br />

skeletal muscle force and pHi is affected by the<br />

depletion of Pi.<br />

REFERENCES<br />

(1) Adams et al. 1991. AJP, 260 (29), C805 -<br />

C812.<br />

(2) Chase & Kushmerick. 1988. Biophysical<br />

J., 53,935-946.<br />

(3) Dawson et al. 1986. Biophysical J., 49,<br />

286a.<br />

(4) Kushmerick et al. (in press) PNAS.<br />

(5) Wiseman et al. 1990. Biophysical J., 59(2),<br />

517a.<br />

(6) Wiseman et al. 1991. SMRM Abstracts,<br />

10th Annual Mtg., 284.


126 Bulletin of Magnetic Resonance<br />

Introduction<br />

A simple model for the influence of motion<br />

on the NMR line shape.<br />

M. Goldman, T. Tabti, C. Fermon, J.F. Jacquinot and G. Saux.<br />

A problem as old as NMR is that of the influence<br />

of motion on the shape of the NMR absorption<br />

signal, or equivalently the shape of the FID.<br />

Although it is known that motion means<br />

narrowing, only two limiting cases are well<br />

defined.<br />

In the rigid solid, one has a clean theoretical<br />

expression for the FID function, in terms of the<br />

secular part 3TSS of the spin-spin interactions, most<br />

of the time essentially dipolar (1). This is<br />

insufficient to know the FID shape, nor that of its<br />

Fourier transform, the absorption f signal, but one<br />

can compute the first few moments of the latter<br />

and obtain reasonably good value for its frequency<br />

width Ad)0 (1-3).<br />

The other extreme is that of a fast motion whereby<br />

the spin-spin interactions have a vanishing average<br />

value and are modulated at a rate x' 1 which is fast<br />

on the time scale of the solid-state FID decay rate<br />

AGV.<br />

ACOOTC [1]<br />

Service de Physique de FEtat Condense,<br />

C.E.N. SACLAY, 91191 GIF SUR YVETTE<br />

FRANCE.<br />

In that case, relaxation theory predicts that the<br />

transverse magnetization decay is exponential,<br />

with a relaxation time T2 related in a precise way<br />

to the magnitude of the spin-spin interactions and<br />

the nature of the motion (1,4). The absorption line<br />

is then Lorentzian, and its half-width at half<br />

intensity T^ x is of the order of:<br />

2<br />

IC<br />

[2]<br />

Things are much more confuse in the intermediate<br />

case, that is for motions rates for which<br />

Ao)oxc«l. In that domain tjiere exists only<br />

qualitative models of the Anderson-Weiss (5) or of<br />

the Kubo-Tomita type (6), which are admittedly<br />

very crude and whose sole ambition is to provide a<br />

general physical feeling of how the signals are<br />

evolving from slow to fast motion.<br />

In this article, we attempt to correlate the FID<br />

shape under intermediate rate motion to that of the<br />

rigid solid. We use for that purpose a simpleminded<br />

model that is not new: it is the so-called<br />

Strong Collision Model (7). It has been used in<br />

particular for (i.SR studies (8) but in conjunction<br />

with other approximation. This is at variance with<br />

our approach, which treats the rigid lattice FID as<br />

an empirical information. This model is not


Vol. 14, No. 1-4<br />

rigorous either, but it fits remarkably well the<br />

experimental results, as will be shown later.<br />

The theoretical model<br />

The scenario of the Strong Collision Model is the<br />

following. Let us consider a spin system whose<br />

FID signal is Gr(t) when it is rigid, with the usual<br />

normalization:<br />

G,(0) = l [3]<br />

This system undergoes sudden motions at random<br />

time intervals with probability X per unit time.<br />

After each motion, the FID starts anew with the<br />

same shape as initially, but with an initial value<br />

equal to Gr immediately before the jump. In other<br />

words all correlations developed in the system by<br />

its previous evolution are lost, and the only<br />

"memory" left is that of the transverse polarization<br />

at the time of the sudden motion. This results in a<br />

FID signal Gm(t) (where m stands for motion)<br />

which is the weighted average of all possible<br />

random successions of initial rigid-state FBD's.<br />

A pulse being applied at time 0, the possibilities at<br />

time t are:<br />

i) No jump between 0 and t. This corresponds to a<br />

FED signal shape Gr(t) and its probability to occur<br />

is equal to exp(-Ai). Its contribution to Gm(t) is<br />

then:<br />

Gr(f)«p(-X0<br />

ii) One jump only at a time between f, and /, +dtv<br />

The corresponding probability is:<br />

x exp[-X(t - tx)] = X exp(-Xt)dt1<br />

and the signal at time t is then:<br />

mt egrating over the time f, we obtain the<br />

contribution to Gm(t):<br />

127<br />

iii) Two jumps between 0 and /. By an extension of<br />

the preceding argument, the contribution to Gm(t)<br />

is:<br />

JT Gr{t2)Gr(t-tx-t2)dt2<br />

etc...<br />

By taking the Laplace transforms:<br />

we obtain:<br />

that is:<br />

e<br />

= [G(t)exp(-zt)dt<br />

n=0<br />

[4]<br />

[5]<br />

[6]<br />

which is identical with Eq.(18) of ref.(8).<br />

The same result can be obtained much more simply<br />

by noting that after a jump the subsequent FID is<br />

that of the mobile system normalized to the<br />

transverse polarization at the time of the jump.<br />

The two possibilities being no jump between 0 and<br />

/, or at least one jump, we obtain :<br />

[7]<br />

whence the following relations between Laplace<br />

transforms:<br />

[8]<br />

from which Eq.[6] follows.<br />

The advantage of the first treatment is to make<br />

explicit use of the assumption that each partial FID<br />

between jumps has the same shape. The simplest<br />

way of interpreting Eq.[6] is through the<br />

consideration of memory functions (2,3,9). The<br />

memory function K of a function G(t) is defined<br />

through:


128 Bulletin of Magnetic Resonance<br />

[9]<br />

This form yields remarkably simple expressions for<br />

the Laplace transforms. That of the left-hand side<br />

is equal to: z0(z)-G(O) = z8(z)-l, where we<br />

have used Eq.[3], and that of the right-hand side is<br />

equal to -cp(z)9(z), where cp(z) is the Laplace<br />

transform of K(t):<br />

Then, Eq.[9] yields:<br />

z8(z)-l = -(p(z)8(z)<br />

Now, Eq.[6] yields:<br />

1 1<br />

or else:<br />

1<br />

1<br />

[10]<br />

[11]<br />

"•— A* [12]<br />

-z-X<br />

whence, according to Eq.[ll]:<br />

[13]<br />

[14]<br />

This corresponds to the following relations<br />

between memory functions:<br />

[15]<br />

We obtain the remarkably simple result that in the<br />

Strong Collision Model, the rate of decay of the<br />

"memory" shows up simply by an extra<br />

exponential decay of the memory function.<br />

Constraints and limitations to the model<br />

The most questionable assumption underlying the<br />

model is that following a motion, everything is lost<br />

but the transverse magnetization. This will be<br />

discussed in a forthcoming article.<br />

A key assumption in the formulation of the model<br />

is that the shape of the rigid-state FID is not<br />

modified by a motion, which implies that the<br />

motions do not change the form of the spin-spin<br />

interactions. This is only possible if the nuclear<br />

environment of each individual spin is the same<br />

after as before a motion, the only change being<br />

that it is not the same nuclei that occupy the same<br />

relative sites. This case corresponds to atomic or<br />

molecular motion induced by the diffusion of<br />

vacancies at low concentrations in a single crystal.<br />

One must make the distinction, first introduced by<br />

Eisenstadt and Redfield (10), between a jump and<br />

an encounter. A given portion of the crystal<br />

experiences the sudden arrival of a vacancy, which<br />

performs many jumps before disappearing for<br />

away (This is the standard expression. It is evident<br />

that it is the atoms, or molecules, that jump into<br />

vacant sites.). The whole process is sudden, in the<br />

sense that it takes place in a time too short for the<br />

spin system to undergo any significant evolution.<br />

That portion of the crystal was vacancy-free at the<br />

end of the encounter, so that the form of the spinspin<br />

interactions is indeed the same. The motions<br />

referred to in describing the model are in fact<br />

encounters, and not individual jumps.<br />

In a powder, each crystallite obeys the relation<br />

[15], but with different rigid-lattice memory<br />

functions Kr(t). The motion rate X being<br />

independent of crystal orientation, we have on the<br />

average:<br />

and:<br />

[151<br />

[14 1 ]<br />

These relations being linear, it seems that we might<br />

use powder samples, and not solely single crystals<br />

to test the model. Unfortunately, this is not the ,<br />

case because all one can observe in a powder is the<br />

average FID (or average absorption signal). We<br />

have, in place of Eq.[8]:


Vol. 14, No. 1-4<br />

This is a non-linear relation, and we have in<br />

general:<br />

[16]<br />

It is therefore impossible to test the model with<br />

powders.<br />

Experimental study<br />

We have tested the model with a single crystal of<br />

Hexamethylethane (HME) which is the most<br />

symmetrical octane molecule.<br />

Single crystals of HME are obtained from the<br />

liquid state by the Stockbarger method. Due to the<br />

high vapour pressure of the solid, the crystal is<br />

enclosed in a sealed tube.<br />

This molecular crystal experiences a first-order<br />

phase transition at 152.5 K whereby the molecules<br />

undergo rapid reorientation up to the melting point<br />

at 374 K. As a consequence, the inter-molecular<br />

dipolar interactions average to zero, whereas the<br />

average dipolar interactions between protons of<br />

.1<br />

129<br />

different molecules are the same as if each proton<br />

was located at the centre of gravity of its<br />

molecule. These centres form a body centred cubic<br />

structure with a unit cell of size 7.69 A.<br />

NMR relaxation measurements reveal the<br />

existence in this phase of a thermally activated<br />

translational diffusion of the molecules through the<br />

motion of vacancies (11). Analysis of these<br />

measurements yields the value of the average time<br />

x between successive motions, as a function of<br />

temperature (12).<br />

The proton FID's were observed at 91 MHz with a<br />

home-made pulse spectro-meter. We have used<br />

0.5 (is pulses with a repetition time of 0.3 sec. The<br />

FID signal were sampled in 2048 channels with a<br />

time of O.3(xs per channel. The FID signals were<br />

extrapolated back to the origin through a fourthorder<br />

Taylor expansion approxi-mation:<br />

4!<br />

The main test consists in checking whether the<br />

model is actually able to account for the<br />

variation of the FID shape with temperature. This<br />

is done with the help of Eq.[6] for z = i(0. The<br />

results of the fits are given Figure 1.<br />

100 200 300 400 500 600*<br />

time (jxsec)<br />

figure 1: Free Induction Decay Signals at different temperatures: open circles: 283 K, open squares<br />

292.5 K, full circles: 303 K, full squares : 311.3 K. Solid curves are calculated with our model.


130 Bulletin of Magnetic Resonance<br />

The only fit parameter is X,. The values of X are<br />

plotted as a function of 1/T in figure 2, together<br />

with values of 1/T taken from ref.(12) . It is seen<br />

that both X and 1/x obey an Arrhenius law with the<br />

same activation energy.<br />

The ratio (Xi)' 1 ~ 1.6 is compatible with X being<br />

equal to the decay rate of the secular dipolar autocorrelation<br />

function.<br />

2<br />

§<br />

I<br />

1E+06 T<br />

1E+05 •:<br />

1E+04 -:<br />

1E+03 -:<br />

1E+02<br />

3.1 32 3.3 3.4 35 3.6 3.7 3.8<br />

1000 ,„_..<br />

figure 2 : Memory decay rates A. (white circles, present<br />

work) and average jump rates f* (black circles, deduced<br />

fromref. (12)).<br />

Conclusion<br />

The naive strong collision model is remarkably<br />

successful in accounting for the variation of the<br />

FID with motion of intermediate rate. There is<br />

however a surprising and unexplained fact : the<br />

rate X entering eq.[7] is smaller than the average<br />

rate between encounters. This rate X is found<br />

approximately equal to the rate of decay of the<br />

dipolar auto-correlation function, which is well<br />

known from theory to involve several encounters<br />

(13).<br />

References<br />

(1) A. ABRAGAM, The principles of nuclear magnetism,<br />

Oxford University Press, Oxford (1961).<br />

(2) M. MEHRING, High resolution NMR in solids, 2 nd Ed.<br />

Springer-Verlag, Berlin (1983). Appendix G.<br />

(3) A. ABRAGAM and M. GOLDMAN, Nuclear<br />

magnetism: order and disorder, Oxford University Press,<br />

Oxford (1982) chap 1.<br />

(4) N. BLOEMBERGEN, E.M. PURCELL and R.V.<br />

POUND, Phys. Rev. 73 679 (1948).<br />

(5) P. W. ANDERSON and P.R. WEISS, Rev. Mod. Phys.<br />

25, 269 (1953).<br />

(6) R. KUBO and K. TOMTTA, J. Phys. Soc. Japan 9, 888<br />

(1954).<br />

(7) R. KUBO, J. Phys. Soc. Japan 9, 935 (1954).<br />

(8) R. S. HAYANO, Y. J. UEMURA, J. IMAZATO, N.<br />

NISHJDA, T. YAMAZAKI and R. KUBO, Phys. Rev. B20,<br />

850 (1979).<br />

(9) H. MORI, Prog. Theor. Phys. 33,423 (1965).<br />

(10) M. EISENSTADT and A.G. REDFTELD Phys. Rev.<br />

132, 635 (1963).<br />

(11) J.M. CHEZEAU, J. DUFOURQ and J.H. STRANGE,<br />

Molec. Phys. 20, 305 (1971).<br />

(12) A.R. BRICHTER and J.H. STRANGE, Molec. Phys.<br />

37, 181 (1979).<br />

(13) D. WOLF, Phys.Rev. B10,2710 (1974).


Vol. 14, No. 1-4 131<br />

1. Introduction<br />

THE EFFECT ON Ti OF CORRELATED WATER<br />

MOTIONS IN THE POLAR PHASE OF COLEMANITE<br />

Proton magnetic resonance in the ferroelectric<br />

colemanite, CaB3O4(OH)3.H2O, is dominated by<br />

the dynamical motion of the water molecules.<br />

Absorption lineshape [1] and spin lattice<br />

relaxation studies [2,3] have shown that in the non<br />

polar phase (above 270 K) the water molecules<br />

undergo two kinds of reorientational motion; a<br />

180° flipping motion about the H-O-H bisectrix<br />

which merely exchanges the hydrogen positions,<br />

together with a jumping motion in which one of<br />

the water hydrogens takes up one of two possible<br />

sites. This latter motion is also accompanied by a<br />

jump of some of the hydroxyl hydrogens between<br />

two possible sites.<br />

H',<br />

H'84(O —<br />

^N 2.16 A<br />

J. Sun and A. Watton<br />

Dept. of Physics and Astronomy,<br />

University of Victoria<br />

Victoria, BC<br />

Canada V8W 3P6<br />

He*<br />

l g- 1: Projection along the b axis of colemanite<br />

mowing the relative configuration of a pair of<br />

[ater molecules and neighbouring hydroxyl<br />

jrorogens. The smaller water hydrogen -<br />

f?S? yl hydrogen distances, which make a<br />

111 d Dt contribution t0 the relaxation, are<br />

In the polar phase, as illustrated by the<br />

hydrogen positions in Fig. 1, the jumping motion<br />

freezes out accompanied by a slight<br />

rearrangement of the heavy atom network [4]-<br />

The predominant lineshape and relaxation<br />

mechanism in this phase is the remaining 180°<br />

flipping motion of the water groups. This<br />

accounts very well for most of the observed<br />

features of the absorption lineshape and BPP type<br />

temperature dependence of the spin lattice<br />

relaxation time (Fig. 2). However there still<br />

remains a small discrepancy between the<br />

0.13±0.01 s value observed for the Ti minimum<br />

at a Larmor frequency of 30 MHz and the 0.09 s<br />

value predicted by this model.<br />

T (K)<br />

500 300 200 150 125 100<br />

100<br />

i r<br />

0.1<br />

Fig. 2: Temperature dependence of the proton<br />

spin lattice relaxation time in powdered<br />

colemanite at a Larmor frequency of 30 MHz.


132<br />

Although experimental values for the Ti minimum<br />

smaller than model predictions can usually<br />

be accounted for by additional mechanisms not<br />

included in the model, this is clearly not the<br />

situation here and the interpretation is not so<br />

straightforward. It appears that the flipping<br />

mechanism in the theoretical model, assuming this<br />

motion is an appropriate one, is not as effective as<br />

it should be. One possible reason for this reduced<br />

efficiency is that since the water molecules occur<br />

in adjacent pairs their flips are correlated with<br />

each other. That is to say that a flip of one water<br />

molecule is accompanied by a simultaneous flip<br />

of the other one in the pair. Such a correlation<br />

would result in the number of intermolecular<br />

configurations available to the proton pairs being<br />

reduced over those available in uncorrelated<br />

motion. It would seem reasonable then that the<br />

corresponding reduction in the fluctuations of the<br />

intermolecular dipole interaction could lead to a<br />

reduction in relaxation time. We have therefore<br />

extended the theoretical model for water flipping<br />

to include correlations within adjacent water pairs.<br />

The goal was to see whether such correlations<br />

would lead to an increase of the theoretical Ti and<br />

if so whether this could, by itself, account for the<br />

larger value observed.<br />

2. Model Calculation<br />

The spin lattice relaxation time resulting from<br />

the reorientations of the water molecules is given<br />

by [5,6]<br />

£ = f y 4 * 2 X<br />

where the J(co)'s represent the spectral density of<br />

the fluctuating dipolar interaction between protons<br />

i and j, and coo is th e Larmor frequency.<br />

In the formalism which we will adopt the water<br />

dynamics are described by a dimensionless<br />

transition matrix [7]<br />

Y--A-<br />

" 0<br />

where the matrix A has elements Xjj, i^ j, which<br />

are the probabilities/unit time of the interproton<br />

vector ?j- making a transition to r • among the n<br />

possible position vectors (k[[ is chosen so that<br />

Bulletin of Magnetic Resonance<br />

In a similar way the proton dipolar interaction is<br />

described by a geometrical matrix A with<br />

elements given by<br />

l-3cos 2 a;,-<br />

3 3<br />

where r^, r • =1 r^ I, I r • I and ay is the angle between<br />

r- andr: .<br />

X has eigenvalues XJ and is diagonalized with a<br />

matrix T.i.e.<br />

jj<br />

If A is then transformed by the same matrix T,<br />

AT = TAT~ l<br />

the spin lattice relaxation time can be written as<br />

where<br />

M(©0)= X -^<br />

In a 180° flip of a water molecule the<br />

intramolecular proton-proton vector merely<br />

changes sign leaving its contribution to the dipolar<br />

interaction invariant and hence contributing<br />

nothing to the relaxation. As a result only the<br />

intermolecular couplings need be considered, of<br />

which there are two dominant kinds; the<br />

interaction between water pairs and the waterhydroxyl<br />

interaction.<br />

In uncorrelated water motions each water-water<br />

interproton vector can occupy the four positions<br />

illustrated schematically in Fig. 3 and labelled,<br />

arbitrarily 1,2,3 and 4. The transition matrix is<br />

then<br />

-2<br />

0<br />

1<br />

0<br />

-2<br />

1<br />

1<br />

1<br />

-2<br />

1 1 0 - 2<br />

where X is the probability/unit time of any water a<br />

molecule, all assumed dynamically equivalent, jf<br />

making a 180° flip. However, if the water -<br />

motions are correlated in the sense that the water<br />

molecules in each pair reorient in unison then the<br />

transition matrix for the same inter-water proton<br />

vectors becomes<br />

1<br />

1<br />

0


Vol. 14, No. 1-4<br />

"-1 1 0 0"<br />

1 - 1 0 0<br />

0 0 - 1 1<br />

0 0 1-1<br />

Fig. 3: Schematic representation of a pair of<br />

water molecules showing the four positions<br />

occupied by each inter proton vector between<br />

water molecules during uncorrelated motions.<br />

The choice of label numbers is arbitrary.<br />

In either case the water-hydroxyl interproton<br />

vectors have only two values for each waterhydroxyl<br />

group and their corresponding transition<br />

matrix is given by<br />

In fact, for correlated water motions the waterwater<br />

interproton vectors can be partitioned into<br />

two sets {1,2} and {3,4} each of which has a<br />

transition matrix of the simple 2x2 form above.<br />

Applying the diagonalization and transformation<br />

^procedure described previously to these matrices<br />

^produces the following expression for Ti resulting<br />

ifrpm correlated 180° flips of the water molecules,<br />

= 7.0<br />

4coQr<br />

where x=l/\ is the correlation time for the 180°<br />

up motion.<br />

his gives a Ti minimum of 0.10s for a Larmor<br />

fluency of 30 MHz which is greater than that<br />

^"jcorrelated motion by about 10% but still<br />

nilicantly short of the 0.13s observed. It<br />

133<br />

appears that at least the simple correlated model<br />

considered here is inadequate to fully account for<br />

the minimum in Ti and that some other<br />

mechanism is presumably in effect.<br />

In colemanite, as shown by Fig. 1, some<br />

hydroxyl hydrogens are quite close to the water<br />

molecules and there are only two water hydrogens<br />

for every three hydroxyl hydrogens in each<br />

molecular unit. As a result the hydroxyl-water<br />

coupling and the water-water coupling make<br />

comparable contributions to the proton relaxation.<br />

Since the hydroxyl-water contribution is<br />

independent of any correlations in the water<br />

dynamics it seems reasonable that there is not a<br />

large difference in the overall Ti between the<br />

correlated and uncorrelated motions, but it is<br />

gratifying that the change is in the right direction<br />

and is consistent with the reduction in the dipolar<br />

fluctuation discussed earlier.<br />

References<br />

1. A. Watton, H.E. Petch and M.M. Pintar,<br />

Can. J. Phys. 48, 1081 (1970)<br />

2. R. Blinc, M. Brenman, S.R. Miller and J.S.<br />

Waugh, J. Phys. Chem. Solids 23_, 156<br />

(1962)<br />

3. A. Watton, H.E. Petch and M.M. Pintar,<br />

Can. J. Phys. 51, 1005 (1973)<br />

4. F.N. Hainsworth and H.E. Petch, Can. J.<br />

Phys. 44, 2083, (1966)<br />

5. N. Bloembergen, E.M. Purcell and R.V.<br />

Pound; Phys Rev. 73, 679 (1948)<br />

6. A. Abragam, The Principles of Nuclear<br />

Magnetisim (Oxford University Press, New<br />

York, 1961)<br />

7. A. Watton, Phys. Rev. BI7, 945 (1978)


134 Bulletin of Magnetic Resonance<br />

MEASUREMENT OF DEUTERON SPIN RELAXATION<br />

TIMES IN LIQUID CRYSTALS by a BROADBAND<br />

EXCITATION SEQUENCE<br />

I. Introduction<br />

Ronald Y. Dong<br />

Department of Physics and Astronomy, Brandon University<br />

Brandon, Manitoba R7A 6A9<br />

Liquid crystals are composed of flexible organic<br />

molecules and capable of forming different<br />

ordered structures in their mesophases. Nuclear<br />

spin relaxation [1], [2] is a powerful technique<br />

that provides useful information on the<br />

molecular dynamics of liquid crystals. There<br />

are collective director fluctuations, molecular<br />

reorientation and internal rotations in flexible<br />

end chains. Recently internal dynamics<br />

of mesogenic molecules has attracted much attention.<br />

Both theoretical [3], [4] and experimental<br />

[5], [6] studies have been carried out.<br />

Experimentally carbon-13 and deuteron may<br />

be used to probe internal dynamics of flexible<br />

mesogens. In aligned samples of deuterated<br />

liquid crystals, deuterium NMR spectroscopy<br />

yields well-resolved spectral lines having different<br />

quadrupolar splittings for various atomic<br />

sites. These quadrupolar splittings result from<br />

incomplete averaging by anisotropic reorientation<br />

of molecules in the mesophases. For a single<br />

deuteron spin (1=1), there are five independent<br />

spin relaxation times [7]. These are two<br />

spin-lattice relaxation times and three independent<br />

spin-spin relaxation times. Since the deuterium<br />

Zeeman (T\z) and quadrupolar (T\Q)<br />

spin-lattice relaxation times are given by [7],<br />

[8]<br />

rp -l<br />

~3Ji(u;0),<br />

4J2(2w0)<br />

they can be used to separate the two spectral<br />

densities of motion Ji(wo) and J2(2u>o)- Accurate<br />

determination of these spectral parameters<br />

as a function of temperature and the Larmor<br />

frequency (u>o) is necessary for testing various<br />

motional models.<br />

Both T\z and T\Q can be simultaneously measured<br />

by the Jeener-Broekaert (J-B) method [9]<br />

or the selective-inversion method [5]. However a<br />

separate experiment has to be performed for the<br />

deuterons on each labelled site in order to maximize<br />

their quadrupolar order for better signalto-noise<br />

considerations. The J-B sequence has<br />

been used to determine T\z and T\Q in several<br />

nematogens [10]. The pulse sequence was modified<br />

using an additional 45° pulse to produce<br />

the net effect of subtracting the equilibrium Moc<br />

signal from the J-B signal. Here we examine the<br />

modification of the J-B sequence (Table 1) to<br />

produce [11] a broadband excitation sequence<br />

(Table 2) in order to minimize the number of<br />

separate experiments required to give T\z an< l<br />

T\Q for various deuterons on an alkyl chain. Recently<br />

this broadband J-B excitation sequence<br />

was vised to create quadrupolar order with same<br />

efficiency on all the labelled sites in a liquid -j<br />

crystal [12].


Vol. 14,i No. 1-4<br />

\<br />

X<br />

-y<br />

X<br />

-y<br />

X<br />

-y<br />

X<br />

-y<br />

TABLE 1<br />

J-B Sequence with Phase-cycling<br />

Receiver<br />

7 3 Aqu T 2 4>s Aqu T Aqu T Phase<br />

-y y y X +•<br />

0<br />

-X X X y<br />

90<br />

-y y y -X<br />

0<br />

-X X X -y +<br />

y +<br />

90<br />

90<br />

X<br />

0<br />

-y<br />

90<br />

-x +<br />

0<br />

-x +<br />

0<br />

-y<br />

90<br />

X<br />

0<br />

y + 90<br />

y -y -y<br />

90<br />

X -X -X<br />

0<br />

y -y -y<br />

90<br />

)X -X -X<br />

0<br />

-y +<br />

-X<br />

y<br />

x +<br />

2. Experimental<br />

; The deuterium Txz and T\Q were measured on<br />

a home-built superheterodyne coherent pulsed<br />

|INMR spectrometer operating at 15.3 and 46.05<br />

^MHz for deuteron with a Varian 15 in electromagnet<br />

and a 7.1 Tesla Oxford superconductmagnet.<br />

The TT/2 pulse width of ca. 4.5 //s<br />

^produced by an Amplifier Research Model<br />

power amplifier. Pulse control, signal coltiou,<br />

Fourier transformation and data pro-<br />

«mg were performed using a General Electric<br />

1280 computer [10]. Both the J-B sequence and<br />

the broadband J-B excitation sequence (Figure<br />

1) were used with the appropriate phase-cycling<br />

[9] of RF and receiver phases to rid of the unwanted<br />

double quantum coherence (see Tables 1<br />

and 2). Several nematogens (5CB, MBBA and<br />

60 CB) were employed to check the spin relaxation<br />

times obtained by the two different multipulse<br />

sequences.<br />

135<br />

3. Results and Discussion<br />

In figure 2 we show a comparison of the<br />

J-B sequence (2(a)) and the broadband J-B<br />

excitation sequence (2(b)) for a set of partially<br />

relaxed spectra at 15.3 MHz in the nematic<br />

phase of 4-n-pentyl-dn-4'-cyano-2,3,5,6d4-biphenyl<br />

(5CB-di5). Minimal phase correction<br />

was required and the baseline of each spectrum<br />

has been corrected. In figure 2( a) the J-B<br />

sequence was set to maximize the quadrupolar<br />

order of the C4 methylene deuterons. In comparison<br />

with the J-B sequence, we found that<br />

maximum quadrupolar order was created for all<br />

the chain deuterons with r = 5/j.s (or an excitation<br />

bandwith of ca. 75 kHz). In table 3 we<br />

summarize the T\z and T\Q measured at 33.2°C<br />

in 5CB-di5 by the two different J-B sequences.<br />

As seen in this table, their values agree with<br />

each other for all the labelled sites within experimental<br />

errors.<br />

a)<br />

b)<br />

90, 67.5, 45, 45,<br />

9 ^ 4>


136 Bulletin of Magnetic Resonance<br />

a)<br />

1 32 4 5R<br />

R>-c5oll<br />

A 70m ^ h Hijft A, A M A 70m<br />

.50 ft W iiii A A I il 60<br />

A 45<br />

Figure 2 Plots of partially relaxed spectra at 33.2°C and 15.3 MHz. (a) Using the J-B sequence, (b) Using the<br />

broadband J-B excitation sequence.<br />

b)<br />

a)<br />

12 34 5<br />

0 0 DO<br />

6R<br />

R<br />

_ ! , , 1 1 , 1 1 1 r<br />

40000 20000<br />

= 36.2°c<br />

, . 1 1 . r—j-<br />

0 -20000 -40000 Hz<br />

Figure 3 (a) A typical deuterium NMR spectrum of 60CB-d2i showing the peak assignments; (b) A schematic<br />

diagram of a 60CB-d2i molecule.


Vol. 14, No. 1-4 137<br />

>N<br />

CO<br />

c<br />

0)<br />

D<br />

o<br />

20<br />

15 -<br />

10 -<br />

0<br />

o J1<br />

o<br />

oJ2<br />

o<br />

•o<br />

AA J<br />

AA<br />

J 2<br />

AA AA<br />

0<br />

O #)<br />

A Ai<br />

A j^\<br />

t<br />

o<br />

A A A<br />

A A A<br />

40 50 60 70<br />

T(°C)<br />

.We 4 Plots of spectral densities versus temperature in the nematic phase of 60CB-d2i. Open symbols are obtained<br />

^ e broadband J-B excitation sequence, while closed symbols by the J-B sequence. O and A denote data for C4<br />

[a '^6, respectively.<br />

1


138<br />

TABLE 3<br />

Comparison of T\ z and Tiq in ms measured<br />

at 15.3 MHz and 33.2°C<br />

for 5CB-d^<br />

12.6<br />

(13.5)<br />

10.4<br />

(10.1)<br />

c2<br />

26.9<br />

(25.8)<br />

21.5<br />

(21.8)<br />

c3<br />

30.0<br />

(30.7)<br />

22.8<br />

(26.4)<br />

ct<br />

55.4<br />

(48.9)<br />

51.9<br />

(46.1)<br />

159<br />

(137)<br />

107<br />

(104)<br />

Ring<br />

8.5<br />

(9.1)<br />

10.9<br />

(11.4)<br />

* Ti values in parentheses were obtained by the J-B sequence,<br />

while those without parentheses were obtained by the broadband<br />

J-B excitation sequence.<br />

Figure 3 shows the molecular structure of<br />

60CB-d2i and a typical deuterium NMR spectrum<br />

for this mesogen at 15.3 MHz. We have<br />

used both pulse sequences to measure spectral<br />

densities JI(JJJQ) and J2(2u;o) for all the labelled<br />

sites except the ring R!, because of excessive<br />

overlapping of its signal with that from the<br />

methyl (C6) at high temperatures. The agreement<br />

between the two methods are extremely<br />

good. As an example, we show in figure 4<br />

plots of spectral densities versus temperature<br />

for C\ and C&. Since the relaxation times of the<br />

methyl deuterons are much longer than those of<br />

the ring R! deuterons, the overlapped doublet<br />

signals can still be used to determine T\z and<br />

T\Q for the methyl deuterons as long as t is chosen<br />

larger than 40 ms in the pulse sequence.<br />

Finally MBBA-di3 has been studied at 15.3<br />

MHz using the J-B sequence [10]. For comparison<br />

we summarize in Table 4 the results obtained<br />

using the broadband J-B excitation sequence<br />

at 26°C and at 15.3 and 46 MHz. Thus<br />

both Ji(wo) and J2(2u>o) show frequency dependence.<br />

The frequency dependence of J2(2w0) is<br />

weaker; it is negligible for the methine deuteron<br />

(Co). Currently we are analyzing the temperature<br />

and frequency dependences of the measured<br />

spectral parameters in MBBA using models<br />

[3], [4] proposed for flexible mesogens.<br />

In conclusion, the relaxation data can be effectively<br />

obtained in liquid crystals by using the<br />

broadband J-B excitation sequence.<br />

Bulletin of Magnetic Resonance<br />

TABLE 4<br />

Measurements of Ji(u>o) an d ^2(2wo) ins ' for<br />

MBBA-di3 at 26°C (u>o = Larmor frequency<br />

in MHz) using broadband excitation<br />

15.3 46<br />

Co 49.8<br />

40.7<br />

14.6<br />

13.7<br />

1.73<br />

36.2<br />

21.0<br />

9.0<br />

7.5<br />

1.58<br />

15.3 46<br />

Acknowledgments<br />

24.9 27.7<br />

15.1 10.2<br />

6.7 5.4<br />

5.1 4.0<br />

1.05 0.98<br />

The financial support of the Natural Sciences<br />

and Engineering Council of Canada is gratefully<br />

acknowledged. We thank Ms. L. Friesen for her<br />

assistance in carrying out some experiments.<br />

References<br />

1. C.G. Wade, Annu. Rev. Phys. Chem. 28,<br />

47 (1977) and references therein.<br />

2. R.R. Void, in "Nuclear Magnetic Resonance<br />

of Liqud Crystals" J.W. Emsley,<br />

Ed., Reidel, Dordrecht (1985).<br />

3. A. Ferrarini, G.J. Moro and P.L. Nordio,<br />

Liq. Cryst. 8, 593(1990).<br />

4. R.Y. Dong, Phys. Rev. A 43, 4310 (1991).<br />

5. P.A. Beckmann, J.W. Emsley, G.R. Luckhurst<br />

and D.L. Turner, Mol. Phys. 50, 699<br />

(1983); 59, 97 (1986).<br />

6. R.Y. Dong and G.M. Richards, J. Chem.<br />

Soc. Faraday Trans. 88, ni the press.<br />

7. J.P. Jacobsen, H.K. Bildsor and K. Schumburg,<br />

J. Magn. Reson. 23, 153 (1976).<br />

8. S.B. Ahmad, K.J. Packer and J.M. Ramsden,<br />

Mol. Phys. 33, 857 (1977); R.R. Void<br />

and R.L. Void, J. Chem. Phys. 66, 4018<br />

(1977).<br />

9. R.L. Void, W.H. Dickerson and R.R. Void,<br />

J. Magn. Reson. 43, 213 (1981).<br />

10. R.Y. Dong and G.M. Richards, J. Chem.<br />

Soc. Faraday Trans. 84, 1053 (1988) and<br />

references therein.<br />

11. S. Wimperis, J. Magn. Reson. 86, 46<br />

(1990).<br />

12. G.L. Hoatson, J. Magn. Reson. 94, 152<br />

(1991).


Vol. 14, No. 1-4 139<br />

Introduction<br />

CARBON-13 RELAXATION MECHANISMS AND MOTIONAL STUDIES<br />

The carbon-13 relaxation rates of several symmetric<br />

top halomethane species, CH3I [1], CClBr3<br />

[2], CHBr3 [3] and three others that have been<br />

shown to be quasi-symmetric top, CH2Br2 [4,5],<br />

CH2C12 [6], CH2I2[7] were determined. These<br />

results were evaluated at two field strengths,<br />

2.1 and4.7Tesla, at UNC-Wilmington and East<br />

Carolina University, respectively, and at an identical<br />

set of temperatures at each site. With<br />

these data and several theoretical models we<br />

were able to determine, or calculate, the contributions<br />

for all plausible relaxation modes.<br />

IN SELECTED HALOMETHANE MOLECULES<br />

Art A. Rodriguez , Tim Davis<br />

Stokes-Einstein-Debye (SED) [8,9] theory of<br />

rotational diffusion and several variants to characterize<br />

the anisotropic reorientation of spheroids<br />

were also used to investigate for goodness of fit<br />

for hydrodynamically controlled rotational motion.<br />

In the hydrodynamic model, terms called<br />

"stick " and "slip" that attempt to describe<br />

the involvement of probe molecules and the<br />

solvent are exploited. The stick limit is normally<br />

encountered where the solute molecular radius is<br />

'•much larger than that of the solvent, while the<br />

^slip condition is approached as solvent radius<br />

lears or exceeds that of the solute molecule. The<br />

'-extended diffusion model is used to deterinertial<br />

properties of rotational diffusion<br />

Department of Chemistry, East Carolina University<br />

Greenville, NC 27858, USA<br />

and<br />

Lewis E. Nance.<br />

Department of Chemistry, UNC-Wilmington<br />

Wilmington, NC 28403, USA<br />

[4] from J-diffusion to free rotation of the molecule.<br />

Separation of R, 510 from R/ 01 allowed cal-<br />

culation of Tin, Q for Brand 'Jr.Rr for CClBr,,<br />

IBr<br />

C-Br<br />

CH2Br2, and CHBr3 by use of plots of derived<br />

sc 2<br />

T^vs. -s.A(O .<br />

Experimental Section<br />

Separation of Relaxation Mechanisms<br />

Tj values were obtained by the inversion recovery<br />

method using (Mo,cos6, T\), a three parameter<br />

fit for magnetization, M(x), as shown<br />

below.<br />

M{x) = Mo [ 1 - (1 - cos G) exp(-T/ri)] (1)<br />

Partitioning of relaxation rates into contributions<br />

by specific mechanisms was generally<br />

made with the following set of relationships in<br />

mind: When present, contribution to the relaxation<br />

rate by the scalar mechanism of the "second<br />

kind" (SC) is greater at 2.1 T than at 4.7<br />

T while the reverse will be true for chemical<br />

shift anisotropy (CSA). Relaxation is more efficient<br />

at higher temperatures for SC while it is<br />

less so for CSA. Dipole-dipole (DD) relaxation<br />

is more rapid at lower temperatures while that of<br />

spin-rotation (SR) is less. These contributions


140<br />

can be summed as follows where R, tot is the experimental<br />

rate of relaxation:<br />

JL<br />

tot<br />

CSA SC<br />

jCSA 4.<br />

(2)<br />

Integration of the C 13 signals with and without<br />

decoupling (pulse delay set to a full 10 Tt<br />

[10]) for the protonated species, CH3I, CH2I2,<br />

CHBr3, CH2Br2, and CH2C12 permitted determination<br />

of NOE values, r\, and allowed direct<br />

calculation of the H contribution to R, DD [10] by<br />

the following relationship.<br />

The evaluation of %C{C-H) [11] follows immediately<br />

from equation (4).<br />

(3)<br />

XC(C-H) = (y 2 cfHh 2 /2ii 2 r 6 CH)R° D {C-H) (4)<br />

The dipole-dipole contribution to C fromBr<br />

can then be calculated from the relationship below<br />

[12].<br />

•)DD, •>DD<<br />

R" U (C-H)IR\ )U {C-Bf) =<br />

2 (rCH) 6<br />

(5)<br />

A value for the quadrupole coupling constant,<br />

(2ne 2 Qq/h), orT, Q for Br allows a calculation<br />

of xc(C-Br) in equation (6) [13] or (7) [14] respectively.<br />

Rr~-<br />

3 2/+3<br />

40<br />

If<br />

(6)<br />

(7)<br />

R, DD subtracted from R, tot leaves R, 0 *" to be<br />

partitioned among the other contributions. The<br />

correlation time, xc, that was acquired from<br />

R, DD (C-H)canbe used in calculating Ri CSA by<br />

Bulletin of Magnetic Resonance<br />

equation (8) [12] if aper and cpar, the perpendicular<br />

and parallel components of the sheilding<br />

tensor, are known.<br />

(8)<br />

A classic case for relaxation by the CSA<br />

mode was that of CH3I [1]. We found that the<br />

CSA component to relaxation was significant<br />

since Rj tot was much greater at 4.7 T than at 2.1<br />

T and that this value decreased with an increase<br />

in temperature. Using the relationship<br />

CSA _ CSA<br />

(9)<br />

we utilized the fact that the variance in Rj tot<br />

going from 2.1 to 4.7 T must be due to the CSA<br />

mechanism since the SR mechanism is independent<br />

of field strength. A plot of R, tot vs. v 2 at<br />

303K has at the intercept R, tot = R, 1 and at this<br />

point the contribution of R, CSA to the value of<br />

R, tot vanishes. These results show the R, CSA<br />

contributions at 303K of 7.52 x 10" 3 (T, CSA =<br />

133s) at 4.7 T and 1.52 xlO" 3 (T, CSA =659s) at<br />

2.1 T.<br />

The Tc(C- H) values were also employed in<br />

conjunction with the J-extended diffusion theory<br />

computer program by McClung to calculate Xj,<br />

and its reduced form, x}, for symmetric tops<br />

[15,16,17 ]. The equation for R, is as<br />

follows [12]:<br />

Rf t = (2KlkT/h 2 )C 2 effxJ<br />

(10)<br />

A second program written by McClung relates<br />

R, SR to xj. This program requires values<br />

for Ix,Iz,Xj and the spin-rotation tensor components<br />

, Cx and Cz, which are not generally available.<br />

These spin-rotation components can<br />

be approximated by the method of Flygare [18],<br />

which uses the facts that op( n CO) = -259.5 |<br />

ppm and 8( 13 CO) - -182.2 ppm, and that the |


Vol. 14, No. 1-4 141<br />

difference between Adp and A5 for n C0 and<br />

the particular halomethane of interest, for<br />

instance, (Aap = ap(halomethane) - ap(CO),<br />

etc.) are the same . In general one must assume<br />

that Ac is zero. In this manner the relationship<br />

IZCZ ~ IXCX holds if Aa is small compared to<br />

aP.<br />

Scalar relaxation of the "second kind" is expected<br />

to be a prevalent relaxation mechanism in<br />

the Br bearing halomethanes. It is predominantly<br />

the 79 Br isotope instead of 81 Br that makes<br />

the greatest contribution to relaxation of C-13<br />

[19,20]. This is due to the fact that the<br />

(G)/-G)S) term in equation (7) is smaller for<br />

79 Br. Scalar coupling is greater at 2.1 T than at<br />

4.7 T since the Aw term is smaller in the lower<br />

magnetic field. R, sc also increases with temperature.<br />

A plot of Aco 2 vs. T, sc gives<br />

rrQ _<br />

IBr = Jm/b<br />

and<br />

J= \)m)<br />

(11)<br />

(12)<br />

where m and b are the slope and intercept, and<br />

N is the number of Br atoms attached to the<br />

halomethane carbon atom.<br />

Rotational Motion<br />

In the limit of extreme narrowing , the small-step<br />

diffusion theory [13,21,22] predicts the relationship<br />

between xc and the diffusion constants<br />

Dx and Dz in a symmetric top environment by<br />

the following equation [4]<br />

X - °- 3sin 2 6cos 2 e 0.7Ssin 4 6<br />

5DX+D2 2<br />

where 8 is the angle between the reorienting<br />

vector and the unique axis of rotation. This is<br />

the axis with the lowest moment of inertia. For<br />

example, this is the C3 axis in CH3I, but not in<br />

CHBr3 where the C-H vector is at 90° to this<br />

unique axis.<br />

( We have applied the J-extended diffusion<br />

.theory of Gordon [23], as expanded to symmet-<br />

ric tops byMcClung [15, 16], an inertial model,<br />

to generate Xj values using our previously<br />

determined xc(C-H), or xc, parameters. In this<br />

model molecules undergo a period of free rotation<br />

generating angular momentum, then upon<br />

hard molecular collision randomize both magnitude<br />

and direction of this vector. In the smallstep<br />

diffusion limit, Xj « xc • Here the reduced<br />

correlation time, x}, is found to obey the<br />

following equation:<br />

t} = Ty(^-)T«l (14)<br />

In addition, under these conditions, the rotational<br />

diffusion constants, DzandDx of equation (13)<br />

are related to Xy by<br />

Dz = Z-Xj and Dx = 4H (15)<br />

In the inertial model Xy = xc. The period of<br />

time for rotation through 1 radian, X/, as determined<br />

by the equipartition principle is given as<br />

[24]:<br />

y=(~)^ (16)<br />

Thus one is able to follow with this program<br />

by McClung the limits from small-step diffusion<br />

to free rotation where the rotations are inertially<br />

controlled.<br />

At the other extreme, molecular shape instead<br />

of inertial effects may dictate the mode of rotational<br />

diffusion of a molecular species. Extension<br />

of the SED theory to prolate and oblate<br />

spheroids has produced many equations as<br />

boundary conditions for stick and slip rotational<br />

motion are considered for Dx and Dz. For<br />

instance, Perrin [25] was able to show a relationship<br />

be tween the Stokes diffusion constant,<br />

Ds, by solving the Navier-Stokes equation<br />

assuming a stick boundary condition.<br />

Dt = A-Ds = A. kT (17)<br />

The factors fPiX and fPtZ are functions of p = bla<br />

(1 for oblate spheroid). The


142 Bulletin of Magnetic Resonance<br />

average molecular radius is used for r and the<br />

bulk viscosity for T\.<br />

Hu and Zwanzig [26] tackled the rotational diffusion<br />

problem by assessing the fact that the<br />

Perrin values , using the stick boundary condition,<br />

did not fit the experimental work well.<br />

They instead assumed a slip boundary condition<br />

in solving the Navier-Stokes equation. A<br />

separate value for a friction coefficient, £*,<br />

was obtained and tabulated for each axis ratio<br />

for prolate and oblate spheroids. Under this<br />

treatment, motion parallel to the top axis would<br />

experience no tangential stress and would have<br />

the motion of a free rotor.<br />

180<br />

(18)<br />

There would be some solvent displacement possible<br />

about the x-axis for perpendicular rotational<br />

motion.. The expression used here is<br />

Dx = ±LDs<br />

(19)<br />

where in our work /#z = ^*/8.<br />

Gillen [27] and Griffiths [28] note that the<br />

Dx motion in some molecules is diffusionally<br />

controlled and as such may be treated by<br />

the Gierer- Wirtz micro viscosity model [29]. The<br />

parallel motion, Dz, is treated in the slip boundary<br />

conditions, as essentially frictionless rotation.<br />

(20)<br />

The Gierer-Wirtz factor, fGw, is 0.1633 for neat<br />

liquids.<br />

Tanabe [30, 31] has extended the Hynes, Kapral,<br />

Weinberg (HKW) model [32] to include<br />

nonspherical molecules in solution and in neat<br />

form.<br />

(21)<br />

The a factor is zero for Dz but is fitted to a Hu-<br />

Zwanzig coefficient for Dx since even in the<br />

case of slip ((3=0) there is a finite friction coefficient.<br />

Results<br />

In CH2Br2 [4, 5] and CHBr3 [3 ], DD and SC of<br />

the second kind were shown to be the important<br />

relaxation modes. The dipolar contribution to R,<br />

fell off as the temperature was raised, while the<br />

scalar coupling rate increased. There was a<br />

larger value forRj at any particular temperature<br />

at 2.1 T than found at 4.7 T. R, SR and R, CSA were<br />

calculated and found to be negligible . The dipole-<br />

dipole relaxation contribution from Br to<br />

C-13 was calculated by equation (5) and was<br />

much less than the experimental error. In CQBr3<br />

[3 ] the relaxation results were clearly scalar.<br />

Average values of l JcBr and T\Br for CQBr3,<br />

CH2Br2, and CHBr3 are respectively: 75 Hz and<br />

4.3 x 10 6 ; 53 Hz and 4.3 x 10' 6 ; 49.7 Hz and<br />

8.03 x 10" 7 .<br />

The rotational diffusional motion of CH2Br2<br />

andCHBr3 as treated by the several models<br />

above, matches very closely that predicted by<br />

the J-extended diffusion model.<br />

The trend towards faster relaxation rates in<br />

CH3I [1] at 4.7 T compared to 2.1 furnishes positive<br />

evidence for the importance of CSA rather<br />

than SC in this molecular species. SRandDD<br />

are also found to significant.<br />

The approach taken here was that of Gillen [27]<br />

and Griffiths [28]. Motion about the top axis in<br />

CH3I fit closely the Gierer-Wirfz microviscosity<br />

model with an average difference of less than<br />

1% compared to experimental results. The excellent<br />

correlation indicates that the tumbling motion<br />

is hindered by viscous drag while the motion<br />

about the top axis is dominated by inertial<br />

effects.<br />

The modes of relaxation in CH2I2 [7] were as<br />

follows: DD; the dominant mechanism, SR; a<br />

very small contribution, and SC; about 18% at all<br />

temperatures.<br />

The J-extended diffusion model was a good<br />

predictor of rotational motion for both CH2I2<br />

andCH2Cl2[4,6].


Vol. 14, No. 1-4 143<br />

References<br />

1. Manuscript in preparation.<br />

2. Submitted to /. ofMolec. Spectroscopy.<br />

3. Manuscript in preparation.<br />

4. D. N. Dixon and A. A. Rodriquez, J. of<br />

Molec. Liq. 44, 79 (1990).<br />

5. P. B. Simcox, A. A. Rodriguez, L. E, Nance,<br />

The J. of Physical Chem. 96, (1992).<br />

6. A. A. Rodriguez, S. J. H. Chen and M.<br />

Schwartz,/. Magn. Reson. 74, 114 (1987).<br />

7. L.E. Nance, M. R. Nealey, and A. A.<br />

Rodriguez, Magn. Reson. Chem. 28, 11<br />

(1990).<br />

8. R. T. Boere 1 and R. G. Kidd, Annu. Rep.<br />

NMR<br />

Spectrosc, edited by G. A. Webb, Academic<br />

Press, New York, 13, 319 (1982).<br />

9. P. Debye, Polar Molecules, Dover, New<br />

York 1929.<br />

10. M. L. Martin, J.-J. Delpuech and G. L.<br />

Martin, Practical NMR Spectroscopy,<br />

Heyden, London (1980).<br />

11. T. C. Farrar,and E. D. Becker, Pulse and<br />

Fourier Transform NMR, Chap. 4,<br />

Academic Press, New York, (1971).<br />

12. E. D. Becker, High Resolution NMR:<br />

Theory and Applications, 2nd ed., Chap. 8.<br />

Academic Press, New York (1980).<br />

13. W. T. Huntress, J. Chem. Phys. 48, 3524<br />

(1968).<br />

14. A. Abragam, Principles of Nuclear<br />

Magnetism, Chap. 8. Oxford University<br />

Press, Oxford (1961).<br />

15. R. E. D. McClung, /. Chem. Phys. 57, 5478<br />

(1972).<br />

16. R. E. D. McClung, Adv. Mol. Relaxation<br />

Interact. Processes 10, 83 (1977).<br />

17. R. E. D. McClung, J. Chem. Phys. 51, 3842<br />

(1969).<br />

18. W. H. Flygare, J. Chem. Phys. 41,7903<br />

(1964).<br />

19. C. R. Lassigne and E. J. Wells, J. Magn.<br />

Reson. 27, 215 (1977).<br />

.20. T. C. Farrar, S. J. Druck, R. R. Shoup and E.<br />

D. Becker, J. Am. Chem. Soc. 94, 669<br />

(1972).<br />

1- H. Shmizu, J. Chem. Phys. 40, 754 (1964).<br />

22. D. E. Woessner, J. Chem. Phys. 37, 647<br />

(1962).<br />

23. R. G. Gordon,/. Chem. Phys. 44, 1830<br />

(1966).<br />

24. George C. Levy, Topics in Carbon-13 NMR<br />

Spectroscopy, Vol 1, Chapter 3,<br />

Wiley-Interscience, New York, (1974).<br />

25. E. Perrin, /. Phys. Radium, 5,497 (1934).<br />

26. C. M. Hu and R. Zwanzig, J. Chem. Phys.,<br />

60 4354 (1974).<br />

27. K. T. Gillen and J. E. Griffiths, Chem. Phys.<br />

Lett., 17 359 (1972).<br />

28. J. E. Griffiths, Chem. Phys. Lett., 21 354<br />

(1973).<br />

29. A. Gierer and K. Wirtz, Naturforsch, A8<br />

532 (1953).<br />

30. K. Tanabe, Chem. Phys., 31 319 (1978).<br />

31. K. Tanabe and J. Hiraishi, Molec. Phys., 39<br />

493 (1980).<br />

32. J. T. Hynes, R. Kapral and M. Weinberg, /.<br />

Chem. Phys., 69 2725 (1978).


144<br />

An Efficient Large Sample Volume System<br />

for Solid State NMR<br />

Ronald J. Pugmire, Yi Jin Jiang, Mark S. Solum and David M. Grant<br />

Department of Chemistry and Fuels Engineering<br />

University of Utah<br />

Salt Lake City, Utah 84112<br />

ABSTRACT<br />

U.S.A.<br />

An efficient large sample volume<br />

system has been developed to carry out<br />

MAS solid state NMR experiments. The<br />

system components are primarily<br />

zirconia and macor and no background<br />

1 3 C is observed. The stator design<br />

employs separate air bearing and drive<br />

systems and is run using dry air. At a<br />

bearing pressure of about 32 psi, the<br />

rotor can be spun in a stable manner<br />

from less than one hundred Hz (with a<br />

driving pressure of 5 psi) to 4.3 KHz (24<br />

psi driving pressure). This low gas<br />

pressure feature makes the system easy<br />

to operate. The volume of the rotor is<br />

1.8 cm 3 and it can hold 1.1 g of HMB.<br />

The S/N ratio obtained is a factor of 4.6<br />

better than the rotor previously<br />

designed and used in our laboratory<br />

(volume 0.6 cm 3 : 0.28g HMB). This<br />

increased sample size allows us to obtain<br />

the same S/N ratio in a MAS spectrum<br />

with a factor of 21 saving in<br />

spectrometer time. The time saving<br />

achieved with this rotor system is<br />

extremely useful in obtaining data on<br />

biological samples and polymers, and is<br />

especially useful when experiments on<br />

fossil fuels require the use of the Bloch<br />

delay technique. Examples of relevant<br />

applications will be discussed.<br />

Bulletin of Magnetic Resonance<br />

INTRODUCTION<br />

A large MAS spinner system has been<br />

designed to achieve spinning speeds<br />

from less than a hundred Hz to<br />

approximately 4.3 KHz with a relatively<br />

large sample volume of about 1.8 cm^.<br />

The reason for employing a large<br />

sample volume was to improve the S/N<br />

ratio in l3 c NMR experiments. This *is<br />

extremely important for obtaining<br />

quantitative 13 C Bloch decay (BD)<br />

spectra on coal samples with an<br />

acceptable amount of spectrometer<br />

time, due to their reasonably long T\<br />

relaxation times.<br />

In order to avoid a * 3 C background<br />

signal the spinner and stator system<br />

are constructed from zirconia and<br />

macor respectively both of which<br />

contain no carbon.[l] Attention has<br />

also been focused on increasing the S/N<br />

ratio by using a probe electronic circuit<br />

which optimizes the sensitivity of the<br />

l3 C observe channel.[2,3] The result is<br />

a system which gives a S/N ratio for a<br />

* 3 C CP/MAS spectrum of l.lOg sample of<br />

hexamethylbenzene (HMB) that has a<br />

factor of 4.6 improvement over the S/N<br />

ratio of a spectrum taken on our<br />

traditional spinner system, using 0.28g<br />

of HMB and the same spectral<br />

parameters.


t.!-.<br />

* * • •<br />

K<br />

Vol. 14, No. 1-4 145<br />

DESIGN AND CONSTRUCTION OF THE<br />

SPINNER<br />

A cross-sectional drawing of the<br />

spinner system is shown in Figure 1.<br />

There are eighteen flutes in the Up of<br />

the rotor. They are driven by a set of<br />

eight air jets (with a diameter of 0.025<br />

in.) that are mounted on the stator,<br />

along with a set of eight bearing holes<br />

of the same diameter located in the<br />

middle of the stator. These air holes are<br />

spaced at 45' intervals around the<br />

circumference of the stator. The<br />

clearance between the rotor and stator<br />

is approximately 0.002 in. The stator<br />

housing is machined from Kel-F which<br />

has a stepped surface, as shown in<br />

Figure 1, in order to separate the<br />

driving and bearing gases. This design<br />

leads to ease in assembling the system,<br />

and also increases operational safety.<br />

The spinning speed of the rotor as a<br />

function of the driving gas pressure,<br />

with the rotor containing 1.1 Og of HMB<br />

and a bearing gas pressure of 32 psi is<br />

shown in Figure 2. Figures 3 through 5<br />

are a series of photographs of this<br />

system. Due to the high rotational<br />

energy of the spinning system, the<br />

probe must either be in the bore of the<br />

magnet or behind an explosion screen<br />

when in operation.<br />

EXPERIMENTAL RESULTS<br />

Some typical results of data obtained<br />

with the large sample spinning system<br />

are shown below. Figure 7 shows the<br />

CP/MAS 13 C spectrum of 1.1 Og of HMB<br />

taken with only 12 scans and with 21 Hz<br />

line broadening applied. The S/N ratio<br />

of this spectrum is approximately 517.<br />

Figure 8 is a comparison of the spectral<br />

results using this new large rotor and<br />

the standard rotor system previously<br />

used in our laboratory. The size of the<br />

standard rotor (it holds about 0.28g of<br />

HMB) is fairly typical of those<br />

commercially available. In Figure 8a,<br />

CP/MAS spectrum is shown of l.lOg of<br />

HMB consisting of 20 scans with no line<br />

broadening. A S/N ratio of 130 is<br />

obtained. Figure 8b shows a 20 scan<br />

spectrum taken with the smaller<br />

standard rotor system which exhibits a<br />

S/N ratio of 28, again, with no line<br />

broadening. The comparison of Figures<br />

8a and 8b demonstrate that an<br />

improvement in the S/N ratio of about<br />

4.6 can be achieved which provides a<br />

factor of 21 in terms of spectrometer<br />

time. Figure 9 demonstrates that there<br />

are no l^C background signals in BD<br />

and cross polarization (CP) experiments.<br />

This is an important feature in the<br />

quantitative application of CP NMR<br />

experiments. This rotor/stator system<br />

has been used to carefully compare the<br />

BD and CP spectra of the eight coals in<br />

the Argonne Premium Coal Safhple<br />

Bank and these data are presented<br />

elsewhere.[4]<br />

ACKNOWLEDGMENT<br />

This work was supported through the<br />

Advanced Combustion Engineering<br />

Research Center at the University of<br />

Utah and Brigham Young University<br />

which is supported by the NSF, 23<br />

industrial firms and DOE/PETC.<br />

Additional support was provided<br />

through the Consortium for Fossil Fuel<br />

Liquefaction Science at the Univeristy<br />

of Kentucky, Auburn University,<br />

University of Pittsburgh, University of<br />

West Virginia, and the University of<br />

Utah.


1<br />

i<br />

146<br />

REFERENCES<br />

1. Ming Zhang and Gary E. Maciel; J.<br />

Magn. Reson., 85, 156 (1989).<br />

2. Yi Jin Jiang, Ronald J. Pugmire,<br />

and David M. Grant; /. Magn.<br />

Reson.,11, 485 (1987).<br />

3 Yi Jin Jiang, Warner R.<br />

Woolfenden, Anita M. Orendt,<br />

Karen L. Anderson, Ronald J.<br />

Pugmire, and David M. Grant;<br />

Poster MB, 30th ENC, Asilomar,<br />

California, April 1989.<br />

4. Mark S. S,olum, Yi Jin Jiang,<br />

Ronald J. Pugmire, and David M.<br />

Grant, submitted for publication.<br />

Figure 1: Cross-sectional view of the spinner system:<br />

I. Stator, 2. Rotor. 3. Turbine driving jet, 4. Housing<br />

- upper pan, 5. NMR coil terminal plug, 6. Bearing gas<br />

inlet plug, 7. Bearing gas orifice, 8. Light fiber mount.<br />

9. Screw for fixing light fiber. 10. Light fiber.<br />

II. Housing - lower pan, 12. Driving gas inlet plug.<br />

13. Bearing gas exit holes, 14. Mounting nng.<br />

5000<br />

Bulletin of Magnetic Resonance<br />

Driving Gas Pressure (PSD<br />

Figure 2" Plot of the spinning speed of :he<br />

arge volume rotor system as a function ot<br />

he driving pressure used. Bearing pressure<br />

vas 32 psi.<br />

Figure 3: Photograph of the large sample rotor (a) as<br />

compared to the typical rotor (b) and the /mm<br />

rotor supplied by Doty Scientific (c).


if<br />

&<br />

Figure 4: Photograph of the large volume rotor and the<br />

stator.<br />

Figure 5: Photograph of the spinner system.<br />

£' Figure 6: Photograph of the probe containing this<br />

spinning system.<br />

A<br />

B<br />

Figure 7: CP/MAS spectrum of l.lOg of HMB. taken with 12 sc<br />

5ms contact time. Is delay time. 21 Hz line broadening<br />

Spectrum at 25.12 MHz on a Bruker CXP-100.<br />

(b)<br />

CP<br />

2280 scans<br />

lms contact time<br />

1 sec. recycle time<br />

BD<br />

944 scans<br />

180 sec. recycle<br />

delay<br />

CP/MAS spectra of HMB taken at 25.12 MHz:<br />

(t)l«r|e volume (1.8 cm 3 ) system, l.lOg HMB.<br />

20 scans, 3m* contact time. Is delay, no line<br />

broadening, S/N ratio 130; (b) small (0.63 cm 3 )<br />

rotor system. 0.28| HMB, 20 scan*. 3ms contact<br />

time. Is delay, no line broadening. S/N ratio 28.<br />

330 230 130 30<br />

PPM from TMS<br />

-70<br />

Figure »• Comparison of the spectral<br />

Iineshapes obtained for BD and CP<br />

experiments on Pittsburgh #8 Argonne<br />

Premium Coal. A is the spectrum and B is the<br />

background obtained by taking the spectrum<br />

with an empty rotor under identical<br />

conditions to A.<br />

147


148<br />

Bulletin of Magnetic Resonance<br />

Magnetic Resonance Spectroscopic Investigations<br />

of Poly(p-Phenylene Sulfide/Disulfide), PPS/DS<br />

Douglas W. Lowman and David R. Fagerburg<br />

Research Laboratories, Eastman Chemical Company<br />

P. O. Box 1972, Kingsport, TN 37662-5150 USA<br />

Introduction<br />

Recently a new process for the<br />

synthesis of the commercially important<br />

semi-crystalline, engineering<br />

thermoplastic poly(p-phenylene<br />

sulfide/disulfide), PPS/DS, was presented<br />

[1]. This process, based on the reaction of<br />

p-diiodobenzene (DIB) and sulfur at<br />

elevated temperatures (Figure 1),<br />

generates a polymeric material containing<br />

para-substituted aromatic groups<br />

connected by sulfide and disulfide<br />

linkages.<br />

+ S<br />

230 to 300°C<br />

I + S »•<br />

Melt, air bleed<br />

230 to 250°C<br />

• —<br />

Melt, air bleed<br />

Pro-Polymer + '2<br />

PPS/DS<br />

Solid-Stats<br />

Build-Up<br />

(240°C, Nj)<br />

Figure 1. Synthesis of PPS/DS<br />

PPS/DS is not easily characterized by<br />

conventional magnetic resonance<br />

techniques operating under normal<br />

conditions due to its high thermal stability<br />

and solvent resistance. Wade and<br />

coworkers [2] recently reported on their<br />

application of high-temperature solution-<br />

state carbon-13 NMR to commercially<br />

available poly(phenylene sulfide), PPS We<br />

have extended this study by examining<br />

PPS/DS with epr spectroscopy as well as<br />

high-temperature solution-state carbon-13<br />

NMR.<br />

The PPS/DS chemistry is accomplished<br />

under conditions that are in certain<br />

respects considerably less stringent than<br />

those previously employed for other PPS<br />

synthetic routes [3,4]. Free radicals have<br />

been observed in PPS previously by<br />

several workers [5-8] but not under the<br />

conditions of synthesis. Since it was<br />

anticipated that radicals would play an<br />

important role in our chemistry, we<br />

examined this reaction process directly<br />

employing high-temperature epr<br />

spectroscopy under conditions of the<br />

reaction. In this report we present<br />

evidence for sulfur and carbon radical<br />

formation during PPS/DS synthesis.<br />

Experimental<br />

EPR spectra were collected on a Bruker<br />

ER 200D SRC EPR spectrometer operating<br />

on X-band (9.65 GHz) in the general<br />

temperature range of 230 to 300°C and<br />

using 3-mm O.D. glass epr tubes with a 200<br />

gauss sweep width, 100 sec sweep time,<br />

modulation frequency of 100 KHz and<br />

modulation of 40 milligauss. Measurement<br />

of g values was accomplished in a dual<br />

cavity with the resonance of<br />

diphenylpicrylhydrazyl (g value = 2.0037)<br />

as the reference.<br />

PPS/DS synthesis in the epr tube was<br />

accomplished in a manner similar to that


Vol. 14, No. 1-4 149<br />

20 G<br />

Figure 2. EPR spectra of perdeuterated PPS/DS<br />

at (A) 235°C, g(iso) = 2.0073, and (B) 290°C, g =<br />

2.0042.<br />

Figure 3. EPR spectra of melting<br />

perdeuterated PPS/DS<br />

Table 1. Linkage and end group<br />

species found in PPS/DS. Letters<br />

and numbers are used in Figure 4<br />

for resonance assignments.<br />

B<br />

D<br />

G PPS<br />

Table 2. Chemical shift assignments<br />

in ppm for iodo-terminated PPS<br />

(structure shown in Figure 5)<br />

Cl<br />

C2<br />

C3<br />

C4<br />

C5<br />

C6<br />

92.3<br />

138.4<br />

132.7<br />

135.7<br />

135.1<br />

131.8


150<br />

A1<br />

140 138 136 134 132<br />

Chemical Shift, ppm<br />

130 128<br />

Figure 4. Solution-state carbon-13 NMR spectrum of PPS/DS<br />

previously reported [1]. Perdeuterated<br />

PPS/DS (PPS-d4) (degree of polymerization,<br />

DP, of 25) was synthesized in a manner<br />

similar to that previously reported [1].<br />

Iodo-terminated PPS (DP of 9) was prepared<br />

according to the previously reported<br />

method [1] but employing a 53 mole %<br />

excess of DIB.<br />

High-temperature solution-state<br />

carbon-13 NMR employed a hightemperature<br />

10-mm probe system from<br />

Doty Scientific, Inc. (Columbia, SC) on a<br />

JEOL Model GX-400 NMR spectrometer.<br />

Samples were dissolved in N-cyclohexylpyrrolidinone<br />

(CHP) under a blanket of<br />

argon and examined at either 230 or 260°C.<br />

Field/frequency stabilization was<br />

accomplished with glyme-d6 ne ld<br />

concentrically in a 5-mm capillary.<br />

Solid-state NMR spectra were collected<br />

on a Varian XL-300 NMR spectrometer.<br />

Discussion<br />

EPR Spectroscopy. During the<br />

synthesis of PPS/DS in the epr<br />

spectrometer, two distinctly different<br />

resonances are observed (Figure 2). At<br />

235°C a triplet resonance is observed, g<br />

Values for the components of the triplet<br />

resonance are 2.0033 ± 0.0002 (gi), 2.0072 ±<br />

0.0001 (g2) and 2.0113 ± 0.0001 (53). The<br />

Bulletin of Magnetic Resonance<br />

126 124<br />

isotropic g value is 2.0073 This resonance<br />

is assigned to a sulfur-centered radical<br />

cation. The isotropic g value is in excellent<br />

agreement with g values of 2.0079 and<br />

2.0076 reported by Murray and coworkers<br />

[9] for a sulfur-centered radical cation in<br />

two ASF5-doped PPS samples.<br />

Above 280°C, the triplet resonance is<br />

replaced by a singlet resonance with a g<br />

value of 2.0042. This resonance is assigned<br />

to an aryl radical.<br />

It is interesting that the melting<br />

process for PPS/DS can be easily monitored<br />

by epr (Figure 3). Using PPS-d4, two<br />

overlapping triplet resonances are<br />

observed at 230°C. The triplet resonance<br />

arises from hyperfine coupling between<br />

the carbon-centered radical and the<br />

deuterium attached to the carbon. The two<br />

triplet resonances arise from free radicals<br />

being in two different environments. At<br />

230°C, the larger triplet resonance arises<br />

from radicals in the solid phase while the<br />

smaller triplet resonance arises from<br />

radicals in a melted phase. As the<br />

temperature is increased, the solid-phase<br />

component decreases in intensity until it<br />

disappears just above 280°C. The DSC<br />

melting point for this polymer is 279°C.<br />

It is unlikely that the carbon radical<br />

observed here is mechanistically involved<br />

in the synthesis of linear para-substituted<br />

PPS/DS since these radicals would produce


Vol. 14, No. 1-4<br />

J8 3 2<br />

180 160 140 120 100<br />

CHP<br />

Carbonyl<br />

160 140<br />

Chemical Shift, ppm<br />

151<br />

Solid-State NMR Spectrum<br />

Solution-State NMR Spectrum<br />

120 100<br />

Figure 5. Solid-State and Solution-State Carbon-13 NMR<br />

spectra of an Iodo-Terminated PPS<br />

1


152<br />

aromatic substitution other than parasubstitution.<br />

Our PPS/DS chemistry has<br />

been shown to produce exclusively parasubstituted<br />

polymer [10].<br />

NMR Spectroscopv. Solution-state<br />

carbon-13 NMR at 260°C in CHP provides<br />

high-resolution NMR spectra of PPS/DS<br />

that enable an analysis for various linkage<br />

groups, such as sulfide and disulfide, as<br />

well as numerous end groups of the type p-<br />

X-phenyl, where X = H, SH or I. Resonances<br />

for the species in Table 1 are assigned in<br />

the spectrum shown in Figure 4.<br />

Assignments are based on comparison with<br />

spectra for model compounds and analysis<br />

of several PPS/DS spectra. Species A, B and<br />

C were assigned by comparison of spectra<br />

from PPS or PPS/DS with and without .<br />

excess sulfur as well as model compounds<br />

such as diphenyl sulfide and diphenyl<br />

disulfide. Based on the thermal history of<br />

the synthesis, excess sulfur is expected to<br />

be present only as disulfide linkages.<br />

Species D was assigned by comparison with<br />

the spectrum of thianthrene. Species E<br />

assignments are discussed below. Species F<br />

and G were assigned based on previous<br />

assignments [2]. Several resonance<br />

assignments in Figure 4, labelled with "?",<br />

have not been made conclusively.<br />

Complete assignments await synthesis of<br />

appropriate model compounds.<br />

Solution-state carbon-13 NMR of an<br />

iodo-terminated PPS polymer (Species E,<br />

Table 1) at 230°C in CHP clearly shows<br />

sufficient detail to confirm the degree of<br />

polymerization of 9 and to assign all<br />

resonances in the terminal iodophenyl<br />

group (Figure 5, Bottom). The assignments<br />

for these carbons are shown in Table 2. By<br />

comparison, the room temperature solidstate<br />

NMR spectrum of this polymer is not<br />

very informative (Figure 5, top).<br />

Conclusions<br />

PPS/DS does not lend itself to study by<br />

conventional magnetic resonance<br />

techniques operating under normal<br />

conditions. Through the use of several<br />

magnetic resonance techniques, we have<br />

begun to understand some of the<br />

limitations to studying PPS/DS by these<br />

techniques and to enhance our<br />

understanding of the chemistry of this<br />

Bulletin of Magnetic Resonance<br />

unusual polymer. We have presented here<br />

data on the first direct observation of free<br />

radicals generated under conditions of<br />

synthesis of PPS/DS and have expanded the<br />

applicability of the high-temperature<br />

solution-state carbon-13 NMR technique.<br />

Solution-state NMR has provided evidence<br />

for several linkage groups and end groups<br />

in PPS/DS enabling us to better appreciate<br />

the chemistry of this polymer. Room<br />

temperature solid-state NMR has not<br />

proven to be useful in these studies.<br />

References<br />

1. Rule, M.; Fagerburg, D. R.; Watkins, J. J.;<br />

Lawrence, P. B.; Makromol. Chem.,<br />

Rapid Commun. 1991. 12, 221<br />

2. Wade, B.; Abhiraman, A. S.; Wharry, S.;<br />

and Sutherlin, D.; J. Poly. Sci: Pt B:<br />

Polym. Phys., 28 1233 (1990)<br />

3. Fahey, D. R.; Ash, C. E. Macromolecules<br />

1991,24, 4242<br />

4. Lopez, L. C; Wilkes, G. L. J. Macromol.<br />

Sci., Rev. Macromol. Chem. Phys. 1989,<br />

C29, 83<br />

5. Ma, C.-C. M.; Hsiue, L.-T.; Liu, W.-L. J.<br />

Appl. Polm. Sci. 1990,39, 1399<br />

6. Hill, D. J. T.; Hunter, D. S.; Lewis, D. A.;<br />

O'Donnell, J. H.; Pomery, P. J. Radiat.<br />

Phys. Chem. 1990, 36, 559<br />

7. Kreja, L.; Rozploch, F.; Warszawski, A.<br />

Angew. Makromol. Chem. 1988,160, 163<br />

8. Kispert, L. D.; Files, L. A.; Frommer, J. E.;<br />

Shacklette, L. W.; Chance, R. R. J. Chem.<br />

Phys. 1983,57, 4858<br />

9. Murray, D. P.; Kispert, L. D.; Frommer, J.<br />

E. J. Chem. Phys. 1985, 83, 3681<br />

10. Rule, M.; Fagerburg, D. R.; Watkins, J.<br />

J.; Lawrence, P. B.; Zimmerman, R. L.;<br />

Cloyd, J. D.; Makromol. Chem.,<br />

Macromol. Symp. 1992,54/55, 233


Vol. 14, No. 1-4 153<br />

Application of 2-D HETCOR NMR to Investigate Polymer Blend Heterogeneity<br />

Introduction:<br />

Most NMR techniques for determining<br />

domain structures in polymeric systems are<br />

based upon the classic Goldman-Shen<br />

experiment [1]. Domain sizes are calculated<br />

from the time for spin diffusion to transfer<br />

magnetization from one region of the sample<br />

to another. These experiments generally<br />

require resolution of the individual<br />

components in the proton spectrum,<br />

although recent experiments [2] demonstrate<br />

that, in some cases where there is<br />

inadequate proton chemical shift resolution,<br />

it is possible to follow the evolution of spin<br />

diffusion when the constituents have significant<br />

differences in proton lineshapes.<br />

An alternative approach to analyzing<br />

domain structures entails application of<br />

heteronuclear ^C- 1 !! NMR. Of the several<br />

techniques introduced to measure heteronuclear<br />

correlated spectra in the solid state,<br />

a particularly effective method has been proposed<br />

by Burum and Bielecki [3]. In the<br />

basic 2-D experiment, crosspeaks occur primarily<br />

for carbon-proton distances of less<br />

than ~3 A. By incorporation of a spin<br />

diffusion evolution time, crosspeak intensities<br />

reflect longer range couplings, thus<br />

I &.. 1- Goldman, M.; Shen, L. Phys. Rev. 1966,<br />

^ 144,^21. J<br />

U. 2 - C^P^ll, G.C.; VanderHart, D.L. J.<br />

Magn. Reson. 1992,96,69-93.<br />

\ 3 - Burum, D.P.; Bielecki, A. J. Magn.<br />

Reson. 1991,94,645-652.<br />

S. Kaplan<br />

Xerox "Webster Research Center<br />

800 Phillips Road 1004-39D<br />

Webster, NY 14580, USA<br />

enabling conformational and domain size<br />

analyses [4]. This latter experiment consists<br />

(Fig. 1) of an evolution period during which<br />

homo- and heteronuclear interactions are<br />

suppressed by simultaneous application of<br />

BLEW-24 ( J H) and BB-24 ( 13 C) pulse<br />

sequences, a waiting period without rf<br />

during which exchange of magnetization<br />

(spin diffusion) takes place, a mixing period<br />

for isotropic cross polarization transfer of<br />

proton magnetization to carbons utilizing<br />

WIM-24 sequences on both nuclei, and<br />

finally, carbon acquisition with proton<br />

decoupling. In this paper we examine the<br />

90° 63° 90° 90°<br />

BLEW-24 WIM-24<br />

13C BB-24 WIM-24<br />

Figure 1. The HETCOR Experiment<br />

cw<br />

Decoupling<br />

applicability of the HETCOR technique to<br />

investigate domain sizes in a two component<br />

blend.<br />

Results and Discussion:<br />

We have chosen for the present study a blend<br />

of an aromatic diamine, N,N'-diphenyl-N-<br />

4. Simpson, J.H.; Ruggeri, G.; Rice, D.M.;<br />

Karasz, F.E., submitted.


154<br />

N'-bis(3-methylphenyl)-[l,l'-biphenyl]-4,4'diamine<br />

(TPD) in a bisphenol A polycarbonate<br />

matrix. Pure TPD is highly crystal-<br />

TPD<br />

line, with a melting point of Tm = 167 °C and<br />

a glass transition temperature of Tg=63 °C<br />

[5]. A well mixed 50/50 blend with polycarbonate<br />

(Tg=137 °C) cast from methylene<br />

chloride is amorphous and shows a single Tg<br />

of ~89 °C. Since the aromatic region of the<br />

13 C spectrum of the blend has severe overlap,<br />

we have focused our analysis on the aliphatic<br />

region, where the individual component<br />

methyl resonances are resolved.<br />

Figure 2 shows contour plots of the<br />

carbon aliphatic region in the two dimensional<br />

HETCOR spectra as a function of the<br />

spin diffusion mixing time. The 13 C peaks at<br />

22 ppm and 32 ppm are from TPD and polycarbonate<br />

methyl carbons, respectively.<br />

Crosspeaks in the 20 us mixing time plot are<br />

due primarily to short range directly bonded<br />

carbon-proton couplings. However, with<br />

increasing mixing times, longer range correlations,<br />

e.g., between the methyl carbons<br />

and aromatic protons, intensify. Ultimately,<br />

for very long spin diffusion times, the<br />

relative signal intensities for the contours<br />

associated with a particular carbon will<br />

reflect quantitatively the chemical shift distribution<br />

of protons that are within an<br />

effective range of spin diffusion from the protons<br />

directly bonded to that carbon. Figure 3<br />

shows the volume integral fraction of<br />

aliphatic protons for the polycarbonate and<br />

TPD methyl carbons as a function of spin<br />

diffusion time. Both curves asymptote to an<br />

5. Prest, W.M., unpublished data.<br />

Bulletin of Magnetic Resonance<br />

40 30 20 40 30 20<br />

ppm (13Q<br />

ppm OH)<br />

Figure 2. HETCOR contour plots of the aliphatic<br />

carbon region of a 50/50 (wt/wt) blend<br />

as a function of spin diffusion time.<br />

aliphatic volume fraction of 0.3 (point c),<br />

which corresponds exactly to the fraction of<br />

protons in the entire sample that are from<br />

methyl groups. The results are very<br />

different from measurements on a physical<br />

mixture of the same components, where<br />

separate asymptotes, corresponding to the<br />

individual aliphatic proton fractions, are<br />

observed for each component (polycarbonate<br />

at point a and TPD at point b). Also shown<br />

in Figure 2 is the difference response<br />

between these two curves. The short term<br />

initial rise can be attributed to<br />

intramolecular and the long term decay to<br />

intermolecular proton spin diffusion. From<br />

the ratio of these rates (-10) and the known<br />

intramolecular proton-proton distances (0.3<br />

nm) the intermolecular distances can be


Vol. 14, No. 1-4 155<br />

Aliphatic Fraction<br />

1.0<br />

0.8'<br />

0.6<br />

0.4 -<br />

: •fW*t<br />

0.2 r.<br />

•* * •<br />

—O— PC Icompatible 1<br />

-••-TPD J Blend i<br />

D PC 1 Physical z .<br />

• TPD J Blend j<br />

• 4i<br />

m :<br />

(a)<br />

(c)<br />

.. Difference (X5) =<br />

(b)<br />

0<br />

0 0.5 1.0 1.5 2.0<br />

Spin Diffusion Time (ms)<br />

"I I 1 1 1 1 1 I • 1 1 i • 1 > • i i f 1 I I i t 1 i i i i 1 i i I i 1 f i i i 1 11 i r<br />

Figure 3. TPD and polycarbonate methyl<br />

carbon correlations with protons.<br />

estimated to be about 1 nm (0.3XV10),<br />

indicative of intimate molecular level<br />

mixing. This treatment of the data permits<br />

estimation of interdomain separation<br />

without knowledge of the spin diffusion<br />

constant, which is only assumed to be equal<br />

in both domains.<br />

An alternative approach to analyzing the<br />

results of this experiment is shown in Figure<br />

4. Plotted here is a measure of the similarity<br />

(mean square difference) of the proton slices<br />

for each of the methyl carbon resonances as a<br />

function of spin diffusion time. This plot is a<br />

direct measure of the diffusion of spin order<br />

between the two blend constituents. In spite<br />

of scatter, the best fit for the data is a single<br />

exponential with a decay rate of 350 ms.<br />

Employing the standard equation for<br />

determining domain sizes from spin diffusion<br />

i- times, r = (nDt)- 1/2 , where n represents the<br />

domain dimensionality, t is the measured<br />

if', spin diffusion time, and D is the spin diffusion<br />

constant, typically 5X10" 12 cm2 sec" 1 ,<br />

• domain sizes in the range of 0.8-1.4 nm are<br />

• estimated.<br />

0.2 r<br />

0 -<br />

0 0.5 1.0 1.5 2.0<br />

Spin Diffusion Time (ms)<br />

Figure 4. Intermolecular correlations between<br />

TPD & polycarbonate methyl protons.<br />

Heating the blend overnight at 110 °C induced<br />

some phase separation, as evidenced<br />

by a change in appearance from clear to<br />

opaque. HETCOR diffusion spectra of this<br />

sample showed negligible differences from<br />

the data of the clear films. We conclude that<br />

a very small amount of phase separation<br />

occurred, e.g., near impurities or on the<br />

surface only.<br />

In summary, it is shown that the two<br />

dimensional heteronuclear correlated spin<br />

diffusion experiment is capable of extending<br />

the range of applicability of domain size<br />

measurements to blends comprised of<br />

components that are structurally very<br />

similar.<br />

Acknowledgements:<br />

The author wishes to thank J. Yanus for<br />

supplying the blend sample used for the<br />

current study and D. Rice (Univ. of<br />

Massachusetts) and D. Burum (Bruker<br />

Instruments, Inc.) for helpful discussions on<br />

the HETCOR experiment.


156 Bulletin of Magnetic Resonance<br />

NEW HIGH RESOLUTION NMR STUDIES<br />

IN<br />

POLYCRYSTALLINE TETRACYANOQUINODIMETHANE<br />

M.T. Nunes<br />

ICTPOL/CFMC-INIC, Av.Prof. Gama Pinto 2, 1699 Lisboa Codex, Portugal;<br />

Dep. Quimica, ICEN, LNETI 2686 Sacavem Codex Portugal<br />

A. Vainrub<br />

Institute of Chemical and Biophysics, Estonian Academy of Sciences,<br />

200001 Tallinn, Estonia<br />

M. Ribet, F. Rachdi, P. Bernier<br />

Groupe de Dynamique des Phases Condensees, U.S.T.L., 34095 Montpellier<br />

Cedex 05, France<br />

M. Almeida<br />

Dep. Quimica, ICEN, LNETI 2686 Sacavem Codex, Portugal<br />

and<br />

G.Feio<br />

ICTPOL/CFMC-INIC,Av.Prof. Gama Pinto 2,1699 Lisboa Codex,Portugal<br />

Introduction<br />

Tetracyanoquinodimethane (TCNQ) is<br />

a well-known electron acceptor<br />

that has been widely used to<br />

prepare charge transfer complexes,<br />

some of them exhibiting very high<br />

electrical conductivity; locally<br />

resolved "C Knight shifts [1] were<br />

measured, accordingly [2].<br />

NC<br />

NC<br />

CH==CH<br />

CH=CH<br />

TCNQ<br />

\c==c.<br />

CN<br />

CN<br />

In a static magnetic field of<br />

4.7T, high resolution NMR studies<br />

were performed on neutral TCNQ<br />

using different polycrystalline<br />

samples: natural isotopic<br />

abundant, selectively enriched on<br />

13 15<br />

C isotope content and N enriched<br />

[3]. Consequently, a complete<br />

assignment of the 13 C resonances<br />

was reported; the principal<br />

components of the chemical<br />

shielding tensor, the shielding<br />

anisotropy and the shielding<br />

asymmetry factor were also<br />

obtained for the 13 C nuclei in the<br />

fragments C-(CN)2, by graphical<br />

analysis and by a computer fit of<br />

the line intensities of CP/MAS<br />

spectra [4]. However, large<br />

uncertainties were reported on CN<br />

groups data. 14 N quadrupole effects<br />

were observed on 13 C spectra of CN<br />

groups and the "N quadrupole<br />

coupling constant was obtained (*"<br />

3.84±0.12MHz). Using different<br />

experimental conditions (like a<br />

static magnetic field of 7.0T) we<br />

highlight here<br />

cyano groups.<br />

new<br />

3<br />

C data on


Vol. 14, No. 1-4 157<br />

RESULTS and DISCUSSION<br />

When applying the numerical method<br />

[4] to the spectra obtained at<br />

4.7T, two different data sets were<br />

found for the principal components<br />

of the 13 C shielding tensor in TCNQ<br />

cyano groups [ 3 ]; only one of the<br />

corresponding asymmetry factors<br />

(t]= 0 and 0.4) was characteristic<br />

of an axially symmetric system.<br />

Aiming to a clarification of this<br />

problem, the improvement of the<br />

experimental data for the<br />

application of the numerical<br />

method seems to be the next step;<br />

in particular, the acquisition of<br />

13 C spectra displaying an higher<br />

number of spinning sidebands and a<br />

better signal/noise is required.<br />

These conditions are fulfilled by<br />

running the spectra at higher<br />

static magnetic field but<br />

selecting the same MAS rate and a<br />

longer recovering delay (600s);<br />

Figure 1 shows typical 13 C CP/MAS<br />

high resolution NMR spectra of<br />

selectively enriched TCNQ to 99%<br />

13 C isotope content on CN groups.<br />

Again, two data sets are obtained<br />

by a computer fit of the<br />

intensities of the lines, one of<br />

them corresponding to an axial<br />

symmetry for the cyano groups, in<br />

close agreement with that obtained<br />

at 4.7T [3 ] :<br />

ffll<br />

ppm<br />

228<br />

210<br />

a22<br />

ppm<br />

190<br />

208<br />

ppm<br />

-80<br />

-80<br />

Aa<br />

ppm<br />

-289<br />

-289<br />

0.2<br />

0.0<br />

However, regarding the other set,<br />

a much lower deviation from axial<br />

symmetry is now obtained (axx-o22<br />

equal to 38ppm instead of 69ppm).<br />

Indeed, Clayden and co-workers<br />

already pointed out that similar<br />

simulated MAS spectra<br />

Too 1«5 us ito iio M o -So<br />

FIG.l. 13 C CP/MAS high resolution<br />

NMR spectra of TCNQ cyano groups<br />

obtained at 75.47 (on the top) and<br />

50.13 MHz (on the bottom) with the<br />

same spinning rate: 3.01 kHz.<br />

are obtained when the chemical<br />

shift tensor is axial or nearaxial<br />

(ii0.2. On the<br />

light of these conclusions the<br />

present results are definitely<br />

acceptable.<br />

For the acquisition of the<br />

spectra previously reported [ 3 ],<br />

too short relaxation delays were<br />

used possibly inducing a distorted<br />

lineshape for the sidebands<br />

envelope; anisotropic molecular<br />

motion would influence distinctly<br />

the relaxation times of magnetic<br />

equivalent 13 C nuclei on TCNQ<br />

molecules perpendicular and


158<br />

parallel to the static magnetic<br />

field. In fact, if 1.5s is now<br />

used instead of 600s for that<br />

duration period, at least three<br />

fits are obtained (depending on<br />

the sidebands used) with much<br />

higher residual values; comparing<br />

the data sets in this case, large<br />

uncertainties are observed for all<br />

the principal components of the<br />

shielding tensor.<br />

The study of 13 C TCNQ relaxation<br />

in a wide temperature range are<br />

now in progress and will certainly<br />

contribute to the understanding of<br />

these results; also, additional<br />

insight on the role of TCNQ on<br />

charge transfer complexes (like<br />

DMTM(TCNQ)2 [2]) will be obtained.<br />

To obtain the 14 N quadrupole<br />

coupling constant from X3 C CP/MAS<br />

spectra, run at 7.0T, the<br />

procedure previously described was<br />

used [3]; a C-N dipolar splitting<br />

close to 230 Hz should be measured<br />

now, which is in fact the case.<br />

References<br />

1 M.Mehring and J.Spengler,<br />

Phys.Rev.Lett. ,53., 2441 (1984)<br />

2 for a recent nmr study on these<br />

materials see: F.Rachdi, T.Nunes,<br />

M.Ribet, P.Bernier, M.Helmle,<br />

M.Mehring and M.Almeida,<br />

Phys.Rev.B, 45(14), 8134 (1992)<br />

3 T.Nunes, A.Vainrub, M.Ribet,<br />

F.Rachdi, P.Bernier and M.Almeida,<br />

J.Chem.Phys.,96(11) , 8021 (1992)<br />

4 J.Herzfeld and A.E.Berger,<br />

J.Chem.Phys., 73_, 6021 (1980)<br />

5 N.J.Clayden, C.M.Dobson, L.Lian<br />

and D.J.Smith,<br />

J.Magn.Reson.,69/476 (1986)<br />

Bulletin of Magnetic Resonance


Vol. 14, No. 1-4 159<br />

USE OF NMR RELAXATION<br />

MEASUREMENTS TO DERIVE THE<br />

BINDING SITE OF PLASTOCYANIN<br />

IN COMPLEXES WITH<br />

CYTOCHROME-F AND C<br />

Sandeep Modi 1 , Ewen McLaughlin 1 , Derek S. Bendall 1 ,<br />

S. He 2 and J.C. Gray 2<br />

Departments of Biochemistry 1 and Plant Sciences 2<br />

University of Cambridge, Cambridge, CB2 1QW<br />

England (U.K.)<br />

1 Introduction<br />

Plastocyanin (PC) is a small (Mr 10<br />

500), 'blue' copper protein which<br />

transfers electrons from cytochrome<br />

/ to the primary electron<br />

donor of photosystem I (P700) in<br />

the photosynthetic electron transport<br />

chain. It reacts rapidly with<br />

cytochrome / in vitro and also<br />

with several other cytochromes,<br />

although somewhat more slowly<br />

[1].. The crystal structures of both<br />

oxidized and reduced poplar<br />

plastocyanin have been determined<br />

and show that redox changes<br />

cause only small changes at the Cu<br />

site, leaving the structure of the<br />

rest of the molecule essentially<br />

unchanged [2,3]. The Cu atom and<br />

its ligands (His-37, Cys-84, His-87<br />

and Met-92) are located in a<br />

hydrophobic pocket near one end<br />

of the molecule (the "northern"<br />

end) such that only the imidazole<br />

ring of His-87 (the northern<br />

histidine) is accessible to solvent<br />

(Fig 1).<br />

Studies with small molecules,<br />

such as [Fe(CN)6] 3 " and<br />

[Co(phen)3]3+, have identified two<br />

reaction sites on plastocyanin; one<br />

close to the copper ligand His-87<br />

at the northern hydrophobic patch<br />

and the other close to the more<br />

remote Tyr-83 at the eastern<br />

acidic patch [4-8]. Chemical modification<br />

of the acidic amino acid<br />

residues, inhibitory effects of<br />

small molecules and ionic strength


160 Bulletin of Magnetic Resonance<br />

effects on electron transfer<br />

strongly suggest that cytochrome /<br />

binds at the remote eastern site<br />

[9,10]. However the pathway of<br />

electron transfer from cytochrome<br />

/ to the copper site in plastocyanin<br />

C<br />

L12 H87<br />

E60<br />

Fig. 1 : Structure of Plastocyanin. The Cu<br />

ligands and the side-chains of Tyr83 and the<br />

residues of the eastern acidic patch are shown<br />

here with the coordinates of poplar<br />

plastocyanin from the Brookhaven Database,<br />

except for substitution of Glu45 (as in pea<br />

plastocyanin) for Ser45. ^-Strands are<br />

shaded.<br />

is not clear. One possibility is that<br />

cytochrome / binds to the negative<br />

charges of the eastern acidic patch<br />

and donates electrons via a<br />

tunneling pathway starting at<br />

Tyr-83 [5,11]. However, the<br />

distance from Tyr-83 to the<br />

copper ion is approximately 12A,<br />

whereas at the northern site the<br />

copper ion is only 6A from the<br />

surface of the molecule [11].<br />

The recent development of<br />

expression systems for the small<br />

blue copper protein, plastocyanin<br />

[12-15], has provided a valuable<br />

tool for study of the molecular<br />

details of its interaction with its<br />

native reaction partners (cytochrome<br />

/ and photosystem I in the<br />

photosynthetic electron transport<br />

chain). To examine the pathway of<br />

electron transfer from cytochrome<br />

/ to plastocyanin we have altered<br />

Tyr-83 to Phe-83 and Leu-83 by<br />

site-directed mutagenesis of the<br />

pea plastocyanin gene [13-14].<br />

Measurements of binding constants<br />

and electron transfer rates<br />

indicate, that Tyr-83 not only<br />

forms part of the main route of<br />

electron transfer from cytochrome<br />

/ to plastocyanin but is also involved<br />

in binding to cytochrome /.<br />

Nuclear magnetic resonance<br />

spectroscopy has been established<br />

as a very convenient and effective<br />

technique for structural studies of<br />

proteins. For haemproteins, the<br />

characteristic hyperfine-shifted<br />

NMR spectrum of a paramagnetic<br />

haemprotein carries the signature<br />

of the electronic and structural<br />

properties of the haem group.<br />

Measurements of spin - lattice


Vol. 14, No. 1-4 161<br />

relaxation (Tj) and spin-spin<br />

relaxation (T2) times provides<br />

useful methods for the<br />

determination of dissociation<br />

constants and distances of various<br />

nuclei from the paramagnetic<br />

centre in a protein-protein<br />

complex. Relaxation measurements<br />

were carried out to get<br />

information about the relative<br />

dispositions of the two proteins<br />

(PC and cytochromes) in their<br />

complexes.<br />

2 Assignment of<br />

Proton NMR resonances<br />

for pea<br />

Plastocyanin<br />

Proton NMR measurements were<br />

carried out on a Bruker AM 500-<br />

MHz FT-NMR spectrometer at<br />

300K in 50 raM phosphate buffer<br />

(pH 6.0). Proton chemical shifts<br />

were referred to a proton signal of<br />

dioxan as a reference at 3.74 ppm.<br />

Proton NMR resonances for pea PC<br />

were assigned using 2D-NMR<br />

spectroscopy at 300K (pH = 6.0).<br />

The conformation of the Phe-83<br />

plastocyanin was examined by<br />

^-NMR. A ID spectrum in H2O,<br />

compared with that of the wildtype<br />

protein, clearly demonstrated<br />

the disappearance of the amide<br />

resonance of Tyr-83 at 9.40 ppm,<br />

and the appearance of a new<br />

resonance at 9.37 ppm which is<br />

likely to be that of Phe-83 [13],<br />

although positive identification<br />

must await further 2D<br />

experiments. Overall the close<br />

similarity between the spectra of<br />

the mutant and wild-type proteins<br />

indicated that the replacement of<br />

Tyr-83 with Phe-83 had not led to<br />

major conformational changes<br />

throughout the protein. Insufficient<br />

protein of the Leu-83<br />

mutant was available for NMR<br />

analysis, so its conformation was<br />

examined by CD between 190 and<br />

260 nm. The spectrum obtained<br />

for the oxidized protein was<br />

essentially identical to that of the<br />

wild-type. The Phe-83 mutant<br />

protein also gave a closely similar<br />

spectrum. These results confirm<br />

that the three proteins had<br />

identical gross conformations.<br />

3 Kinetics<br />

surements<br />

Mea-<br />

Electron transfer from reduced<br />

cytochrome to oxidized plastocyanin<br />

was monitored at 422 nm<br />

with an Applied Photophysics<br />

stopped-flow spectrophotometer<br />

(SF.17MV). The rate of binding of<br />

plastocyanin and cytochrome was<br />

measured by following the


162 Bulletin of Magnetic Resonance<br />

increase in absorbance of oxidized<br />

cytochrome at 410 nm in the<br />

stopped-flow spectrophotometer.<br />

K& was determined by taking<br />

advantage of the increased absorbance<br />

of the Soret band of<br />

cytochrome on binding to<br />

plastocyanin<br />

The results reported in Table I<br />

[13,14] demonstrate convincingly<br />

that Tyr-83 of plastocyanin is part<br />

of the main tunneling pathway for<br />

the electron between the haem<br />

rings of both cytochrome c and<br />

cytochrome / and the Cu atom of<br />

plastocyanin. A leucine residue in<br />

this position is much less effective<br />

and it seems likely that the<br />

facilitation of electron transfer by<br />

tyrosine or phenylalanine is due<br />

to the aromatic nature of the ring,<br />

as has been proposed in other<br />

proteins. A striking difference<br />

between the kinetics of reduction<br />

of plastocyanin by the two<br />

cytochromes is that the wild-type<br />

protein and the Phe-83 mutant<br />

behave identically towards<br />

cytochrome c, but not towards<br />

cytochrome /. In the latter case<br />

the rate of reduction of the<br />

mutant protein is about seven<br />

times slower, a difference that can<br />

be ascribed entirely to weaker<br />

Table I<br />

Kinetic parameters for reduction of pea plastocyanin by cytochromes c and/<br />

Plastocyanin<br />

Cytochrome<br />

Wild type<br />

Phe-83<br />

Leu-83<br />

k2 (x 10-6)<br />

(M-V 1 )<br />

c as donor<br />

3.26<br />

3.28<br />

0.421<br />

Cytochrome/as donor<br />

Wild type<br />

Phe-83<br />

Leu-83<br />

40.6<br />

5.43<br />

0.955<br />

*A<br />

(M-l)<br />

1253<br />

1295<br />

1260<br />

9890<br />

1270<br />

968<br />

•*a(xlO-6)<br />

(MrV 1 )<br />

20.0<br />

22.7<br />

21.7<br />

43.5<br />

5.86<br />

1.27<br />

*.a(xl0< *) k( (x 10-3)<br />

(s- 1 )<br />

16.0<br />

17.5<br />

17.2<br />

4.40<br />

4.61<br />

1.31<br />

(s- 1 )<br />

3.11<br />

2.96<br />

0.340<br />

Values are given as mean + standard deviation. Data for cytochrome/and c are from<br />

[13].and [14] respectively.<br />

62<br />

58<br />

4.0


Vol. 14, No. 1-4 163<br />

binding. We therefore predict that<br />

when the structure of cytochrome<br />

/ becomes known it will reveal a<br />

surface residue in the region of<br />

the exposed haem edge which is<br />

capable of hydrogen bonding to<br />

the -OH of Tyr-83.<br />

4 NMR relaxation<br />

(Ti) measurements<br />

Protein concentrations were<br />

determined from the following<br />

absorption coefficients: reduced<br />

horse heart cytochrome c, £550nm<br />

= 2.76 x 10 4 M-icm" 1 ; reduced<br />

oil-seed rape cytochrome /,<br />

£554nm = 2.6 x 10 4 M-icm-l;<br />

oxidised plastocyanin, £597nm =<br />

4.7 x 103 M-lcm-1. Proton NMR<br />

relaxation measurements were<br />

carried out on a Bruker AM 500-<br />

MHz FT NMR spectrometer at<br />

300K. The samples were in 0.01 M<br />

phosphate buffer (containing 90<br />

raM NaCl) at pH 6.0 (volume, 0.4<br />

ml). Proton NMR spectra of Cd-PC<br />

were obtained by accumulation of<br />

about 160 transients at 16K data<br />

points in quadrature mode. To<br />

facilitate relaxation measurements,<br />

a redox-inactive form of<br />

plastocyanin was prepared, in<br />

which Cu was replaced by Cd. 2D<br />

NMR spectra of Cd-PC showed that<br />

the conformations of the two<br />

proteins are essentially the same.<br />

For the relaxation time<br />

measurements, samples were<br />

treated with Chelex 100 (Bio-Rad)<br />

to remove any traces of free metal<br />

ions. To obtain the longitudinal<br />

relaxation time (Tiobs), the<br />

inversion recovery method with<br />

180^-T -90° pulse sequence was<br />

used.<br />

5 Determination of<br />

the Apparent Dissociation<br />

Constant of<br />

Cd-PC Binding to<br />

Cytochromes using<br />

1H-NMR Tl Measurements<br />

Longitudinal proton relaxation<br />

times of Cd-PC were measured in<br />

the presence of various<br />

concentrations of cytochrome (cor<br />

f) to find the binding constants<br />

and the distances from various<br />

protons of Cd-PC to the<br />

cytochrome iron atom. Observed<br />

longitudinal relaxation time<br />

( T lobs) of Cd-PC proton<br />

resonances can be considered as<br />

the sum of the relaxation rates of<br />

the bound and free Cd-PC<br />

fractions and is related to Kj), T\ 5<br />

and Tif, where Kj) is the apparent<br />

dissociation constant of the<br />

Cyt/Cd-PC complex, Tjb is the Ti


164<br />

of the Cyt/Cd-PC complex, and Tif<br />

is the Ti of the Cd-PC in the<br />

absence of the cytochrome. KTJ and<br />

Tib for Cd-PC binding to<br />

cytochrome was obtained from the<br />

above data. Kj) obtained from NMR<br />

relaxation measurements agreed<br />

very well with the value obtained<br />

from optical spectroscopy (Table<br />

I).<br />

6 Determination of<br />

Distance using<br />

Measurements<br />

The Solomon and Bloembergen<br />

equations were used to determine the<br />

distance (r) of individual protons of<br />

bound Cd-PC from the ferric centre<br />

of cytochromes. The distances for<br />

these protons were used to get the<br />

relative position and conformation of<br />

PC with respect to the ferric ion of<br />

cytochromes. Our initial results show<br />

that the ferric ion of cytochrome is<br />

very near to the Tyr-83 residue of<br />

PC, which is consistent with our<br />

kinetics studies. This work is still in<br />

progress.<br />

7 References<br />

1. P.M. Wood, Biochim. Biophys.<br />

Acta 357, 370 (1974).<br />

2. J.M. Guss, and H.C. Freeman, J.<br />

Mol. Biol. 169, 521 (1983).<br />

Bulletin of Magnetic Resonance<br />

3. J.M. Guss, P.R. Harrowell, M.<br />

Murata, V.A. Norris and H.C. Freeman,<br />

J. Mol. Biol. 192, 361 (1986).<br />

4. DJ. Cookson, M.T. Hayes and P.E.<br />

Wright, Biochim. Biophys. Acta<br />

591, 162 (1980).<br />

5. A.G. Sykes, Chem. Soc. Rev. 14,<br />

283 (1985).<br />

6. A.G. Sykes, Struct. Bonding, 75,<br />

175 (1990).<br />

7. F.A. Armstrong, H.A.O. Hill and<br />

C. Redfield, J. Inorg. Biochem. 28,<br />

171 (1986).<br />

8. J.D. Sinclair-Day and A.G. Sykes<br />

J. Chem. Soc. Dalton Trans. 2069<br />

(1986).<br />

9. G.P. Anderson, D.G. Sanderson<br />

and E.L. Gross, Biochim. Biophys.<br />

Acta, 894, 386 (1987).<br />

10. C. Beoku-Betts, S.K. Chapman<br />

and A.G. Sykes Inorg. Chem. 24,<br />

1677 (1985).<br />

11. P.M. Colman, H.C. Freeman,<br />

J.M. Guss, M. Murata, V.A. Norris,<br />

J.A.M. Ramshaw and M.P. Venkatappa,<br />

Nature, 272, 319 (1978).<br />

12. M. Nordling, T. Olausson and<br />

L.G. Lundberg, FEBS Lett., 276, 98<br />

(1990).<br />

13. S. He, S. Modi, D.S. Bendall and<br />

J.C. Gray, EMBO J, 10, 4011 (1991).<br />

14. S. Modi, S. He, D.S. Bendall and<br />

J.C. Gray, Biochim. Biophys. Acta<br />

(in press).<br />

15. S. Modi, M. Nordling, L.G. Lundberg,<br />

O. Hansson and D.S. Bendall,<br />

Biochim. Biophys. Acta (in press).


Vol. 14, No. 1-4<br />

1 Introduction<br />

Metal-peptide Interaction:<br />

Influence of the Aminoacid Sequence on the Binding of Co(II)<br />

to Glycyltryptophan and Tryptophylglycine<br />

Studied by ^NMR and Fluorescence<br />

A. Spisni, G. Sartor, L. Franzoni<br />

Institute of Biological Chemistry, University of Parma<br />

Via Gramsci, 14,<br />

43100 Parma, Italy<br />

A. Orsolini, P. Cavatorta<br />

Department of Physics, Section of Biophysics, University of Parma<br />

Viale delle Scienze,<br />

43100 Parma, Italy<br />

and<br />

M. Tabak<br />

Instituto de Fisica e Quimica de Sao Carlos, University of Sao Paulo<br />

Av. Dr. Carlos Botelho, 1465,<br />

13560 Sao Carlos (SP), Brazil<br />

Aim of this work is to investigate the nature of the<br />

interactions between transition metals and peptides.<br />

Peptides bind metal ions in various manner<br />

depending on their aminoacid composition, on the pH of<br />

the solution as well as on the metal/peptide ratio. Among<br />

the various aminoacids those with a charged side-chain<br />

are the most efficient for metal binding, though, ions can<br />

also be coordinated by the peptidic nitrogen as well as by<br />

the terminal arnino group. It is well known that the<br />

binding of transition metals to peptides decreases the<br />

apparent pKa of both the terminal amino group and of the<br />

peptidic NH. The fluorescence of the aromatic<br />

aminoacids (Trp, Tyr and Phe) can be used to monitor<br />

those changes. It is well recognized that fluorescence<br />

spectroscopy is suitable for the study of peptides and<br />

proteins, and that Trp is the best internal probe due to its<br />

' Hgh quantum efficiency and molar absorption as<br />

compared to Tyr and Phe. Stemming from these<br />

considerations, conformational changes of proteins or<br />

Peptides can be monitored following the modifications of<br />

.the Trp fluorescence. Similarly, the interaction of Trp<br />

•*ith transition metal ions can be easily detected by<br />

measuring the fluorescence quenching induced by the<br />

metal ions themselves.<br />

Recently [1], [2] we have been studying the<br />

interaction of Cu(II) and Ni(II) with Trp and Gly-Trp<br />

and we found not only that the quenching of Trp<br />

fluorescence is mainly due to a ground state interaction<br />

but also that, for the two metals, the formation of the<br />

metal complex with the dipeptide and the AA is<br />

different, both in terms of binding constants and<br />

stoichiometry.<br />

Another technique that can be used for the study of<br />

metal-peptide complexes is high resolution 'HNMR.<br />

The possibility of a transition metal to act as a line<br />

broadening or as a shift reagent is strictly associated to<br />

its electronic relaxation time. Paramagnetic ions with<br />

a short relaxation time are responsible for changes in<br />

chemical shift without line broadening while, if the<br />

relaxation time is long enough, the effect is a strong<br />

line broadening with no changes of the chemical shift.<br />

Therefore the study of the chemical shift variation as<br />

a function of the metal concentration can lead to<br />

interesting results relevant for a better understanding of<br />

the complexe's stoichiometry. At the same time<br />

relaxation studies will provide valuable information on<br />

their geometry.<br />

165


166<br />

2 Materials and Methods<br />

Tryptophylglycine (Trp-Gly) and glycyltryptophan (Gly-<br />

Trp) were obtained from Sigma Co. S. Louis, MO, their<br />

purity was checked by gas-chromatography. CoCl2-6H2O<br />

was obtained from Merk, Darmstad, Germany and used<br />

without further purification.<br />

*HNMR experiments were carried out using a Bruker<br />

AMX 400 spectrometer, operating at 9.41 T, static<br />

fluorescence experiments were carried out using a Perkin<br />

Elmer MPF 44A spectrofluorimeter, time resolved<br />

fluorescence experiments were carried out with a time<br />

correlated single photon counter equipped with an<br />

Edinburgh F199 nanosecond flash lamp operating in a N2<br />

flux of 1 2/min, a Philips XP2020Q fast photomultiplier<br />

and an EG&G Ortec fast NIM electronics. Time resolved<br />

fluorescence data analysis was carried out using the<br />

Global Analysis [3].<br />

The temperature for all the experiments was 25°C.<br />

Peptide's solutions for *HNMR were freshly prepared<br />

in double distilled water, with 10% D2O, at a final<br />

concentration of 10 mM. Solutions for fluorescence<br />

experiments were obtained by appropriate dilution of 1<br />

mM stock solution in double distilled water. No buffers<br />

were used. CoCl2 solutions were 3 M for 'HNMR and<br />

1 M for fluorescence experiments.<br />

Distinct protocols were used to obtain the pH<br />

titration of the two peptides in : HNMR and fluorescence<br />

experiments. In 'HNMR experiments small aliquots (y.£)<br />

of 0.1 M HC1 or 0.1 M NaOH were added to 0.5 m£ of<br />

the sample in order to obtain a given pH. As for the<br />

fluorescence measurements, additions of 1 M HCl or 1 M<br />

NaOH were made on 50 m£ in order to avoid dilution<br />

artifacts, the same pHmeter was used. The data from pH<br />

titrations, for both *HNMR and fluorescence, were fitted<br />

using the Henderson-Hasselbach equation in order to<br />

obtain the pK, values .<br />

In the case of fluorescence experiments correction for<br />

the inner filter effect was made using Parker's equation<br />

[4]-<br />

3 Results and Discussion<br />

Fluorescence quenching experiments demonstrate that the<br />

complexes fonned by Gly-Trp and Trp-Gly with Co(II)<br />

have distinct properties. The binding of the metal ion is<br />

strictly influenced by the aminoacid sequence and by pH.<br />

Moreover Co(II)like other paramagnetic transition metals<br />

such as Cu(II) and Ni(II), is known to influence the pK,<br />

of the ionizable groups in aminoacids and peptides [1],<br />

[2].<br />

Bulletin of Magnetic Resonance<br />

In the case of the binding of metal ions to<br />

fluorescent aminoacids a non fluorescent ground state<br />

complex is formed, thus, it is possible to calculate the<br />

binding constant from the variation of the fluorescence<br />

intensity. Co(II) binding to Trp-Gly and Gly-Trp<br />

produces biphasic quenching curves [5], indicating that<br />

at least two different complexes are formed in each<br />

case. In table I the binding constants of Co(II) for the<br />

two peptides are reported together with the fraction of<br />

the two complexes present at the given pHs.<br />

Unfortunately, fluorescence spectroscopy is not<br />

able to give the required informations for the direct<br />

determination of the stoichiometry of the complexes<br />

and of their geometry. To overcome these limitations<br />

J HNMR has been used.<br />

The pK, values of the ionizable groups<br />

(carboxylate and amino group) in the absence and in<br />

the presence of Co(II) have been obtained from<br />

fluorescence data. These values turn out to be quite<br />

different with respect to those obtained using the same<br />

technique and reported in a recent publication [6].<br />

To verify these values, the determination of the<br />

two pK,s was carried out by means of 'HNMR. We<br />

studied the pH dependence of the proton chemical<br />

shifts for the two dipeptides in the absence of the metal<br />

ion. Figure 1A reports the plot of the data for the two<br />

a protons of glycine and figure IB for the NH proton<br />

and for the indolic one. From these data the pK,s of<br />

the carboxylate and of the amino group (Table II) were<br />

calculated. The pK,s values of the amino group<br />

obtained from pH dependence of the chemical shift oi<br />

the various protons were in good agreement with those<br />

obtained from fluorescence experiments. In the case oi<br />

the indolic and the NH proton resonances, as they<br />

disappear above pH 7.5, probably because of their fasi<br />

exchange with H2O, the fitting of the data was obtained<br />

imposing, as pK^ values, the average values obtained<br />

from all the other protons. As can be seen the fitting is<br />

quite satisfactory.<br />

When Co(II) is present, while with fluorescence<br />

spectroscopy it is possible to carry out a pH titration oi<br />

the dipeptide-metal complex up to pH 12, NMR i*<br />

limited to pH 7. In fact, due to its high sensitivity<br />

fluorescence spectroscopy allows to operate at (iw<br />

concentration for the dipeptides and for Co(II), thus<br />

avoiding the precipitation of the complex at high pn<br />

NMR, being less sensitive, requires concentrations ii<br />

the range of 10 mM, therefore, above pH 7.5 w<<br />

observe the formation of a mixed complex with OH<br />

that precipitates, with a consequent disappearing of th<<br />

signal.


Vol. 14, No. 1-4 167<br />

•>"• .<br />

TABLE I: Binding constants (K} and K^) obtained from fluorescence quenching experiments, /, and f2 represent the<br />

fraction of complex formed<br />

Dipeptides<br />

Trp-Gly<br />

pH 3.2<br />

Gly-Trp<br />

pH 3.2<br />

Trp-Gly<br />

pH 8.2<br />

Gly-Trp<br />

pH 8.2<br />

3.9 -<br />

fi<br />

1<br />

0.36<br />

0.22<br />

0.55<br />

0<br />

2810.6<br />

896.6<br />

900.6<br />

0 2 4 8 8 10 12 14 0 2 4<br />

f2<br />

—<br />

0.64<br />

0.78<br />

0.45<br />

K, (M- 1 )<br />

Figure 1 A Glycine a protons chemical shifts of glycyltryptophan (•,•) and tryptophylglycine (O,E).<br />

B Indolic proton (», v) and NH(*, A) chemical shifts of glycyltryptophan ( T , A ) and tryptophylglycine (V,A).<br />

—<br />

33.5<br />

131.3<br />

12.1


168<br />

TABLE II: pK,,s obtained from fluorescence and from 1 H NMR experiments<br />

pKal<br />

P*^<br />

P*^<br />

pK,,<br />

P*^<br />

4.0<br />

3.8<br />

3.7<br />

s 3 - 8<br />

a, ' 3.5<br />

3.3<br />

3.Z<br />

Gly-Trp Gly-Trp:Co ++<br />

2.74<br />

8.27<br />

—<br />

3.13<br />

8.30<br />

Fluorescence intensities<br />

1.85<br />

7.46<br />

9.66<br />

a H NMR Chemical Shifts<br />

2.89<br />

3.1<br />

0 1 2 3 4 5 6 7<br />

pH<br />

—<br />

Bulletin of Magnetic Resonance<br />

Trp-Gly Trp-Gly:Co ++<br />

2.59<br />

7.73<br />

—<br />

2.86<br />

7.74<br />

1 ; I I I<br />

2 3 4<br />

PH<br />

B 10.2<br />

Figure 2 A Glycine a protons chemical shifts of glycyltryptophan (•,•) and tryptophylglycine (O,Q) in presence q<br />

Co(II) (1:5 peptide metal ratio). ><br />

B Indolic proton (^.v) and NH (^,A)chemical shifts of glycyltryptophan (•,*) and tryptophylglyctoe (v,i<br />

in presence of Co(II) (1:5 peptide metal ratio).<br />

10.0<br />

8.8<br />

K><br />

8.0<br />

7.8<br />

7.6<br />

7.4<br />

1Z<br />

2.51<br />

7.05<br />

9.15<br />

2.88<br />


Vol. 14, No. 1-4 169<br />

Figure 3 Top<br />

Bottom<br />

""I"'<br />

-SO<br />

WG ilOmM) CO • 1; 5 pH-7.15<br />

SW-30


170<br />

In figure 2A and 2B the 'HNMR titration of the glycine<br />

a protons and of the indolic and NH protons in the<br />

presence of Co(II) (1:5 ratio) are reported. It can be seen<br />

that above pH 7.5 no signal was detectable due to the<br />

formation of a bluish precipitate and that, as a<br />

consequence only the carboxylate pK, has been calculated<br />

and reported in Table II.<br />

Despite these limitations, we have been able to detect<br />

the proton NMR spectrum for the two dipeptides<br />

complexes (Figure 3). A group of resonances is shifted<br />

between -50 ppm to -100 ppm and two or three peaks<br />

appear between 80 ppm and 130 ppm. Interestingly,<br />

though the assignment of the peaks is still to be<br />

completed, it can be seen that the NMR profile for the<br />

two complexes is quite different suggesting that Trp-Gly<br />

and Gly-Trp, indeed, are forming two distinct complexes.<br />

The half line width is approximatively 200 Hz for Gly-<br />

Trp and 1000 Hz for Trp-Gly indicating that Co(II) is<br />

either closest or more tightly bound to Trp-Gly as<br />

compared to the Gly-Trp. Indeed the Trp-Gly binding<br />

constants suggests an average high affinit of this peptide<br />

respect to Gly-Trp. The NMR spectra of the two<br />

peptide-metal complexes present one single peak with<br />

the same chemica shift.A that peak tend to disappear<br />

upon D2O addition we believe it is a proton bound to a<br />

nitrogen atom. The NMR spectra of the two complexes<br />

posses other peculiar characteristics. The chemical shifts<br />

of their resonance lines are insensitive to the pH variation<br />

suggesting a high stability of the complexes. Moreover,<br />

the NMR spectra disappear below pH 5.2 for Trp-Gly<br />

and below pH 4.15 for Gly-Trp. We do believe these<br />

evidences are an indication both of the need for the<br />

carboxylic group to be ionized as well as of the<br />

requirement for a small fraction (1/1000) of NH2 (the pK,<br />

for the amino group are 8.27 and 7.73 respectively) in<br />

order to have metal binding.<br />

In conclusion, these preliminary results indicate that,<br />

because of the complementarity of NMR and<br />

fluorescence spectroscopy, it is possible to better evaluate<br />

the pK,s of the dissociable groups and the metal binding<br />

properties of small peptides. We believe that such an<br />

integrated approach can be relevant for the study of more<br />

complex macromolecules such as polypeptides and<br />

proteins.<br />

4. References<br />

Bulletin of Magnetic Resonance<br />

1. Tabak M., Sartor G. and Cavatorta P., /. of<br />

Luminescence, 43., 355 (1989).<br />

2. Tabak M., Sartor G., Neyroz P. and Cavatorta P.,<br />

J. of Luminescence, 46, 291 (1990).<br />

3. Knutson J.R., Beechem J.M. and Brand L., Chem.<br />

Phys. Lett. 102, 501 (1983).<br />

4. Parker C.A. Photoluminescence of solutions,<br />

Elsevier, (1968).<br />

5. Sartor G. Franzoni L., Cavatorta P., Tabak M. and<br />

Spisni A., manuscrip in preparation.<br />

6. Chen F., Knutson J.R., Ziffer H. and Porter D.,<br />

Biochemistry, 2Q, 5184 (1991).<br />

Acknowledgments This work was supported<br />

by a CNR Grants #92.00752.CT04 and #92.02243.ctl4,<br />

by MURST 60% (SG) and MURST 40% (SA)


Vol. 14, No. 1-4 171<br />

INTRODUCTION<br />

ASSIGNMENTS OF THE *H NMR SPECTRUM OF A CONSENSUS<br />

DNA-BINDING PEPTIDE FROM THE HMG-I PROTEIN<br />

Jeremy N. S. Evans§$*, Mark S. Nissen§ and Raymond Reeves§<br />

Departments of Biochemistry/Biophysics§ and Chemistry*,<br />

Washington State University, Pullman, WA 99164-4660. U. S. A.<br />

The HMG-I subfamily [1-3] of high mobility group<br />

(HMG 1 ) chromatin proteins [4] consists of DNA-binding<br />

proteins that preferentially bind to stretches of A'T-rich sequence<br />

both in vitro [5-8] and in vivo [9]. Recently, members<br />

of the HMG-I family have been suggested to bind in<br />

vitro to the narrow minor groove of A*T-DNA by means of<br />

an 11 amino acid peptide binding domain (BD) which, because<br />

of its predicted structure is called the "A*T-hook motif<br />

[10]. The HMG-I proteins are specific substrates for the<br />

ii i i • / \ -, Acdc2/cdc28 , . , ,<br />

cell cycle regulating enzyme(s) p34 kmase (also<br />

known as histone HI kinase) both in vivo and in vitro [11-<br />

13]. The sites of phosphorylation by cdc2 kinase are the<br />

threonine residues at the amino terminal ends of the A*Thook<br />

motifs and such modifications have been demonstrated<br />

to reduce markedly the affinity of binding of the HMG-I proteins<br />

to DNA [12,13].<br />

HMG-I proteins are also of considerable biological interest<br />

because they are expressed at elevated levels in actively<br />

proliferating cells and have been observed to be a<br />

characteristic feature of undifferentiated [1,7] or neoplastically<br />

transformed cellular phenotypes [14,15]. High HMG-I<br />

levels have been found to be a consistent feature of rat and<br />

mouse malignant cells [14-17] and have been suggested to<br />

be protein markers for both neoplastic transformation [15]<br />

and metastatic potential [18]. The HMG-I proteins have also<br />

been implicated in control of DNA replication [19,20] and<br />

the regulation of gene transcription [8,21,22]. It is known<br />

from their primary sequences that the HMG-I proteins have<br />

the overall structure typical of Ptashne-type transcriptional<br />

activator proteins possessing both a DNA binding domain(s)<br />

and a highly acidic COOH terminus [23]. In vitro HMG-I-<br />

Ijke proteins have been demonstrated to increase transcription<br />

of isolated ribosomal genes [21] and to alter the conformation<br />

and stability of A-T-rich regions of DNA [24]<br />

properties often associated with DNA-binding gene regulatory<br />

proteins.<br />

As an initial attempt to determine in molecular detail the<br />

interaction of the BD peptide with A»T rich DNA, we have<br />

examined a synthetic 13 residue BD peptide by NMR<br />

spectroscopy. In this paper we report the assignments of the<br />

j<br />

HMG - Wfi 11 mobility group; BD,<br />

domain; DQF, double quantum filtered;<br />

C0Trelated<br />

spectroscopy; NOESY, nuclear<br />

! effCt spectroscopy; ROESY, rotatingei<br />

VT 3USer effect spectroscopy; TOCSY, total<br />

elated spectroscopy.<br />

resonances for the peptide at 295 K and pH 3.4, and provide<br />

preliminary evidence on the peptide structure.<br />

MATERIALS AND METHODS<br />

Peptide Synthesis. The 13mer BD peptide<br />

(VPTPKRPRGRPKK) was synthesized by solid-phase synthesis<br />

(on a Departmental Applied Biosystems model 431A<br />

peptide synthesizer), and purified by reverse-phase HPLC on<br />

a Vydak C4 column (1 x 25 cm) using, a water (containing<br />

0.1% trifluoroacetic acid)-acetonitrile gradient under standard<br />

conditions. The 13mer BD peptide eluted at 15% acetonitrile<br />

(Rt = 13 mins @ 1.5 mL min" 1 ).<br />

NMR Spectroscopy. High field Fourier transform (FT)<br />

NMR studies were performed on a Varian VXR-500S (11.75<br />

T, 500MHz J H) NMR spectrometer. Deuterium was used<br />

for locking the field. *H NMR chemical shifts were referenced<br />

externally to samples of similar dielectric constant<br />

containing sodium 3-(trimethylsilyl) propanoate-2,2,3,3-d4<br />

(TSP) in D2O buffer (5H = 0.00 ppm). Sample temperature<br />

was maintained with a Varian variable temperature control<br />

unit, using gaseous nitrogen (from boil-off liquid nitrogen)<br />

cooled using an FTS XR-85 cryo-cooler. The majority of<br />

samples of peptide were maintained at 295 K, except where<br />

the amide exchange rates were being measured, when the<br />

sample was maintained at 277 K. Data was downloaded to<br />

either a Silicon Graphics 4D25TG or a-4D70GT workstation,<br />

and converted from Varian format to FELIX format using<br />

the VNMR2FELIX conversion program (a gift from<br />

Darrell Davies, University of Utah). The output from this<br />

was processed using FELIX (Hare Research Ltd.). All 2D<br />

data was obtained using the hypercomplex phase sensitive<br />

method [25] and processed as 2K x 2K complex data sets<br />

with baseline correction and sine-bell squared weighting<br />

functions in both dimensions.<br />

DQF-COSY were recorded with the pulse sequence tQ-<br />

90°-f7-90 o -S-90 o -f2, where tj is the evolution time, t2 is<br />

the acquisition time, and 8 is a fixed delay of 3 \is [26].<br />

TOCSY spectra were recorded with the pulse sequence /#-<br />

90°-fi-SLx-(MLEV-17)-SLx-/2, where SLX was a 4 ms<br />

trim pulse along the x axis [27]. The MLEV-17 spin-locking<br />

pulse sequence was repeated to give a mixing time of<br />

40 ms. ROESY spectra were recorded with the pulse sequence<br />

/o-90 0 -r;-90 0 -SLx(30°)-90°-t2, where SLx(30°) is a<br />

small spin-lock pulse repeated to give a mixing time of 200<br />

ms [28]. NOESY spectra were recorded with the pulse


172 Bulletin of Magnetic Resonance<br />

TABLE 1 Sequential Assignments of HMG-I 13mer Binding Domain Peptide, pH 3.4,295K<br />

Residue<br />

VI<br />

P2<br />

T3<br />

P4<br />

K5<br />

R6<br />

P7<br />

R8<br />

G9<br />

RIO<br />

Pll<br />

K12<br />

K13<br />

NH<br />

_<br />

_ 8.44<br />

• _<br />

8.10<br />

8.35<br />

8.54<br />

8.41<br />

8.28<br />

— 8.44<br />

7.24<br />

otCH<br />

4.23<br />

4.58<br />

4.60<br />

4.33<br />

4.23<br />

4.68<br />

4.46<br />

4.34<br />

3.92, 4.04<br />

4.68<br />

4.45<br />

4.32<br />

3.27<br />

sequence ta-90°-tj-90°-Tm-90 o -t2 [29,30] with mixing<br />

times, TW, of 100, 300, 400, and 600 ms. Solvent suppression<br />

was achieved by presaturation of the H2O resonance for<br />

all the 2D experiments.<br />

Sample Preparation. The BD peptide was dried with successive<br />

cycles of lyophilization and re-hydration with either<br />

H2O or D2O and then dissolved in one of the following:<br />

Buffer (i) 25 mM potassium phosphate, 0.01% (w/v) NaN3,<br />

in 10% v/v D2O/H2O, pH 3.4; or Buffer (ii) 25 mM potassium<br />

phosphate, 0.01% (w/v) NaN3, in 99% v/v<br />

D2O/H2O, pH 3.4. pH titrations were carried out by careful<br />

4.0 3.0 2.0<br />

(ppm)<br />

tPx<br />

m 6*<br />

f\j<br />

I<br />

&CH<br />

2.36<br />

1.91, 2.37<br />

4.16<br />

1.91, 2.35<br />

1.76, 1.87<br />

1.77, 1.87<br />

1.93, 2.34<br />

1.73, 1.87<br />

1.76, 1.88<br />

1.94, 2.34<br />

1.79, 1.84<br />

1.74, 1.89<br />

Others<br />

1.03, 1.14<br />

2.08<br />

1.31<br />

2.04<br />

1.43<br />

1.71<br />

2.05<br />

1.44, 1.50<br />

1.73<br />

2.05<br />

1.51<br />

1.49<br />

3.67, 3.80<br />

3.73, 3.90<br />

3.03, 3.25<br />

3.05, 3.25, 7.55<br />

3.68, 3.87<br />

3.05, 3.25<br />

3.25<br />

3.66, 3.84<br />

3.05<br />

3.05<br />

addition of small quantities of HC1 or NaOH and measurement<br />

with a 4 mm pH electrode (Ingold Co.).<br />

RESULTS<br />

The 13mer peptide was studied by NMR at 277 and 295 K<br />

and at a variety of pH values. The linewidths of the ID 500<br />

MHz NMR spectra were not sensitive to concentration in<br />

the range 0.1-20 mM, indicating negligible aggregation of<br />

the peptide [31]. The use of 2-dimensional double-quantum<br />

filtered correlated spectroscopy (DQF-COSY) and total<br />

correlated spectroscopy (TOCSY) has enabled us to assign<br />

resonances to amino acid spin types, although distinctions<br />

1.0 4.0 3.0 2.0<br />

(ppm)<br />

Figure 1 500 MHz *H DQF-COSY (A) and TOCSY (B) NMR spectra of the aliphatic resonances of the 13mer BD<br />

(20 mM) in 90%H2O/10%D2O phosphate (25 mM) buffer, pH 3.4,295K. Cross-peak assignments are indicated.


Vol. 14, No. 1-4 173<br />

G9NH<br />

•°KI2NHa<br />

R8nHa<br />

• K-JNHa<br />

»T3NHO<br />

* »Rl 6 hours.<br />

Interestingly, the slowly exchanging amide resonances also<br />

do not titrate over the pH range 2.2 to 7.1, with the exception<br />

of the K5 NH, and all appear to be located along one<br />

side of the peptide molecule.<br />

DISCUSSION<br />

A model for how the consensus binding domain peptide<br />

from the HMG-I protein binds to the minor groove of A«T<br />

rich DNA has been proposed by this laboratory [10] on the<br />

basis of molecular modelling. In order to test this model, we<br />

have initiated NMR studies of a 13mer BD peptide in<br />

solution. While no


HI<br />

i ill,<br />

I 1<br />

174<br />

the 13mer peptide). These amides might be expected to be<br />

less accessible to solvent, and potentially may participate in<br />

hydrogen-bonds with the carbonyl groups from (i, i + 2)<br />

residues. At this stage, there is no additional corroborating<br />

evidence from medium-range or long-range NOEs either to<br />

support or refute these speculations. However, in the<br />

presence of A«T rich DNA, the conformational lability of<br />

the peptide backbone and side-chains would be expected to be<br />

reduced significantly, and longer range NOEs detectable.<br />

ACKNOWLEDGEMENTS<br />

We should like to acknowledge Gerhard Munske for the<br />

synthesis of the 13mer peptide, Wendy Shuttleworth for<br />

help with some sample preparations, Darrell Davies<br />

(University of Utah) for the VNMR2FELIX conversion<br />

software. Supported in part by an American Cancer Society<br />

Institutional Grant IGR-119L (JNSE), NIH Grant R01<br />

GM46352 (RR), and the WSU NMR Center is supported by<br />

NIH grant RR 0631401, NSF grant CHE 9115282 and<br />

Battelle Pacific Northwest Laboratories Contract No. 12-<br />

097718-A-L2.<br />

REFERENCES<br />

[ 1] Lund, T., Holtlund, J., Fredriksen, M. & Laland, S.<br />

(1983) FEBS Lett. 152, 163-167.<br />

t 2] Johnson, K., Lehn, D., Elton, T., Barr, P. & Reeves,<br />

R. (1988) J. Biol. Chem. 263, 18338-42.<br />

[ 3] Johnson, K., Lehn, D. & Reeves, R. (1989) Mol.<br />

Cell. Biol. 9,2114-23.<br />

[ 4] Goodwin, G. & Bustin, M. (1988) in Kahl, G.<br />

Architecture of Eukaryotic Genes, 187-205, VCH<br />

Weinheim, Germany.<br />

[ 5] Strauss, F. & Varshavsky, A. (1984) Cell 37, 889-<br />

901.<br />

[ 6] Solomon, M., Strauss, F. & Varshavsky, A. (1986)<br />

Proc. Nail. Acad. Sci. U.SA. 83, 1276-1280.<br />

[ 7] Elton, T., Nissen, M. & Reeves, R. (1987)<br />

Biochem. Biophys. Res. Commun. 143, 260-265.<br />

[ 8] Reeves, R., Elton, T., Nissen, M., Lehn, D. &<br />

Johnson, K. (1987) Proc. Natl. Acad. Sci. U.SA.<br />

84, 6531-6535.<br />

[ 9] Disney, J., Johnson, K., Magnuson, N., Sylvester,<br />

S. & Reeves, R. (1989) / Cell Biol. 109,1975-82.<br />

[10] Reeves, R. & Nissen, M. (1990) J. Biol. Chem.<br />

265, 8573-8582.<br />

[11] Lund, T. & Laland, S. (1990) Biochem. Biophys.<br />

Res. Commun. 171, 342-7.<br />

[12] Reeves, R., Langan, T. & Nissen, M. (1991) Proc.<br />

Natl. Acad. Sci. USA88, 1671-5.<br />

[13] Nissen, M., Langan, T. & Reeves, R. (1991) /.<br />

Biol. Chem. 266, 19945-19952.<br />

Bulletin of Magnetic Resonance<br />

[14] Giancotti, V., Pani-D'Andrea, P., Berlingieri, M. T.,<br />

Di Fiore, P. P., Fusco, A., Veccio, G., Philip, R.,<br />

Crane-Robinson, C, Nicolas, R. H., Wright, C. A., &<br />

Goodwin, G. H. (1987) EMBO J. 6,1981-1987.<br />

[15] Giancotti, V., Buratti, E., Perissin, L., Zorzet, S.,<br />

Balmain, A., Portella, G., Fusco, A., & Goodwin,<br />

G. H. (1989) Exp. Cell Res. 184, 538-45.<br />

[16] Elton, T. & Reeves, R. (1986) Anal. Biochem.157, 53-<br />

62.<br />

[17] Vartiainen, E., Palvimo, J., Mahonen, A., Linnala-<br />

Kankkunen, A. & Maenpaa, P. (1988) FEBS Lett<br />

228, 45-48.<br />

[18] Bussemakers, M., van de Ven, W., Debruyne, F. &<br />

Schalken, J. (1991) Cancer Res. 51,606-611.<br />

[19] Grummt, F., Hoist, A., Muller, F., Wegner, F.,<br />

Schwender, S., Luksza, H., Zastow, G., & Klavinius,<br />

A. (1988) Cancer Cells 6,463-466.<br />

[20] Wegner, M., Zastrow, G., Klavinius, A., Schwender,<br />

S., Muller, F., Luksza, J., Hoppe, J., Wienberg, J. &<br />

Grummt, F. (1989) Nucleic Acids Res. 17, 9909-9932<br />

[21] Yang-Yen, H. & Rothblum, L. (1988) Mol. Cell. Bio<br />

8, 3406-3414.<br />

[22] Eckner, R. & Bimstiel, M. (1989) Nucleic Acids Res.<br />

17, 5947-59.<br />

[23] Ptashne, M. (1988) Nature 335,683-689.<br />

[24] Lehn, D., Elton, T., Johnson, K. & Reeves, R. (198*<br />

Biochem. Int. 16, 963-971.<br />

[25] States, D. J., Haberkorn, R. A., & Ruben, D. J. (198:<br />

/. Magn. Reson. 48, 286-293.<br />

[26] Ranee, M., S0rensen, O.W., Bodenhausen, G., Wagne<br />

G., Ernst, R. R., & Wuthrich, K. (1983) Biochem.<br />

Biophys. Res. Commun. 117, 479-485.<br />

[27] Bax, A., & Davis, D. G. (1985) J. Magn. Reson. 65,<br />

355-360.<br />

[28] Kessler, H., Griesinger, C, Kerssebaum, R., Wagner,<br />

R., & Ernst, R. R. (1987) J. Am. Chem. Soc. 109,<br />

607-609.<br />

[29] Jeener, J., Meier, B. H., Bachmann, P., & Ernst, R. 1<br />

(1979) /. Chem. Phys. 71, 4546-4553.<br />

[30] Bodenhausen, G., Kogler, H., & Ernst, R. R. (1984;<br />

/. Magn. Reson. 58, 370.<br />

[31] Dyson, H. J., & Wright, P. E. (1991) Annu. Rev.<br />

Biophys. Biophys. Chem. 20, 519-538.<br />

[32] Wuthrich, K., Billeter, M., & Braun, W. (1984) /.<br />

Mol. Biol. 180, 715-740.<br />

[33] Mayo, K. M., Parra-Diaz, D., McCarthy, J. B., &<br />

Chelberg, M. (1991) Biochemistry 30, 8251-8267.<br />

[34] Grathwohl, C. & Wuthrich, K. (1976) Biopolymers<br />

15, 2025-2041.<br />

[35] Grathwohl, C. & Wuthrich, K. (1981) Biopolymers<br />

20,2623-2633.


Vol. 14, No. 1-4 175<br />

Solution Structure of the<br />

DNA-binding Domain of GAL4<br />

from Saccharomyces cerevisiae<br />

James D. Baleja, V. Thanabal, Ted Mau, and Gerhard Wagner<br />

Department of Biological Chemistry and Molecular Pharmacology<br />

1 Introduction<br />

The GAL4 transcriptional<br />

activator protein has long been a<br />

favorite for the study of<br />

transcription in eukaryotic biology.<br />

Genetic studies reveal a modular<br />

architecture for the protein with<br />

different functions associated with<br />

each module [1]. A DNA-binding<br />

domain of the protein recognizes and<br />

binds to a sequence of DNA termed<br />

the Upstream Activating Sequence<br />

(UASQ). Other parts of the protein are<br />

relevant for activation. They interact<br />

with the transcriptional machinery<br />

including RNA polymerase to activate<br />

transcription. The UASQ is near the<br />

genes that encode the proteins<br />

required for galactose utilization.<br />

Upon presentation of galactose to the<br />

yeast cell, this DNA site is specifically<br />

bound by GAL4, the transcription<br />

function of RNA polymerase is<br />

activated, and enzymes required for<br />

galactose utilization are produced [2].<br />

As a dimer of 881 amino acids,<br />

GAL4 is too large for the<br />

determination of a high-resolution<br />

NMR structure and we have instead<br />

studied a fragment containing the Nterminal<br />

65 amino acid residues<br />

Harvard Medical School<br />

Boston, Massachusetts 02115<br />

including its DNA-binding domain<br />

[3]. In the absence of DNA, GAL4(65)<br />

is monomeric in solution,<br />

presumably because it does not have<br />

the amino acid residues necessary for<br />

dimerization [4]. GAL4(65) is dimeric<br />

when bound to any of four DNA sites<br />

present in the UAS [5]. Each of these<br />

binding sites is approximately twofold<br />

palindromic, and like many<br />

other dimeric DNA-binding proteins,<br />

each monomer of the GAL4 dimer<br />

interacts with a half-site of DNA.<br />

The structure and dynamics of the<br />

monomeric DNA-binding domain of<br />

GAL4 (residues 1-65; Figure 1), an<br />

investigation of the binding to a DNA<br />

half-site (Figure 2), and the structure<br />

of the resultant protein-DNA complex<br />

are presented in this paper.<br />

1 11 21<br />

MKLLSSIEQA CDICRLKKLK CSKEKPKCAK<br />

31 41 51<br />

CLKNNWECRY SPKTKRSPLT RAHLTEVESR<br />

61<br />

LERLE<br />

Figure 1. Primary amino acid<br />

sequence for the DNA-binding<br />

domain of GAL4.


176<br />

1 2 3 4 5 6 7 3 9 10<br />

C C G G A G G A C T<br />

G G C C T C C T G A<br />

20 19 18 17 16 15 14 13 12 11<br />

Figure 2. Nucleotide sequence for the<br />

one-half DNA-binding site of GAL4.<br />

2 1H, 15N, and n^Cd NMR<br />

resonance assignments<br />

Assignment of specific nuclei to<br />

the observed NMR resonance<br />

frequencies is the first step in<br />

determining the structure of a<br />

protein using NMR techniques [6].<br />

Assignments for the 1 H, ^$N, and<br />

113 Cd NMR resonances of GAL4(65)<br />

were made using homonuclear and<br />

hetero-nuclear NMR experiments<br />

and by following standard protocol<br />

([6], Baleja and Wagner, unpublished<br />

results).<br />

Zinc is required for the DNAbinding<br />

activity of GAL4 [4]. It can be<br />

replaced by NMR-active 113 Cadmium<br />

without loss of DNA-binding. The<br />

113 Cd NMR spectrum (Figure 3)<br />

shows that two metal ions are<br />

coordinated by a Cys - (X)2 - Cys - (X>6<br />

- Cys - (X)6 - Cys - (X)2 - Cys - (X)6 -<br />

Cys motif using six cysteines and<br />

forming a bimetal-thiolate cluster<br />

[7]. Hetero-nuclear correlation<br />

experiments between the l* 3 Cd and<br />

*H define the liganding of the two<br />

central metal ions (Figure 3).<br />

Having assigned resonance frequencies<br />

to specific nuclei, the<br />

correspondence of cross peaks to the<br />

protons can be made and local<br />

structural information was can then<br />

be determined. A section of the twodimensional<br />

NOESY spectrum is<br />

shown in Figure 4. The region shows<br />

the cross peaks among the amide<br />

protons of the protein backbone and<br />

protons of the aromatic sidechains.<br />

BuJietin of Magnetic Resonance<br />

Figure 3. Heteronuclear ^ C d H<br />

correlation experiments. The standard<br />

reverse INEPT pulse sequence<br />

[8] was followed by a short MLEV-17<br />

TOCSY transfer [9] and proton<br />

detection./<br />

Indicating the presence of two ahelices,<br />

there are a series of close<br />

approaches between amide proton<br />

resonance frequencies for several<br />

sequential residues. Other NOEs arise<br />

between amino acid residues more<br />

distant in sequence and define the<br />

unique topological features for the<br />

protein. NOE intensities at 56, 150, and<br />

250, and 500 millisecond mixing times<br />

were converted to the corresponding<br />

interproton distances using the usual<br />

inverse r 6 relationship [3]. In case of<br />

overlap in the homonuclear NOESY,<br />

spectrum analysis was also made<br />

using a three-dimensional ^JSJ.IJJ.IH<br />

NOESY-HMQC spectrum recorded on a<br />

uniformly ^N-labeled protein<br />

(Figure 5, [10]). From the NOESY data,


Vol. 14, No. 1-4<br />

residues 9-40 are observed to form a<br />

well-defined, compact globular cluster.<br />

The terminal residues 1-8 and 41-<br />

66 show little persistent structure in<br />

solution. There are many interresidue<br />

NOE crosspeaks for the<br />

central core (30 per residue, on<br />

average), but only a few weak NOE<br />

cross peaks for protons in the<br />

disordered region. In addition, the<br />

amide protons of the flexible trails<br />

are in rapid exchange with solvent<br />

[10] and have narrow resonance<br />

lines, indicating that these residues<br />

have considerable conformational<br />

mobility in the absence of DNA. This<br />

unstructured character for parts of<br />

the GAL4 protein may be vital to its<br />

biological function. On the other<br />

hand, at least some aspects of this<br />

mobility may merely be a<br />

consequence of the truncation used<br />

for GAL4 in this study. Nonetheless,<br />

our picture of GAL4 is one in which<br />

two N-terminal recognition modules<br />

are connected by flexible linkers to a<br />

dimeric core [4],<br />

-©V<br />

o<br />

K23<br />

o<br />

J28<br />

R15<br />

D1i! L1<br />

O<br />

E37<br />

OC11<br />

S41<br />

DS22<br />

>C21<br />

C38,<br />

9.0 8.5 8.0 7.5 7.0<br />

6)2 (ppm)<br />

6.5 6.0<br />

Figure 4. NOESY spectrum of amide<br />

and aromatic protons of GAL4(65).<br />

Sample concentration was 1.5 mM in<br />

0.2 M NaCl, 20 mM sodium phosphate,<br />

pH 7.0 at 25°C. Numbers indicate the<br />

residues for which sequential strong<br />

amide to amide cross peaks are<br />

observed which are indicative of ahelices.<br />

K&O T»K27<br />

,©113<br />

C31


178<br />

3 Structure of the GAL4<br />

DNA-binding domain<br />

Sets of interproton distances and<br />

torsion angles formed the basis for<br />

structure determination. Distances<br />

were determined from the NOESY data<br />

[3]. § torsion angles were interpreted<br />

from measured NH-oc coupling constants<br />

and x 1 angles were derived<br />

from ^N-j} and a-(5 coupling constants<br />

[3]. Sulfur-Cd liganding distances<br />

were imposed to be 2.35 to 2.45 A, Cys<br />

Cp-Cd distances to be less than 3.4 A,<br />

and sulphurs liganding the same Cd<br />

to be between 3.3 and 4.2 A. 614<br />

distance and 41 torsion angle<br />

measurements were used with the<br />

distance geometry package DG-II to<br />

generate a set of structures (0.6 A<br />

backbone atom rmsd) in agreement<br />

with the NMR data [3]. A schematic of<br />

the structure (Figure 6) shows the<br />

two central metal ions coordinated by<br />

the six cysteines. The DNA-binding<br />

domain consists of an a-helix and an<br />

extended structure, then a sharp turn<br />

that contains a cis-proline bond, and<br />

then a 2nd a helix followed by an<br />

extended structure. If the central<br />

metal binding subdomain of GAL4 is<br />

split, and the two halves<br />

superimposed, a striking correspondence<br />

between each part is revealed.<br />

The conformation of the 13 residue<br />

segment from residues 10 to 22 is<br />

almost identical to that of residues 27<br />

to 39, with an rmsd of 0.8 A for the<br />

backbone atoms. Although the<br />

structural integrity of the protein<br />

would be provided mainly by the way<br />

in which the two metal ions are<br />

liganded, some hydrophobic packing<br />

is observed with the side chains of<br />

W36 and Y40 [3]. Slowly exchanging<br />

amides of GAL4, in both the free and<br />

DNA-bound forms [10], can be<br />

attributed, in part, to hydrogen<br />

bonding to carbonyl oxygens within<br />

the a-helices, and to sulfurs of the<br />

cysteinyl side-chains [12].<br />

Bulletin of Magnetic Resonance<br />

Figure 6. Structure of the central<br />

core for the DNA-binding domain of<br />

GAL4. Cysteines that ligand the two<br />

central metal ions are numbered.<br />

4 Formation of a GAL4-DNA<br />

complex<br />

Imino protons of one half of the<br />

consensus dimeric GAL4 DNAbinding<br />

were monitored in forming<br />

the complex between the protein and<br />

DNA (Figure 7). One imino proton is<br />

present for each base-pair of a 10<br />

base-pair DNA duplex.<br />

_JV<br />

13.8 13.6 13.4 13.2 13.0 12.8 12.6 ppm<br />

Figure 7. Titration of DNA with GAL4.<br />

Experimentally, eight imino proton<br />

resonances are observed, since the<br />

imino protons on the base-pairs at<br />

each end of the duplex exchange<br />

rapidly with the bulk solvent at room<br />

temperature. Each imino resonance<br />

has been assigned to a specific base-


Vol. 14, No. 1-4<br />

pair. Resonances shift as the<br />

environment around each proton<br />

changes upon addition of GAL4. All<br />

resonances broaden as the more<br />

slowly protein-DNA complex is<br />

formed. The equilibrium dissociation<br />

constant is 169± 13 |J.M. Resonances<br />

are in fast exchange between the<br />

protein-DNA complex and the free<br />

components indicating rapid equilibrium<br />

between free and bound<br />

forms.<br />

NOESY and TOCSY two-dimensional<br />

data for the protein-DNA complex<br />

14.0 12.0 10.0 8.0<br />

F2<br />

6.0<br />

(ppm )<br />

have been collected (Figure 8). These<br />

spectra are promising for full<br />

analysis since the resonance lines<br />

are not too broad for resonance<br />

assignment, chemical shift dispersion<br />

is good, and solubility of the<br />

protein-DNA complex is adequate. The<br />

DNA resonances have been completely<br />

re-assigned and the protein<br />

resonances have been partially reassigned<br />

within the protein-DNA<br />

complex.<br />

4.0 2.0<br />

Figure 8. NOESY spectrum of the GAL4:DNA complex. The concentration of the<br />

complex was approximately 0.5 mM. Buffer conditions were 0.15 M NaCl, 20 mM<br />

PO4, pH 7, 25°C.<br />

5 NMR structure of<br />

GAL4-DNA complex<br />

the<br />

Several NOE contacts have been<br />

observed between protein and DNA<br />

(dashed lines, Figure 9) yielding a<br />

preliminary stucture for this GAL4-<br />

DNA complex. The recognition helix<br />

for the free form of the protein was<br />

docked onto a B-DNA. The amino acid<br />

residues responsible for recognizing<br />

DNA are part of the metal-binding<br />

cluster and interact with edges of<br />

'base-pairs exposed in the<br />

major<br />

groove of the DNA.<br />

The structure of the DNA-binding<br />

domain of GAL4 (residues 1-65) bound<br />

to full site DNA containing two GAL4<br />

binding sites has been determined<br />

crystallographically. In agreement<br />

with our observed intermolecular<br />

NOE cross peaks (Figure 8), the Cterminal<br />

end of the first a-helix of<br />

the metal binding cluster (residues 9-<br />

40) provides for sequence-specific<br />

re-cognition of DNA by GAL4. When<br />

bound to full site DNA, parts of<br />

GAL4(65), before unstructured in<br />

solution, adopt a regular conformation.<br />

Residues 41 to 49 form a<br />

linker region which interacts with<br />

179


180<br />

the backbone of the DNA. In addition,<br />

residues 50-64 form a small dimerization<br />

domain using a coiled-coil type<br />

packing arrangement [4]. Our preliminary<br />

results on an intact<br />

dimerization element for GAL4<br />

(residues 50-106) indicate that the ahelical<br />

character for the coiled-coil<br />

of the protein extends beyond residue<br />

64 (Baleja, Marmorstein, and Wagner,<br />

unpublished results). The same<br />

conformation (1.1 A rmsd) is observed<br />

for the recognition module of<br />

GAL4(65) in solution using NMR<br />

techniques as for the central metalbinding<br />

cluster of GAL4(65) bound to<br />

DNA using crystallographic techniques.<br />

Thus the core of the DNAbinding<br />

domain changes little upon<br />

binding DNA.<br />

Figure 9. Preliminary structure for a<br />

GAL4-DNA complex.<br />

6 Summary<br />

The DNA-binding domain of GAL4<br />

has many interesting features. It is a<br />

novel DNA-binding motif with a<br />

globular two-metal cluster that has<br />

hydrogen-bonds to sulfur and twofold<br />

internal symmetry. GAL4 reads<br />

the sequence of DNA through the<br />

amino acids present at the C-terminal<br />

Bulletin of Magnetic Resonance<br />

end of the first a-helix. The two DNAreading<br />

modules of intact GAL4 are<br />

tethered by flexible linkers to the<br />

central body of the protein. Once<br />

bound to DNA through the Nterminal<br />

recognition domains, the Cterminal<br />

portion of GAL4 is in the<br />

correct position to interact with the<br />

components of the transcriptional<br />

machinery to bring about transcriptional<br />

activation.<br />

7 References<br />

1. Keegan, L., Gill, G., and Ptashne,<br />

M., Science 231, 699 (1986).<br />

2. Johnston, M., Micro. Rev. 51, 458<br />

(1987).<br />

3. Baleja, J. D., Marmorstein, R.,<br />

Harrison, S. C, and Wagner G.,<br />

Nature 356, 448 (1992).<br />

4. Marmorstein, R., Carey, M.,<br />

Ptashne, M., and Harrison, S. C,<br />

Nature 356, 408 (1992).<br />

5. Carey, M., Kakidani, H.,<br />

Leatherwood, J., Mostashari, F.,<br />

and Ptashne, M., J. Mol. Biol. 209,<br />

423 (1989).<br />

6. Wuthrich, K. NMR of Proteins<br />

and Nucleic Acids (Wiley, New<br />

York, 1986).<br />

7. Pan, T., and Coleman, J. E.,<br />

Biochemistry 209, 3023 (1990).<br />

8. Briihwiler, D., and Wagner, G., J.<br />

Magn. Reson. 69, 546 (1986).<br />

9. Bax, A., and Davis, D. G., J. Magn.<br />

Reson. 65, 355 (1985).<br />

10. Mau, T., Baleja, J. D., and Wagner,<br />

G., Protein Science (submitted).<br />

11. Bodenhausen, G., and Ruben, D. J.,<br />

Chem. Phys. Lett. 69, 185 (1980).<br />

12. Kraulis, P., Raine, A. R. C<br />

Gadhavi, P. L., and Laue, E. D.,<br />

Nature 356, 448 (1992).


Vol. 14, No. 1-4<br />

STRUCTURAL INVESTIGATION OF FOLIC ACID BY<br />

NMR PROTON RELAXATION AND MOLECULAR<br />

MECHANICS ANALYSIS<br />

1 Introduction<br />

Claudio Rossi, Alessandro Donati, Sergio Ulgiati*<br />

and Maria Rosaria Sansoni<br />

Department of Chemistry, University of Siena, Pian dei Mantellini, 44<br />

53100 Siena ITALY<br />

•Department of Chemistry, University of Sassari, Via Vienna, 2<br />

07100 Sassari ITALY<br />

Folic acid N-[4-(2-Amino-4-hydroxypteridinyl-(6)-methylamino)-benzoyl]-Laminoglutaric<br />

acid (Figure 1) is a<br />

fundamental coenzyme involved in onecarbon<br />

unit transfer processes 1 . The solid<br />

state conformation of folic acid has been<br />

defined but there have been few<br />

investigations on the structure of this<br />

coenzyme in solution^.<br />

12<br />

by Nuclear Magnetic Resonance (NMR).<br />

Information on dynamical motion,<br />

magnetic dipolar connectivities and<br />

energy minimization calculations are<br />

combined in order to define the solution<br />

structure of folic acid.<br />

2 Experimental<br />

Two-dimensional COSY, NOESY and Hetcor<br />

Figure 1: Structure and numbering of folic acid.<br />

f ole is also fundamental in reductive<br />

zymatic processes in which<br />

fchydrofolate (the reduced form of folic<br />

) is oxidized to dihydrofolate and folate.<br />

•aihydrofolate reductase NADPH-<br />

•ndent enzyme controls the biological<br />

of tetrahydrofolate 3 .<br />

present paper the conformational<br />

of this molecule were analyzed<br />

experiments were obtained by (3C/2-ti -7C/2-<br />

AT)n, 4 (rc/2-tl-TC/2-tm-JU/2AT)n 5 a n d<br />

pulse sequences respectively. Spin-lattice<br />

relaxation rates were measured using the<br />

(180°-T-90°-t)n pulse sequence. The NMR<br />

measurements were performed using a 0.12<br />

mol.dm' 3 DMSO-d6 solution at 27°C. NMR<br />

spectra were recorded on a Varian XL-200<br />

181


182<br />

and a Bruker AMX-600 spectrometers<br />

operating at 200 and 600 MHz respectively.<br />

Molecular mechanics calculations were<br />

computed by the MacroModel program,<br />

version 2.5^ implemented on a Vax 11/750<br />

computer. The force field used was that<br />

reported by Weiner et al^.<br />

3 Results and Discussion<br />

In the present paper the NMR properties of<br />

folic acid were investigated in depth in<br />

order to define the molecular structure in<br />

DMSO-d6 solution.<br />

As the proton and carbon assignments of<br />

folic acid refer only to water solution at<br />

basic pH^^O, both proton and carbon<br />

chemical shifts in DMSO-d6 were<br />

determined by a method based on<br />

conventional two-dimensional COSY and<br />

Hetcor experiments. Figures 2 and 3 show<br />

the COSY and Hetcor spectra of folic acid.<br />

The complete proton assignments and<br />

chemical shifts of protonated carbons can<br />

be determined from these sets of data. The<br />

quaternary carbons were assigned by a<br />

frequency-dependent selective protoncarbon<br />

NOE experiment 11 > 1 ^.<br />

The results obtained are reported in Table 1.<br />

The strategy for structural analysis was<br />

based on a combined approach. First we<br />

studied the dynamical properties of folic<br />

acid in solution which led to two possible<br />

scenarios:<br />

i) the molecular motion is subject to<br />

different degrees of freedom, in which case<br />

each molecular moiety behaves<br />

independently as a consequence of the lack<br />

of "ordered" elements;<br />

ii) the molecule is subject to overall<br />

dynamical reorientation, characterized by<br />

a single rotational correlation time, %c.<br />

These conditions, for molecules of the size<br />

of folic acid, are verified whenever noncovalent<br />

interactions stabilize specific<br />

conformations.<br />

The appropriate method for dynamical<br />

investigation is based on analysis of the<br />

carbon spin-lattice relaxation rate, Ric.<br />

The experimental Ric> calculated for<br />

protonated carbons, are reported in Table 1.<br />

From these data and using the Allerhand's<br />

approach 1 3 a unique correlation time<br />

value of 3x10'^ s was calculated. In Table 1<br />

the selective and non-selective proton<br />

Bulletin of Magnetic Resonance<br />

relaxation rates are also reported. These<br />

experimental values confirm the dynamical<br />

region of the isotropic molecular motion of<br />

folic acid in solution. A second set of<br />

structural information can be derived from<br />

the study of the extent of the dipolar<br />

magnetization transfer in the protonic<br />

environment.<br />

In this case the NOESY spectrum can<br />

provide the complete network of the<br />

proton-proton dipolar interactions, which<br />

is related to internuclear distances. The<br />

analysis of the NOESY spectrum enables us<br />

to identify proton pairs in which the<br />

cross-relaxation contribution is<br />

significant. These include the NH(ifj)-<br />

H(12/16). NH(io)-H(9), H(i2/16)-H(9),<br />

H(12/16)-H(13/15) and H( i 8)-H(2 lb).<br />

Different cross-peaks due to exchange<br />

contributions between two different sites<br />

can also be detected in the NOESY spectrum.<br />

These cross-peaks are related to the<br />

NH(1O)/NH2(2) — HOD exchange.<br />

Information on the extent of dipolar<br />

interactions can be used as experimental<br />

"constraints" in theoretical energy<br />

minimization calculations. The presence of<br />

an exchange process selectively restricted<br />

to the NH(iO)/NH2(2) — HOD protons<br />

suggests the involvement of other<br />

exchangeable nuclei such as NH(i8) in non<br />

covalent interactions (e.g. hydrogenbonds),<br />

important for the stabilization of<br />

the conformation of folic acid in solution.<br />

Further evidence of the slow chemical<br />

exchange process of NH(J8) with respect to<br />

NH (10)/NH2(2) can be obtained from<br />

saturation transfer experiments.<br />

By irradiating the HOD resonance for a<br />

sufficient period of time, with a selective<br />

frequency, a strong reduction in signal<br />

intensity is observed on protons involved<br />

in exchange phenomena. The experimental<br />

findings show that the NH(1O)/NH2(2) .<br />

signal is drastically affected by HOD ;<br />

saturation whereas NH(18) does not show ]<br />

significant intensity variation. This is<br />

further evidence of the importance of |<br />

NH(i8) in the stabilization of selective ;<br />

conformation by hydrogen bond with the J<br />

C(23) carboxyl group of glutamic acid. Thi<br />

hypothesis is confirmed by the selectivity j<br />

in NH(i8)-H(2l) dipolar connectivities,<br />

observed in the NOESY spectrum. In fact t<br />

specific NH(i8)-H(21b) cross-peak<br />

observed, suggesting the stabilization of ^<br />

unique conformation in solution.


Vol. 14, No. 1-4<br />

F2 (PPM)<br />

9 8 7 6 5 4 3 2 }<br />

Figure 2: COSY spectrum of 0.12 mol.dm-3 folic acid DMSO-d6 solution at 27°C.<br />

12S 1M<br />

Figure 3: 1H-13C Hetero correlation<br />

' solution at 27°C.<br />

I I '—I—I—|—i—t-<br />

spectrum of folic acid in DMSO-d6<br />

183


184 Bulletin of Magnetic Resonance<br />

Table 1<br />

Proton and carbon NMR parameters of 0.12 mol.dm'3 folic acid solution at 27°C.<br />

Nuclei<br />

2<br />

4<br />

4a<br />

6<br />

7<br />

8a<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21A<br />

21B<br />

22<br />

23<br />

8 ppm<br />

8.75<br />

. .<br />

4.59<br />

7.02<br />

6.74<br />

7.75<br />

7.75<br />

6.74<br />

8.22<br />

4.44<br />

2.15<br />

2.01<br />

2.42<br />

—<br />

8 ppm<br />

156.160<br />

161.274<br />

127.945<br />

148.610<br />

148.610<br />

153.823<br />

45.922<br />

150.793<br />

111.216<br />

128.998<br />

121.321<br />

128.998<br />

111.216<br />

166.438<br />

51.762<br />

173.744<br />

26.045<br />

26.045<br />

30.439<br />

173.932<br />

Further evidence of the structure assumed<br />

by folic acid can be obtained from<br />

molecular mechanics calculations using<br />

the Macro Model program and experimental<br />

NMR constraints, a low energy<br />

conformation of -135.5 KJ/mol was<br />

computed after several iterative and<br />

minimizzation cycles.<br />

Figure 4: Solution structure of folic acid<br />

as determined by NMR experimental<br />

constraints and subsequent energy<br />

minimization calculations.<br />

Ric<br />

s-1<br />

0.11<br />

0.21<br />

0.12<br />

0.55<br />

11.9<br />

0.73<br />

6.12<br />

5.70<br />

0.43<br />

5.70<br />

6.12<br />

0.38<br />

6.07<br />

0.37<br />

11.75<br />

11.75<br />

11.60<br />

0.30<br />

RlNS<br />

s-1<br />

0.77<br />

5.40<br />

2.80<br />

2.00<br />

1.87<br />

1.87<br />

2.00<br />

4.71<br />

1.62<br />

5.50<br />

5.55<br />

5.00<br />

—<br />

RlSE<br />

s-1<br />

0.75<br />

--<br />

5.30<br />

1.46<br />

1.34<br />

1.34<br />

1.46<br />

1.17<br />

5.40<br />

5.40<br />

4.90<br />

The minimized structure calculated (Figure<br />

4), shows a NH(18)-C(i9)-C(21)-H(21b)<br />

torsion angle of 315°. This is in agreement<br />

with the experimental NOESY data and<br />

confirms the conformation of folic acid<br />

stabilized in solution by the NH(i8)-C(23)<br />

hydrogen bond.<br />

4 References<br />

1) R.L. Blakley; "The Biochemistry of Folic<br />

Acid and Related Pteridines"; North-<br />

Holland publishers, Amsterdam, (1969).<br />

2) D. Mastropaolo, A. Camerman and H.<br />

Camerman; Science 210, 334 (1980).<br />

3) J.M. Blaney, C. Hansch, C. Silipo and A.<br />

Vittoria; J. Am. Chem. Soc. £4, 303 (1984).<br />

4) A.D. Bax, R. Freeman and G. Morris; J-<br />

Magn. Reson. 42, 169 (1981).


Vol. 14, No. 1-4<br />

5) D.J. States, R.A. Haberkorn and D.J.<br />

Ruben; J. Magn. Reson. 4JL 286 (1982).<br />

6) A.D. Bax and G. Morris; J. Magn. Reson.<br />

42, 501 (1981).<br />

7) C. Still; Macromodel, Columbia Universty<br />

Molecular Modelling System (1987).<br />

8) S. Weiner, P.A. Kollman, D.A. Case, U.C.<br />

Singh, C.Ghio, G. Alagona and P.K. Weiner;<br />

J. Am. Chem. Soc. 106, 765 (1984).<br />

185<br />

9) W. Frick, R. Weber and M. Viscontini;<br />

Helv. Chim. Acta 52, 2658 (1974).<br />

10) M. Poe; Method in Enzymology 6JL 483<br />

(1980).<br />

11) N. Niccolai, C. Rossi, V. Brizzi and W.A.<br />

Gibbons; J. Am. Chem. Soc. 106, 5732 (1984).<br />

12) C. Rossi, N. Niccolai and F. Laschi; J.<br />

Phys. Chem. 9_i, 3903 (1987).<br />

13) A. Allerhand, R.A. Komoroski; J. Am.<br />

Chem. Soc. 9_5_, 8828 (1973).


186<br />

Bulletin of Magnetic Resonance<br />

Characterization of Water-in-Bitumen Emulsions<br />

in Model Porous Media by NMR Microscopic Imaging Techniques<br />

1. Introduction<br />

Leslie H. Randall and George E. Sedgwick<br />

Alberta Research Council,<br />

Oil Sands and Hydrocarbon Recovery Division<br />

PO Box 8330, Station F, Edmonton, Alberta, T6H 5X2.<br />

Production from thermal (steam enhanced) oil<br />

recovery processes is-complicated by the presence of<br />

water-in-oil emulsions. [1], [2] Critical to monitoring<br />

the in-situ formation and flow of these water-in-oil<br />

emulsions is the ability to distinguish between the<br />

various components (bitumen, water/steam and<br />

emulsified water) present in a core flood experiment.<br />

In principle, NMR imaging is ideally suited to<br />

monitor the spatial distribution of absorbed fluids, and<br />

in this regard, the viability of the NMR imaging<br />

technique to examine the distribution of fluids in<br />

reservoir rock samples has recently been<br />

demonstrated. [3] - [15]. In general, several fluids or<br />

phases may be present and the ability to distinguish<br />

these components is of prime interest.<br />

In the majority of these studies, the NMR<br />

imaging technique has been applied to samples which<br />

contain a low viscosity crude oil and water/brine. In<br />

the case of heavy oil production, the fluids have a<br />

high and varying viscosity, (ie: bitumen, water and<br />

emulsions of bitumen and water). Under these<br />

circumstances, it is possible that these differences in<br />

viscosity will lead to the ability to discriminate<br />

between phases via differences in NMR relaxation<br />

behaviour. In addition, the influence that the solid<br />

matrix has on the relaxation behaviour of water<br />

dispersed as an emulsion may be quite different from<br />

water absorbed into the porous medium. In an effort<br />

to evaluate the feasibility of characterizing water-inbitumen<br />

emulsions by NMR microimaging techniques,<br />

and<br />

Colin A. Fyfe<br />

University of British Columbia,<br />

Dept. of Chemistry and Pathology,<br />

Vancouver, B.C., V6T 1Y6.<br />

the one-dimensional NMR spectra, the relaxation time<br />

constants and the spin-echo images for a series of samples<br />

consisting of water, bitumen and water-in-bitumen emulsions<br />

absorbed into glass beads were examined and are presented<br />

herein.<br />

2. Experimental<br />

All water-in-oil (bitumen) emulsions were prepared from Cold<br />

Lake bitumen which has been ultracentrifuged to remove solid<br />

particles. The emulsion samples were created by passing a<br />

heated mixture of bitumen and water through an auxiliary sand<br />

column at a suitable flow rate. The emulsions were checked<br />

under an optical microscope to ensure that the water phase was<br />

well dispersed prior to packing. Eight samples were prepared<br />

each consisting of a different fluid/porous medium mixture.<br />

Sample 1 contained 10 mL of Cold Lake Bitumen. Sample 2<br />

contained 10 mL of a 20 % (w/w) water-in-bitumen. The<br />

remaining samples contained fluid absorbed into a glass bead<br />

matrix. Two different sizes of glass beads were used. The<br />

small glass beads were determined to be 88 -104 pm in<br />

diameter which represents fine sand. The large glass beads<br />

were 0.8 - 1.0 mm in diameter. Sample 3 contained<br />

approximately 10 g of the large glass beads with 5 mL ol<br />

distilled water. Sample 4 had a similar composition to sampl<<br />

3 except that the smaller glass beads were used. Sample -<br />

contained 10 g of the large glass beads with 5 mL of a water<br />

in-oil emulsion (20% w/w water). Sample 6 had a simila<br />

composition except that the smaller glass beads were used. T<<br />

ensure that the composition of the water-in-bitumen emulsio)<br />

was maintained in the glass bead matrix, samples 5 and 6 wep<br />

prepared by mixing the emulsion with the glass beads by haflj


Vol. 14, No. 1-4 187<br />

and then placing the appropriate amount of the<br />

mixture at the bottom of a 10 mm NMR tube. The<br />

samples were then spun at low speeds on a bench-top<br />

centrifuge for 10 minutes to obtain a uniform packing.<br />

In an effort to compare the NMR behaviour of the<br />

two types of fluids directly, Sample 7 was composed<br />

of two regions, the bottom layer contained emulsion<br />

in large glass beads and the top layer contained<br />

distilled water in large glass beads. Sample 8 was<br />

identical in composition to sample 7 except that the<br />

smaller glass beads were used. The contents of the<br />

samples are summarized in Table 1.<br />

Table 1. Composition of Samples<br />

Sample Fluid Matrix<br />

1 Bitumen<br />

2 Emulsion<br />

3 Distilled Water<br />

4 Distilled Water<br />

5 Emulsion<br />

6 Emulsion<br />

7 Emulsion/Water<br />

8 Emulsion/Water<br />

None<br />

None<br />

88 - 104 urn<br />

0.8 - 1.0 mm<br />

0.8 - 1.0 mm<br />

88 - 104 urn<br />

0.8 - 1.0 mm<br />

88 - 104 um<br />

All NMR measurements were made on a Bruker<br />

MSL 400 spectrometer equipped with a microimaging<br />

system using the proton microimaging probe<br />

equipped with a vertical 12 mm saddle coil. The<br />

nonselective 90° rf pulse length was 14.5 ps.<br />

Quadrature phase cycling was used in all the<br />

spectroscopic measurements. ID J H NMR spectra and<br />

Carr-Purcell spin-echo (90-tau-180) NMR spectra [16]<br />

were obtained to characterize the samples. The spinecho<br />

sequence was also used to determine the average<br />

spin-spin relaxation times (T2). The inversionrecovery<br />

sequence [17] was used to determine the<br />

average Tj spin-lattice relaxation times.<br />

After evaluating the relaxation time constants<br />

and the NMR lineshapes for several of the samples,<br />

was determined that the spin-echo imaging<br />

luence [18] was an appropriate choice. The spinio<br />

imaging pulse sequence employs a 90-tau-180 rf<br />

sequence in which a hard 180° pulse sequence<br />

* 110 refocus ^ eff ects of field inhomogeneity.<br />

selection was performed by using a selective<br />

Pulse with an appropriate Gz gradient. Echo<br />

fes for the spin-echo imaging experiments varied<br />

from 4.5 ms to over 100 ms and the actual echo times are<br />

indicated in the text. The slice thickness was typically 2.1<br />

mm. The phase encoding gradient was incremented through<br />

256 experiments. The frequency encode gradient was 5.8<br />

G/cm resulting in an in-plane resolution of 95 um. The<br />

samples used in this study had a porosity of approximately 30<br />

-35 % and the imaging experiments required 1-4 hours to<br />

acquire. Images presented in this paper follow the convention<br />

in which an inverse gray scale is used to indicate relative<br />

intensity. The darker the region on the image, the higher the<br />

concentration of water.<br />

3. Results and Discussion<br />

The relaxation behaviour and linewidths were investigated<br />

(Table 2) to determine the appropriate imaging sequence for<br />

the bitumen and water-in-bitumen emulsions. A ID *H NMR<br />

spectrum of Cold Lake Bitumen consisted of a large peak (Vj^<br />

= 690 Hz) which is assigned to the heavy oil component<br />

(bitumen) and a small shoulder which is attributed to trace<br />

amounts of connate water. The spin-spin relaxation time (T2)<br />

of the oil component was determined to be 1.3 ms which<br />

indicates that quantitative spin-echo imaging of the bitumen<br />

component in the samples will not be possible.<br />

The linewidth determined for distilled water placed in the<br />

small glass beads (Sample 3) is approximately 1000 Hz, an<br />

increase of more than 2 orders of magnitude over that observed<br />

for a distilled water phantom. Similar line broadening was<br />

observed for the other samples (Table 2). The line broadening<br />

is a result of the magnetic susceptibility differences between<br />

the solid matrix and the absorbed fluids. These line-widths<br />

indicate that the application of the gradient echo sequence [18]<br />

would be difficult due to the short T2* values (T2* = 300-800<br />

ps). The spin-spin relaxation parameter for the distilled water<br />

absorbed into either glass bead matrix (Table 2) is markedly<br />

reduced from that observed for bulk water. An NMR image<br />

of Sample 4 with an echo time of 4.5 ms demonstrates that the<br />

distribution of water in these types of samples are possible.<br />

(Figure la) The image with an echo time of 104 ms (Figure<br />

lb) is more interesting. It shows a large decrease in signal<br />

intensity, particularly where the distilled water is in contact<br />

with the glass beads. Due to the non-uniform packing of the<br />

large glass beads, small pockets of water on the order of 100 -<br />

400 um in diameter can be observed.<br />

An examination of the NMR relaxation parameters of the<br />

20% (w/w) water-in-bitumen emulsion (Sample 5) absorbed<br />

into the large glass bead matrix reveals that the T2 relaxation<br />

time constant of the water phase is unaffected by placing the<br />

emulsion into the glass bead matrix. The image obtained with<br />

a 104 ms echo time (Figure 2) displays a strong uniform signal<br />

intensity. The bitumen component of the sample has a T2<br />

relaxation time constant on the order of a millisecond, and thus<br />

does not contribute to the intensity of the image.


i ii<br />

'I,' , ',<br />

il i I<br />

i i<br />

188 Bulletin of Magnetic Resonance<br />

Figure 1 (a) Spin echo images of sample 4.<br />

a) Echo time = 4.5 ms. (b) Echo time = 104 ms<br />

During a core-flood experiment it is likely that both<br />

water and emulsified water phases will be present.<br />

The ability to distinguish between bulk absorbed<br />

water and emulsified water is considered essential to<br />

analyzing the formation and flow of the emulsions in<br />

sand packs. To this end, a sample which contains<br />

both distilled water and water-in-bitumen emulsion in<br />

a glass bead pack was examined (Samples 7 and 8).<br />

The conditions of the imaging experiments were<br />

chosen such that only the water component of the<br />

samples were observed. Using an echo time of 4.5<br />

ms the distribution of water in sample 7 (Figure 3a)<br />

can be obtained. The distilled water plus large glass<br />

beads are in the top half of the NMR tube and this<br />

results in a much higher signal intensity due to the<br />

higher water concentration. To discriminate between<br />

the emulsified water phase and the absorbed bulk<br />

water, an echo time of 104 ms was used. (Figure 3b)<br />

Under these conditions, only water which has a high<br />

mobility or low surface contact will be observed (ie:<br />

water which is emulsified will be favoured). This<br />

results in an image in which small pockets of water<br />

are observed in the distilled water region, while the<br />

emulsion containing region appears nearly uniform.<br />

The effect of the smaller grain size of the<br />

relaxation parameters was examined using sample 8.<br />

The smaller glass beads have a more uniform pore<br />

size distribution and are more representative of the<br />

matrix used in core flood experiments. In such<br />

samples, the pore size is typically on the order of 30<br />

pm which means that all of the water will be in<br />

intimate contact with the solid matrix.<br />

Figure 2. Spin-echo image of sample 5.<br />

Echo time = 104 ms.<br />

Figure 3. Spin echo images of sample 7.<br />

a) Echo time = 4.5 ms. (b) Echo time = 104 ms<br />

The T2 of the distilled water component is reduced to 4.9 ms,<br />

on the same order of magnitude as the echo time (4.5 ms).<br />

This means that the intensity of the water will be strongly T2<br />

weighted and the region which contains the distilled water will<br />

no longer have an intensity which is much higher than the<br />

emulsion. (Figure 4a) When the echo time is increased to 34<br />

ms, (Figure 4b) the region which contains the distilled water<br />

disappears in a uniform manner. The smaller grain size of the<br />

solid matrix has enlarged the relaxation time differences<br />

between the bulk absorbed water and the emulsified water and<br />

thus the contrast between the two physical states are enlarged.


Vol. 14, No. 1-4 189<br />

Figure 4. Spin echo images of sample 8.<br />

a) Echo time = 4.5 ms. (b) Echo time = 34.5 ms<br />

3. Summary<br />

In the present study, the feasibility of examining<br />

heavy oil emulsion samples has been demonstrated.<br />

Bitumen, water and emulsified water placed into glass<br />

beads are easily distinguished in an NMR imaging<br />

experiment on the basis of their relaxation times. The<br />

spin-spin relaxation time constant, T2 observed for the<br />

water component of the emulsion samples placed in<br />

contact with glass beads was dramatically different<br />

than that found for distilled water. The NMR<br />

experiment is therefore sensitive to the physical state<br />

of the water, ie: whether the water is emulsified and<br />

therefore 'protected' from the solid matrix. This<br />

difference in relaxation times was exploited to provide<br />

water-selective images of the emulsified water phase<br />

in the presence of oil and non-emulsified water<br />

phases. Good S/N images can be obtained with high<br />

in-plane resolution (50-100 pm) can be obtained on a<br />

microimaging system in reasonable time periods.<br />

Saturation profiles can be obtained in a manner of<br />

seconds which would allow for the continuous<br />

monitoring of the formation of an emulsion under<br />

flow conditions. The ability to monitor the formation<br />

and flow of water emulsions will be important to<br />

understanding enhanced oil recovery processes.<br />

Table 2: Proton Relaxation Times and Linewidths of Water<br />

and Bitumen<br />

Sample<br />

1<br />

2 water<br />

bitumen<br />

3<br />

4<br />

5 water<br />

bitumen<br />

6 water<br />

bitumen<br />

Ti(s)<br />

0.71<br />

2.2<br />

0.62<br />

2.4<br />

2.3<br />

1.6<br />

0.58<br />

1.5<br />

0.45<br />

T2 (ms)i<br />

1.3<br />

200*<br />

1.8<br />

29<br />

4.9<br />

495<br />

1.3<br />

484<br />

1.4<br />

v l/2<br />

690<br />

150<br />

650<br />

350<br />

1000<br />

The water component of emulsion samples show a variation<br />

in T2. The reasons for this variation are under investigation<br />

but are believed to be due to the size distribution of the water<br />

droplets.<br />

Acknowledgements<br />

The authors wish to acknowledge G. A. Kissel and D. Vu of<br />

the Alberta Research Council for the invaluable technical<br />

assistance in preparing the samples. The financial assistance<br />

of NSERC is acknowledged (CAF).<br />

4. References<br />

1. Thermal Recovery of Oil and Bitumen, Butler, R. B.<br />

Prentice Hall, New Jersey, 1991.<br />

2. Emulsions; Fundamentals and Applications in the Petroleum<br />

Industry, Schram, L. L., Ed. American Chemical Society, 1992.<br />

3. Rothwell, R. P.; Vinegar, H. J. Applied Optics, 24, 1985,<br />

3969-3972.<br />

4. Blackband, S.; Mansfield, P.; Barnes, J. R.; Clague, A. D.<br />

H.; Rice, S. A. Soc. Pet. Eng. Form. Evaln.l, 1986, 31-34.<br />

5. Hall, L. D.; Rajanayzgmi, V.; Hall, C. /. Magn. Res. 68,<br />

1986, 185-188.<br />

6. Baldwin, B. A.; Yamanashi, W. S. Mag. Res. Imaging, 6,<br />

1988, 493-500.


190 Bulletin of Magnetic Resonance<br />

7. Hall, L. D.; Rajanayagmi, V. /. Magn. Res. 74,<br />

1987, 139-147.<br />

8. Chen, J. D.; Dias, M. M.; Patz, S.; Schwartz, L. M.<br />

Phys. Rev. Lett. 61, 1988, 1489-1492.<br />

9. Edelstein, W. A.; Vinegar, H. J.; Tutunjian, P. N.;<br />

Roemer, P. B.; Mueller, O. M. SPE preprint 18272,<br />

63rd Annual Technical Conference, Houston, 1988,<br />

101-112.<br />

10. Mandava, S. S.; Watson, A. T.; Edwards, C. M.<br />

Amer. Inst. Chem. Eng. 36, 1990, 1680-1686<br />

11. Majors, P. D.; Smith, J. L.; Kovarik, F. S.;<br />

Fukushima, E. /. Magn. Res. 89, 1990, 470-478.<br />

12. Woessner, D. E.; Gleeson, J. W.; Jordan, C. F.<br />

SPE preprint 20493, 65th Annual Technical<br />

Conference, New Orleans, 1990, 247-253.<br />

13. Osment, P. A.; Packer, K. J.; Taylor, M. J.;<br />

Attard, J. J.; Carpenter, T. A.; Hall, L. D.; Herrod, N.<br />

J.; Doran, S. J. Phil. Trans. R. Soc. Lond. A. 333,<br />

1990, 441-452<br />

14. Dereppe, J. M.; Moreaux, C; Schenker, K. J.<br />

Magn. Res. 91, 1991, 596-603.<br />

15. Dechter, J. J.; Komoroski, R. A.; Ramaprasad, S.<br />

/. Magn Res. 93, 1991, 142-150.<br />

16. H. Y. Carr and E. M. Purcell, Phys. Rev., 94,630<br />

(1954).<br />

17. R. L. Void, J. S. Waugh, M. P. Klein and D. E.<br />

Phelps, J. Chem. Phys., 48, 383 (1968).<br />

18. Edelstein, W.A.; Hutchinson, J.M.S.; Johnson G.;<br />

Redpath, T. Phys. Med. Bioi, 25, 751 (1980).<br />

19. A. Haase, J. Frahm, D. Matthaei, W. Hanicke and<br />

K. D. Merboldt, /. Magn. Reson., 67, 258 (1986).


Vol. 14, No. 1-4 191<br />

COMPUTER GRAPHICS FOR PULSE SEQUENCE<br />

ANALYSIS<br />

Introduction<br />

Jonathan Callahan, Debbie Mattiello and Gary P. Drobny<br />

The development of software for<br />

the simulation of NMR experiments<br />

has increased at an enormous pace in<br />

the last decade. Its usefulness has<br />

spread far beyond the analysis of<br />

lineshapes and spectra. Today,<br />

search-and-optimize strategies are<br />

used to develop new pulse sequences<br />

while other programs measure the<br />

performance of pulse sequences on<br />

spin systems of interest [1, 2]. Up to<br />

now, most of the output from these<br />

programs has been displayed as twodimensional<br />

hard-copy output. With<br />

the arrival of relatively inexpensive<br />

graphics workstations, the possibility<br />

of visualizing the time development of<br />

the density operator has spurred us to<br />

develop graphics software in<br />

conjunction with our ongoing<br />

development of simulation software.<br />

Current projects which benefit<br />

from graphical analysis include the<br />

development of "time-suspension"<br />

sequences for use with solids imaging,<br />

development of improved sequences<br />

for the creation of Zeeman or<br />

quadrupolar order in deuterium NMR,<br />

and analysis of artefacts seen in<br />

imaging experiments in the presence<br />

of flow.<br />

Chemistry Dept. University of Washington<br />

Seattle, Washington 98195 USA<br />

Time Suspension<br />

"Time suspension" sequences are<br />

multi-pulse sequences which remove<br />

the effects of both the dipolar coupling<br />

and the chemical shift Hamiltonians<br />

[3-5]. They are useful in pulsedgradient<br />

imaging experiments where<br />

imaging gradients are applied only<br />

during the multi-pulse windows [6-8].<br />

Currently implemented pulse<br />

sequences apply average Hamiltonian<br />

theory and use carefully cycled<br />

"wahuha" subcycles to achieve the<br />

suppression of internal Hamiltonians<br />

[3]. Like other multi-pulse<br />

experiments these sequences perform<br />

best at or near resonance and show<br />

decreased line-nawowing as one<br />

moves off resonance. A new "timesuspension"<br />

sequence developed with<br />

computer search-and-optimization<br />

techniques (CDIS-4) [9] achieves<br />

comparable reduction of the internal<br />

Hamiltonians but shows<br />

complementary behavior: poor line<br />

narrowing on resonance but improved<br />

performance off resonance. In an<br />

effort to understand this behavior we<br />

calculate the evolution of the density<br />

operator for a two-spin system under<br />

the influence of effective Hamiltonians<br />

defined by the multi-pulse


192 Bulletin of Magnetic Resonance<br />

propagators associated with each<br />

sequence.<br />

With perfect wahuha type<br />

sequences the trajectory of the net<br />

magnetization should be around the<br />

base of a cone whose axis is along the<br />

cube diagonal in spin space (ie. in the<br />

nodal plane of the coupling ternsor).<br />

When chemical shift offsets are small<br />

(


Vol. 14, No. 1-4 193<br />

greater than 10KHz the conventional<br />

sequence shows modulated rotation<br />

(Fig. 1C) whereas the new sequence<br />

traces out a somewhat bent circle<br />

lying in the nodal plane of the<br />

chemical shift tensor (Fig.ID). From<br />

such pictures we hope to gain a better<br />

understanding of the effect of error<br />

terms which are not easily amenable<br />

to analytical treatment.<br />

Deuterium<br />

The deuterium quadrupole is an<br />

excellent probe of dynamics and as<br />

such is synthetically incorporated into<br />

DNA and other biologically important<br />

molecules [10]. The same quadrupole<br />

which allows one to use deuterium as<br />

a probe also presents formidable<br />

experimental difficulties when the<br />

strength of the quadrupole approaches<br />

the strength of the rf field. This<br />

situation is realized in some of our<br />

labeled oligonucleotides. In this<br />

regime, the effective axis about which<br />

the magnetization is rotated (the sum<br />

of rf and quadrupolar terms) is<br />

substantially different from the rf<br />

axis.<br />

Current pulse sequences which<br />

convert Iz magnetization into -Iz or<br />

into quadrupolar order are found to<br />

be inefficient at higher values of the<br />

quadrupole coupling. For an inversion<br />

pulse this means incomplete inversion<br />

at the shoulders of the powder<br />

pattern. This is unacceptable in<br />

experiments where one attempts to<br />

measure the orientation dependence<br />

of spin-lattice relaxation time. Such<br />

measurements are necessary to prove<br />

or disprove particular dynamical<br />

models.<br />

An example of non-uniform<br />

excitation is given in Fig. 2) which<br />

shows the development of spin<br />

coherences during a 180 degree pulse<br />

at five different values of the<br />

quadrupole coupling (OKHz, +/- 62.5<br />

KHz and +/- 125KHz). The operator<br />

basis of Vega and Luz [11] is used<br />

because of the simple form of the<br />

quadrupolar operator in that basis:<br />

Ix<br />

Jx=IyIZ+IzIy<br />

Kx=Iy 2 -Iz 2 Ky=Iz IVy=lZ —1X<br />

x-lx -1 ^y-Iy -1<br />

2 -Ix 2 K2=Ix2-Iy 2<br />

QZ=Iz 2 -I 2<br />

Rotation about Ix during the<br />

pulse is modified by (0q dependent<br />

rotation about the Qz axis. This causes<br />

a buildup of "antiphase" Jy and zeroquantum<br />

Jz. By the end of the pulse it<br />

is clear that a simple n pulse is<br />

ineffective at creating -Iz over the<br />

entire range of couplings (+/-125 KHz).<br />

We are again using search and<br />

optimize strategies to find pulse<br />

sequences which create -Iz or Qz<br />

evenly over a broad range of<br />

quadrupole couplings [12]. With its<br />

small operator basis, deuterium NMR<br />

provides us with an excellent system<br />

on which to develop interactive<br />

computer aided pulse sequence<br />

design. With user control of pulse<br />

amplitude and phase, rapid calculation<br />

of the time evolution of the spin 1<br />

density operator and a graphical view<br />

of Hilbert space we will soon have the<br />

opportunity to design pulse<br />

sequences and shaped pulses<br />

intelligently rather than relinquishing<br />

our insight to the cpu.


194<br />

Qz<br />

\lx<br />

Jx<br />

Iz<br />

iy<br />

Bulletin of Magnetic Resonance<br />

Figure 2) Evolution of spin coherences for deuterium with ()(•), +/- 62.5(AY) and +/-<br />

KHz quadrupole coupling. The evolution is depicted in the operator basis of Vega and Lus<br />

Rotation about Ix due to the rf pulse is seen but rotation about quadrupole operator Qz is als<br />

evident. For +/- 125 KHz coupling the simple K pulse is very inefficient at creating -Iz.<br />

Flow Imaging<br />

Another simulation program<br />

which benefits from graphical display<br />

calculates the response of flowing<br />

spins in an NMR imaging experiment.<br />

Our interest in this area focuses on the<br />

artefacts that arise when spins move<br />

from a region with one gradient<br />

strength to a region with anothe<br />

during an imaging experiment. Ou<br />

simulation incorporates the effects c<br />

flow and of arbitrarily comple<br />

imaging sequences on an array c<br />

spins 1/2. When following an<br />

particular spin through time we kee<br />

track of its three-dimensional positio<br />

and its magnetization vector. Thus w


Vol. 14, No. 1-4 195<br />

have for each spin seven parameters<br />

(3 space, 3 spin and time) which<br />

describe its state.<br />

In order to understand how<br />

artefacts develop during the<br />

experiment we display the system as<br />

an array of vectors whose position<br />

corresponds to spatial position and<br />

whose orientation corresponds to<br />

magnetization state. From the z-axis<br />

we observe spatial flow of spins in the<br />

spatial x-y plane and also the<br />

magnitude and phase of transverse<br />

magnetization in the superposed spin<br />

x-y plane.<br />

A<br />

B<br />

c<br />

/*-<br />

In Fig. 3) we see how the sliceselect<br />

portion of an imaging<br />

experiment can be perturbed by flow<br />

in the direction of the slice-select<br />

gradient. As the rate of flow<br />

increases, the width of the slice<br />

remains fairly constant but the phase<br />

order of the spins in that slice<br />

deteriorates. Such phase disorder will<br />

lead to artefacts along the phaseencode<br />

dimension at the edges of the<br />

slice. As it now exists our simulation<br />

will allow us to evaluate imaging<br />

sequences on flow geometries<br />

Flow / Imaging Gradient<br />

Figure 3) Excitation profiles for the slice-select portion of an imaging experiment. In this<br />

simulation the direction of flow is along the slice-select gradient causing spins to change<br />

their frequency during the sine excitation pulse. The flow velocities are in arbitrary units<br />

but three regimes are displayed: A) static spins; B) slow flow; C) moderate flow.


iii.:<br />

196<br />

of interest. A better understanding of<br />

the formation of such errors will aid<br />

us in the development of improved<br />

imaging sequences.<br />

Conclusion<br />

We have extended the computer<br />

techniques available to the NMR<br />

spectroscopist by presenting<br />

simulated data in multi-dimensional<br />

animations. With these animations it<br />

is much easier to see the development<br />

of spin coherences which are the<br />

result of experimental imperfection or<br />

which are intended by design. Our<br />

original goal with computer graphics<br />

was to enhance our own<br />

understanding of the experiments we<br />

perform and to aid us in experimental<br />

development In the process we have<br />

found animated simulations to be a<br />

generally useful pedagogical tool for<br />

explaining all types of NMR<br />

experiments. With computer designed<br />

pulse sequences containing unusual<br />

phases and non-analytical shapes<br />

becoming more and more common we<br />

hope to bring some of the intuition<br />

back to experimental design.<br />

References<br />

[1] S. J. Glaser and G. P. Drobny,<br />

.Adv. Mag. Res., 14, 35 (1990)<br />

[2] H. Liu, S. J. Glaser, and G. P.<br />

Drobny, J. Chem. Phys., 93(111),<br />

7543 (1990)<br />

[3] D. G. Cory, J. B. Miller, and A. N.<br />

Garroway, /. Mag. Res.,9Q, 205<br />

(1990)<br />

Bulletin of Magnetic Resonance<br />

[4] P. Caravatti, L. Braunschweiler,<br />

and R. R. Ernst, Chem. Phys. Lett.,<br />

100(4), 305 (1983)<br />

[5] P. Mansfield and P. K. Grannell,<br />

Phys. Rev. B., 12(9), 3618<br />

(1975)<br />

[6] D. G. Cory and W. S. Veeman, J.<br />

Mag. Res., 84, 392 (1989)<br />

[7] J. B. Miller, D. G. Cory, and A. N.<br />

Garroway, Chem. Phys. Lett.,<br />

164(1), 1 (1989)<br />

[8] J. B. Miller, D. G. Cory, and A. N.<br />

Garroway, Philos. Trans. R. Soc.<br />

London, Ser. A., 333(1632), 413<br />

(1990)<br />

[9] J. Iwamiya, S. Sinton, J. Callahan,<br />

and G. P. Drobny, in abstracts of<br />

the 33 rd ENC, Asilomar, CA USA,<br />

221 (1991)<br />

[10] T. M. Alam and G. P. Drobny,<br />

Chem. Rev., 91, 1545 (1991)<br />

[11] A. J. Vega and Z. Luz, /. Chem.<br />

Phys., $6, 1803 (1987)<br />

[12] D. Mattiello, J. Callahan, T. M<br />

Alam, and G. P. Drobny, in<br />

Proceedings of the 13 th <strong>ISMAR</strong><br />

Meeting, Vancouver, B.C Canada<br />

(1992)


Vol. 14, No. 1-4 197<br />

NMR INVESTIGATION OF THE SIMULTANEOUS<br />

FERMENTATION OF XYLOSE AND GLUCOSE BY A<br />

SELECTED STRAIN OF KLEBSIELLA PLANTICOLA (Gil).<br />

C.Rossi*, A.Lepri*, M.P.Picchi*, S.Bastianoni*, D.Medaglini 0 ,<br />

M.Vanassina 0 and E.Cresta*.<br />

•Department of Chemistry, University of Siena, Pian dei Mantellini<br />

44, 53100 Siena, ITALY.<br />

°Department of Molecular Biology, University of Siena, 53100<br />

Siena, ITALY<br />

1. INTRODUCTION<br />

The hydrolysis of hemicellulose<br />

yields a mixture of sugars of which<br />

D-xylose and D-glucose are the<br />

major constituents.O) This sugar<br />

mixture is an excellent substrate for<br />

growing microorganisms and yields<br />

high energy products such as<br />

ethanoH 2 ' 3 ). In developing this<br />

project two main problems have to<br />

be analyzed in detail: i) the isolation<br />

and identification of microorganisms<br />

whose metabolism can be sustained<br />

by the hemicellulose-derived sugar<br />

mixture, ii) the characterization of<br />

the metabolism, and the selection of<br />

specific metabolic pathways of<br />

microorganisms growing on sugar<br />

mixtures. It has already been shown<br />

that the use of selectively carbon-13<br />

enriched substrates enables the "in<br />

vivo" metabolization process of<br />

microorganisms and tissues^ 4 ' 5 ) to be<br />

studied by NMR. This technique was<br />

applied to the study of the<br />

simultaneous fermentation of xylose<br />

and glucose by a newly isolated<br />

Klebsiella planticola (G 11) strain.<br />

In the present investigation [2- 13 C]glucose<br />

and [l- 13 C]-xylose were<br />

used as carbon-13 enriched<br />

substrates. Isotopic enrichment in<br />

different positions of the sugar chain<br />

enabled us to: i) separate the xylose<br />

from the glucose signals in the<br />

carbon spectrum and ii) calculate<br />

the contribution of each sugar to<br />

end-product yield.<br />

2. EXPERIMENTAL<br />

Klebsiella planticola Gil was<br />

isolated from the soil of a corn field<br />

and selected for its capacity of<br />

growing on a mixed sugar<br />

substrate^ 6 ). Identification of the<br />

bacterium was performed on the<br />

basis of taxonomic characters and<br />

biochemical behaviour. The<br />

microorganism was cultivated in a<br />

nitrogen atmosphere at pH 7.5 and<br />

35°C on a mineral medium<br />

containing 0.2 g/1 of yeast extract.<br />

The metabolism of the bacterium<br />

was investigated by "in vivo" NMR<br />

spectroscopy, using selectively


198<br />

carbon-13 enriched sugar<br />

substrates. The cell culture used for<br />

microbatch NMR experiments had an<br />

initial Optical Density (O D) of 0.5.<br />

The microorganism was grown on a<br />

NMR coaxial tube containing D2O as<br />

lock signal on the external section<br />

and positioned permanently in the<br />

magnetic cavity.during fermentation<br />

[l-^CJ-xylose and R-^CJ-glucose<br />

obtained from Cambridge Isotope<br />

Laboratories were used as enriched<br />

substrates.<br />

c)<br />

Bulletin of Magnetic Resonance<br />

The sugar metabolic process was<br />

followed by recording carbon<br />

spectra at 30 minute intervals until<br />

the end fermentation.<br />

3. RESULTS AND DISCUSSION<br />

The NMR spectra of the enriched<br />

sugar and the end-products of sugar<br />

fermentation are shown in Figure 1.<br />

The metabolization of xylose and<br />

glucose followed two different and<br />

iBU|iiii|Dii|iuuiiiijiiii|uiijuii|au|iiiiiniijiiiiiau|iiiiiiiii|iiiijuii|in inii|iui IIIII|INI IIIII|IIII jaiijiiimiii|iiiiiuii|iiii iu<br />

180 160 140 120 100 80 60 40 20 PPM 180 160 140 120 100 80 60 40 20 PPM<br />

Figure 1 - 1 3 C-NMR spectra obtained at the begining (left) and at the end<br />

(right) of fermentation by K. planticola Gil. Total sugar lOg/1; pH<br />

7.5; Temperature 35 °C. a) [2- 13 C]-glucose fermentation; b) [1- 13 C]xylose<br />

fermentation and c) simultaneous [l-l 3 C]-xylose and [2-<br />

* 3 C]-glucose fermentation. End-product signals detected by * 3 C-<br />

NMR were:l) [2- 13 C]-lactic acid, 2) [2- 13 C]-ethanol, 3) formic acid,<br />

4) [2- 13 C]-succinic acid 5) [l- 13 C]-lactic acid 6) [I- 13 C]-acetic acid<br />

and 7) [l- 13 C]-ethanol.<br />

The end-products of fermentation<br />

were identified on the basis of the<br />

NMR chemical shifts. Analysis of the<br />

dependence of chemical shift on pH<br />

enabled the preliminary<br />

identification of carboxylic and non<br />

carboxylic end-products.<br />

independent pathways: "ethanol and<br />

mixed acids" for xylose( ? ) and<br />

"Emboden-Meyerhof" for glucose.<br />

Diagrams of the sugar pathways are<br />

shown in Figure 2. Glucose<br />

metabolism was also analyzed using<br />

[1- 13 C] enrichment. The same end-


Vol. 14, No. 1-4<br />

products as<br />

enrichment<br />

detected.<br />

a)<br />

for [2- 13 C]-glucose<br />

(Figure 1A) are<br />

i<br />

Xylulose<br />

5, phosphate<br />

Pentose phosphate pathway<br />

biomass have been detected. The<br />

increase in uptake rate with<br />

increasing xylose concentration in<br />

b)<br />

Fructose<br />

1,6 diphosphote<br />

Dihydroxyaceton<br />

phosphate<br />

Figure 2 - Metabolic pathways of xylose (a) and glucose (b) fermentation<br />

by Klebsiella planticola Gil.<br />

The metabolization of xylose by<br />

Klebsiella planticola Gil was<br />

interesting from the view points of<br />

the metabolization rate and the endproducts<br />

of the process. Following<br />

the xylose metabolization process by<br />

NMR in a range of sugar<br />

concentration from 5 to 100 g/1, the<br />

uptake rate can be calculated. This<br />

parameter it is very important<br />

because is correlated with the<br />

efficiency of the process of sugar<br />

transport through the cell<br />

membrane. Table 1 shows the xylose<br />

uptake rate during the first three<br />

hours of fermentation. Sugar<br />

consuption rates for Klebsiella<br />

planticola ATCC 33531, in the range<br />

of 0.8-1.6 g.l^.h" 1 per gram of<br />

AoeBc<br />

Acid<br />

our case is indicative of a "low<br />

affinity" mechanism (not previously<br />

detected in Klebsiella species)<br />

similar to the "facilitated diffusion"<br />

transport process described in<br />

Candida sheabetaeS 9 ^<br />

Figure 3 A shows the effect of<br />

glucose on the metabolization of<br />

xylose. The NMR spectra show that<br />

glucose does not interfere with<br />

xylose metabolism and its uptake<br />

rate. Figure 3B reports the results of<br />

a similar experiment, in which<br />

glucose was added after two hours<br />

of xylose fermentation. The addition<br />

of glucose did not affect xylose<br />

fermentation. It is very unusual for<br />

two substrates to be metabolized at<br />

199


200 Bulletin of Magnetic Resonance<br />

rm 111T111111 nr [ 1111 n 11111 rn iTTT I I I I M I I I I I I I I I I I I I M I I<br />

170 165 160 155 150PPM45 95 90 85 BO 75 PPH7C<br />

Figure 3)- ^C NMR spectra recorded at different stages of sugar<br />

fermentation, a) Simultaneous fermentation of xylose<br />

and glucose b) The effect of addition of glucose after 2<br />

hours of xylose metabolization. Spectra were recorded<br />

every 30 minutes.<br />

TABLE 1<br />

Xylose uptake rate calculated from NMR spectra during the<br />

first three hours of fermentation.<br />

Xylose<br />

concentration<br />

g/1<br />

5<br />

10<br />

20<br />

40<br />

50<br />

80<br />

100<br />

a) per gram of dry weight biomass<br />

Uptake rate a<br />

g.l-i.h-l<br />

1<br />

2.1<br />

3.5<br />

4.6<br />

5.1<br />

6.2<br />

7.0


Vol. 14, No. 1-4<br />

the same time. The phenomenon has<br />

been hypothesized before but has<br />

only ever been demonstrated in<br />

Candida sheabetae. The non diauxic<br />

growth of Klebsiella planticola Gil is<br />

very important because it could be<br />

used to ferment sugar mixtures like<br />

that obtained from the hydrolysis of<br />

hemicellulose.<br />

If the property of good xylose<br />

uptake can be combined with good<br />

end-product yields by selection or<br />

genetic engineering^ 10 ' 11 ), it will<br />

constitute a further advance in the<br />

development of bioethanol<br />

production.<br />

4. REFERENCES<br />

1) T.E.Timell; Adv. Carbohydr. Chem.<br />

19, 247 (1964).<br />

2) K.Skoog and B.Hahn-Hgerdal; Enz.<br />

Microb. Technol., 10, 66 (1988).<br />

3) H.Shneider; Critical Review in<br />

Biotecnology, 9, 18 (1989).<br />

4) K.Ugurbil, T.R.Brown, J.A.Den<br />

Hollander, P.Glynn and R.Shulman;<br />

Proc. Acad.<br />

(1978).<br />

Sci. USA, 75., 3742<br />

5) J.A.Den Hollander, T.R.Brown,<br />

K.Ugurbil and R.Shulman; Proc. Acad.<br />

Sci. USA, 76, 6096 (1979).<br />

6) E.Cresta, SJez, A.Lepri A.Pisani,<br />

C.Rossi and G.Sabatini; "NMR<br />

investigation of xylose bioconversion<br />

by Klebsiella sp. strain isolated from<br />

soil". In "Biomass for Energy and<br />

Industry" G.Grassi, G.Gosse and G.dos<br />

Santos eds., Elsevier Applied<br />

Sciences, 2, 253 (1990).<br />

7) C.S.Gong, L.F.Chen, M.C.Flikinger<br />

and G.T.Tsao; Adv. Biochem. Eng.,<br />

20,93 (1981).<br />

8) J.Tolan and R.K. Finn; Appl.<br />

Environ. Microbiol., 53_, 2039 (1987).<br />

9) C.Lucas and N.van Uden; Appl.<br />

Microbiol. Botechnol., 2J3_, 491<br />

(1986).<br />

10) F.Alterthun and L.O.Ingram;<br />

Appl. Environ. Microbiol., 55_ 1943<br />

(1989).<br />

11) J.S.Tolan and R.K.Finn; Appl.<br />

Environ. Microbiol., 51 2039 (1987).<br />

201


202 Bulletin of Magnetic Resonance<br />

Interleukin-1 Receptor Antagonist Protein:<br />

Solution Secondary Structure from NOE's and<br />

1H« and 13C« Chemical Shifts<br />

Brian J. Stockman, Terrence A. Scahill, Annica Euvrard, Nancy A. Strakalaitis,<br />

David P. Brunner, Anthony W. Yem, and Martin R. Deibel, Jr.<br />

1 Introduction<br />

The Upjohn Company, 301 Henrietta St., Kalamazoo, MI 49007<br />

Interleukin-la and interleukin-ip are two<br />

polypeptides which share a significant<br />

number of inflammatory, immunological<br />

and pathological properties (for a review<br />

see [1]). Importantly, these dissimilar 17<br />

kDa proteins bind to two classes of interleukin-1<br />

receptors, resulting in the<br />

mediation of several immune and inflammatory<br />

responses and in the induction of a<br />

variety of biological changes in neurologic,<br />

metabolic, hematologic, and endocrinologic<br />

systems [1]. In addition to IL-loc and IL-1(3,<br />

an interleukin-1 receptor antagonist<br />

protein (termed either IRAP or IL-lra) has<br />

been isolated, characterized, cloned and<br />

expressed in E. coli [2-4]. This newer<br />

member of the IL-1 gene family is a<br />

naturally occurring inhibitor of the<br />

interleukin-1 receptor [2,4], and represents<br />

the first described naturally occurring<br />

cytokine that functions entirely as a<br />

specific receptor antagonist.<br />

Site-directed mutagenesis [5-7] and<br />

protein modification [6] studies have<br />

identified three regions of IL-1 that are<br />

involved in either receptor binding or<br />

transmission of the biological response<br />

upon binding. For IRAP, it can be hypothesized<br />

that the regions of structure important<br />

for receptor binding are maintained, but<br />

that the region or regions responsible for<br />

eliciting the response are somehow<br />

different. To this end, we have begun an<br />

intensive program to determine the solution<br />

structure of IRAP using NMR spectroscopy.<br />

Since the solution [8-12] and crystalline<br />

[13,14] structures of IL-lp have been<br />

determined, direct comparisons can be made<br />

between IRAP and IL-lp. This may lead to a<br />

correlation between structural and biological<br />

differences.<br />

2 Methods<br />

Expression of IRAP was carried out using E.<br />

coli K-12 strain DU379. Fermentation media<br />

were supplemented with (15NH4)2SO4, [ 13 Ci]and/or<br />

[i5N]-L-methionine, 1 5 NH4C1, and<br />

[ 13 C]-D-glucose (stable isotopes were<br />

obtained from Cambridge Isotope Laboratories,<br />

Isotec, and/or MSD Isotopes) as<br />

required to produce either 15 N- or doubly<br />

13 C/ 15 N-enriched IRAP. Analysis of<br />

resolved *H resonances indicated that both<br />

13 C and 15 N were incorporated at an<br />

enrichment level greater than 95%.<br />

Samples for NMR spectroscopy contained 2<br />

mM IRAP, 50 mM 2H4-ethanolamine and 300<br />

mM NaCl at pH 6.4. Trace amounts of PMSF<br />

and NaN3 were added to prevent any<br />

protease digestion or bacterial growth in<br />

the sample.<br />

All NMR spectra were recorded at<br />

27 °C on a Bruker AMX-600 spectrometer<br />

equipped with a triple-resonance probe and<br />

a multi-channel interface. Threedimensional<br />

1H-15N NOESY-HMQC and TOCSY-<br />

HMQC experiments were recorded with<br />

slight modification of the methods of<br />

Zuiderweg and Fesik [15] and Marion et al.<br />

[16]. Three-dimensional 1H-15N-13C HNCA<br />

and HN(CO)CA triple resonance experiments<br />

were recorded with constant-time 15 N<br />

evolution as described by Grzesiek and Bax<br />

[17]. Detailed acquisition parameters have<br />

been described elsewhere [18]. Three-


Vol. 14, No. 1-4 203<br />

6.0<br />

Figure 1. Region of the slice corresponding the 1H-13H-13C-1H TOCSY spectrum of IRAP.<br />

to 13 C frequencies of 18.7 and 51.8 ppm in Several assigned correlations are labeled.<br />

dimensional 1H-13C-13C-1H COSY [19], m<br />

13C-1H TOCSY [20], and 1H-13C NOESY-HMQC<br />

[21] experiments were also recorded.<br />

3 Results<br />

Assignment of the majority of the backbone<br />

!H, 13 c, and 15 N resonances of IRAP was<br />

accomplished by analysis of four threedimensional<br />

data sets. First, 1H-15N NOESY-<br />

HMQC and TOCSY-HMQC experiments were<br />

recorded on uniformly 15 N-enriched IRAP.<br />

Then, two *H- 15 N- 1 3 C triple resonance<br />

experiments were recorded, the so-called<br />

HNCA and HN(CO)CA experiments [22,23].<br />

Redundant sequential connectivities<br />

obtained from the heteronuclear data sets<br />

simplified and increased the reliability of<br />

the assignments. During the assignment<br />

process, NOE's indicative of secondary<br />

structure were identified.<br />

For many residues, magnetization<br />

transfer in the 1H-15N TOCSY-HMQC spectrum<br />

extended resonance assignments to at<br />

least one X HP resonance and sometimes even<br />

further down the side chain. In cases of<br />

favorable resolution, such as for high-field<br />

shifted resonances, two-dimensional DQF-<br />

COSY and TOCSY spectra confirmed and/or<br />

extended these side-chain assignments.<br />

Extensive side-chain assignments, however,<br />

will require ^C-directed strategies [19,20]<br />

and are currently in progress. A represen-<br />

tative slice from the 1H-13C-«C-»H TOCSY<br />

spectrum of IRAP is shown in Figure 1.<br />

Once the majority of correlations<br />

were assigned in the !H- 15 N HSQC spectrum<br />

recorded in J H2O, an identical spectrum was<br />

recorded after exchanging the protein into<br />

2 H2O. Only 50 1H- 15 N correlations remained<br />

after six hours in 2 H2O solvent. As discussed<br />

below, each of these residues was found to<br />

participate in the p-sheet framework of<br />

IRAP.<br />

4 Discussion<br />

During analysis of the 1H-15N NOESY-HMQC<br />

data set, NOE's indicative of the solution<br />

secondary structure [24] of IRAP were<br />

identified. The majority of these- were<br />

classified as cross-strand NOE's between<br />

residues involved in ($ -sheet structure. They<br />

are manifested in the NOESY-HMQC spectrum<br />

as a third 1 H a NOE to an amide proton (the<br />

others being the interresidue and intraresidue<br />

1 H« *s), or as weak, non-sequential !H N -<br />

!H N NOE's. Stretches of residues giving rise<br />

to these types of NOE's also had other<br />

characteristics associated with p-sheet<br />

residues: low field 1 H a and !H N chemical<br />

shifts, strong ^N- 1 !!" coupling (manifested<br />

by intense DQF-COSY and TOCSY correlations),<br />

and reduced } H N exchange rates. In<br />

addition, 22 intense iHa-iH« NOE's, characteristic<br />

of antiparallel p-sheet [24], were


204 Bulletin of Magnetic Resonance<br />

-iY^k XII 146-152<br />

119-122<br />

X 120-125<br />

III30_34<br />

II23-27<br />

VI.5-73<br />

VH76.84<br />

Figure 2. Schematic diagram of the topology strand NOE's. Dashed lines indicate interof<br />

the p-sheet framework of IRAP. Double- strand hydrogen bonds inferred from<br />

arrowhead lines identify assigned inter- analysis of *H N exchange rates.<br />

11-19


Vol. 14, No. 1-4 205<br />

1.20<br />

1.00<br />

0.80<br />

0.60<br />

0.40<br />

A5 0.20<br />

0.00<br />

-0.20<br />

-0.40<br />

-0.60<br />

-0.80 -L i i....?...J I !.........! n i [in! i.J<br />

-0.50 --<br />

-1.00 -•;•<br />

-1.50 --<br />

-2.00 --<br />

10 20 30<br />

Alpha proton<br />

iv I | v | [ vi i | vii<br />

40 50 60 70 80 90<br />

Residue<br />

Alpha carbon<br />

: IXj<br />

100 110 120 130 140 150<br />

-2.50 --<br />

rrni_n mm —yyj |_yj L^L<br />

-3. 1<br />

DQ DL J US — IEJ CO OD<br />

10 20 30 40 50 60 70 80 90 140 150<br />

Residue<br />

Figure 3. Comparison of 'H" (top) and 13 C a shown at the bottom of each plot, p-strands<br />

(bottom) A5 chemical shifts for IRAP and IL- are boxed, while helical regions are denoted<br />

ip. Locations of NOE-defined secondary by lines,<br />

structure elements in both proteins are


206 Bulletin of Magnetic Resonance<br />

identified in the two-dimensional NOESY<br />

spectrum recorded in 2 H2O and confirmed in<br />

the 1H-13C NOESY-HMQC spectrum. Analysis<br />

of this pattern of NOE's results in alignment<br />

of the 12 p-sheet strands as shown in Figure<br />

2. Arrows indicate observed cross-strand<br />

NOE's, while dashed lines indicate hydrogen<br />

bonds inferred from !H N exchange rates.<br />

The p-sheet strands have been<br />

presented in Figure 2 in a manner that<br />

allows easy comparison to the P-sheet<br />

framework elucidated for IL-lp in solution<br />

by Driscoll et al. ([9], Figure 5). Comparison<br />

of the two figures illustrates how the<br />

overall topology of the two proteins is<br />

identical, but in several regions is composed<br />

of different stretches of the primary<br />

sequence. Strands II and III, which are<br />

adjacent strands connected by a fiveresidue<br />

turn in IL-lp, are shifted by six<br />

residues in the primary sequence and are<br />

connected by a four-residue turn in IRAP.<br />

Similarly, strands I and IV in IRAP are<br />

shifted by six arid five residues, respectively.<br />

The consequence of shifting the<br />

residues that comprise these portions of the<br />

P-sheet is that the N-terminal six residues of<br />

IRAP have no structural counterpart in the<br />

IL-ip structure. Structurally significant<br />

shifts of one residue are seen for strands VI,<br />

VII, and XII.<br />

As expected for predominantly Psheet<br />

proteins, large positive (downfield)<br />

deviations from random coil chemical shifts<br />

are observed for the l H a resonances [25],<br />

and large negative (upfield) deviations<br />

from random coil chemical shifts are<br />

observed for the !3C a resonances [25,26].<br />

Comparison of the secondary 'H a and 13 C a<br />

chemical shifts of IRAP and IL-ip also<br />

illustrates the differences and similarities<br />

in location of secondary structure elements,<br />

as shown in Figure 3. Note the excellent<br />

agreement of the out-of-phase appearance<br />

of the plots over the first 50 residues with<br />

the five- or six-residue offset in location of<br />

the P-strands. Also note how the in-phase<br />

sections of the plots correspond to P-strands<br />

at identical positions in both proteins.<br />

While the solution secondary<br />

structure of IRAP is dominated by antiparallel<br />

p-sheet, short stretches of strong !H N -<br />

!H N NOE's between adjacent residues,<br />

indicative of a helical conformation, were<br />

also observed. As shown in Figure 2, these<br />

regions involve residues 40-45 and 86-89.<br />

In addition, NOE's from 1HN of L89 to 1H« of<br />

186 and from IH* of E44 to 1H« of V41, both<br />

medium-range NOE's characteristic of a<br />

helical conformation [24], were observed.<br />

These were the only NOE's of this type<br />

unambiguously assigned. In addition, the<br />

i 3 C a chemical shifts in these two stretches<br />

are shifted downfield slightly compared to<br />

their random coil values, as would be<br />

expected for a helical conformation [26].<br />

These residues correspond to residues 35-40<br />

and 87-90 in IL-ip, the former of which is a<br />

3 io helix in solution [9]. Isolated strong !H N -<br />

!H N NOE's between two or three sequential<br />

residues were also observed, and locate turn<br />

conformations at positions: 18-22, 27-30, 62-<br />

65, and 116-118.<br />

5 Acknowledgements<br />

We thank Dr. Paul E. Fagerness, Dr. Eldon L.<br />

Ulrich, Dr. Melinda Roy, Kathleen A. Farley,<br />

Ron L. VanZanten, and Cindy A. Granatir for<br />

their continuing contributions. We thank<br />

Dr. Ad Bax (NIH) for advice and for supplying<br />

preprints of the constant-time HNCA<br />

and HN(CO)CA pulse sequences.<br />

6 References<br />

1. Dinarello, C. A. (1989) Adv. Immunology<br />

44, 153-205.<br />

2. Hannum, C. H., et al. (1990) Nature 343,<br />

336-340.<br />

3. Eisenberg, S. P., et al. (1990) Nature 343,<br />

341-346.<br />

4. Carter, D. B., et al. (1990) Nature 344, 633-<br />

638.<br />

5. MacDonald, H. R., et al. (1986) FEBS Lett.<br />

209, 295-298.<br />

6. Wingfield, P., et al. (1989) Eur. J. Biochem.<br />

179, 565-571.<br />

7. Gehrke, L., et al. (1990) J. Biol. Chem. 265,<br />

5922-5925.<br />

8. Driscoll, P.C., et al. (1990a) Biochemistry<br />

29, 3542-3556.<br />

9. Driscoll, P. C, et al. (1990b) Biochemistry<br />

29, 4668-4682.<br />

10. Clore, G. M., Wingfield, P. T., &<br />

Gronenborn, A. M. (1991) Biochemistry 30,


Vol. 14, No. 1-4 207<br />

2315-2323.<br />

11 Clore, G. M, &. Gronenborn, A. M. (1991)<br />

/. Mol. Biol. 221, 47-53.<br />

12. Tate, S., et al. (1992) Biochemistry 31,<br />

2435-2442.<br />

13. Finzel, B. C, et al. (1989) J. Mol. Biol. 209,<br />

779-791.<br />

14. Priestle, J. P., Schar, H. P., & Griitter, M.<br />

G. (1989) Proc. Natl. Acad. Sci. U.S.A. 86, 9667-<br />

9671.<br />

15. Zuiderweg, E. R. P., & Fesik, S. W. (1989)<br />

Biochemistry 28, 2387-2391.<br />

16. Marion, D., et al. (1989) Biochemistry 28,<br />

6150-6156.<br />

17. Grzesiek, S., & Bax, A. (1992)7. Magn.<br />

Reson. 96, 432-440.<br />

18. Stockman, B. J., et al. (1992)<br />

Biochemistry 31, 5237-5245.<br />

19. Ikura, M., Kay, L. E., & Bax, A. (1991) J.<br />

Biomol. NMR 1, 299-304.<br />

20. Clore, G. M, et al. (1990) Biochemistry<br />

29, 8172-8184.<br />

21. Ikura, M., et al. (1990) J. Magn. Reson.<br />

86, 204-209.<br />

22. Ikura, M., Kay, L. E., & Bax, A. (1990)<br />

Biochemistry 29, 4659-4667.<br />

23. Bax, A., & Ikura, M. (1991) J. Biomol.<br />

NMR 1, 99-104.<br />

24. Wiithrich, K. (1986) NMR of Proteins and<br />

Nucleic Acids, Wiley, New York.<br />

25. Wishart, D. S., Sykes, B. D., & Richards, F.<br />

M. (1991) /. Mol. Biol. 222, 311-333.<br />

26. Spera, S., & Bax, A. (1991) J. Am. Chem.<br />

Soc. 113, 5490-5492.


208<br />

1 Introduction<br />

Green's Function Calculation of<br />

Effective Nuclear Relaxation Times in<br />

Metals and Disordered Metals<br />

M. Martin-Landrovc and J A. Moreno<br />

Departamento de Fisica and Centro de Resonancia Magnetica,<br />

Facultad de Cicncias, Universidad Central de Venezuela,<br />

Apartado 47586, Caracas 1041-A, Venezuela.<br />

The theoretical derivation of expressions for die<br />

magnetic nuclear relaxation times at very low<br />

temperatures and the description of the behaviour<br />

with temperature for such relaxation times,<br />

has been of major interest among the researchers<br />

in the field, specially because of the recent experimental<br />

possibility to obtain measurements of<br />

nuclear magnetic properties at such low temperatures.<br />

There has been a considerable amount of<br />

work in the area of nuclear magnetism [1], but a<br />

comprehensive theoretical interpretation of<br />

NMR relaxation times, at arbitrary temperatures,<br />

is still lacking. Recently, Sbibata et al. has published<br />

a series of papers [2], [3] concerning the<br />

theoretical determination of the nuclear spin lattice<br />

relaxation time for a system of nuclear spins<br />

interacting with conduction electrons ina a metal.<br />

Using a theory of nonlinear spin relaxation [4], [5]<br />

they predicted a multicxponential spin-lattice relaxation<br />

behaviour.<br />

In the case of disordered metals and high temperatures,<br />

where a Korringa law is aplicable, Warren<br />

[6] predicted an enhancement of the relaxation<br />

rate, which in some cases [7] could be as large<br />

as 6,500. More recently [8], Gdtze and Ketterle<br />

derived expressions for die Warren enhancement<br />

factor by means of normalized Kubo response<br />

functions [9].<br />

ED the present work, we make use of the two-times<br />

Green's function formalism in the regime of<br />

the Linear Response Theory to derive the temperature<br />

behaviour of nuclear relaxation times [10]<br />

for nuclei in metals and disordered metals. The<br />

Bulletin of Magnetic Resonance<br />

results obtained are in complete agreement with<br />

those derived by Sbibata in the assumption of an<br />

effective unique relaxation time [2], [3] and with<br />

experimental evidence [11], [12]. Also, there is<br />

agreement between our results and those derived<br />

by Gdtze and Ketterle [8] in the high temperature<br />

regime, where Korringa law is valid but additionally<br />

we obtained expressions for the enhanced<br />

relaxation rate which are valid in the whole temperature<br />

range. The organization of this paper is<br />

as follows, in section 2, the general formalism is<br />

derived, in section 3, we work out the Hamiltonian<br />

of die system from which the equation for the<br />

Green's function < < I°/I°> > to, which contains<br />

all the information relevant to the spin-lattice<br />

relaxation, is derived. This equation is then<br />

solved including terms up to second order in the<br />

electron nucleus interaction and the disorder parameters.<br />

Finally in section 4 we discuss the relaxation<br />

times formulas.<br />

2 Green's Functions and the Relaxation<br />

Rate<br />

We will consider a system which can be modelled<br />

by the total Hamiltonian:<br />

H = Hs + HSL + HL (1)<br />

where Hs represents the nuclear spin Hamiltonian,<br />

HL is the Hamiltonian for the heat bath and<br />

HSL represents the couplig between both<br />

systems, usually under the condition<br />

HL > Hs > HSL which occurs commonly in NMR<br />

experiments. In order to consider the evolution of


Vol. 14, No. 1-4 209<br />

the system toward thermal equilibrium, it is necessary<br />

to prepare the system in an initial nonequilibrium<br />

state. This can be achieved by the<br />

adiabatic switching on of a pertrubation, which in<br />

our case is a magnetic field along a particular<br />

direction, and suddenly at time t—0, this perturbation<br />

is switched off letting the system to evolve<br />

freely according to die Hamiltonian H. The perturbation<br />

can be written as:<br />

where:<br />

Hi' = - M .<br />

H\* = Q(-t)e Et Hi<br />

(2)<br />

(3)<br />

Within the linear response theory, the magnetization<br />

for t>0 is given by the following expression<br />

[13]:<br />

H\ r dt<br />

which is written in terms of the two times retarded<br />

Green's function. By using the causality property<br />

[13] of the retarded Green's function, it is<br />

posible to write the equation for the displacement<br />

of magnetization from the equilibrium situation<br />

as:<br />

1 [+OO > lT ) H\ _ lm ,<br />

where < 8Af(/> = - o.Itcan<br />

be shown that:<br />

(4)<br />

. H\ (6)<br />

so that taking the Laplace transform of eq. (5), it<br />

can be written as:<br />

/(z) . (7)<br />

(5)<br />

where:<br />

with GM ( ><br />

- iz)<br />

We are interested in the asymptotic behaviour<br />

of < 8 M(f > since in that regime is that the relaxation<br />

rates are experimentally measured. According<br />

to Tauber's theorem [14], [15] an asymptotic<br />

espanskm of f(t) can be written as:<br />

*« 0<br />

where Z\ represents the poles of the function<br />

f(z), n* ( v) the order of the pole and Ck (y) the<br />

coefficient in a Laurent expansion of f(z) around<br />

the pole. In the particular case of f(z) having only<br />

first order poles, eq. (9) can be written:<br />

(8)<br />

(10)<br />

Let us suppose that the Green's function can be<br />

written in the general form:<br />

)<br />

o> - acoo -<br />

(11)<br />

where Qxand Waare complexfunctions. According<br />

to eq. (8) the poles of the function f(z) are<br />

related to the poles of the Oieen's function so that<br />

only those poles that are located at the upper<br />

complex plane must be considered [16]. In general<br />

it is necessary to make a complete analysis of<br />

the pole structure of the Green's function at the<br />

upper complex plane in order to describe the total<br />

aymptotic behaviour of the function f(t). As a simplifying<br />

assumption, we will consider the following<br />

zeroth-order approximation to the pole<br />

structure. We will first assume that the function<br />

Ox(ci>) does not contribute with any pole. This<br />

(V)


210 Bulletin of Magnetic Resonance<br />

analyzed with some care. Secondly that the function<br />

Wo. can be approximated by its value at die<br />

frequency aoo. This second assumption ussually<br />

is a very strong one since the whole pole structure<br />

is sometimes collapsed to a single first order pole,<br />

but nevertheless in many cases, this approximation<br />

gives the right relaxation behaviour correlating<br />

quite well with the experimental results. The<br />

relaxation rates obtained under this assumptions<br />

can be written in terms of the imaginary part of<br />

the function Wa evaluated at the frequency aoao,<br />

that is:<br />

and:<br />

T2<br />

j - ImWo(O) (12)<br />

too) - (- too)<br />

(13)<br />

3 Hamiltonian and Equation for the<br />

Green's Function Go (coX<br />

For tile system considered in this work, the Hamiltonian<br />

can be written as in equation (1) where<br />

the different terms are:<br />

vv'tt<br />

X Vz «*v*+ ~ «v'f-<br />

with:<br />

r<br />

(14)<br />

(15)<br />

The lattice Hamiltonian can be written as the<br />

sum of two terms [8]:<br />

HL = 2- £v (ks) at k s av k s<br />

\ks<br />

where:<br />

p + (q)V(q) (17)<br />

X at- §.*«* + (18)<br />

and represents the electron density fluctuations<br />

for wavevector q. The disorder in the lattice is<br />

represented by the Fourier transform U(q) of the<br />

random potential. This coefficient satisfies the<br />

symmetry property:<br />

U*(q)= V(-q) (19)<br />

The equation for the Green function Go(


Vol. 14, No. 1-4 211<br />

and:<br />

with:<br />

W(to) = § **'<br />

Att - A** A** -<br />

At* -<br />

Gf-+ (0 JU<br />

Ajt* -<br />

oi<br />

(22)<br />

(23)<br />

X {»(* *-)(!- *(*+)) + «(lt+)(i- «(*:*-))}<br />

At* -<br />

2<br />

- OJO)+<br />

at-- at+ ,<br />

i<br />

(24)<br />

o<br />

(25)<br />

(26)<br />

(27)<br />

(28)<br />

21 1 + I<br />

\k-g At •+,.*]<br />

(29)<br />

As it was discussed previously [17], the longitudinal<br />

relaxation rate is proportional to the imaginary<br />

part of the function W ( 0<br />

ImAt*(w)= * q> 0<br />

V(q)\ Z ,<br />

&t',k-<br />

(30)<br />

(31)<br />

(32)<br />

Equation (31) represents the dynamical shift in<br />

electronic energy due to the disorder and equation<br />

(32) corresponds to the electronic relaxation<br />

rate function, which in the limit of a>-»O,gives the<br />

electronic relaxation rate 1/J> . The imaginary<br />

part of W (to ) now become s:<br />

1ml<br />

x r<br />

kk<br />

XI-<br />

(At*-ReAt*) 2 +(/mAt*) 2<br />

(A**- RcAt*) 2 + (<br />

I<br />

(33)<br />

In the case of a perfect metal, that is for U(q) -<br />

0, we get from equation (33), the following expression<br />

for the relaxation rate [17]:


212<br />

k* '<br />

[n(*+)(!-<br />

+ n(*'-)(!-«(*+))]5( At*•) (34)<br />

which exhibits a Korringa behaviour at high<br />

temperatures and attains a maximum for the relaxation<br />

time at very low temperatures, comparable<br />

with nuclear spin energies [17]. in the case of<br />

disorder we get for the relaxation rate:<br />

njk - XI- n(k+<br />

(At*- ReAt*) 2 + (ImA**) 2<br />

I (A*i- ReAi"*) 2 -<br />

in the limit goes to 0.<br />

4 Conclusions<br />

(35)<br />

In die whole temperature range, equation (34)<br />

can be calculated numerically, and the result for<br />

the case of a perfect metal was obtained in reference<br />

17, where it can be appreciated that T i<br />

shows a maximum at a temperature that is approximately<br />

half of the nuclear spin temperature,<br />

which is consistent with the result obtained by<br />

Shibata and co-workers in the supposition that<br />

the relaxation process can be described by a unique<br />

effective relaxation time T l as the observed<br />

experimental behaviour is. Also there is a correspondence<br />

between both results in the limit of low<br />

temperature, leading us to think that our calculation,<br />

even in its simpler approximation, is quantitatively<br />

correct From the experimental point of<br />

view there are not enough data to decide whether<br />

a single or a multiexponential relaxation takes<br />

place, but the general tendency is to believe that<br />

even though the process seems actually to be multiexponential,<br />

it could be described by an effective<br />

relaxation time T 'i, which is die time that characterizes<br />

the evolution of observables, in particu-<br />

Bulletin of Magnetic Resonance<br />

lar, the longitudinal magnetization [11], [12]. The<br />

temperature dependence shown experimentally<br />

by this time T 'i, agrees completely with the behaviour<br />

calculated in reference 17. Also, equation<br />

(35) takes into account the enhancement of the<br />

relaxation rate for the whole temperature range,<br />

showing at high temperatures a departure from<br />

Korringa's law proportional to the electronic relaxation<br />

time as Warren proposed. The result obtained<br />

for the enhanced relaxation rate shows<br />

that this enhancement will be present even at low<br />

temperatures, as equation (35) is valid in the whole<br />

temperature range. The approximation assumed<br />

in this work, besides its simplicity, takes into<br />

account the main features present in die temperature<br />

behaviour of relaxation times, witiiin the<br />

limits of Linear Response Theory, and it can be<br />

extended to consider more realistic models or<br />

systems.<br />

Acknowledgments<br />

This work was partially supported by CONICTT<br />

and die Consejo de Desarrollo Cientffico y Humanistico<br />

of die Universidad Central de Venezuela.<br />

We also dutnk die collaboration of Fundacion<br />

Polar for die necessary funding to present<br />

diis work.<br />

References<br />

[1] Abragam A. and Goldman M., in Nuclear<br />

Magnetism: Order and Disorder (Oxford, Clarendon,<br />

1982).<br />

[2] Shibata, F. and Hamano, Y., Solid St. Cotnmun.,<br />

44,921 (1982).<br />

[3] Shibata, F. and Hamano, Y., J. Phys. Soc.<br />

Japan, 52,1410(1983).<br />

[4] Shibata, F., J. Phys. Soc. Japan, 49,15 (1980).<br />

[5] Shibata, F. and Asou, M., J. Phys. Soc. Japan,<br />

49,1234(1980).<br />

[6] Warren, W.W., Phys. Rev. B, 3,3708 (1971).<br />

[7] Warren, W.W. and Brennerrt, G.F., Amorphous<br />

and Liquid Semiconductors, (London, Taylor<br />

and Francis, 1974), 2, p. 1074.<br />

[8] Gdtze, W. and Ketterle, W., Z. Phys. B, 54,49<br />

(1983).<br />

[9] Kubo, R., J. Phys. Soc. Japan, 12,570 (1957).


Vol. 14, No. 1-4 213<br />

[10] Mardn Landrove, M. and Moreno, J.A., Acta.<br />

Oent Venez., 37,387 (1986).<br />

[11] Bacon, F., Barclay, A., Brewer, W.D., Shirley,<br />

D.A. and Templeton, JJE., Phys. Rev. B, 5,<br />

2397 (1972).<br />

[12] Brewer, WJD., Shirley, D.A. and Tcmplcton,<br />

J.E., Phys. Lett. A, 27,81 (1968).<br />

[13] Zubarev, D.N., Nonequilibrium Statistical<br />

Thermodynamics (Consultants Bureau, New<br />

York, 1974).<br />

[14] Berg, h., Introduction to the Operational<br />

Calculus (John WHey, New York, 1967).<br />

[15] Krasnov, M.L., Kiselev, A.I. and Makarenko,<br />

G.I., Functions of Complex Variable, Operational<br />

Calculus and Stability Theory (Mir Publishers,<br />

Moskow, 1983).<br />

[16] Paley, R.E.A.C. and Wiener, W., Fourier<br />

Transforms in the Complex Domain (A.M.S., New<br />

York, 1934).<br />

[17] Martin Landrove, M. and Moreno, J.A.,<br />

Phil. Mag., 58,103 (1988).


214<br />

1 Introduction<br />

The extraction of molecular<br />

potentials and rates by modeling the<br />

dependence of spectra on thermodynamic<br />

state is one of the major contributions of<br />

magnetic resonance to molecular physics.<br />

We argue here that this entire endeavor is<br />

conceptually suspect due to the implicit<br />

factorization of spin and spatial degrees of<br />

freedom in calculating stochastic averages.<br />

All such averages have heretofore neglected<br />

spin—dependent energies in the potential for<br />

nuclear motion and are therefore not results<br />

of equilibrium statistical mechanics. We<br />

develop an averaging procedure from<br />

equilibrium statistical mechanics and find<br />

that it predicts large, previously<br />

unrecognized contributions to motionally<br />

averaged spin Hamiltonians which are<br />

directly proportional to spatial terms in the<br />

energy. We discuss the present state of the<br />

experimental evidence for the accepted<br />

average, with particular emphasis on<br />

indirect scalar couplings averaged by<br />

conformer equilibria, and find that the<br />

quality of presently available data cannot<br />

resolve the issue conclusively. The weakness<br />

of the theoretical basis of the accepted<br />

theory is also outlined.<br />

2 The Traditional Stochastic Average<br />

Stochastic Averaging Revisited<br />

D.H. Jones, N.D. Kurur, and D.P. Weitekamp<br />

Arthur Amos Noyes Laboratory of Chemical Physics<br />

California Institute of Technology<br />

Pasadena, CA 91125, USA<br />

For over forty years it has been<br />

accepted 1 " 36 that spin states are transported<br />

between spatial states with spinindependent<br />

rates. This unexamined<br />

assumption was clearly stated in the seminal<br />

work of Bloembergen, Purcell and Pound: 1<br />

"The atom or molecule is simply a vehicle<br />

Contribution No. 8724<br />

Bulletin of Magnetic Resonance<br />

by which the nucleus is conveyed from point<br />

to point. We thus neglect the reaction of<br />

magnetic moments of the nuclei upon the<br />

motion." This notion is the" basis for all<br />

existing formalisms 1 " 36 for calculating the<br />

magnetic resonance lineshapes of spin<br />

systems undergoing spatial rate processes —<br />

most importantly, chemical exchange. A<br />

well—known consequence is that the average<br />

value of a spin—Hamiltonian parameter<br />

(chemical shift, scalar coupling, dipolar<br />

coupling, etc.) in the fast—exchange limit is<br />

given by<br />

= SpnX:<br />

n n, (1)<br />

where Xn is the value of this parameter in<br />

the spin Hamiltonian of the nth spatial<br />

manifold and pn is viewed as the probability<br />

of the system being in that manifold,<br />

irrespective of spin state. The sum may be<br />

over molecular eigenstates or over' large<br />

groups of them, as when n indexes molecular<br />

conformers. The traditional prescription,<br />

which neglects spin energies, is to express<br />

the molecular partition function q as a sum<br />

of parts qn associated with each indexed<br />

manifold. The probability for each manifold<br />

is pn = qn/ and the ratio of two such<br />

probabilities is (in the absence of work<br />

terms)<br />

Pn/pn' = exp[-AAnn' /RT] (2a)<br />

= exp[(-AUnn'+TASnn')/RT], (2b)<br />

where AAnn'= -RTln(qn/qn') is the<br />

difference in the molar Helmholtz free<br />

energies of the manifolds. In Eq. 2b, the free<br />

energy differences have been divided into


Vol. 14, No. 1-4 215<br />

differences in energy AUnn' and entropy<br />

ASnn'- The connection to molecular<br />

energies is AUnn' = N^(En-En'), where En<br />

is the common spatial contribution to free<br />

energy of the n th manifold of spin states and<br />

N 4 is Avogadro's number.<br />

3 An Alternative Stochastic Average<br />

The actual molecular energies may be<br />

written as E? = En+E (n), where 7 indexes<br />

a spin eigenstate within the n th manifold.<br />

Since spin Hamiltonians are constructed as<br />

traceless, the spin—dependent contributions<br />

E (n) sum to zero in each manifold.<br />

Nevertheless, the absence of the spin energy<br />

terms in the pn indicates unambiguously<br />

that Eq. 1 is not derivable without<br />

approximation from equilibrium statistical<br />

mechanics.<br />

We address the following questions.<br />

What is the exact equilibrium expression for<br />

the stochastically averaged parameters? For<br />

which systems will it differ measurably from<br />

Eq. 1? Do existing experiments decide the<br />

issue conclusively?<br />

To proceed, we specialize to the case<br />

that the spin Hamiltonians in all<br />

significantly occupied spatial manifolds are<br />

mutually commuting. Then the spin<br />

eigenbasis {I7)} is independent of spatial<br />

state and is the basis needed to describe the<br />

stochastically averaged spectrum. The<br />

spatially-averaged energy of a particular<br />

such spin eigenstate 17) is<br />

E =SE2exp(-A2/RT)/Sexp(-A2/RT) (3a)<br />

I n n<br />

= SpM (3b)<br />

n<br />

Note that Eq. 3 uses the complete<br />

manifold free energy, A 2 = NAEZ-TSn,<br />

including spin terms, according to the<br />

prescription of equilibrium statistical<br />

mechanics. Each such summation is<br />

restricted to a particular spin state and the<br />

distribution among spatial states for each<br />

spin state is assumed to be the equilibrium<br />

distribution at the lattice temperature.<br />

Thus p2 is the conditional probability of<br />

being in the n th spatial manifold, given that<br />

the spin state is [7). As in the traditional<br />

formulation, no specification of the<br />

distribution of population among spin states<br />

is needed. Our hypothesis is that the<br />

spectral line positions for sufficiently fast<br />

exchange between spatial manifolds are the<br />

Bohr frequencies, ^^(E^-E^/h<br />

corresponding to differences between the<br />

average energies of Eq. 3. The<br />

corresponding motionally averaged<br />

spin—Hamiltonian parameters are those<br />

which generate this spectrum.<br />

As a simple illustration, consider a<br />

spin Hamiltonian with only two distinct<br />

eigenvalues, ±Xn/2 (in Hz), for each n.<br />

Then the proposed alternative to Eq. (1) is<br />

(with 7 = ±)<br />

(X) =<br />

(<br />

= S [(p£ -<br />

n<br />

(4a)<br />

(4b)<br />

(4c)<br />

For the usual high—temperature case where<br />

hXn


i "i<br />

t! I<br />

m<br />

216<br />

depends on spin, spatial energy differences<br />

between manifolds contribute to the<br />

statistically averaged frequencies for spin<br />

transitions. The product of a small<br />

spin—dependent probability difference<br />

multiplied by a large spatial energy<br />

difference gives a readily measured<br />

contribution to the magnetic resonance line<br />

position.<br />

4 Test Systems<br />

In both formulations it is possible<br />

ideally to predict the fast exchange<br />

observations without adjustable parameters.<br />

This would constitute a purely experimental<br />

test, the quality of which depends only on<br />

the experimental uncertainties. In practice<br />

there seems to be no case where NMR<br />

spectra of individual molecular eigenstates<br />

have been obtained separately and also as a<br />

thermal average.<br />

Thus, it seems necessary to look to<br />

those situations where n indexes conformers.<br />

The experimental concept is simple and<br />

well-known. At low temperature, the spin<br />

Hamiltonian for each conformer can be<br />

determined, since, in the limit of negligible<br />

chemical exchange between them, separate<br />

spectra are seen for each. The relative areas<br />

of these spectra at each slow—exchange<br />

temperature provide the relative populations<br />

and thus the free—energy difference between<br />

conformers. A linear fit to the temperature<br />

dependence of this free—energy difference<br />

allows it to be decomposed into two terms,<br />

which can be viewed as an energy difference<br />

and an entropy difference if one additionally<br />

assumes that the temperature dependence of<br />

these is negligible over the experimental<br />

range.<br />

Since a conformer is a set of<br />

molecular eigenstates, additional dependence<br />

on thermodynamic state (eg. temperature<br />

dependence) is possible due to averaging<br />

within this set. Theoretically, one has<br />

precisely the same problem in deciding how<br />

to take this average as for the averaging over<br />

conformers. However, if one can measure any<br />

such temperature dependence within the<br />

slow—exchange regime and extrapolate to<br />

fast—exchange, then the stochastic theory<br />

need not enter at this level.<br />

Thus, we arrive at a set of criteria<br />

which need to be met for a compelling, fully<br />

experimental test of any theory relating<br />

Bulletin of Magnetic Resonance<br />

slow-exchange and fast—exchange spectra:<br />

i) The system must have<br />

state—dependent rates such that<br />

measurements in both the slow— and<br />

fast-exchange regimes are possible. For<br />

fluids this typically requires barriers between<br />

conformers on the order of 10 kcal/mole.<br />

ii) The state dependence of the<br />

conformer spin Hamiltonians and the<br />

free-energy differences must be measured in<br />

the slow—exchange region to allow<br />

extrapolation through the fast—exchange<br />

region. This is often the major source of<br />

uncertainty because of the small<br />

temperature range corresponding to<br />

slow-exchange. The difference in<br />

thermodynamic state between these regimes<br />

ideally is small or even zero, so as to<br />

minimize the propagation of errors due to<br />

phenomenological extrapolation. Using<br />

different NMR transitions or field conditions<br />

to measure the same spin Hamiltonian<br />

parameter can help in this regard; since the<br />

criterion for motional collapse varies with<br />

the transition observed, there is no minimum<br />

8.<br />

8 1 4.<br />

o<br />

•a o<br />

GO<br />

6. -<br />

1 1<br />

- 'V^<br />

1 __l<br />

—<br />

0.<br />

f1 1 1 1<br />

0. 200. 400. 600. 800. 1000.<br />

Temperature (K)<br />

Figure 1. Traditional (dashed) and<br />

alternative (solid) stochastic averages for a<br />

simple two—conformer, two—spin system.<br />

Simulations have the general features of<br />

some substituted ethanes: a gauche coupling<br />

of 2.0 Hz, a trans coupling of 20.0 Hz, the<br />

trans conformer with a free energy 200<br />

cal/mol greater than the doubly—degenerate<br />

gauche conformer.


Vol. 14, No. 1-4 217<br />

difference in thermodynamic state between<br />

fast and slow exchange.<br />

iii) Some or all of the fast-exchange<br />

data should fall outside the error bars on the<br />

predictions of one of the theories, thereby<br />

disproving it. The accepted and alternative<br />

theories embodied in Eqs. 1 and 3,<br />

respectively, have identical predictions for<br />

mutual exchange and whenever the occupied<br />

conformers are degenerate in spatial energy<br />

or in spin Hamiltonians. For two—site<br />

problems, the theories will typically differ<br />

measurably when the conformer free energies<br />

differ by > 102 cal/mole. When this<br />

difference exceeds 10 3 cal/mole, sensitivity<br />

will usually preclude observing the<br />

slow-exchange spectrum of the minor<br />

conformer. Figure 1 is a numerical<br />

comparison of the two theories for the simple<br />

case of two conformers and a two—spin<br />

system. It demonstrates that the predictions<br />

of the theories are different by a magnitude<br />

that should be measurable. There is also a<br />

qualitative difference: the sign of the<br />

temperature dependence of the<br />

fast—exchange spin Hamiltonian parameter<br />

is opposite to that of the accepted theory<br />

over part of the temperature range. This<br />

behavior is inconsistent with Eq. 1 regardless<br />

of how the pn are calculated.<br />

The above conditions are not<br />

extremely restrictive; a substantial fraction<br />

of the molecules whose conformer equilibria<br />

have been studied by solution-state NMR<br />

fall into this range of free—energy differences.<br />

Since the accepted theory has been in<br />

increasing use for four decades, it might be<br />

expected that it would have substantial and<br />

diverse experimental support. While it is<br />

difficult to have confidence in the<br />

completeness of a search through such a<br />

large literature, we are as yet unaware of<br />

any data set that meets the criteria above.<br />

Thus, no theory has presently been<br />

evaluated by this seemingly reasonable<br />

standard.<br />

The only theories ever considered<br />

previously are of the form of Eq. 1.<br />

Numerous authors have noted failures in its<br />

application, 19 " 21 but these have usually been<br />

plausibly attributed to inadequacies in the<br />

data, most commonly uncertainties of<br />

conformer assignment or unmeasured<br />

temperature dependence of a conformer spin<br />

Hamiltonian.<br />

5 Substituted Ethanes<br />

Substituted ethanes in solution 14 " 24<br />

are the most studied systems and include<br />

cases which nearly meet the criteria of the<br />

previous section. Figure 2 illustrates, with<br />

the example of 1-fluoro-l,1,2,2—<br />

tetrachloroethane, the trans and gauche<br />

rotational isomers which interconvert at<br />

convenient rates. Since the two gauche<br />

conformers are mirror—images, there are<br />

only two magnetically distinct conformers,<br />

one with the H and F atoms trans to one<br />

another and the other a degenerate pair of<br />

gauche rotamers at the other two staggered<br />

positions of the dihedral angle of rotation<br />

about the carbon—carbon single bond.<br />

Increasing temperature carries the system<br />

from the slow—exchange limit to the<br />

fast—exchange limit without a change in<br />

composition or phase. The fast—exchange,<br />

three—bond vicinal coupling (Jrrp) is known<br />

to be temperature dependent and this has<br />

been attributed to the averaging, according<br />

to Eqs. 1 and 2, between distinct values Jt<br />

and Jg in the trans and gauche conformers,<br />

respectively. 15 " 16 Unlike chemical shifts,<br />

scalar couplings do not require a nominally<br />

temperature—independent reference<br />

resonance to compensate for the usual<br />

uncontrolled shifts of internal field with<br />

temperature. Also, scalar couplings are<br />

generally believed to be less sensitive to<br />

intermolecular interactions which could<br />

provide a confounding mechanism of<br />

temperature dependence.<br />

For the particular case of<br />

1—fluoro—1,1,2,2—tetrachloroethane, a test as<br />

E(


218 Bulletin of Magnetic Resonance<br />

described of the stochastic averaging theories<br />

is prevented by the failure to resolve the<br />

coupling Jg. 37 This quantity has an<br />

experimental upper bound of 2 Hz. If it is<br />

assumed that Jg is independent of<br />

temperature, then the data are consistent<br />

with the accepted theory. 16 However a<br />

temperature dependence of 0.005-O.01 Hz/K<br />

between 150 K, where slow-exchange<br />

observations have been made on the 19 F<br />

resonances, and 300 K, where the *H<br />

spectrum is motionally averaged, would<br />

allow our alternative stochastic average in<br />

the form<br />

(5)<br />

to fit the data as well or better.<br />

Reason to suspect such a temperature<br />

dependence can be found from a close<br />

reading of the literature on the related<br />

system 1—fluoro—1,1,2,2—tetrabromo—<br />

ethane. 22 A value of Jg = 2.4 ± 0.3 Hz in<br />

dimethylether at 188 K can be measured<br />

from published data, 39 but has been reported<br />

as 1.7 Hz at 180 K in the same solvent. 23 In<br />

CFCI3 this coupling has been tabulated as<br />

1.15 Hz from 171 to 178 K, 22 but our recent<br />

fits of the spectrum (at 171 K only) in<br />

reference 23 indicate Jg = 1.5 ± 0.3 Hz.<br />

Thus the reported absence of temperature<br />

dependence to three significant figures is<br />

dubious and further experimental work is<br />

needed.<br />

6 Discussion<br />

The present conformer model is by no<br />

means the most refined version of Eq. 3<br />

possible, but has the advantage of employing<br />

only quantities that are experimentally<br />

measured on the same sample. Continuous<br />

classical—mechanical forms of the present<br />

theory are readily written down and the<br />

differences from the accepted theory persist.<br />

Alternatively, the quantum partition<br />

functions could be modeled. Such<br />

modifications would introduce unmeasured<br />

parameters. Such extensions might be<br />

warranted after improvements in the<br />

experimental data base.<br />

It is of interest to note that the same<br />

theoretical prediction for (Jjrp) obtained by<br />

use of Eqs. 3 and 5 also results if the same<br />

conformer model is evaluated using Eq. 4<br />

with Xn = Jn and + and — indicating,<br />

respectively, triplet and singlet zero-field<br />

eigenstates. Thus, no measurable field<br />

dependence of (Jjjp) is predicted by the<br />

alternative theory, which is not obvious<br />

because of the entanglement of different spin<br />

Hamiltonian parameters in the spin energies.<br />

The theoretical justification for Eq. 1<br />

and related propositions is also weaker than<br />

has been appreciated. Such an average<br />

follows from dynamic models 1 " 36 based on<br />

the assumption that spin states in<br />

superposition are transported between<br />

different spatial states in perfect concert.<br />

Any such model exists in a truncated<br />

Liouville space that excludes superpositions<br />

of states that differ in both their spin and<br />

spatial factors. Whether such a truncated<br />

space suffices to describe magnetic resonance<br />

lineshapes is an open question. What is<br />

clear is that such a space cannot describe the<br />

approach to equilibrium of the total system,<br />

since this requires spin—dependent rates<br />

between spatial manifolds. Thus a full<br />

dynamic solution is needed in this complete<br />

Liouville space. One result of such a full<br />

solution will be the equilibrium average spin<br />

energies of Eq. 3. Less clear is under what<br />

dynamic assumptions either these energies or<br />

those that follow from Eq. 1 will describe the<br />

fast—exchange spectrum. In any case the<br />

problem is richer than has been appreciated.<br />

The accepted idea of how to calculate<br />

a stochastically averaged spectrum is<br />

universal in the literature of magnetic<br />

resonance, underlying the interpretation of<br />

average chemical shifts, dipolar couplings,<br />

tunnel splittings, quadrupole couplings, and<br />

hyper-fine interactions. In most situations<br />

the number of unknowns is such that it is<br />

not possible to verify the form of the<br />

stochastic average, but valuable information<br />

could be obtained if the correct form were<br />

known. If the traditional ideas are generally<br />

incorrect, many thousands of experiments<br />

need to be reinterpreted with forms of the<br />

present theory to in fact obtain the<br />

quantitative information on molecular<br />

structure that they were designed to yield.<br />

Ultimately, the choice of theory will be<br />

decided by a preponderance of data. The<br />

present work indicates clearly that the issue<br />

must be reopened and is a first step in the<br />

reexamination of the experimental basis of<br />

Eq. 1. Accurate experimental measurements


Vol. 14, No. 1-4<br />

of systems such as those discussed here will<br />

be needed in order to discriminate between<br />

the stochastic theories, as opposed to fitting<br />

free parameters according to one or the other<br />

theory. Precisely the same issue of how to<br />

calculate a stochastic average also arises in<br />

the (ab initio) theoretical calculation of a<br />

measurable spin Hamiltonian from<br />

expectation values of the underlying<br />

molecular eigenstates, which are almost<br />

never sufficiently long—lived to measure<br />

individually by magnetic resonance.<br />

Application of the correct statistical<br />

prescription will often be needed to in fact<br />

test experimentally whether the quantummechanical<br />

part of the calculation is<br />

adequate.<br />

This work was supported by the<br />

National Science Foundation<br />

(CHE-9005964). DHJ holds a Department of<br />

Education Graduate Fellowship. DPW is a<br />

Camille and Henry Dreyfus<br />

Teacher—Scholar.<br />

*N. Bloembergen, E.M. Purcell, and R.V.<br />

Pound, Phys. Rev. 73, 679 (1948).<br />

2<br />

E.R. Andrew and R. Bersohn, J. Chem.<br />

Phys. 18, 159 (1950).<br />

3<br />

E.L. Hahn and D.E. Maxwell, Phys. Rev.<br />

88, 1070 (1952).<br />


220<br />

1 Introduction<br />

Magnetic Resonance of Trapped Ions<br />

by Spin-Dependent<br />

Cyclotron Acceleration<br />

Pedro J. Pizarro and Daniel P. Weitekamp<br />

Arthur Amos Noyes Laboratory of Chemical Physics<br />

California Institute of Technology, 127-72<br />

Pasadena, CA 91125 USA<br />

The frequencies of motion of ions trapped by static<br />

magnetic and electric fields may be detected through the<br />

charge induced on the trap plates. In a homogeneous<br />

magnetic field, these frequencies have long been used for<br />

high resolution measurements of the charge-to-mass ratio.<br />

The most important chemical application of this is mass<br />

spectroscopy via Fourier transform ion cyclotron resonance<br />

(ICR). 1 ' 2 ' 3 Single-ion sensitivity in the detection of the<br />

axial trapping frequency has also been demonstrated<br />

recently for masses of chemical interest by D. Pritchard. 4 - 5<br />

We have recently proposed methods for transferring this<br />

exquisite sensitivity to the detection of the internal<br />

spectroscopy of ions, in particular the magnetic resonance<br />

spectra. 6 ' 7 This presentation focuses on design issues for<br />

the most promising such method, in which spin-dependent<br />

cyclotron acceleration is imposed and the resulting change<br />

in ion orbit is detected as a change in the axial trapping<br />

frequency.<br />

As in the electron g-factor measurement of H.<br />

Dehmelt, 8 the shift in axial trapping frequency is<br />

proportional to the strength of a static magnetic bottle field<br />

gradient. However, that direct effect of a spin flip is<br />

impractically small for nuclear spin flips of ions. In the<br />

present case, the transverse magnetic moment is coupled to<br />

a radiofrequency gradient to provide an accelerating force.<br />

A precedent is M. Bloom's deflection of neutral molecular<br />

beams by radiofrequency field gradients (the "transverse<br />

Stern-Gerlach effect"). 9 ' 10<br />

We have derived both semiclassically and quantummechanically<br />

the conditions under which a magnetic field<br />

gradient modulated at both the Larmor and cyclotron<br />

Contribution No. 8714<br />

Bulletin of Magnetic Resonance<br />

frequencies will lead to cyclotron acceleration proportional<br />

to the transverse magnetic moment of a coherent state of<br />

the particle and radiation field. In the presence of a<br />

magnetic bottle, the corresponding shift in the axial<br />

trapping frequency due to this spin-dependent work can be<br />

made much larger than the shift due directly to a spin flip.<br />

This effect has been incorporated into a proposed<br />

experimental procedure in which the spin-flip probability,<br />

resulting from a period of high-resolution magnetic<br />

resonance, controls the presence or absence of a net axial<br />

frequency shift between two detection periods. A data<br />

reduction algorithm based on the fast Fourier transform<br />

allows rapid conversion of the "before" and "after" signals<br />

from one or many trapped ions into a point of the magnetic<br />

resonance spectrum or interferogram. Simulated signals,<br />

including the anticipated noise from both the detection<br />

circuit and intrinsic quantum fluctuations in the number of<br />

spins flipped, indicate the method is practical.<br />

2 Ion Confinement in a Penning Trap and a<br />

Magnetic Bottle<br />

An ion in a homogeneous static magnetic field Bo,<br />

whose direction defines the z-axis, will circle in the<br />

transverse (x-y) plane at frequency ©c=qB0/m. Axial<br />

(z-axis) trapping can be obtained by adding a static electric<br />

field E = (vo/d 2 )(xx/2+yy/2-zz). Here Vo is a DC<br />

trapping potential and d is a characteristic linear<br />

dimension dependent on the details of the electrode<br />

geometry (e.g., hyperbolic, cubic, cylindrical). An ion so<br />

trapped undergoes three different types of harmonic<br />

translational motion: 11 axial oscillation at frequency


Vol. 14, No. 1-4 221<br />

a, = ^qV0/md 2 , rapid cyclotron motion in the<br />

transverse plane at frequency ©+, and slower magnetron<br />

motion of frequency co_, with ©± = y{ ©c + J© 2 ~2© 2 I.<br />

In ICR, one monitors the radiofrequency voltage at © +<br />

induced on a capacitor by the cyclotron orbit of a group of<br />

ions coherently excited by an oscillating electric field<br />

resonant with the cyclotron motion. Single ion sensitivity<br />

has been achieved for detection of both ©z 4 and, indirectly,<br />

©+, 5 using axial detection. The charge-to-mass ratio is the<br />

only structural quantity measured on trapped ions by ICR<br />

to date.<br />

In order to describe how other quantities, in particular<br />

the NMR spectrum, may be encoded into trapped ion<br />

signals, it is necessary to analyze the case of ion motion in<br />

the presence of a magnetic field gradient. In particular we<br />

consider the introduction of a magnetic bottle field of the<br />

form (in cylindrical coordinates, f = zz+pp+)<br />

AB = B2f(z 2 -p 2 /2)z-zpp]. n > 12 The presence of AB<br />

couples the axial, cyclotron and magnetron motions to the<br />

spin. A classical analysis of the ion motion in a known<br />

spin state is adequate though the principal results can be<br />

confirmed with the trapping frequencies from first-order<br />

quantum perturbation theory. 11 To a very good<br />

approximation the axial amplitude is a simple onedimensional<br />

sinusoid even in the presence of the magnetic<br />

bottle. The potential energy for the axial motion is<br />

The spin magnetic moment operator has been replaced by<br />

its two high-field eigenvalues (upper and lower signs)<br />

assuming a single spin 1/2 nucleus with gyromagnetic ratio<br />

y. The only coupling to the transverse motion is through<br />

the mechanical magnetic moment \xm = (q/2)[x(dy/dt)y(dx/dt)].<br />

This quantity is however a constant of the<br />

motion, 13 so that the axial motion is separable with a<br />

parametric dependence on the ion's transverse orbit. The<br />

ellipsis denotes terms quadratic or higher in B2/Bo. We<br />

have performed exact three-dimensional trajectory<br />

calculations to confirm that Eq. 1 suffices for the times and<br />

orbits of the numerical examples discussed later. The axial<br />

frequency for each ion, including corrections to ©z due to<br />

the magnetic bottle, can be written from Eq. 1 as<br />

mco,<br />

A key illustration of this coupling was the use of the spindependent<br />

shift of the axial frequency to measure the g-<br />

(i)<br />

factor of the electron 8 ' 11 cooled at 4 K to the ground state<br />

of its cyclotron motion. The straightforward generalization<br />

of this to ions is not practical, since the shift is inversely<br />

proportional to mass at fixed observation frequency (oz.<br />

Extending this shift to a 100 amu ion with a proton<br />

magnetic moment under conditions similar to those used in<br />

the single electron experiments yields an unpractically<br />

small 4 nHz shift. A more difficult problem is that the<br />

three trapping frequencies and the Larmor frequency<br />

become inhomogeneously broadened due to the wide range<br />

of nm values present in a thermal ensemble. Minimizing<br />

this by ion cooling methods would be time-consuming and<br />

reducing this range to be less than the spin magnetic<br />

moment requires lower temperatures as mass increases,<br />

since this quantum decreases inversely with mass.<br />

3 Spin-Dependent Cyclotron Acceleration:<br />

IRICE<br />

Rather than attempt to measure the small spindependent<br />

term in the axial motion, we derive a form of<br />

spin-dependent cyclotron acceleration analogous to<br />

cyclotron excitation via ICR: this is mternally resonant /'on<br />

cyclotron excitation (IRICE). The resonant electric field of<br />

ICR is replaced by an oscillating magnetic gradient with<br />

components at the cyclotron and Larmor (©0) frequencies.<br />

This field is constructed by arranging two orthogonal<br />

quadrupole coils: one is parallel to the x-axis with current<br />

proportional to cos(©0t+n/2)cos(©+t), and the other is<br />

directed along the y-axis with current proportional to<br />

cos(©0t-Hi)cos(©+t-Hc/2). 14 With gradient field strength G<br />

for each, the total magnetic field is<br />

t = {B0 +Gzsin[(©0+© +<br />

-Gcos(oooO sm ( to +<br />

A quantum-mechanical description of a spin-1/2 magnetic<br />

moment in this field shows that the eigenstates of spin lie<br />

in the transverse plane, aligned such that the spin-<br />

dependent force Fs=(fi»v)B resonant with the ion<br />

cyclotron motion is<br />

Like the force in ICR excitation, this is resonant with the<br />

cyclotron motion, but with explicit spin dependence due to<br />

a gradient dipole force. 9 ' 10 * 15<br />

Neglecting the magnetron mode, the transverse ion<br />

motion is described as a cyclotron oscillation of radius p+<br />

and phase + (the sense of rotation used here is appropriate<br />

for a positively charged ion):<br />

(3)<br />

(4)<br />

(5)


222<br />

—- = -ra+p+fsin((a+t+


pa*:'<br />

Vol. 14, No. 1-4 223<br />

' 4 R.M. Weisskoff, G.P. Lafyatis, K.R. Boyce, E.A. Cornell,<br />

R.W. Flanagan, Jr., and D.E. Pritchard, J. Appl. Phys. 63,<br />

4599 (1988).<br />

5 E.A. Cornell, R.M. Weisskoff, K.R. Boyce, R.W.<br />

Flanagan, Jr., G.P. Lafyatis, and D.E. Pritchard, Phys. Rev.<br />

Lett. 63, 1674 (1989).<br />

6 D.P. Weitekamp and P.J. Pizarro, U.S. Patent No.<br />

4,982,088 (1991).<br />

7 C.R. Bowers, S.K. Buratto, P.J. Carson, H.M. Cho, J.Y.<br />

Hwang, L. J. Mueller, P.J. Pizarro, D.N. Shykind, and D.P.<br />

Weitekamp, SPIE Proc. 1435,36 (1991).<br />

8 R.S. Van Dyck, Jr., P.B. Schwinberg, and H.G. Dehmelt,<br />

in New Frontiers in High Energy Physics, edited by B.<br />

Kursunoglu, A. Perlmutter, and L. Scott (Plenum, New<br />

York, 1978); in Atomic Physics 9, edited by R.S. Van<br />

Dyck, Jr. and E.N. Fortson (World Scientific, Singapore,<br />

1984).<br />

9 M. Bloom and K. Erdman, Can. J. Phys. 40, 179 (1962).<br />

10 M. Bloom, E. Enga, and H. Lew, Can. J. Phys. 45, 1481<br />

(1967).<br />

n L.S. Brown and G. Gabrielse, Rev. Mod. Phys. 58, 233<br />

(1986).<br />

12 R.S. Van Dyck, Jr., F.L. Moore, D.L. Farnham, and P.B.<br />

Schwinberg, Rev. Sci. Instrum. 57, 593 (1986).<br />

13 G. Schmidt, Physics of High Temperature Plasmas. 2nd.<br />

ed. (Academic Press, New York, 1979).<br />

14 Note that this field configuration has different symmetry<br />

from that suggested in Ref. 9; the "Transverse Stern-<br />

Gerlach" experiment on ions suggested there uses a field<br />

whose symmetry does not in fact lead to a linear change in<br />

the cyclotron radius over an extended period of time.<br />

15 J.P. Gordon and A. Ashkin, Phys. Rev. A 21, 1606<br />

(1980).


224<br />

Bulletin of Magnetic Resonance<br />

Coordination Modes of Histidine Moiety in Copper (II) Dipeptide<br />

Complexes Detected by Multifrequency ESR<br />

1 Introduction<br />

R.Basosi, R.Pogni, G.Delia Lunga<br />

Department of Chemistry - University of Siena<br />

Pian dei Mantellini 44,53100 Siena, Italy<br />

The copper complexes of histidinecontaining<br />

peptides naturally occurring in<br />

blood plasma [1,2] can be used as a model<br />

for metal/protein interactions.In fact a low<br />

molecular weight peptide containing<br />

histidine or even histidine itself may<br />

compete with ceruloplasmin for copper in<br />

the blood, significantly altering free ligand<br />

levels. For this reason, the coordination<br />

behaviour of the histidine moiety in copper<br />

(II) dipeptides has been the subject of<br />

much investigation [3-7]. From this point<br />

of view the study under physiological<br />

conditions, namely room temperature and<br />

neutral pH, is crucial for maintaining the<br />

biological significance of the model.<br />

Unfortunately a definitive speciation has so<br />

far been hindered by the strong<br />

dependence on the ligand-to-metal ion ratio<br />

and the multiplicity of possible equilibria,<br />

function of pH, in solution.<br />

ESR may be used to study<br />

paramagnetic metal complexes, and<br />

multifrequency ESR in combination with<br />

computer simulation is a decisive tool for<br />

determining the chemical structure of<br />

copper biosystems under physiological<br />

conditions. For ESR spectra the<br />

requirement of a good fit at different<br />

frequencies put constraints on the precision<br />

of magnetic parameters and allows<br />

definitive assignments, despite the lack of<br />

resolution typical of room temperature<br />

spectra. In this paper the above method<br />

was used in the study of complexes of<br />

copper(II) and Glycylhistidine (GlyHis),<br />

Histidylglycine (HisGly) and p-Alanyl-L-<br />

Histidine or Carnosine (Cam) formed in<br />

solution in a large range of pH,<br />

concentration and metal/ligand ratios. The<br />

aim was to show how a computer-aided<br />

multifrequency ESR approach can provide<br />

valuable information on the systems under<br />

investigation, avoiding undesirable effects<br />

due to changes in the physical state of<br />

samples.<br />

2 Material and methods<br />

Glycyl-L-Histidine, L-Histidylglycine and<br />

p-Alanyl-Histidine (Carnosine) from<br />

Sigma Chemical Co. were used without<br />

further purification.<br />

Solutions were made with distilled<br />

water and the uncorrected pH was adjusted<br />

with HC1 or NaOH and determined with<br />

an "LCD" Model pH meter. Isotopically<br />

pure 63 CuO (from Oak Ridge National<br />

Laboratory, Oak Ridge , TN) was used for<br />

the EPR experiments. Stock solutions<br />

with 1:2 molar ratio were prepared, [Cu]<br />

= 10- 2 M, [peptide] = 2 xlO" 2 M.<br />

X-band EPR spectra were obtained<br />

with a Bruker 200D SRC X-band<br />

spectrometer, and S-band spectra were<br />

obtained with a microwave bridge by<br />

Medical Advances Inc., Milwaukee, USA.<br />

All the bridges were equipped with loopgap<br />

resonators (Jagmar, Krakow, Poland)<br />

operating at v = 9.5 GHz for the X-band<br />

bridge and v = 4.04 GHz for the S-band<br />

bridge. Microwave frequencies were<br />

measured with an XL Microwave Model<br />

3120 counter. The spectrometer was<br />

interfaced with a Compaq Deskpro<br />

486/50L computer with an 8-megabyte<br />

memory and a 50 MHz clock. The data<br />

were acquired using the EPR data system<br />

CS-EPR produced by Stelar Inc., Mede,<br />

Italy.<br />

The CUKOS program for the<br />

simulation of EPR spectra of fast tumbling<br />

copper complexes in an isotropic<br />

environment, written in QuickBasic, is<br />

based on Kivelson's theory of linewidth<br />

[8] and includes the second order shift<br />

equation of Bruno et al. [9] and the further<br />

assumption of Lorentzian lineshapes. A<br />

Monte Carlo calculation method was added


Vol. 14, No. 1-4 225<br />

to the program. The Monte Carlo method<br />

randomly varied selected spectral<br />

parameters within defined limits in order to<br />

fit experimental data.<br />

3 Results and Discussion<br />

In recent years, data for copper(II)<br />

complexes of histidine and histidine<br />

containing peptides has been generated by<br />

a variety of techniques such as infraredspectroscopy<br />

[10], x-ray diffraction [11],<br />

nuclear magnetic resonance [12], circular<br />

dichroism [13,14], thermodynamic [3] and<br />

potentiometric studies [15,16] and ESR at<br />

liquid nitrogen temperature [17,18].<br />

Table I<br />

between two species. Aggregation can<br />

occur [21] and often tedious empirical<br />

approaches for the formulation of good<br />

glasses must be employed. In water,<br />

freezing can change the local pH [20,22],<br />

and Wilson and Kivelson [8] found that<br />

the isotropic g- and A-values of copper<br />

acetylacetonate are actually temperature<br />

dependent. All of these complications can<br />

be avoided by the use of multifrequency<br />

ESR of copper complexes above 0 °C.<br />

ESR spectra obtained for Cu-Histidine<br />

complexes with a 1:1,1:2 and 1:100 molar<br />

ratios, in the pH range 5-9 are reported in<br />

Refs. [17,18]. At pH= 7.4 and high molar<br />

ratios the ESR pattern shows a<br />

Parameters Used to Simulate ESR Spectra for Cupric Ion complex with Dipeptides a<br />

system giso Ag AisO( 63 Cu) AA A N isO xc(ps)b Ref.<br />

63 Cu-His<br />

3N<br />

4N<br />

63 Cu-GlyHis<br />

63 Cu-(GlyHis)2<br />

63 Cu-(HisGly)2<br />

63 Cu-B-Ala-His<br />

2.098<br />

2.104<br />

2.1134<br />

2.107<br />

2.114<br />

2.1468<br />

0.162<br />

0.159<br />

0.171<br />

0.166<br />

0.235<br />

0.221<br />

61.50<br />

61.60<br />

69.32<br />

80.36<br />

54.4<br />

62.91<br />

137.4<br />

143.0<br />

174.95<br />

195.87<br />

193.79<br />

190.00<br />

a Values for the hyperfine and superhyperfine splittings arc given in G.<br />

11.1<br />

11.0<br />

12.4<br />

13.5<br />

11.03<br />

12.5<br />

b Correlation times in ps (10~ 12 sec.). 63 Cu magnetogyric ratio 0.70904 s^<br />

A conventional ESR spectroscopic<br />

approach is to freeze the sample and extract<br />

the magnetic parameters by analysis of the<br />

frozen solution powder pattern. There are<br />

many reasons why this is undesirable.<br />

Vanng&rd [19] called attention to changes<br />

in coordination. Falk et al. [20] observed<br />

the temperature dependence of equilibrium<br />

85.00<br />

85.00<br />

50.64<br />

88.23<br />

75.00<br />

45.00<br />

18<br />

18<br />

this work<br />

this work<br />

this work<br />

this work<br />

septet 1:3:6:7:6:3:1 consistent with mixed<br />

glycine-like and histamine-like<br />

coordination. At higher pH, this species is<br />

in equilibrium with a complex in which<br />

both the histidines are coordinated in the<br />

histamine way (4N). The magnetic<br />

parameters for the two species are<br />

reported in Table 1.


226 Bulletin of Magnetic Resonance<br />

Fig. 1 shows ESR spectra at X-band for<br />

the complex Cu-GlyHis in a fast tumbling<br />

regime. All the experimental spectra are<br />

paired with the simulations for different<br />

species.<br />

100 G<br />

Fig. 1. Exp. ( ) and simulated (- • - ) ESR<br />

spectra for Cu-GlyHis at pH = 7.3. a) monomeric<br />

species, b) 62 % monomeric species + 38 % biscomplex,<br />

c) bis-complex.<br />

Fig. la represents the monomeric species<br />

obtained for a 1:1 molar ratio and pH=7.3.<br />

This species is readly obtained for a molar<br />

ratio of 1:2 at lower pH. Fig. lc shows the<br />

ESR spectrum of the bis-complex. This<br />

species is obtained when there is a large<br />

excess of the ligand with respect to the<br />

metal. Magnetic parameters for the above<br />

two species are reported in Table 1. With a<br />

molar ratio of 1:2 and physiological<br />

conditions, the two species occur<br />

simultaneously present in solution and the<br />

overall ESR spectrum is reported in Fig.<br />

lb paired with the simulated spectra. The<br />

best fit in Fig. lb is obtained for a pattern<br />

in which 62% of the monomeric species is<br />

considered together with 38% of the biscomplex.<br />

As the fit is reasonable, the<br />

concomitant presence of the two species<br />

can be supposed. A similar best fit<br />

procedure was performed for S-band. Fig.<br />

2 shows the expansion of the second<br />

derivative of the mi = +3/2 component for<br />

the monomeric species of Cu-GlyHis in<br />

the X and S-bands.<br />

A<br />

Fig. 2. Expanded 2nd derivative experimental mj =<br />

+3/2 component ( ~) paired with<br />

simulation (-••-) for the monomeric species Cu-<br />

GlyHis at: a) X-band, b) S-band.<br />

The procedure proposed [23] is based on<br />

the relative intensities of the patterns of the<br />

three center lines in order to discriminate<br />

between three and four nitrogen<br />

coordination. The simulations are adjusted<br />

until the best fit is obtained. In this case a


Vol. 14, No. 1-4 227<br />

three nitrogens coordination (amino,<br />

peptide and imidazole) from the ligand<br />

molecule accounts for a very stable<br />

monomeric species. The process of<br />

simulation was initiated by reading the<br />

approximate values of all parameters from<br />

the spectra. The g-tensor values were<br />

adjusted by fitting the spectra at X-band<br />

with the highest sensitivity to these<br />

changes. The simulations for liquid phase<br />

EPR spectra started with a Monte Carlo<br />

calculation method. This method randomly<br />

varied selected spectral parameters within<br />

defined limits in order to fit experimental<br />

data. When a good fit was obtained at one<br />

frequency the magnetic parameters were<br />

tested at the other frequency and the<br />

parameters were varied by an iterative<br />

process until the best fit was obtained in<br />

the X- and S-bands for the two physical<br />

states of the samples. The use of second<br />

derivative displays was a crucial step in<br />

obtaining a good set of parameters. In<br />

Fig. 2, second derivative, (or second<br />

harmonic), display emphasizes sharp<br />

features and discriminates against broad<br />

features. This display is particularly useful<br />

for analyzing superhyperfine patterns.<br />

Shoulders in the spectrum become peaks<br />

with precisely defined turning points that<br />

are useful for accurate measurements of<br />

coupling constants. The correlation times<br />

(xc) reported in Table 1 are consistent with<br />

the proposed speciation at pH=7.3.<br />

Despite the fact that HisGly is chemically<br />

similar to GlyHis, ESR spectra obtained<br />

under similar conditions are strikingly<br />

different. The relative parameters are<br />

reported in Table 1. A big difference in<br />

Aiso is evident: 54.4 G for Cu-(HisGly)2<br />

and 80.4 G for Cu-(GlyHis)2. Data<br />

obtained at low temperature confirmed this<br />

assignment. In the case of the Cu-HisGly<br />

complex, if excess HisGly is present, there<br />

may be bis-complex formation in which<br />

deprotonation of the peptide-NH linkage<br />

is suppressed [24]. Voelter et al.<br />

confirmed this hypothesis with 13 C NMR<br />

experiments at pH 7, at which the glycine<br />

moiety is not involved in copper ligation<br />

[25]. The ligation is pure histamine-like<br />

coordination with two imidazole nitrogens<br />

and two amine nitrogens in the first<br />

coordination sphere of copper. This<br />

arrangement is in agreement with the EPR<br />

results previously obtained for histidine<br />

[17,18]. In fact in the case of Cu-HisGly,<br />

the carboxyl group is blocked by peptide<br />

bonding in favour of a total histamine-like<br />

coordination whereas in the histidine<br />

complex a mixture of histamine-like<br />

glycine-like structures has been proposed<br />

on the basis of experimental findings<br />

[17,18,25].<br />

Fig. 3 shows the ESR spectra obtained for<br />

Cu-Carn with a 1:2 molar ratio at pH=5.6<br />

paired with its simulation.<br />

Fig. 3. Experimental ESR spectrum for Cu-Carn at pH<br />

= 5.6 (——) paired with simulation (-•-).<br />

Low pH was explored because at pH=7 a<br />

very distorted ESR pattern, previously<br />

attributed to a predominant dimeric<br />

species, is obtained. If we increase the<br />

ligand concentration up to 1:100 molar<br />

ratio, a very different spectrum arises<br />

which is shown in Fig. 4 with its<br />

simulation.<br />

Fig. 4. Experimental Cu-Carn ESR spectrum ( )<br />

for 1:100 molar ratio paired with<br />

simulation (-•-).<br />

This spectrum is attributed to the Cu-<br />

(Carn)4 complex and has very different


228 Bulletin of Magnetic Resonance<br />

magnetic parameters. In this case, our data<br />

confirms results reported in the literature<br />

[26] for experiments at low temperature.<br />

4 Conclusions<br />

The ESR features of the above copperdipeptide<br />

complexes show very different<br />

coordination behaviour for homologue<br />

Species under similar solution conditions.<br />

At physiological pH, the predominant<br />

histidine species present in solution are a<br />

histamine-like and a mixed histamine-like<br />

glycine-like complex. Under the same<br />

conditions HisGly only shows the<br />

histamine-like coordination whereas for<br />

GtyHis, two species (monomeric and the<br />

bis-complex) characterized by different<br />

magnetic parameters occur simultaneously.<br />

For Carnosine, a dimeric species is<br />

dominant and is possibly in equilibrium<br />

with a monomeric form.<br />

A good characterization of the real species<br />

present in solution can be achieved only by<br />

exploiting the sensitivity of a computer<br />

aided multifrequency ESR approach. A<br />

crucial role is played by a procedure based<br />

on the high selectivity of the second<br />

derivative display.<br />

5 References<br />

1) CJ. Gubler, M.E. Lahey, G.E.<br />

Cartwright, M.M. Wintrobe,<br />

J.ClinJnvest., 32, 405, 1987<br />

2) D.R. Williams, C. Furnival, P.M. May<br />

in "Inflammatory Diseases and Copper",<br />

J.RJ. Sorenson Ed., Humana Press,<br />

Clifton, New Jersey, 45,1982<br />

3) G. Brookes, L.D.Pettit, J.C.S.Dalton<br />

1975,2112<br />

4) R.P.Agarwal, D.D.Perrin, J.C.S.<br />

Dalton 1975,268<br />

5) RJ. Sundberg, R.B. Martin,<br />

ChemRev. 1974, 74(4), 472<br />

6) H. Sigel, "Metal ions in biological<br />

systems", 1973, 63<br />

7) D.B. McPhail, B.A. Goodman,<br />

J.Chem.Soc.Faraday Trans. 1, 1987,<br />

83(12), 3683<br />

8) R. Wilson, D. Kivelson, /. Chem.<br />

Phys., 1966, 44, 4445; 1966, 44, 154<br />

9) G.V. Bruno, J.K. Harrington, M.P.<br />

Eastman, J.Phys.Chem., 1977, 81, 1111<br />

10) R.H.Carlson, T.L.Brown, Inorg.<br />

Chem., 1966, 5, 268<br />

11) N.Camerman, J.K.Fawcett, T.P.A.<br />

Kruck, et al., J.A.C.S., 1978, 100:9,<br />

2690<br />

12) H.Sigel, B.McCormick, J.A.C.S.,<br />

1971,93:8,2041<br />

13) L.Casella, M.Gullotti, J.Inorg.<br />

Biochem., 1983, 18,19<br />

14) E.W.Wilson, M.H.Kasperian,<br />

R.B.Martin, J.A.C.S., 1970, 92:18, 5365<br />

15) D.D.Perrin, V.S.Sharma, J.Chem.<br />

Soc. (A), 1967, 724<br />

16) T.P.A. Kruck, B.Sarkar, Can.J.<br />

Chem., 1973, 51, 3563<br />

17) R.Basosi, G.Valensin, E.Gaggelli et<br />

al., Inorg.Chem., 1986, 25, 3006<br />

18) M.Pasenkiewicz-Gierula, R.Basosi et<br />

al., Inorg. Chem., 1987, 26(6), 801<br />

19) T.Vanngard, in: Biological<br />

Applications of Electron Spin Resonance<br />

(H.M. Swartz, J.R.Bolton and D.C.Borg,<br />

eds) p. 411, Wiley-Interscience, New<br />

York (1972<br />

20) K.E.Falk, E.Ivanova, B.Roos,<br />

T.Vanngard, Inorg.Chem., 1970, 9, 556<br />

21) A.Saryan, K.Mailer, C.Krishnaruti, et<br />

al., Biochem. Pharmacol., 1981, 30, 1595<br />

22) Y.Orii, M.Morita, /. Biochem.<br />

(Tokyo), 1977, 81, 163<br />

23) J.S. Hyde, W.E. Antholine, W.<br />

Froncisz, R. Basosi, In Advanced<br />

Magnetic Resonance Techniques in<br />

Systems of High Molecular Complexity;<br />

N. Niccolai, G. Valensin, Eds.;<br />

Birkhausen: Boston, 1986; p 363<br />

24) I. Sdvagd, E. Farkas, A. Gergely<br />

J.Chem.Soc.Dalton Trans., 1982, 2159<br />

25) W. Voelter, G. Sokolowski, U.<br />

Weber, U. Weser, Eur.J.Biochem.,<br />

1975, 58, 159<br />

26) C.E. Brown, W.E. Antholine, W.<br />

Froncisz, J.C.S.Dalton, 1980, 590


Vol. 14, No. 1-4<br />

An EPR and ab initio Study of a Phosphaalkene Radical Anion, and Comparison<br />

with other Phosphorus-Containing Radical Ions<br />

M. Geoffroy*, G. Terron, A. Jouaiti<br />

Departement de Chimie Physique, Universite de Geneve, 30 Quai E. Ansermet, 1211, Geneve<br />

(Switzerland)<br />

P. Tordo<br />

Universite de Provence, CNRS-URA 1412,13397 Marseille Cedex 13 (France)<br />

and<br />

Y. Ellinger<br />

Ecole Normale Superieure et Observatoire de Paris, 24 rue Lhomond, 75005 (France).<br />

1 Introduction<br />

Electron Paramagnetic Resonance (EPR) is an<br />

efficient method to obtain information about the<br />

spin densities in organic radicals containing a<br />

heteroatom and, when used in conjonction with<br />

ab initio calculations, this spectroscopy can<br />

yield a precise description of the structure of<br />

these species. In the present study, our purpose<br />

is to compare the spin delocalization on the fol-<br />

lowing three paramagnetic moieties (-P= 13 C


230<br />

a P=P bond and to record the EPR spectrum of<br />

the corresponding frozen solution. Diphosphines<br />

are well known stable molecules and we have<br />

measured the 31 P anisotropic hyperfine<br />

interaction after freezing a solution of<br />

tetra,2,4,6,trimethylphenyldiphosphine [3] HI<br />

previously oxidized in an electrolytic cell.<br />

2 Experimental<br />

The various compounds have been obtained by<br />

adapting already published syntheses:<br />

ArP=C(H)Ar' [1], ArP=PAr [2] and Ar"2P-PAr"2<br />

[3] (where Ar: trit-t-butyl phenyl., An phenyl,<br />

Ar": 2,4,6, trimethylphenyl ). The electrolyses<br />

were performed in the EPR cavity of an X-band<br />

Bruker spectrometer by using solutions of the<br />

organophosphorus compound in presence of<br />

tetra-n-butyl amonium hexafluor phosphate. The<br />

31 P, 13 C and *H isotropic and anisotropic<br />

coupling constants were obtained after<br />

simulation of the experimental spectrum with a<br />

program using second order perturbation.<br />

The ab initio calculations were carried out on<br />

a Silicon Graphics (Iris 4D) and a Vax 6700<br />

Bulletin of Magnetic Resonance<br />

computer by using G-82 and G-90 Gaussian<br />

programs [4]. The 6-31G* basis set was<br />

generally used and annihilation of the spin<br />

contamination was performed for the UHF<br />

calculations. The coupling constants were<br />

obtained by calculating the expectation values<br />

of the Fermi contact interaction and of the<br />

hyperfine dipolar interaction. ROHF calculations<br />

were also carried out, in particular when the<br />

UHF method led to final values which<br />

were not equal to 0.75.<br />

3 Results and Discussion<br />

EPR spectra..<br />

(ArP=C(H)Ar')~. The liquid solution spectrum<br />

is characterized by a large splitting of 152 MHz<br />

and a rather broad linewidth exhibiting a poorly<br />

resolved additional structure [5]. Full deuteration<br />

of the C(H)C6H5 fragment only affects the<br />

linewidth which is then equal to 8 MHz . The<br />

spectrum obtained with ArP=C(H)C6D5 clearly<br />

shows a coupling of 11 MHz with the ethylenic<br />

proton while the spectrum obtained with<br />

ArP= 13 C(H)C6H5 indicates a 13 C coupling equal<br />

to 16 MHz. From these couplings and from the<br />

observed linewidth one can conclude that an<br />

appreciable spin delocalization occurs onto the<br />

phenyl ring and the simulated spectra are<br />

consistent with the following three additional<br />

proton couplings: 11MHz, 7MHz and 7MHz.<br />

The frozen solution spectrum obtained with<br />

(ArP=C(H)Ar')" is characterized by an axial<br />

hyperfine coupling with 31 P: T//=455MHz and<br />

Tx =1 MHz; the full deuteration of the<br />

fragment only decreases the


Vol. 14, No. 1-4 231<br />

linewidth whereas the use of the 13 C(H)C6H5<br />

fragment leads to an additional hyperfine<br />

interaction: 13 C-T// =47 MHz, 13 C-T±=1 MHz.<br />

(ArP=PAr)~ : As already mentioned [6,7,8], the<br />

liquid solution spectrum exhibits a 31 P isotropic<br />

coupling constant equal to 158 MHz. The<br />

frozen solution spectrum has still not been<br />

reported; it is characterized (Fig. 1) by an axial<br />

coupling tensor : 31 P: T//=458 MHz and T± =10<br />

MHz. No additional J H coupling is observed<br />

with this compound.<br />

100G<br />

Fig 1. EPR spectrum obtained with a frozen solution of<br />

(ArP=PAr)"<br />

(Ar"2P-PAr"2) + .Only the liquid solution spectrum<br />

has been previously reported [9], the<br />

splittings observed with the frozen solution<br />

spectrum (Fig. 2) show that the two phosphorus<br />

nuclei are equivalent and that their hyperfine<br />

tensors exhibit an axial symmetry: T/y= 796<br />

MHz and T± =356 MHz.<br />

100 G<br />

Fig. 2. EPR spectrum obtained with a frozen solution of<br />

(Ar"2P-PAr"2) + .<br />

A'<br />

Assuming a positive sign for all the hyperfine<br />

eigenvalues leads to the isotropic and<br />

anisotropic coupling constants shown in Table 1<br />

together with the spin densities calculated by<br />

using the atomic parameters obtained from [10].<br />

Table 1. Experimental coupling constants (MHz) and<br />

spin densities<br />

radical<br />

T<br />

il 7± Ps V<br />

ArP=CHR- 31p 152 303 -152 0.01 0.41<br />

ArP=PAT<br />

Ar'2PPAr' 2<br />

13 C<br />

31p<br />

31p<br />

31p<br />

31p<br />

16<br />

158<br />

158<br />

502<br />

502<br />

ab initio calculations.<br />

31<br />

300<br />

300<br />

293<br />

293<br />

-15<br />

-148<br />

-148<br />

-146<br />

146<br />

0.00<br />

0.01<br />

0.01<br />

0.04<br />

0.04<br />

0.18<br />

0.41<br />

0.41<br />

0.39<br />

0.39<br />

Phosphaalkene radical onion: In the Cs<br />

symmetry, the UHF optimized structure for<br />

(HP=CH2)- is characterized by HPC=97.5°, P-C<br />

=1.791 A, P-H= 1.428 A; using the ROHF<br />

method does not significantly affect these<br />

parameters. The structure of (HP=C(H)Ar')-<br />

was optimized (ROHF calculations) by fixing<br />

the geometry of the phenyl group and by<br />

assuming a planar structure: HPC=96.6°, H-P=<br />

1.420 A and P-C =1.761 A.<br />

Diphosphene onion and diphosphine cation.<br />

The optimized structures of these two species<br />

are known [11,12,13] and agree with our UHF<br />

results : for (HP=PH)": HPH=95.8°, P-P=<br />

2.133A (Cs symmetry) and for (H2P-PH2)+:<br />

HPH= 102.2°, PPH=104.09° and P-P= 2.164 A


232<br />

=0.750<br />

symmetry). For these two ions<br />

The various hyperfine coupling constants<br />

resulting from UHF calculations are given in<br />

Table 2 . For both the phosphaalkene and the<br />

diphosphene anions, the "parallel" 31 P coupling<br />

eigenvectors are oriented perpendicular to the<br />

molecular plane (x structure), whereas for the<br />

diphosphine cation these eigenvectors make an<br />

angle of 45° with the P-P bond direction in<br />

accordance with the orientation of the n" orbital.<br />

Table 2. Calculated 31 P hyperfine coupling constants<br />

(MHz).<br />

radical<br />

(HP=CH2)-<br />

(HP=PH)-<br />

(H2PPH2)+<br />

^iso<br />

25<br />

70<br />

445<br />

T ll<br />

130<br />

279<br />

328<br />

T il.<br />

-60<br />

-136<br />

-153<br />

T ±2<br />

-70<br />

-142<br />

-174<br />

For the radical ions containing no phenyl ring,<br />

the various spin densities calculated by using<br />

the UHF method are very similar to those<br />

obtained from ROHF calculations. These spin<br />

densities are shown in Table 3.<br />

Table 3. Calculated spin densities.<br />

radical<br />

(HP=CH2)" a<br />

(HP=CHAiO" a<br />

(HP=PH)' a<br />

+ b<br />

(H2P-PH2)<br />

phosphorus<br />

Ps Pp<br />

0.00 0.22<br />

0.00 0.44<br />

0.00 0.49<br />

0.06 0.42<br />

adjacent carbon or<br />

phosphorus<br />

Ps Pp<br />

0.00 0.77<br />

0.00 0.33<br />

0.00 0.49<br />

0.06 0.42<br />

a ) ROHF calculations, b ) UHF calculations.<br />

phenyl<br />

0.22<br />

Bulletin of Magnetic Resonance<br />

These data show that the substitution, in the<br />

phosphaalkene anion, of an ethylenic hydrogen<br />

atom by a phenyl ring drastically decreases the<br />

spin density on the carbon atom and increases<br />

the spin density on the phosphorus atom. The<br />

spin delocalisation onto the phenyl ring is in<br />

good accordance with the variation of the<br />

linewidth observed after deuteration of the<br />

benzene ring.<br />

The calculated hyperfine tensors for (HP=PH)"<br />

and (H2P-PH2) + reasonably agree with the<br />

values measured for (ArP=PAr)~ and<br />

(Ar"2PPAr"2) + . For (HP=CH2 )", the isotropic<br />

and anisotropic coupling constants are quite dif-<br />

ferent from those measured for (ArP=CF£Ar')-,<br />

but, as indicated by the spin densities calculated<br />

for (HP=CHAr')-> this difference is due to the<br />

spin delocalisation induced by the phenyl ring.<br />

0<br />

0 / 0<br />

In summary, the frozen solution EPR spectra<br />

confirm the x* structure for phosphaalkene and<br />

diphosphene radical anions ( negligible contri-<br />

bution of the s orbitals of the phosphorus or<br />

carbon atom to the SOMO) and the non-<br />

bonding character of the SOMO for the<br />

diphosphine cation.


Vol. 14, No. 1-4 233<br />

References<br />

* R. Appel, J. Menzel, F. Knoch and P.Volz, Z. anorg.<br />

allg. chem. 534,100 (1986).<br />

2 M. Yoshifuji, I. Shima, N. Inamoto, K.Hirotsu and T.<br />

Higuchi, J. Am. Chem. Soc. 103, 4587 (1981).<br />

3 B. I. Stepanov, E. N. Karpova and A. I. Bokanov, Zh.<br />

Obshch. Khim., 39, 1544 (1969).<br />

4 F.M.J. Frisch, M. Head-Gordon, G.W. Trucks, J.B.<br />

Foresman, H.B. Schlegel, K. Raghavachari, M. Robb,<br />

J.S. Binkley, C. Gonzalez, D.J. Defrees, D.J. Fox, R.A.<br />

Whiteside, R. Seeger, C.F. Melius, J. Baker, R.L.Martin,<br />

L.R. Kahn,J.J.P. Stewart, S. Topiol and J.A. Pople.<br />

GAUSSIAN 90, Gaussian Inc., Pittsburg, PA (1990).<br />

5 M. Geoffroy, A. Jouaiti, G. Terron, M. Cattani-<br />

Lorente, and Y. Ellinger. J. Phys. Chem. (in press).<br />

6 B. Cetinkaya,A. Hudson, M. F. Lappert and H.<br />

Goldwhite, J. Chem. Soc. Chem. Commun. 609 (1982).<br />

7 A.J. Bard, A.H Cowley, J.E. Kilduff, J. K. Leland,<br />

N.C. Norman and M. Palkulski, J. Chem. Soc.., Dalton<br />

Trans., 249 (1987).<br />

8 M. Calcasi, G. Grouchi, J. Escudie, C. Couret, L. Pujol<br />

and P. Tordo, J. Am. Chem. Soc., 108, 3130 (19886).<br />

9 M. Culcasi, G. Grouchi,and P. Tordo, J. Am. Chem.<br />

Soc.,107, 7191 (1985).<br />

10 J.R. Morton and K.F. Preston. JMagn. Reson., 30,<br />

577 (1978).<br />

11 M.T. Nguyen, J. Chem. Phys., 91, 2679 (1987).<br />

12 T. Clark, J.Am.Chem.Soc, 107,2598,(1985).<br />

13 D. Feller,E.Davidson and W.T.Borden, J. Am.Chem.<br />

Soc, 107, 2596 (1985)


234 Bulletin of Magnetic Resonance<br />

Conformational Substate Distribution in<br />

Myoglobin as studied by EPR Spectroscopy<br />

Anna Rita Bizzarri 1 ^ and Salvatore Cannistraro 1 ' 2 '<br />

' INFM-CNR, Dipartimento di Fisica dell'Universita', Perugia, Italy<br />

2 ) Dipartimento di Scienze Ambientali, Sezione Chimica e Fisica, Universita'<br />

della Tuscia, Viterbo, Italy<br />

Introduction<br />

It has recently been shown that EPR, in<br />

connection with the aid of a computer<br />

simulation approach, can be successfully<br />

applied to investigate the structural<br />

heterogeneity displayed by metallo-proteins [1-<br />

6]. The g-strain effect characterizing the low<br />

temperature EPR spectra of metallo-proteins<br />

can be interpreted by taking into account for the<br />

presence of an ensemble of molecules frozen in<br />

many slightly different structures [6-8].<br />

Different experimental and theoretical<br />

approaches point out that a protein molecule<br />

can assume a very large number of different<br />

substates, called conformational substates (CS)<br />

[9,10] whose sampling is important for the<br />

biological functionality of the protein [11]. At<br />

physiological temperature, proteins fluctuate<br />

among CS; such a behaviour affecting the<br />

kinetic response of the molecules. By<br />

decreasing the temperature, the protein<br />

solution undergoes a glass-like transition and<br />

the fluctuations among CS are progressively<br />

suppressed [12]. Below the glass-temperature,<br />

T „, the molecules are frozen in many different<br />

CS whose distribution may be modulated by<br />

external agents such as pressure, pH, solvent<br />

composition [13-15]. However, the role of the<br />

solvent on the dynamical coupling between the<br />

protein and the CS distribution is still open.<br />

To get further information on this aspect, we<br />

have analyzed the high and low spin ferric<br />

myoglobin (Mb) samples in different<br />

conditions. The EPR spectra of high spin ferric<br />

Mb samples have been interpreted in terms of a<br />

distribution of the crystal field parameters A ,,<br />

A 2 connected with the energy differences of<br />

the low-lying electronic states of the ferric ion;<br />

whereas, the EPR spectra of the low spin Mb<br />

samples have been analyzed in terms of a<br />

distribution of the tetragonal and rhombic<br />

splitting parameters, A and V. An accurate<br />

computer simulation of the spectra has allowed<br />

us to extract the parameters characterizing<br />

these distributions which, in turn, have been put<br />

into a relationship to the CS distribution. The<br />

effects on these distributions as induced by<br />

different solvent compositions and by different<br />

cooling rates are analyzed.<br />

Materials and experimental<br />

methods<br />

Mb EPR samples were prepared by dissolving<br />

commercial (Sigma Chem. Co.) lyophilized<br />

horse skeletal muscle Mb in 0.2 M phosphate<br />

buffer. The highest concentration of Mb in the<br />

solutions was about 5 mM. Final pH for Mb<br />

solutions was about 6.8. Ferricyanide was used<br />

to oxidize the heme iron to the ferric valence<br />

state and solutions were dialysed several times<br />

against buffers to remove the oxidant.<br />

Approximately a twofold molar excess of<br />

sodium azide was used to convert metMb to the<br />

low spin form. Samples in mixed waterglycerol<br />

solvent were prepared by adding<br />

aliquots of glycerol to Mb solutio until the<br />

required concentration was reached. A fast<br />

cooling rate (Fast) has been obtained by<br />

dipping the the samples into liquid nitrogen at<br />

77K; while in the slow cooling rate ( Slow), the<br />

system was cooled with a rate of 0.5 deg/min<br />

from300Kto WOK<br />

All the EPR spectra were recorded at 77 K by an<br />

X-band Varian E109 spectrometer equipped with<br />

a variable temperature control which was also<br />

used to cool the samples in a controlled way. To<br />

calculate the experimental g-values, a<br />

magnetic field calibration was performed with


Vol. 14, No. 1-4<br />

a Magnion Precision . NMR gaussmeter<br />

Mod.G-542; the microwave frequency being<br />

measured with a Marconi 2440 counter.<br />

The acquisition of EPR data was carried out on<br />

a HP 86A personal computer through a home<br />

made interface connected to a IEEE 488 bus [16].<br />

To run both simulations and bestfit programs,<br />

the same microcomputer was switched to an<br />

intelligent terminal of the main frame<br />

computer (VAX 8350), through a serial interface<br />

and an HP terminal emulator.<br />

Analysis of the EPR spectra<br />

It is well-known that the ferric ion, in Mb heme<br />

complexes, is placed in a crystalline electric<br />

field of cubic symmetry in which four ligands<br />

are provided by the four nitrogen atoms of the<br />

porphyrin ring, the fifth ligand is the nitrogen<br />

of the proximal histidine, finally, in the sixth<br />

coordination site different ligands can be<br />

bound. In general, the presence of a weak<br />

ligand causes the ferric ion to assume a high<br />

spin state, S=5/2; while a strong ligand<br />

determines a low spin state, S=l/2. In this paper<br />

we consider metMb in which the weak ligand<br />

HgO + is present, and azide Mb samples with the<br />

strong sixth ligand N% [17].<br />

The EPR spectrum, at 77 K, of metMb is<br />

characterized by two resonances, one at g ~ 6<br />

and a weaker one at g ~2 (spectrum not shown).<br />

This system can be described by the spin<br />

hamiltonian<br />

Hs=ge(l H-S+D[SZ 2 -S (S + 1)/3]<br />

where ge is the value for the free electron; D<br />

and E are the tetragonal and the rhombic zerofield<br />

splittings, respectively. For heme<br />

proteins, the condition of large zero field<br />

splitting is satisfied ( D -10 cm "*) [18] and<br />

only transitions within the lowest Kramers<br />

doublet occur; a fictitious spin S=l/2 can then be<br />

used to fully represent the spin Hamiltonian of<br />

the system, which for axial symmetry (g = g<br />

= gx and gz = g | |) can be expressed by<br />

H S = 9| | P H z SZ+QJLPI H X S x +H y s y) ( 2 )<br />

where g| |~ 2 and g ~ 6 are the g-values<br />

which are observed in the experimental<br />

spectra. Splitting of the in-plane value into two<br />

values, gx and gv, may result in a broadening<br />

(as in our case) or even in a splitting of the g-<br />

235<br />

6 line [19,20]. High order corrections, arising<br />

from spin-orbit mixing of the excited quartet<br />

states into the lowest Kramers doublet lead,<br />

under the assumption of a four-state model<br />

[21,22] (Fig. 1) to the following expression for<br />

gx and gy<br />

gxy=3ge±24E/D-18.7 (E/D) 2 - 12 n 2 (3)<br />

where the tetragonal zero-field splitting D is<br />

given by<br />

D = J L<br />

and the rhombic zero-field splitting E<br />

-2<br />

(4)<br />

2 (5)<br />

the spin-orbit mixing of excited quartet states<br />

into the lowest Kramers doublet is<br />

Jl<br />

~ 2 _ ^ / 1 . 1 ,1\<br />

\ \<br />

(6)<br />

i is the effective spin-orbit coupling constant<br />

(i = 300 cm ** ) which is reduced from the freeion<br />

valued = 420 cm ~* ; A |, A 2 and y are the<br />

energy differences between the low-lying<br />

electronic states of high spin ferric heme (see<br />

Fig. 1). •<br />

V///////////////A<br />

z<br />

V///////777/7//I/<br />

Z '/////// '///////A<br />

777777777777/7////<br />

Figure 1 Energy level diagram of the low-lying electronic<br />

states of high-spin heme. It has been assumed A-p 2000 cm"<br />

1 ,A2 ~ 6000 cm' 1 ' Y -60 cm" 1 . The shaded regions<br />

indicate the variability of the energy levels (not in scale).<br />

The low temperature (77 K) EPR spectra of<br />

azide Mb samples are characterized by three<br />

absorption lines to which three principal<br />

different g- values (about gx = 2.8, gy = 2.2 and


236<br />

gz = 1.7 ) correspond.<br />

Within the Griffith's model [23], the ground<br />

state electronic configuration is a ^T2 state<br />

that can be described by one hole in the shell<br />

made by the iron dxz, dyZ) dxy orbitals.<br />

Owing to the presence of a rhombic distortion,<br />

the orbitals are split into three Kramers<br />

doublets with energies respectively of -V/2,<br />

V/2 and A ( see Fig.2).<br />

7/77/7/7/////////A<br />

7/ 7777777777//////<br />

7/7777/7/777777/<br />

Figure 2 Energy hole levels of the low-lying d-orbitals for<br />

the low spin ferric ion. The shaded regions indicate the<br />

variability of the energy levels (not in scale).<br />

Accordingly, the EPR spectra of azide Mb<br />

samples can be described by the spin<br />

Hamiltonian associated to S=l/2<br />

H S=P ( 9x H x V9y Hy Sy+QzHzSz) ( 7 )<br />

The three principal g-values are given by the<br />

expressions<br />

gx=2[2 AC - B 2 + k B(C- A) (2 1/2 )]<br />

gy= 2[2 AC + B 2 + k B(C+ A) (2 1/2 )]<br />

gz=2[A 2 -B 2 +C 2 +k ( A 2 - C 2 )] (8)<br />

where k is the orbital reduction factor and A,<br />

B and C are the coefficient characterizing the<br />

Kramers doublet of the lowest energy<br />

where 11* >, !, <<br />

wavefunctions within t^ 2 |-1±> are the<br />

T values<br />

for V and A have been assigned, A, B and C<br />

can be calculated by solving for ^/ + and \j/~ ,the<br />

secular equations associated with the matrix<br />

which takes into account for both the spin-orbit<br />

Bulletin of Magnetic Resonance<br />

coupling and the distortion field; then,<br />

assigned a value for k, the g-values can be<br />

determined from eq.(8).<br />

In a general way, once the g values are<br />

known, the EPR spectra can be generated by<br />

computer simulation with the aid of a suitable<br />

model. The derivative field-swept EPR<br />

absorption spectrum, related to randomly<br />

oriented paramagnetic centers with S=l/2, can<br />

be reproduced by the expression<br />

dS(vc,H) vc h p(e, <br />

(10)<br />

where the 1 / g(e.) is the Aasa-Vanngard [24]<br />

correction, C is a constant that encompasses all<br />

the instrumental parameters, P(8, ) is the<br />

orientation dependent transition probability<br />

which, for an S =1/2 system, can be exactly<br />

expressed by [25]<br />

cos 2 + g 4 y sin 2 8 cos 2 + g 4 z cos :<br />

[g 4 x sin 2<br />

(11)<br />

finally f(H ) is the lineshape function<br />

(residual linewidth [26] centered at the<br />

resonance field HQ and with a linewidth<br />

parameter a^ measured in magnetic field<br />

units.<br />

The integration over 8,4> in eq.(10) takes into<br />

account for the random orientation of the<br />

molecular axes with respect to the magnetic<br />

field.<br />

Use of eq.(10) is not, however, sufficient to<br />

reliably reproduce the EPR spectra of metalloproteins.<br />

It is known in fact that EPR spectra<br />

of metallo-proteins are characterized by a<br />

large inhomogeneous broadening resulting<br />

into a spread of the g-tensor values (g-strain)<br />

[1,4,6]. Such an effect can be interpreted by<br />

taking into account the presence of the CS<br />

distribution [8]; the heterogeneity<br />

corresponding to the presence of a frozen<br />

ensemble of molecules in different CS could<br />

entail a spread of the low-lying electronic state<br />

energies of the metal ion and, in turn, a<br />

modulation of the g-values [7,22]. On such a<br />

ground, and accordingly to previous works<br />

[7,8], it has been assumed that the low-lying<br />

electronic state energies of the ferric iron are<br />

distributed.<br />

In definitive, our spectra of metMb samples<br />

have been simulated by introducing two<br />

independent gaussian distributions for the


Vol. 14, No. 1-4<br />

crystal field parameters A j, A2; on the other<br />

hand, the azide Mb samples have been<br />

simulated by considering two independent<br />

gaussian distributions for the energy<br />

differences A and V.<br />

In both cases, the resulting simulated spectra<br />

can be visualized as a superposition, weighed<br />

in a proper way, of different spectra each one of<br />

them corresponds to a different g-tensor: the<br />

final expression of eq.(lO) convoluted with two<br />

gaussians is<br />

dS(vc,<br />

dH r 2%ar dH<br />

r 2- r ;<br />

(12)<br />

where T, and ?2 refer to the gaussian<br />

distributions.<br />

Computer-synthesized spectra have been used to<br />

fit the experimental EPR spectra; a<br />

minimization procedure of the x^-function,<br />

based on a simulated annealing approach [27],<br />

has been followed to extract the parameters A 0 ,,<br />

and and<br />

A2<br />

characterizing the two gaussian distributions<br />

for high and low spin Mb samples, respectively;<br />

the chi-square function being<br />

i = 1 a. (13)<br />

where I^PCHj) is the derivative of the<br />

experimental EPR absorption spectrum<br />

sampled at 500 discrete points of the magnetic<br />

field, I 8im (Hj,p) is the simulated spectrum,<br />

finally cTj is the standard deviation calculated<br />

for the i-th experimental point of the EPR<br />

spectrum by repeated runs.<br />

Results and Discussion<br />

Fig.3 shows two examples of the experimental<br />

and the corresponding simulated spectra for the<br />

analyzed metMb and azide Mb samples. The<br />

parameters A 0 ,, A°2,


238 Bulletin of Magnetic Resonance<br />

TABLE 1 Central values and variances of the gaussian distributions of the crystal field parameters Ax A2<br />

for high spin ferric Mb samples, and A V for low spin ferric Mb samples obtained by simulations (through<br />

eq.(12) ) of the experimental 77 K EPR spectra of Mb frozen solutions. Gly means that the sample has<br />

been prepared in 1:1 (by volume) water-glycerol mixture. Fast means that the sample has been submitted<br />

to a fast cooling rate ( about 50 deg/min). Slow means that the sample has been submitted to a slow<br />

cooling rate ( about 0.4 deg/min ).<br />

SAMPLE<br />

High spin Mb (Fast)<br />

High spin Mb (Slow)<br />

High spin Mb+Gly (Fast)<br />

High spin Mb+Gly (Slow)<br />

Low spin Mb (Fast)<br />

Low spin Mb (Slow)<br />

Low spin Mb+Gly (Fast)<br />

Low spin Mb+Gly (Slow)<br />

A? cm- 1<br />

2266<br />

2250<br />

2194<br />

2248<br />

Ao<br />

3.03<br />

3.01<br />

3.08<br />

3.05<br />

parameters distributions in both the high and<br />

low spin case and in fast and in slow cooled<br />

samples; such an effect can be interpreted in<br />

terms of a decrease in the structural<br />

heterogeneity of the protein as induced by<br />

glycerol [7,8]. Different molecular<br />

mechanisms could be invoked to interpret such<br />

an effect; it is possible that addition of glycerol<br />

could result into a viscosity-induced damping<br />

of the protein motion; on the other hand,<br />

changes in the dielectric properties of the<br />

solvent could result into a different shielding of<br />

the electrical charges of the amino acid<br />

residues [30,31] and then into a modification of<br />

the protein dynamics; moreover, glycerol, by<br />

decreasing the ice-crystal dimensions, could<br />

minimize the freezing strains [32].<br />

In the high spin Mb samples, the slow cooling<br />

rate induces, in presence and in absence of<br />

glycerol, a narrowing of the crystal field<br />

parameters distributions A , and A2- Such an<br />

effect, which has been observed also in high<br />

spin ferric Hb samples [7] can be interpreted in<br />

different ways. First of all, a sort of<br />

"condensation" could take place in the<br />

molecules populating the frozen CS distribution<br />

[7]; moreover, the cooling rate could induce<br />

some modifications in the state of the hydration<br />

water, as it has been observed by calorimetric<br />

259<br />

239<br />

280<br />

232<br />

CTA<br />

0.13<br />

0.14<br />

0.09<br />

0.10<br />

A£ cm- 1<br />

5759<br />

5750<br />

5500<br />

5417<br />

Vo<br />

2.00<br />

2.01<br />

2.03<br />

2.02<br />

aA, cm- 1<br />

936<br />

919<br />

549<br />

516<br />

*v<br />

0.06<br />

0.08<br />

0.05<br />

0.07<br />

measurements [33], and consequently affect the<br />

CS distribution; finally, it cannot be ruled out<br />

the possibility that cooling rate, acting on the<br />

crystal growth, modifies the freezing straininduced<br />

effects that might be present in low<br />

temperature heme-proteins [34,35]. It is aspected<br />

that all these mechanisms should be operative<br />

also in low spin Mb samples, in which,<br />

however, it has been observed that the slow<br />

cooling rate induces an increase of the<br />

variances a ^ and 0y.<br />

The different behaviour registered in this case<br />

requires a deeper investigation of the role<br />

played by the strong sixth ligand in connection<br />

with the freezing procedure.<br />

In particular, we can speculate about the<br />

possibility that the freezing procedure might<br />

affect, in some way, the average position and<br />

also the spread of the N'g ligand. In this<br />

context, it should be noted that the different<br />

number of ligand orientation, as induced by<br />

different cooling rates, have been observed in<br />

oxycobalt Mb [34].Therefore, different cooling<br />

rates could result into different arrangements<br />

of the ligand with respect to the metal ion.


Vol. 14, No. 1-4<br />

Conclusions<br />

EPR results to be a suitable spectroscopy to<br />

study the structural heterogeneity displayed<br />

by the ferric Mb samples in both the high<br />

and the low spin configuration as derived<br />

from an ensemble of different structures. In<br />

particular, it can fruitfully be applied to<br />

investigate the dynamical coupling between<br />

the solvent and the protein CS distribution.<br />

References<br />

[I] D. O. Hearshen, W. R. Hagen, R. H.<br />

Sands, H. J. Grande, H. L. Crespi, I. C.<br />

Gunsalus, W. R. Dunham, J. Magn.<br />

Resonance 69 (1986)440.<br />

[2] C. More, P. Bertrand and J.P. Gayda, J.<br />

Magn. Resonance 73 (1987) 13.<br />

[3] A. S. Yang and B. J. Gaffney, Biophys. J. 51<br />

(1987)55.<br />

[4] J.C. Salerno: Biochem. Soc. Trans. 13<br />

(1985) 611.<br />

[5] A. S. Brill, F. G. Fiamingo, D. A. Hampton,<br />

J. Inorg. Biochem. 28 (1986) 137.<br />

[6] S. Cannistraro, J. Phys. France 51 (1990)<br />

131.<br />

[7] A.R. Bizzarri and S. Cannistraro, Appl.<br />

Magn. Res. 2(1991)627.<br />

[8] A. R. Bizzarri and S. Cannistraro, Biophys.<br />

Chem. 42(1991)79.<br />

[9] H. Frauenfelder, F. Parak, R. D. Young<br />

Ann. Rev. Biophys. Biophys. Chem. 17 (1988)<br />

451.<br />

[10] V.I.Goldanskii and Y. F. Krupyanskii,<br />

Quart. Rev. Biophys. 22 (1989) 39.<br />

[II] A. Ansari, J. Berendzen, S. F. Bowne, H.<br />

Frauenfelder, I. E. T.Iben, T. B. Sauke, E.<br />

Shyamsunder, R. D. Young, PNAS USA 82<br />

(1985) 5000.<br />

[12] H. Frauenfelder, in Proteins and glasses<br />

(E. Luscher, G. Fritsch and G. Jacucci eds.)<br />

Amorphous and Liquid Materials, Nato series,<br />

Martinus Nijhoff Publishers, Dordrecht (1987).<br />

[13] M. K. Hong, D. Braunstein, B. R. Cowen,<br />

H . Frauenfelder, I. E. T. Iben, J. R. Mourant,<br />

P. Ormos, R. Scholl, A. Schulte, P. J. Steinbach,<br />

A. H. Xie, R. D. Young, Biophys. J. 58 (1990)<br />

429.<br />

[14] H. Frauenfelder, N. A. Alberding, A.<br />

Ansari, D. Braunstein, B. R. Cowen, M. K.<br />

Hong, I. E. T. Iben, J. B. Johnson, S. Luck, M.<br />

C. Marden, J. R. Mourant, P. Ormos, L.<br />

Reinisch, R. Scholl, A. Schulte, E.<br />

Shyamsunder, L. B. Sorensen, P. J. Steinbach,<br />

A. H. Xie, R. D. Young, K. T. Yue, J. Phys.<br />

Chem. 94(1990) 1024.<br />

[15]E.E. Di Iorio, U.R. Hiltpold, D. Filipovic,<br />

K.H. Winterhalter, E. Gratton, E. Vitrano, A.<br />

Cupane, M. Leone and L. Cordone, Biophys. J.<br />

59(1991)742.<br />

[16] G. Giugliarelli, P. Tancini and S.<br />

Cannistraro, J. Phys. E (Sci. Instrum) 22<br />

(1989)702.<br />

[17] M. Kotani Adv. Quantum Chem. 4 (1968)<br />

227.<br />

[18] J.F. Gibson, in ESR and NMR of<br />

Paramagnetic Species in Biological and<br />

related Systems, , eds. I. Bertini and R.S.<br />

Drago, Reider Publishing Company (1979).<br />

[19] J. Peisach, W.E. Blumberg, S. Ogawa,<br />

E.A. Rachmilewitz and R. Oltzik, J. Biol.<br />

Chem. 25(1971)3342.<br />

[20] S. Cannistraro, Chem. Phys. Lett. 122<br />

(1985)165.<br />

[21]C.P. Scholes, J. Chem.Phys.52(1970)4890.<br />

[22] F.G. Fiamingo, A. S. Brill, D. A.<br />

Hampton, R. Thorkildsen, Biophys. J. 55<br />

(1989)67.<br />

123] J.H.E. Griffith, Theory of Transition of<br />

Metal Ions, Cambridge Univ. Press, London<br />

and New York (1961).<br />

[24] R. Aasa and T. Vanngard, J. Magn.<br />

Resonance 19(1975) 308.<br />

[25] A. Isomoto, H. Watari, M. Kotani, J.<br />

Phys. Soc. Japan 29 (1970) 1571.<br />

[26] S. Cannistraro and G. Giugliarelli, Mol.<br />

Phys. 58(1986) 173.<br />

[27] S. Kirkpatrick, C. D. Gelatt, M. P. Vecchi,<br />

Science 220(1983)671. .<br />

[28] C. T. Migita, K. Migita and M. Iwaizumi,<br />

Biochim. Biophys. Acta 743(1983)290.<br />

[29] C. More, J.P. Gayda and P. Bertrand J.<br />

Magn. Resonance 90(1990)486.<br />

[30] G.P. Singh, H. J. Schink, H. V.<br />

Lohneysen, F. Parak, S. Hunklinger, Z. Phys.<br />

B55(1984)23.<br />

[31] G. P. Singh, F. Parak, S. Hunklinger, K.<br />

Dransfert, Phys. Rev. Lett. 47(1981) 685.<br />

[32]W.R. Hagen, J. Magn. Resonance44(1981)<br />

447.<br />

[33] W. Doster, A. Bachleitner, R. Dunau, M.<br />

Hiebl, E. Luscher, Biophys. J. 50 (1986) 213.<br />

[34] H. Hori, M. Iketa-DSaito and T.<br />

Yoketani, Nature 288 (1980) 501.<br />

[35]M.R. Ondrias and D.L. Rousseau, Science<br />

213(1981)2066.<br />

239


240 Bulletin of Magnetic Resonance<br />

1 Introduction<br />

Effect of Paramagnetic Ions in<br />

Aqueous Solution<br />

for Precision Measurement<br />

of Proton Gyromagnetic Ratio<br />

Ae Ran Lim, Chang Suk Kim<br />

Korea Research Institute of Standards and Science,<br />

Taejon 305-606, Korea<br />

and<br />

Sung Ho Choh<br />

The measurement of proton gyromagnetic ratio }'p'<br />

has been the object of an intensive experimental<br />

program for several decades [l]-[3]. The<br />

gyromagnetic ratio of the proton is defined as a<br />

resonance frequency «p divided by a magnetic field<br />

Bo [4], when a spherical water sample at 25°C is<br />

applied by a magnetic field. The ?P' for a defined<br />

pure water sample is somewhat difficult to measure<br />

because of the weak absorption due to the long<br />

relaxation time of the proton [5]. In order to<br />

reduce the relaxation time, paramagnetic ions<br />

are added to the water sample [6]. However,<br />

as the concentration of paramagnetic ions<br />

increases, the resonance point of *H shifts and<br />

differs according to the sample shapes.<br />

The purpose of present work is to investigate the<br />

effect of paramagnetic ions(Fe 3 *, Mn 2 *, Co 2 *, and<br />

Cu 2 *) on the l H NMR in paramagnetic aqueous<br />

solutions [Fe(NO3)3- 9H2O, FeCl3, MnCl2-4H2O,<br />

CoCh-2H2O, and CuCl2-2H2O]. !H NMR in<br />

aqueous solution containing paramagnetic ions was<br />

measured as a function of concentration for the<br />

fixed spherical and cylindrical sample shapes, and<br />

the spinning cylindrical sample shape. The<br />

magnetic susceptibility per unit volume of the<br />

paramagnetic solution has also been measured as a<br />

function of concentration at room temperature.<br />

The interaction between the ' H nucleus and<br />

Department of Physics, Korea University,<br />

Seoul 136-701, Korea<br />

paramagnetic ion is discussed in terms of the<br />

shift of >H resonance point measured with two<br />

sample shapes and the magnetic susceptibility of<br />

the solution. From these experimental results,<br />

we discuss the paramagnetic solution having a<br />

short relaxation time and nearly zero shift of<br />

resonance point to implement the precision<br />

measurement of proton gyromagnetic ratio in a<br />

low magnetic field.<br />

2 Experimental Method<br />

*H NMR experiment was performed by a<br />

Brucker model MSL 200 FT pulse spectrometer.<br />

l H NMR of paramagnetic aqueous solutions in<br />

high magnetic field of 4.7 and 11.74 T, and<br />

frequency of 200.13 and 500 MHz was<br />

measured with two sample shapes at room<br />

temperature. The magnetic susceptibility has<br />

been measured using the Gouy magnetic balance.<br />

2.1 *H NMR and Relaxation Time in<br />

Paramagnetic Aqueous Solutions<br />

The paramagnetic aqueous solutions [Fe(NO3)3-<br />

9H2O, FeCb, MnCl2-4H2O, CoCl2-2H2O,<br />

CuCl2-2H2O] were prepared by dissolving<br />

paramagnetic ions of various concentration in


Vol. 14, No. 1-4 241<br />

distilled water. The linewidth and the shift of 'H<br />

resonance point were measured at room<br />

temperature according to the shape of the<br />

sample tube and the paramagnetic ion<br />

concentration of aqueous solution. The<br />

linewidth and the resonance point of *H NMR<br />

in pure water were also measured.<br />

The spin-lattice relaxation time (7"i) and<br />

spin-spin relaxation time (T2*) were determined<br />

from the signal of l H NMR at room<br />

temperature by the inversion recovery method<br />

and the inverse of linewidth, respectively.<br />

2.2 Magnetic Susceptibility of Paramagnetic<br />

Aqueous Solution<br />

The density of paramagnetic aqueous solution<br />

was measured with the mass and volume, and<br />

the magnetic susceptibility per unit volume was<br />

obtained with the susceptibility per unit mass.<br />

3 Experimental Results<br />

3.1 Shift of !H Resonance Point and Linewidth<br />

In case of cylindrical sample tube, the<br />

resonance point of r H NMR was shifted to the<br />

negative direction with respect to that of pure<br />

water according to the concentration of<br />

paramagnetic ion as shown in Figure 1. It was<br />

nearly unchanged with the variation of<br />

concentration of paramagnetic ion in<br />

Fe(NO3>3-9H2O solution. However, the identical<br />

Fe 3+ ion in FeCb shows the largest shift of the<br />

resonance point. Since the distribution of<br />

valence electrons is influenced by the chemical<br />

bqnding of an atom, it could be explained that<br />

the displacement of nuclear magnetic resonance<br />

frequency depends upon the chemical<br />

environment [7].<br />

Figure 2 shows the linewidth of } H NMR<br />

according to the concentration of paramagnetic<br />

ions. Here we have used the Lorentzian<br />

absorption lineshape, and the linewidth<br />

corresponds to the full width at the half<br />

maximum. The Co 2+ and Cu 2+ ions are<br />

almost ineffective to the linewidth. Whereas the<br />

paramagnetic aqueous solutions containing Fe 3+<br />

ion influence the linewidth as a function of<br />

concentration of paramagnetic ions. The trend<br />

in aqueous solution containing the Mn 2+ ion<br />

differs from those in other paramagnetic ions.<br />

For the case of spherical sample shape, the<br />

shift of resonance point as a function of<br />

concentration of paramagnetic ions is shown in<br />

OO<br />

o<br />

5<br />

-10-<br />

> -15-1<br />

B = 4.7 T<br />

gy-;wHyt-w3<br />

Fe(NO3)3-9H2O<br />

FeCb<br />

-20-<br />

0 To" 20 30<br />

Ion Concen.(10 20 ions/cc)<br />

Figure 1. The frequency shift of 'H NMR signal as<br />

a function of concentration of paramagnetic ions in<br />

aqueous solution contained in a fixed cylinder (* is<br />

the resonance point of 'H in pure H2O).<br />

e(NO3) 9H2O<br />

0<br />

Ion Concen.(10 20 ions/cc)<br />

Figure 2. The linewidth of 'H NMR signal as a<br />

function of concentration of paramagnetic ions in<br />

aqueous solution contained in a fixed cylinder (* is<br />

the linewidth of 'H in pure H2O, 0.15 kHz).<br />

Figure 3. The resonance point of 'H NMR was<br />

shifted to the positive direction compared with<br />

that of pure water. The frequency shift was<br />

nearly unchanged with the variation of the<br />

concentration of paramagnetic Cu 2+ ion. The<br />

linewidth of *H NMR in the spherical shape is<br />

the same as that in the cylindrical one as shown<br />

in Figure 2, i.e. the linewidth has no difference<br />

between the cylindrical and spherical samples<br />

within the experimental error.


242 Bulletin of Magnetic Resonance<br />

00<br />

O<br />

o<br />

20-<br />

15-<br />

B = 4.7 T<br />

B"<br />

MnCl2-4H2O<br />

•CuCU-2H;O<br />

0 10 20 30<br />

Ion Concen.(10 20 ions/cc)<br />

Figure 3. The frequency shift of 'H NMR signal as<br />

a function of concentration of paramagnetic ions in<br />

aqueous solution contained in a sphere (* is the<br />

resonance point of 'H in pure H2O).<br />

However, the shift of the resonance point of<br />

l H in case of spinning cylindrical sample<br />

shows the similar trend as that of the fixed<br />

spherical sample, but with the larger shift than<br />

the fixed sphere as shown in Figure 4. Each<br />

100-<br />

B = 11.74 T /A^MnClr4H,O<br />

FeClj<br />

/<br />

Fe(NOj)j-9HzO<br />

CoCl;-2H2O<br />

CuCl2-2H2O<br />

0 10 20 30<br />

Ion Conceti.(10 20 ions/cc)<br />

Figure 4. The frequency shift of *H NMR signal as<br />

a function of concentration of paramagnetic ions in<br />

aqueous solution contained in a spinning cylinder (*<br />

is the resonance point of 'H in pure H2O).<br />

line in these figures was determined by the least<br />

square fit with the experimental data.<br />

3.2 l H Relaxation Time<br />

Figure 5 shows the *H spin-lattice relaxation<br />

time (Ti) for the cylindrical and spherical<br />

samples measured by the inversion recovery<br />

method. As the concentration of paramagnetic<br />

ions increased, the relaxation time was<br />

shortened. The 'H relaxation time of<br />

paramagnetic ions containing Co 2 * or Cu 2+ was<br />

longer than that of Mn 2 + or Fe 3 + . The 'H<br />

spin-lattice relaxation time of 2.51 s measured in<br />

pure water was consistent with the previously<br />

reported value of 2.3 s at 20°C and 29 MHz<br />

[8]. Figure 6 shows the spin-spin relaxation<br />

time (7*2 *) for the cylindrical and spherical<br />

samples obtained with the inverse linewidth of<br />

the resonance line. This result shows the<br />

similar trend as that of Ti. However, the J H<br />

relaxation time 7z* is shorter than Ti in<br />

aqueous solutions.<br />

10-5. 10 20<br />

10 21<br />

! T c(N03)-9HzO<br />

1022<br />

Ion Concentration(ions/cc)<br />

Figure 5. Spin-lattice relaxation time Ti of 'H due<br />

to the paramagnetic ions in aqueous solution<br />

contained in the fixed cylindrical and spherical<br />

shapes. Both shapes have the same values within<br />

experimental error.<br />

3.3 Magnetic Susceptibility<br />

The magnetic susceptibility per unit volume<br />

of the paramagnetic aqueous solution was<br />

obtained by the Gouy magnetic balance as a<br />

function of concentration of paramagnetic ions


Vol. 14, No. 1-4 243<br />

o<br />

GO<br />

"B<br />

H<br />

o<br />

•X<br />

i<br />

'E,<br />

10" 4 -<br />

10 -5-<br />

B = 4.7 T<br />

ESS 3SS1<br />

10 20 10 21<br />

Ion Concentration(ions/cc)<br />

Figure 6. Spin-spin relaxation time r2* of >H due<br />

to the paramagnetic ions in aqueous solution<br />

contained in the cylindrical and spherical shapes.<br />

Both shapes have the same values within<br />

experimental error.<br />

at room temperature as shown in Figure 7.<br />

The susceptibility was proportional to the<br />

concentration of paramagnetic ions.<br />

4 Analysis and Discussion<br />

10-4-<br />

MnCh-4H2O<br />

FeClj<br />

«CoCl2-2H2O<br />

Fe(NO3)3-9H2O<br />

CuCU- 2HjO<br />

10-6-<br />

1020 1021 1022<br />

Ion Concentration(ions/cc)<br />

Figure 7. The magnetic susceptibility of aqueous<br />

solution as a function of concentration of<br />

paramagnetic ions.<br />

The l H resonance point in pure water differs<br />

from that in the paramagnetic aqueous<br />

solution. The paramagnetic aqueous solution<br />

induced the shift of resonance point due to the<br />

presence of paramagnetic ions. In this study,<br />

we have tried to search a suitable paramagnetic<br />

solution, having the short relaxation time and<br />

nearly zero shift of resonance point in order to<br />

obtain the correct proton resonance frequency in<br />

a low magnetic field.<br />

4.1 The Shift of 'H Resonance Point and<br />

Interaction Factor<br />

For a liquid, the time averaged field at a<br />

nucleus may be divided into three significant<br />

components<br />

Bav = Bo B' + B" (1)<br />

where Bo is the external magnetic field, which is<br />

the main component in Bav. B' is the magnetic<br />

shielding field at the nucleus due to the induced<br />

motion of the electrons in the atom or molecule.<br />

B" is the magnetization field due to the<br />

paramagnetic ions to shorten the spin-lattice<br />

relaxation time T\ of the nuclear spin system.<br />

The dipole interaction between the *H nucleus<br />

and paramagnetic ion is given by [9]<br />

(2)<br />

The field B\ is ascribed to the induced magnetic<br />

dipoles on the surface of a small hypothetical<br />

sphere with its center at the nucleus. This is<br />

the so-called Lorentz or cavity field and has the<br />

value (4H/3)M, where W is the magnetization.<br />

The field Bz is the familiar demagnetizing field,<br />

defined by #2 = -aW, where a is the<br />

demagnetizing factor. The value of a is 47?/3 and<br />

2n for the spherical and infinite cylindrical<br />

sample perpendicular to the field, respectively. It<br />

might be expected that the remaining field #3<br />

due to those paramagnetic ions inside the<br />

hypothetical sphere would be exactly zero.<br />

However, it is found experimentally that #3 may<br />

differ significantly from zero. Therefore, we<br />

define an "interaction factor" q=B3/M. The<br />

expression for B" hence becomes [10]<br />

B" = [(4B/3) - a + q]M (3)<br />

The magnetization H was obtained from the<br />

susceptibility per unit volume according to the<br />

concentration of paramagnetic ions. Also, the


V '<br />

244 Bulletin of Magnetic Resonance<br />

shift of resonance point(B") for paramagnetic<br />

ions with respect to resonance point of proton in<br />

pure water was measured from *H NMR<br />

experiment in various paramagnetic solutions.<br />

Using the magnetic field induced to the 'H<br />

nucleus and the value of magnetization, we<br />

calculated the interaction factor q from eq.(3)<br />

for the fixed spherical and cylindrical samples.<br />

A summary for the various paramagnetic ions is<br />

given in Table 1. The consistency of the<br />

Table 1. Experimental values of the interaction<br />

factor q for the fixed spherical and cylindrical<br />

samples, obtained with eq.(3).<br />

paramagnetic<br />

ions<br />

Fe 3 *<br />

Fe 3 '<br />

Mn 2i<br />

Co 2t<br />

Cu 2 '<br />

chemical<br />

compound<br />

Fe(NO3)3-9HjO<br />

FeCl3<br />

MnCI2-4H2O<br />

CoCl2-2H2O<br />

CuCl2-2H2O<br />

Q<br />

cylinder<br />

2.19<br />

0.85<br />

1.70<br />

1.28<br />

0.39<br />

value<br />

sphere<br />

1.77<br />

0.86<br />

1.21<br />

1.06<br />

0.85<br />

experimental data is indicated by the agreement<br />

between the interaction factors for the<br />

corresponding cylindrical and spherical cases.<br />

The amount of disagreement can be attributed<br />

partly to the experimental error and partly to<br />

the meniscus effect and the lack of perfect<br />

sphericity of the spherical sample.<br />

4.2 Relaxation Time<br />

The spin-lattice relaxation time measured by the<br />

inversion recovery method with a pulse sequence<br />

of 180°(2 /is) - t - 90°( 1 us) - 5 fis(Td) - free<br />

induction decay. The ringing down delay-time Ta<br />

was used to remove the effect of the pulse and<br />

the free induction decay was measured with<br />

time t.<br />

The spin-lattice relaxation time obtained with<br />

the inversion recovery method decreases as the<br />

concentration of paramagnetic ions increases.<br />

The relaxation time measured with the<br />

spherical sample is similar to that with the<br />

cylindrical sample. The 'H relaxation time of<br />

paramagnetic solution containing Co 2 + or Cu 2 *<br />

shows longer than that containing Mn 2 * or<br />

Fe 3+ . The spin-lattice relaxation time of J H in<br />

various paramagnetic solutions is shorter than<br />

that in pure water because of the interaction<br />

between the nuclear spin and paramagnetic<br />

ions. When the number of paramagnetic ions<br />

was increased, the shortening mechanism of the<br />

relaxation time could be understood as follows.<br />

If the number of paramagnetic ions is increased,<br />

the nuclear spin is coupled more with the<br />

magnetic field produced by the paramagnetic<br />

ions. This magnetic interaction between the<br />

nuclear spin and magnetic field of the<br />

paramagnetic ions can contribute to the decrease<br />

in the spin-lattice relaxation time [11].<br />

In case of the short spin-lattice relaxation<br />

time Ti, the following relation generally holds<br />

[12] :<br />

(V)" 1 =<br />

(4)<br />

where Tz is the "natural" spin-spin relaxation<br />

time, and Tz' is the time due to the field<br />

inhomogeneity. The value of Tz" is measured<br />

from the full width at half maximum of the<br />

NMR lineshape.<br />

The linewidth of 'H NMR was brodened when<br />

the concentration of paramagnetic ions was<br />

increased. In case of the aqueous solution<br />

containing Fe 3+ ion, the linewidth was<br />

remarkably increased according to the<br />

concentration of paramagnetic ions. The<br />

linewidth could be broadened by the magnetic<br />

dipole field produced by the paramagnetic ions<br />

at the site of *H nucleus. Normally the dipole<br />

field of the paramagnetic ions has the field<br />

strengths of several thousands times greater than<br />

that due to the magnetic monents of the nucleus,<br />

but it is averaged out at the site of ] H nucleus.<br />

Consequently only a small effect, the linewidth<br />

broadening is occured in the magnetic resonance<br />

[13].<br />

5 Conclusion<br />

For the spherical and cylindrical samples, B"<br />

would be always zero and positive,<br />

respectively, if q were zero. The deviation of<br />

the shift of resonance point between the<br />

experimental results and the theoretical<br />

prediction (q = 0) could be understood as an<br />

effect due to an additional interaction between<br />

the paramagnetic ions and the 'H nucleus.<br />

The spin-lattice and spin-spin relaxation times of<br />

l H NMR in paramagnetic aqueous solution were<br />

shortened as the concentration of paramagnetic ions<br />

was increased.<br />

From these experimental results, we found<br />

that the paramagnetic solution having the<br />

short relaxation time and nearly zero shift of<br />

resonance point is CUO22H2O aqueous<br />

solution. Therefore, the aqueous solution<br />

containing Cu 2+ would be the best candidate to<br />

implement the precise determination of the


Vol. 14, No. 1-4 245<br />

proton gyromagnetic ratio.<br />

Acknowledgement<br />

This work was supported by the Ministry of<br />

Science of Technology and in part the KOSEF<br />

through the SRC of Excellence Program<br />

(1991-94).<br />

References<br />

[1] E.R.Williams and P.T.Olsen, Phys. Rev.<br />

Lett. 42, 1575 (1979).<br />

[2] E.R.Williams, G.R.Jones, J.S. Song,<br />

W. D. Phillips, and P. T. Olsen, IEEE<br />

Trans. Instrum. Meas. IM-38(2), 233<br />

(1989).<br />

[3] H. Nakamura, N. Kasai and H. Sasaki,<br />

IEEE Trans. Instrum. Meas. IM-36, 196<br />

(1987).<br />

[4] N. Bloembergen, Nuclear Magnetic<br />

Relaxation (W. A. Benjamin, New York,<br />

1961), Chap. 4.<br />

[5] J. H. Simpson and H. Y. Carr, Phys. Rev.<br />

Ill, 1201 (1958).<br />

[6] N. Bloembergen, E. M. Purcell, and R. V.<br />

[7]<br />

Pound, Phys. Rev. 73, 679 (1948).<br />

J. T. Arnold, S. S. Dharmatti and M. E.<br />

Packard, J. Chem. Phys. 19, 509 (1951).<br />

[8] N. Bloembergen and W. C. Dickinson,<br />

Phys. Rev. 79, 179 (1950).<br />

[9] W. C. Dickinson, Phys. Rev. 77, 736<br />

[10]<br />

(1950).<br />

A. R. Lim, S. H. Choh, Saemulli 26, 381<br />

(1986).<br />

[11] A. Abragam, The Principles of Nuclear<br />

Magnetism(Oxford Univ. Press, Oxford,<br />

1961), Chap. 3.<br />

[12] D. Pines and C. P. Slichter, Phys. Rev.<br />

100, 1014 (1955).<br />

[13] W. C. Dickinson, Phys. Rev. 81, 717<br />

(1951).


246 Bulletin of Magnetic Resonance<br />

1. Introduction<br />

Magnetic Resonances of 23 Na and 14 N Nuclei<br />

in Single and Multi-Domain<br />

.Crystals of Ferroelectric NaN02<br />

Sung Ho Choh and Kee Tae Han<br />

Department of Physics, Korea University, Seoul<br />

136-701, Korea.<br />

Betsuyaku [1] and Kanashiro et al [2] reported<br />

that the two satellite lines of 23 Na NMR<br />

in a ferroelectric NaNCh crystal were asymmetric,<br />

and that even the central line for Bo//a<br />

deviates appreciably from the first derivative of<br />

the Gaussian line shape. Kanashiro et al [2] tried<br />

to explain this effect in terms of the nuclear<br />

dipole coupling between the Na atoms. Hughes<br />

and Pandey [3] also studied the same effect, and<br />

atempted to explain it by means of the magnetic<br />

dipolar third moments. They concluded that the<br />

asymmetry arises from the perturbation of the<br />

magnetic dipolar coupling, and this is the<br />

intrinsic property of NaNO2 crystal from the<br />

similarity of experimental results obtained with<br />

two different crystals.<br />

In connection with these reports [1-3], the<br />

present work is focused on the 23 Na NMR line<br />

shape as well as the electric field gradient<br />

(e.f.g.) at 23 Na and 14 N sites in NaNO2 crystal<br />

by employing the NMR and NQR technique. Two<br />

NaNO2 crystals prepared here have different<br />

domain states each other: one has a multi-domain<br />

and the other has a single domain state,<br />

respectively.<br />

2. Experimental<br />

A. Procedure<br />

The NaNO2 is body centered orthorombic and<br />

its space group is C2v 20 -Im2m in the ferro-<br />

electric phase at room temperature [4],. The used<br />

sample crystals were grown from the melt. Virgin<br />

crystals of NaNC>2 turned out to be in the<br />

multi-domain state (herein named by Sm). The<br />

crystal of single domin (Ss sample) was prepared<br />

by applying an external electric field of 3 kV/cm<br />

along the ferroelectric axis of a virgin crystal for<br />

8 hrs near its Tc of 163.5°C. Their domain states<br />

were confirmed by empolying optical polarizing<br />

microscopy and etching technique [5].<br />

The NMR signals for 23 Na nucleus were observed<br />

with an rf-frequency of 6 MHz by using<br />

the cw-NMR (Varian WL-112). They were recorded<br />

by setting the magnetic field modulation at<br />

35 Hz with an approximate amplitude of less than<br />

one-third of the resonance line width. The<br />

pulse-NMR (Bruker MSL 200) system was also<br />

employed to investigate the exact line shape of<br />

23 Na with an inversion recovery sequence, and<br />

NQR measurements were made by using the<br />

Robinson type spectrometer [6]. All the<br />

measurements were made at room temperature.<br />

B. Results<br />

Since the 23 Na nucleus has a spin of 1 = 3/2,<br />

three resonance lines are typically observed: one<br />

central and two satellite lines. The line shape of<br />

23 Na NMR in two single crystals of NaNO2 had<br />

been measured [7] by using the cw-NMR<br />

spectrometer. In the Ss sample, as shown in Fig.<br />

1, the asymmetric satellite lines were observed<br />

for Bo//a and Bo//c, consistent with the previous<br />

reports [1-3]. However, in the Sm, the observed<br />

satellite lines was nearly symmetric (see Fig. 2).


Vol. 14, No. 1-4<br />

(a) Bo//c-axis<br />

(b) Bo//a-axis<br />

-4— -+-<br />

510.0 511.0 531.0 531.5 552.5 553.5 (mT)<br />

505.0 510.0 530.5 531.5 558.0 559.0 (mT)<br />

Fig. 1. In the Ss sample, asymmetric satellite lines were observed for (a) Bo//c and<br />

(b) Bo //a with the cw-NMR spectrometer.<br />

(a) Bo//c-axis<br />

(b) Bo//a-axis<br />

510.0 511.0 531.0 552.5 553.5 (mT)<br />

505.0 510.0 530.5 531.5 558.0 559.0 (mT)<br />

Fig. 2. In the Sm sample, symmetric satellite lines were observed for (a) Bo//c and<br />

(b) Bo//a with the cw-NMR spectrometer.<br />

247


248 Bulletin of Magnetic Resonance<br />

The line width of the satellite lines (ABS) in Ss<br />

was found to he broader than that in Sm. whereas<br />

the central line in these two samples is symmetric<br />

and its line shape is nearly the same; it was 0.29<br />

mT (0.21 mT) for Bo//a (B0//c). Meanwhile, the<br />

quadrupole coupling constant (Qcc) and asymmetry<br />

parameter (7?) for 23 Na of these two samples are<br />

nearly the same within the experimental accuracy<br />

(AQcc : + 0.003 MHz, AT? : ± 0.003) [7].<br />

For the 14 N NQR measurements in the two<br />

samples, there were no detectable differences in<br />

their line widths (At 1 * and AP~ ) and NQR parameters.<br />

The quadrupole parameters at the 23 Na<br />

and 14 N nuclei in the Sm and Ss samples are<br />

summarized in Table 1 together with the line<br />

width. The only difference between the Sm and Ss<br />

samples is in the line shape of 23 Na NMR and<br />

its line width.<br />

The line shape of 23 Na NMR in the two<br />

samples obtained with the pulse-NMR is shown in<br />

Fig. 3 and 4, respectively. In both of two<br />

samples, the central line has one peak and its<br />

line shape is a Gaussian as well known [1-3]. So<br />

the central line is not shown in Fig. 3 and 4,<br />

where PL S(III) (i4i s(m) ) is the satellite line of low<br />

(high) frequency, and superscript s(m) stands for<br />

the single domain (multi-domain) state. In Fig. 3,<br />

(a) shows that each satellite line obtained with Ss<br />

has two peaks (PL S ' and VL S , m s and v\\ s ' ),<br />

and (b) displays their first derivative curves.<br />

However, in the Sm sample, the satellite line (vi m<br />

or PH" 1 ) has only one peak and it is well fitted<br />

with a Gaussian line shape as shown in Fig. 4,<br />

which displays VL m representatively since the two<br />

satellite lines all are symmetric. The inner set<br />

(vi s and vn s ) of the two couples obtained with<br />

Ss [Fig. 3 (a)] corresponds to those (^L m and fH m )<br />

with the Sm sample (Fig. 4). The outer set (vi s '<br />

and m s ' ) in Ss is newly observed additionally.<br />

3. Discussion<br />

Betsuyaku and Kanashiro et al [1-2] had<br />

measured the NMR line shapes for Bo //a and<br />

Bo//c. Their data for Bo//a was distinctively<br />

asymmetric but those for Bo//c was hardly asymmetric.<br />

They, by considering the dipolar interaction,<br />

argued that this effect arised from the<br />

geometrical arrangement of Na atoms in NaNO2<br />

crystal having two fold symmetry along the<br />

b-axis. Namely, the dipole coupling between the<br />

two Na nuclei along the a-axis contribute considerably<br />

to the asymmetry of the satellite lines<br />

because the interatomic distance along the a-axis<br />

is about two-thirds of that along the c-axis.<br />

Kanashiro et al [2] determined the sign of the<br />

Qcc at Na in NaNCh by comparing the line<br />

shape of the satellites for Bo //a with the<br />

calculated line spectrum of two interacting I =<br />

3/2 spins. However, Hughes et al indicated that<br />

there was some uncertainty in the determination<br />

by Kanashiro et al since the resonances showed a<br />

substantial inhomogeneous quadrupole broadening,<br />

and no steps were taken to confirm that the<br />

asymmetry was indeed associated with the dipolar<br />

interaction and not with the quadrupole<br />

broadening. Hughes et al confirmed the sign of<br />

Qcc determined by Kanashiro et al by comparing<br />

the magnetic dipolar third moment with the<br />

observed asymmetry, and reported that the<br />

asymmetry due to the quadrupole broadening was<br />

dominant at some crystal orientations by measuring<br />

quantitatively the asymmetry of inhomo-<br />

Table 1. The quadrupole parameters and the line widths at room temperature.<br />

samples<br />

Sm<br />

Ss<br />

23 Na NMR<br />

Qcc(MHz) .»?(%) ABs(Bo//a) ABs(Bo//c)<br />

1.094 11.0 0.30 mT 0.23 mT<br />

1.094 11.0 0.38 mT 0.25 mT<br />

14 N NQR<br />

Qcc (MHz) 7?(%) At> + (kHz) Ai""(kHz<br />

5.490 35.7 0.66 0.53<br />

5.490 35.7 0.67 0.53


Vol. 14, No. 1-4<br />

(b)<br />

240 230<br />

(KHz)<br />

-250 (KHz)<br />

240 230 (KHz) -240 -250 (KHz)<br />

Fig. 3. The satellite lines of 23 Na NMR in the Ss sample observed for B0//c with the<br />

pulse-NMR spectrometer, (a) Each satellite line has two peaks, (b) the first derivative<br />

curve of (a)<br />

m<br />

I<br />

240 230 KHz<br />

Fig. 4. The satellite line of 23 Na NMR in the Sm sample observed for Bo//c with the<br />

pulse-NMR spectrometer. The satellite line has only one peak and is well fitted<br />

with a Gaussian line shape, where + denotes signal trace and the dashed line is<br />

a best fitted Gaussian line shape.<br />

\<br />

249


250 Bulletin of Magnetic Resonance<br />

geneous quadrupole broadening. They also suggested<br />

that the asymmetry was the intrinsic property<br />

of this crystal by observing the similar quadrupolar<br />

broadening for two different crystals.<br />

As can be seen in Fig. 1, our NMR results<br />

show that the satellite line shape in Ss is<br />

asymmetric, which is in accordance with the<br />

previous reports [1-3]. However, those in Sm is<br />

nearly symmetric (Fig. 2). The exact line shapes<br />

of Fig. 3 and 4 obtained with the pulse-NMR<br />

system may provide a decisive clue for the origin<br />

of the observed asymmetry. Fig. 3(a) shows that<br />

each satellite line in Ss has two peaks, and the<br />

line intensity of the inner set is stronger than<br />

that of the outer one. Fig. 3(b) displays their<br />

first derivative curves in Fig. 3(a). It can be<br />

expected to obtain the asymmetric satellite line of<br />

Fig. 1 by drawing a curve as an envelope of the<br />

satellite line made of two peaks of Fig. 3(b).<br />

Whereas, in the Sm sample, the satellite line has<br />

only one peak as shown in Fig. 4 and it is found<br />

to be well fitted with a Gaussian line shape.<br />

Our experimental results shown in Fig. 4<br />

suggest that the asymmetry due to the dipole<br />

coupling is quite small or very weak based on the<br />

fact that the satellite line in the Sm is well fitted<br />

with a Gaussian. Two peaks in the satellite line<br />

of Fig. 3 may imply that there are possibly two<br />

different absorption centers in the Ss sample.:<br />

one is a normal set of the satellite lines<br />

corresponding to those in Sm and the other is a<br />

new one (vi s ' and PH S ' ). Such a new set might<br />

have orignated from the strained part of the<br />

crystal during the process of the polarization<br />

reversal to make the virgin crystal into a single<br />

domain state. Meanwhile, as listed in Table 1,<br />

the Qcc and r? for 23 Na in these two samples<br />

with the cw-NMR were the same within the experimental<br />

uncerntainty [7]. NQR results such as<br />

line widths and quadrupole parameters of 14 N in<br />

Sm and Ss are also similar to each other. The<br />

appreciable difference between the Sm and Ss<br />

samples is only in the line shape of 23 Na NMR.<br />

These facts on the e.f.g. at 23 Na and 14 N sites<br />

supports the proposal of the existence of two<br />

absorption regions in the Ss sample. The asymmetry<br />

and broad line width of the satellites in<br />

the Ss sample can be understood by means of the<br />

formation of a new region in the crystal due to<br />

the strain induced by the polarization reversal.<br />

From our experimental results, one can<br />

propose that the asymmetric line shape of the<br />

satellite line of 23 Na is due to an imperfection<br />

caused by the applied electric field rather than<br />

the magnetic dipole coupling in the crystal. In<br />

general, crystal imperfection may be produced<br />

during the crystal growth [8] or by the external<br />

stress [9], and due to the external electric field<br />

as the present case [10].<br />

4. Conclusion<br />

We observed the symmetric line shape as well<br />

as the asymmetric one in the 23 Na NMR in the<br />

ferroelectric NaNO2 crystal by employing the<br />

cw-NMR and pulse NMR technique. The pulse<br />

NMR spectrometer provided a better resolved resonance<br />

line shape than the cw-NMR. Each satellite<br />

line obtained with Ss has two peaks, while<br />

only one peak in the Sm sample. Thus the<br />

observable asymmetry might have originated from<br />

the existence of two absorption regions in the<br />

crystal such as a normal region and a new one<br />

abnormally induced in NaNO2 crystal. Such a<br />

new region seems to be due to the internal strain<br />

induced during the process of polarization reversal<br />

by the external electric field. These results imply<br />

that the asymmetry is not the intrinsic property of<br />

an NaNO2 crystal but due to a kind of crystal<br />

imperfection caused by strain.<br />

Acknowledgements<br />

This work is supported by the Korea Science<br />

and Engineering Foundation through the SRC of<br />

Excellence Program (1991-94). Authors are<br />

grateful to Professor R. Blinc of University of<br />

Ljublijana for his helpful discussion during his<br />

visit to Korea in February 1992.


Vol. 14, No. 1-4<br />

References<br />

[1] H. Betsuyaku, J. Phys. Soc. Japan. 27,<br />

1485 (1969)<br />

[2] T. Kanashiro, T. Ohno, and M. Satoh, J.<br />

Phys. Soc. Japan. 54, 2720 (1985).<br />

[3] D. G. Hughes and L. Pandey, J. Mag.<br />

Reson. 75, 272 (1987)<br />

[4] M. I. Kay and B. C. Frazer, Acta cryst.<br />

J4, 56 (1961)<br />

[5] K. T. Han, Ph. D. thesis, Korea Univ.<br />

(1992)<br />

[6] J. Lee and S. H. Choh, Rev. Sci. Instr.<br />

53, 232 (1982)<br />

[7] K. T. Han, H. W. Shin, I. W. Park and<br />

S. H. Choh, J. Korean. Phys. Soc. 25,<br />

67 (1992)<br />

[8] S. H. Choh, J. Lee and K. H. Kang,<br />

Ferroelectrics 36, 297 (1981)<br />

[9] K. T. Han, T. H. Yeom and S. H. Choh,<br />

Ferroelectrics 107, 349 (1990)<br />

[10} Private communication with Prof. R. Blinc<br />

of University of Ljublijana<br />

251


252<br />

1. Introduction<br />

Knight Shifts and<br />

Spin Dynamics in Disordered<br />

Systems<br />

M.J.R. Hoch and S.T. Stoddart<br />

Department of Physics and<br />

Condensed Matter Physics Research Unit,<br />

University of the Witwatersrand, Johannesburg<br />

The metal insulator (MI) transition is a<br />

problem that has received much attention in<br />

solid state physics. Heavily doped<br />

semiconductors have featured prominently in<br />

this work with Si:P the archetypal system.<br />

In order to explain the observed low<br />

temperature properties, such as the magnetic<br />

susceptibility, of MI systems in the vicinity<br />

of the critical concentration, n(;, of dopant<br />

atoms, a phenomenological two—fluid model<br />

has been proposed [I]. The present work is<br />

concerned with interpreting 29 Si NMR<br />

relaxation time measurements and Knight<br />

shifts for Si:P and Si:(P,B) in the vicinity of<br />

nc. The results are analyzed in terms of<br />

available theory in the context of the twofluid<br />

model.<br />

2. The Two-Fluid Model and the Bhatt-<br />

Lee Theory<br />

For the just metallic or just insulating<br />

phases of MI systems, the two—fluid model<br />

distinguishes between two types of electron<br />

spins. As the transition is traversed, the<br />

proportions of the two fluids change. The<br />

fluids are comprised of localized moments<br />

associated with isolated dopant atoms, or<br />

small clusters of dopant atoms, on the one<br />

hand, and delocalized moments on the other.<br />

In broad terms, the localized moments<br />

dominate in determining the magnetic<br />

properties in the vicinity of nc, while the<br />

delocalized moments determine the electrical<br />

properties, such as the conductivity.<br />

Bulletin of Magnetic Resonance<br />

For n < nc, the localized moments<br />

constitute a disordered antiferromagnetic<br />

system. The exchange Hamiltonian may be<br />

written in the usual way as<br />

H = Hi - 1 ' -J '<br />

where Jij is the exchange coupling between<br />

spins i and j and Si and Sj are the spin<br />

operators.<br />

In order to explain the behaviour of<br />

the magnetic susceptibility x with<br />

temperature for n < nc,Bhatt and Lee [2]<br />

have developed a theory in which the<br />

exchange coupling Jy between nearest<br />

neighbour pairs of localized moments is used<br />

to separate the moments into two groups.<br />

In simple terms, spin pairs with Jjj >> kT<br />

are tightly coupled or frozen in the singlet<br />

state and effectively do not contribute to X-<br />

The remaining spins do contribute and the<br />

susceptibility may be written in terms of the<br />

Curie law susceptibility as<br />

n Curie (1)<br />

Numerical procedures were used to<br />

determine ne(T), the effective number of<br />

spins at temperature T. Good quantitative<br />

agreement was obtained with available<br />

experimental susceptibility data. Bhatt and<br />

Lee used the following form for the J<br />

distribution in their calculations :<br />

P(J) « J-tt, with 0.6 < a < 0.8. This led to<br />

X « T-«, as observed.


Vol. 14, No. 1-4<br />

3.<br />

29 Si Spin Relaxation and Localized<br />

Electron Spin Dynamics (n < nc)<br />

Previous work [3] has provided strong<br />

evidence that localized moments dominate in<br />

determining the 29 Si spin lattice relaxation<br />

times at low temperatures. These moments<br />

constitute an exchange coupled reservoir to<br />

which the nuclear spin system is coupled via<br />

the dipolar interaction. Hoch and Holcomb<br />

[3] have analyzed available Ti results using<br />

a model which allows for spin diffusion<br />

to the localized moments, which exhibit<br />

fluctuations in orientation with a frequency<br />

related to the strength of the exchange<br />

coupling to neighbouring spins. Frozen spin<br />

pairs have been excluded by introducing an<br />

effective number of spins in the spirit of the<br />

Bhatt—Lee approach to the susceptibility.<br />

Available T( data, measured at various<br />

fields B, at 1.5 K and 13.5 mK were fitted<br />

reasonably well by choosing the spectral<br />

function for the spin fluctuations to have the<br />

form f(w) « »/w and using calculated values<br />

for other quantities, such as the spin<br />

diffusion coefficient D.<br />

The best fits were obtained by keeping<br />

the diffusion barrier radius, b, constant<br />

independent of the field used. Calculations,<br />

however, suggest that b should vary as in B.<br />

We now propose that the form of the<br />

spectral function can be obtained from the<br />

form of the J distribution used by Bhatt and<br />

Lee. An outline of the treatment is given<br />

below. Converting the J distribution into a<br />

r distribution and integrating over all<br />

correlation times for the unfrozen spins gives<br />

J(u/) = P(T) dr<br />

with P(T) « V 7 " 2 "" • The form of the J(w,r)<br />

may be chosen in various ways corresponding<br />

to different possible forms for the<br />

correlation function. Our calculations<br />

suggest that the form of J(w) is not very<br />

sensitive to this, and for simplicity, we use<br />

an exponential correlation function, leading<br />

to a Debye form for J(w,r). If we put a - 1,<br />

this leads to J(w) « 70;, as used in the work<br />

referred to above. Numerical integration is,<br />

in general, necessary for other values of a.<br />

For values of a < 1, we obtain a<br />

spectral function which varies as l /uP- and<br />

therefore more slowly with frequency than<br />

the l /w form. This permits b to vary when<br />

fitting the experimental data. Further<br />

details will be published elsewhere.<br />

The approach to the spectral function<br />

for an amorphous antiferromagnet, outlined<br />

above, is consistent with the ideas of the<br />

253<br />

Bhatt-Lee theory. Susceptibility measurements<br />

as a function of n and T and nuclear<br />

relaxation measurements as a function of n,<br />

B and T can be explained in terms of the<br />

properties of the localized fluid component.<br />

4. 29Si Knight Shifts (n > nc)<br />

Knight shifts have previously been measured<br />

as a function of dopant concentration n at<br />

4.3 K and 1.5 K [4,5,6] in various magnetic<br />

fields. At high concentrations the Knight<br />

shifts tend towards the Pauli susceptibility<br />

behaviour (xP « n^*), as expected for a<br />

metal. At lower concentrations the values<br />

fall below the Pauli susceptibility<br />

predictions.<br />

Using a tight binding approximation<br />

based on the approach given by Kaveh<br />

and Liebert [7], we have given a semiquantitative<br />

explanation [8] for the observed<br />

Knight shift behaviour for both Si:P and<br />

Si:(P,B). The delocalized moment fluid<br />

determines the Knight shift.<br />

In order to see whether there is any<br />

temperature dependence of the Knight shift,<br />

we made measurements on two just metallic<br />

samples at temperatures down to 50 mK in<br />

an Oxford dilution refrigerator. The samples<br />

were in the form of a stack of wafers, which<br />

were well anchored thermally to the high<br />

purity copper tail used in the refrigerator.<br />

The field of 1 T was supplied by a high<br />

homogeneity superconducting solenoid.<br />

The results are shown in Figure 1.<br />

10<br />

10<br />

10 -6<br />

10 -2<br />

,Si:P<br />

oSi:P<br />

(n/nc = 1.6)<br />

(n/nc = 1.6) Ref. 5<br />

• Si:(P,8)(n/nc = 1.1)<br />

°Si:(P.BHn/ne a 1.1) Ref. 6<br />

10 1<br />

Temperature (K)<br />

Figure 1<br />

The mean Knight shift for Si:P (n/nc = 1.6)<br />

andSi:(P,B) (n/nc = 1.1) as a function of<br />

temperature down to 50 mK.<br />

Within experimental uncertainty it can be<br />

seen that there is no temperature<br />

dependence of the mean Knight shift <br />

over the temperature range 4 K — 50 mK for<br />

either Si:P or Si:(P,B).<br />

In terms of the two-fluid model this<br />

aro irarxr \irpaV<br />

10


254 Bulletin of Magnetic Resonance<br />

interactions between the localized and<br />

delocalized spins. The local susceptibility of<br />

the delocalized electrons does not change<br />

with temperature.<br />

5. Conclusion<br />

Measurements of the 29 Si relaxation rates<br />

and Knight shifts as a function of donor<br />

concentration, magnetic field and<br />

temperature have provided evidence which<br />

supports the two—fluid model for the MI<br />

transition in Si:P. The Ti measurements<br />

provide information on the spin dynamics of<br />

the localized moments, while the Knight<br />

shifts probe the properties of the delocalized<br />

moments.<br />

The relaxation rate<br />

interpreted using ideas<br />

Bhatt—Lee theory for<br />

susceptibility. The form<br />

results may be<br />

based on the<br />

the magnetic<br />

of the spectral<br />

function for spin fluctuations in the<br />

amorphous antiferromagnet system has been<br />

deduced using an accepted form for the<br />

distribution of exchange couplings.<br />

Knight shifts may be explained using a<br />

tight binding model for the delocalized fluid.<br />

No temperature dependence of has<br />

been found over the range 4 K — 50 mK.<br />

This may be interpreted to mean that<br />

interactions between the two fluids, which<br />

occupy spatially distinct regions in the<br />

sample, are very weak.<br />

6. References<br />

1. H. Alloul and P. Dellouve, Phys. Rev.<br />

Lett. 59, 578 (1987).<br />

S. Sachdev, R.N. Bhatt and<br />

M.A. Paalanen, J. Appl. Phys. 63,<br />

4285 (1988).<br />

2. R.N. Bhatt and P.A. Lee, Phys. Rev.<br />

Lett. 48, 344 (1982).<br />

3. M.J.R. Hoch and D.F. Holcomb, Phys.<br />

Rev. B 38, 10550 (1988).<br />

4. S. Kobayashi, Y. Fukagawa, S. Ikehata<br />

and W. Sasaki, J. Phys. Soc. Jpn. 45,<br />

1276 (1978).<br />

5. M.J. Hirsch and D.F. Holcomb, Phys.<br />

Rev. B 33, 25201 (1986).<br />

6. M.J.R. Hoch, U. Thomanschefsky and<br />

D.F. Holcomb, Physica B 165/166, 305<br />

(1990).<br />

7. M. Kaveh and A. Liebert, Phil. Mag.<br />

Lett. 58, 247 (1988).<br />

8. S.T. Stoddart and M.J.R. Hoch - to be<br />

published in Phys. Rev. B.


Vol. 14, No. 1-4 255<br />

Numerical Design and Evaluation of Broadband<br />

Pulse Sequences for 1=1 spin systems<br />

Debra Lynne Mattiello, Jonathan Callahan, Todd Alam^ and Gary Drobny<br />

Chemistry Department, University of Washington,<br />

Seattle WA USA 98195<br />

INTRODUCTION<br />

Quadrupolar coupling constants of<br />

170 kHz result from reduced motional<br />

averaging in biological polymers. The<br />

decay rates of Zeeman and quadrupolar<br />

order are direct measures of the spectral<br />

densities of motion. Information on the<br />

orientation-dependence of relaxation rates<br />

of Zeeman and quadrupolar order, Tlz and<br />

Tlq respectively, assists in understanding<br />

the dynamics of molecules. 1 Sensitivity and<br />

dynamic range considerations mandate<br />

optimum efficiency and uniformity over<br />

large spectral widths. The design of<br />

composite pulses to create broadband<br />

excitation and inversion is well established<br />

for deuterium NMR. 2 " 11 The numerical<br />

optimization of existing composite inversion<br />

pulse sequences was performed in order to<br />

increase the uniformity of excitation over<br />

spectral widths of 250 kHz. 2 " 7<br />

Pulse sequences to create<br />

quadrupolar order over moderately broad<br />

spectral widths have recently been<br />

proposed. 12 " 14 The pulse sequence design<br />

f Present Address: University of New Mexico,<br />

Albuquerque, NM<br />

of Wimperis is a good starting place from<br />

which to numerically optimize for the<br />

creation of quadrupolar order. 13 Broadband<br />

quadrupolar order is transferred to<br />

detectable magnetization with a 45 degree<br />

pulse and an additional refocusing pulse<br />

eliminates large phase corrections. 15<br />

Numerical optimization of the conversion<br />

from quadrupolar order to well-behaved<br />

transverse magnetization is worthy of<br />

investigation.<br />

A program developed in our<br />

laboratory for the numerical optimization of<br />

pulse sequences has been modified for 1=1<br />

spin systems. 15 " 18 This research is part of an<br />

ongoing investigation into the local and<br />

global dynamics of oligonucleotides utilizing<br />

solid state deuterium NMR. At the present<br />

time, the dynamics of the sugar rings of<br />

DNA are under study. The lower levels of<br />

hydration of DNA exhibit rigid-lattice<br />

lineshapes. Information on orientation<br />

dependence and the substantiation of<br />

motional models requires increased<br />

sensitivity and large spectral windows.


256<br />

METHODS<br />

Solid state deuterium NMR spectra<br />

were obtained at 76.72 MHz on a homebuilt<br />

spectrometer controlled by a DEC<br />

micro VAX II. Both the inversion<br />

sequences and the broadband Jeener-<br />

Broekaert experiments were phase cycled to<br />

eliminate double quantum coherence. 19 ' 20<br />

Phase shifting was accomplished with a<br />

homebuilt digital phase shifter. The sample,<br />

a labeled nucleoside, 2"-deutero-2'deoxyguanosine,<br />

was prepared by Jerome<br />

Shiels at the University of Washington.<br />

Either 2K or 4K scans were taken with a<br />

recycle delay of 5.0 seconds. The field<br />

strength was 100 kHz and the dwell time<br />

was 200 nanoseconds. Each spectrum was<br />

acquired with 4096 points.<br />

Calculations were accomplished on a<br />

DEC UNIX 3100 workstation. The strategy<br />

thus far has been to parameterize the pulse<br />

sequence and generate random trial pulse<br />

sequences. The basis set proposed by Vega<br />

and Luz was used. The coherences of<br />

interest were single basis elements rather<br />

than linear combinations of basis<br />

elements. 19 The expectation value of -Iz<br />

Qz quadrupolar order, or Iy was compared<br />

with the target function over a specific<br />

spectral width. These quality factors<br />

quantify the performance of the pulse<br />

sequence as a function of the quadrupolar<br />

frequency. Excitation profiles of the<br />

expectation value as a function of reduced<br />

frequency illustrate overall smoothness and<br />

breadth. Evolution profiles of the elements<br />

of the density operator as a function of time<br />

demonstrate the effects of rf pulses and<br />

evolution under a strong quadrupolar<br />

Hamiltonian. Additional three-dimensional<br />

graphics developed in our laboratory assist<br />

in visualizing the transfer of coherences. 22 "<br />

RESULTS<br />

Bulletin of Magnetic Resonance<br />

Measurement of spin-lattice<br />

relaxation times and investigation of the<br />

orientation-dependence of Tt in solids are<br />

important tools for understanding<br />

dynamics. 1 Inversion pulse lengths of more<br />

than 4 microseconds compromise the<br />

inversion breadth and uniformity across<br />

wide line deuterium powder patterns- 7 A<br />

variety of composite pulse schemes have<br />

been proposed for spectra with widths<br />

approaching the Rabi frequency of the<br />

radiofrequency pulse. The composite<br />

excitation triplet designed by Levitt, Suter<br />

and Ernst and supercycled in the method<br />

proposed by Levitt in order to create<br />

broadband inversion of deuterium<br />

lineshapes was numerically optimized. •<br />

The sequence consists of the composite<br />

excitation pulse, 45o9O18O135o, supercycled<br />

in the triplet form, O0 O9Q 3>O Only the<br />

flip angles of the excitation triplet were<br />

parametrized. The sequence has already<br />

been shown to invert rigid-lattice deuterium<br />

spectra with an rf field of 139 kHz while a<br />

weaker field led to a small loss in<br />

performance. 9<br />

Higher order pulses would lead to<br />

longer total pulse lengths. The total<br />

duration of the composite pulse should be<br />

short to avoid irreversible loss of<br />

magnetization during the excitation. 9<br />

Experimental spectra acquired with the<br />

optimized composite excitation pulse of<br />

43010018Q142Q, supercycled as above (B)<br />

and the Levitt triplet (A) are displayed in<br />

figure 1.<br />

Recent application of Tycko's use of<br />

the Magnus expansion for the design of<br />

composite pulses has led to the<br />

development of new excitation schemes. 8 ' 11


Vol. 14, No. 1-4 257<br />

B<br />

P<br />

460 n6<br />

kHz<br />

-46a<br />

fig-1<br />

The modified Jeener-Broekaert pulse<br />

sequence was founded on the model of a<br />

composite excitation pulse broadband with<br />

respect to rf field strength 13 ' 14<br />

The multipulse sequence eliminated<br />

the frequency selection of the original<br />

Jeener-Broekaert experiment. The<br />

broadband sequence has been utilized to<br />

measure the spectral densities of liquid<br />

crystals. 21 Increased signal sensitivity<br />

becomes highly desired as one goes to<br />

broader linewidths, lower Larmor<br />

frequencies and smaller biological samples<br />

with dilute nuclei.<br />

Computer search routines to create<br />

broadband quadrupolar order were<br />

performed with 10, 8 and 7 parameters.<br />

The 10 parameter search consisted of 4<br />

pulses, 3 phases, and 3 seperate delays. The<br />

8 parameter search had a single delay<br />

parameter. An assortment of the resulting<br />

sequences were tested experimentally. The<br />

"91" sequence was a 10 parameter search<br />

optimized over a spectral width of 300 kHz.<br />

It proved to perform well in breadth and<br />

sensitivity. Experimental spectra obtained<br />

with the Wimperis sequences A and B and<br />

the "91" pulse sequence, C, found in this<br />

study, are shown in figure 2.<br />

400 -400<br />

A: 900 -4.0p. -67.5270 -4.0n -4590 -2.0p. -4590<br />

B: 900 -4.0(i -75270- 4.0M. -52.590 -2.0M -4590<br />

C: 1130 -5.5n -72 284 -3.2n -52142 -3.9M -59112<br />

CONCLUSION<br />

fig-2<br />

The use of solid state deuterium<br />

NMR as a probe of dynamics in DNA offers<br />

the luxury of selective labeling. The same<br />

luxury introduces some of the limitations of<br />

site-selective wide line spectroscopy. The<br />

technique enables one to focus on the<br />

dynamics of single positions within large<br />

molecules. The small, precious samples are<br />

dilute in the observe nuclei yet one can<br />

monitor the onset of motion in both a local<br />

and global fashion as water is added to the<br />

spaces in DNA. The technique requires high<br />

power, short pulses for broad lineshapes,<br />

resistant samples, and extensive phase<br />

cycling and signal averaging. The dry DNA<br />

has longitudinal relaxation times of several<br />

seconds making signal intensity even more<br />

elusive. For these reasons, this investigation<br />

aims to numerically optimize existing pulse<br />

sequences for the creation of selected<br />

coherences over static deuterium powder<br />

linewidths of 250 kHz. Analysis of the


258<br />

evolution of the system provides clues for<br />

producing the very best tailored excitation.<br />

REFERENCES<br />

1 R.R. Void and R.L. Void, "Advances in Magnetic<br />

and Optical Resonance". Vol. 16, Academic Press,<br />

San Diego (1991)<br />

2 R. Tycko. Phys. Rev. Lett. 51. 775, (1983)<br />

-* M.H. Levitt. D. Suter. and R.R. Ernst, J. Chem.<br />

Phys.,S0, 3064 (1984)<br />

4 R. Tycko, E. Schneider. A. Pines. J. Chem. Phys.,<br />

81, 680 (1984)<br />

5 R. Tycko, H.M. Cho, E. Schneider, A. Pines, J.<br />

Mag. Res., 61, 90 (1985)<br />

6 M.H. Levitt. Prog, in NMR Spec. 18, 61 (1986)<br />

7 D J Simonivitch. DP Raleigh. E.T. Olejniczak.<br />

and R.G. Griffin../. Chem. Phys.,84, 2556 (1986)<br />

1 S. Wimperis and G. Bodenhausen. J. Mag. Res.,<br />

69.264(1986)<br />

9 N.J. Healon. R.R. Void, and R.L.Vold, J. Mag.<br />

Res..17. 572(1988)<br />

10<br />

DP. Raleigh. E.T. Olejniczak and R.G. Griffin,<br />

J. Mag. Res.,$l, 455 (1989)<br />

11 S. Wimperis../. Mag. Res..83.509 (1989)<br />

12 J. Jeener and P. Broekaert. Phys. Rev. , 157, 232<br />

(1967)<br />

13 S. WimperisJ. Mag. Res.. 86, 46 (1990)<br />

S. Wimperis and G. Bodenhausen. Chem. Phys.<br />

Lett. .132. 194(1986)<br />

15<br />

G. Hoatson.J. Mag. Res.. 94. 152-159 (1991)<br />

16<br />

S.J. Glaser and G.P. Drobny. "Advances in<br />

Magnetic Resonance" (W.S. Warren. ED) Vol 14.<br />

Academic Press. San Diego. 1990<br />

17 H. Liu. S.J. Glaser. G.P. Drobny. J. Chem. Phys..<br />

93.7543.(1990)<br />

18 B. Ewing. S.J. Glaser and G.P. Drobny, Chem.<br />

Phys. Lett. J47. 121 (1990)<br />

19 A.J Vega and Z. Luz. ./. Chem. Phys.. 86. 1803-<br />

1813(1987)<br />

Bulletin of Magnetic Resonance<br />

20 R.R. Void and G. Bodenhausen. J. Mag. Res..<br />

39, 363 (1980)<br />

21 G. Hoatson. T. Tse, R.L. Void. J. Mag. Res., 98.<br />

342 (1992)<br />

zl J. Callahan. D. Mattiello and Gary Drobny,<br />

<strong>ISMAR</strong> conference, Vancouver , B.C., July 1992,<br />

Poster 54


Vol. 14, No. 1-4 259<br />

1 Introduction<br />

A BASIC Program to Calculate<br />

the Evolution of Cartesian Product Operators<br />

Stefano Mammi<br />

Biopolymer Research Center, National Research Council<br />

Via Marzolo 1,35131 Padova, Italy<br />

The introduction of the product operator<br />

formalism [1] has greatly improved the<br />

description of multiple-pulse NMR<br />

experiments allowing the understanding of the<br />

fate of the magnetization in a direct manner.<br />

The use of this formalism has become<br />

widespread [2] because of its advantages ower<br />

both the complete density matrix approach<br />

and the method of vector diagrams. It is<br />

much simpler than the former and it permits a<br />

clear description of the results while<br />

remaining rigorous. With respect to the<br />

latter, it allows one to visualize all the states<br />

of the magnetization and to follow their<br />

evolution over very complicated pulse<br />

sequences.<br />

The rules that govern the evolution of<br />

product operators are very simple and lend<br />

themselves to automation by means of<br />

computer programs. Automation is especially<br />

desirable considering that the description of<br />

any two-dimensional experiment leads quickly<br />

to very long expressions which can be affected<br />

by trivial mistakes.<br />

Recently, computer programs that perform<br />

such calculations have been reported in the<br />

literature. Among these are the program by<br />

Nakashima and McClung [3] and the more<br />

recent one by Shriver [4]. The first was<br />

written in FORTRAN 77 and describes the<br />

evolution of product operators in the<br />

spherical basis [3]. While this approach is<br />

extremely useful in following coherence<br />

pathways and thus deriving phase cycling<br />

schemes, many times the use of the cartesian<br />

basis is preferable as in the development of<br />

new pulse sequences.<br />

The second program [4] was written in the<br />

new computer language Mathematica whose<br />

major advantage is the easy simplification of<br />

algebraic expressions. This program was<br />

written for Macintosh systems and is not yet<br />

available for IBM personal computers. The<br />

input seems to be stepwise and rather<br />

cumbersome.<br />

The program presented here runs within<br />

MS-DOS and describes the evolution of<br />

cartesian product operators. The program,<br />

named "EVOLVE", was written using the<br />

QuickBasic (C) 4.50 language. The input is<br />

from a file which contains all the information<br />

pertaining to the spin system and the pulse<br />

sequence, and the output is filed separately.<br />

2 Features of the Program<br />

A sample input file is presented in Fig. 1. In<br />

this example, an HMQC experiment {5] is<br />

applied to a system composed of an X<br />

nucleus and two protons, only one of which is<br />

coupled to the heteronucleus.<br />

The first information in the input file is the<br />

spin system which can be made of up to four<br />

spins, denoted by capital letters. The initial<br />

magnetization is entered next.


260 Bulletin of Magnetic Resonance<br />

mi Spin System<br />

HHX<br />

### Initial Operators CAxis(Sp#)Axis(Sp#) : coeffj<br />

### Terms in sine or cosine must nave exactly 6 characters in parenthesis<br />

z(2): *1<br />

### Coupling Constants (Spin 1,Spin 2,J)<br />

2,3.90<br />

### Sequence (1 Line for Name, N Lines for Pulse Sequence)<br />

HMQC - 2 spins + 1 spin<br />

P(90)H PH1<br />

D1<br />

P(90)X PH2<br />

DO<br />

P(180)H PH3<br />

DO<br />

P(90)X PH4<br />

D1<br />

A0 PH5 DEC(X)<br />

### Phase Cycles (as Brutcer, separated by commas)<br />

PH1,0<br />

PH2,0,1,2,3<br />

PH3.0<br />

PH4.0<br />

PH5,0,3,2,1<br />

### Delays: D#=Num/Oen*J(Sp#,Sp#) (#, Numerator, Denominator, Spin 1, Spin 2)<br />

1,1,2,2,3<br />

### Do you want to skip the Evolution under Chemical Shift?<br />

N<br />

### Print out only the Observable Operators?<br />

Y<br />

Figure 1. Sample input file for an HMQC sequence applied to a (H + HX) spin system.<br />

In listing the operators, the spins are<br />

numbered from one to four to prevent<br />

ambiguities among like spins. Any number of<br />

operators can be listed with appropriate<br />

coefficients as required. It is not necessary to<br />

start with equilibrium magnetization: for<br />

example, the magnetization of any spin can be<br />

neglected. Moreover, it is possible to follow<br />

the evolution of a specific operator, e.g.,<br />

2x(2)z(3), generated after the first Dl in the<br />

sequence of Fig 1, by restricting the initial<br />

input to just that operator with its own<br />

coefficients, e.g., +cos(Q 2*D1), utilizing an<br />

appropriately shortened pulse sequence.<br />

Next, the scalar coupling network is described<br />

in terms of the spins which are coupled and<br />

the relevant coupling constant.<br />

The format for the pulse sequence is very<br />

similar to the Bruker one. This allows for<br />

simple transcription of sequences already in<br />

use and for easy modification of any part of<br />

the phase cycling scheme. Each pulse is<br />

written as a "P" followed by the flip angle in<br />

parenthesis, by the spin(s) to which it is<br />

applied and by the phase cycle to be used.<br />

Only 90° and 180° pulses are currently<br />

accepted. Each delay is written as a "D"<br />

followed by a single digit. The acquisition is<br />

referred to as "AQ" followed by its own phase<br />

cycle. Up to ten different phase cycles can be<br />

utilized each containing up to 128 steps. Each<br />

step is recorded as a number according to<br />

the usual notation: 0= +x; 1 = +y; 2 =<br />

-x; 3 s -y.<br />

Many sequences require decoupling during<br />

specific delays, including the acquisition.


Vol. 14, No. 1-4 261<br />

z(2):<br />

-- 90 (H)+x --><br />

y(2): -1<br />

- D1 --><br />

2x(2)z


262 Bulletin of Magnetic Resonance<br />

EVOLVE will "decouple" any spin if an<br />

appropriate statement is added on the same<br />

line of any delay.<br />

EVOLVE was written specifically for<br />

sequences in which some delays are inversely<br />

proportional to certain scalar coupling<br />

constants as in heteronuclear correlation<br />

experiments. Proper space is provided in the<br />

input file for defining such cases.<br />

The user is then asked to indicate if<br />

evolution under both coupling and chemical<br />

shift should be taken into account or if the<br />

latter should be neglected. Finally, the user<br />

chooses whether all the resulting operators or<br />

just the observable ones should be printed out<br />

after the acquisition step.<br />

The program runs through the pulse<br />

sequence as many times as required by the<br />

phase cycle. If only the observables are<br />

chosen as output, the final intensities of the<br />

signal are written in two separate files, one for<br />

the real part and one for the imaginary part.<br />

At the end of the phase cycle, these files are<br />

read and final simplifications are carried out.<br />

In Fig. 2, a portion of the output obtained<br />

with the input file of Fig. 1 is reported, Le., the<br />

result from the first of the four steps of the<br />

phase cycle and the signal obtained at the end<br />

of the four step cycle. It can be seen that the<br />

term y(2) generated by the first 90° pulse gives<br />

rise only to antiphase terms at the end of the<br />

first Dl, because Dl = 1/2J23.<br />

The signal from the proton not coupled to<br />

the X-nucleus is present in the first<br />

acquisition, but is canceled out at the end of<br />

the four step cycle while the single quantum<br />

terms from the proton coupled to the Xnucleus<br />

add up to generate the final signal.<br />

The BASIC language does not have built-in<br />

routines for the simplification of algebraic<br />

expressions. A sizable portion of the program<br />

is devoted to such routines. Beside the more<br />

trivial expressions containing n/2, EVOLVE<br />

is able to handle terms containing sin(^/4) or<br />

cos(n/4), most commonly encountered in<br />

heteronuclear sequences. The program does<br />

not evaluate these expressions; rather, it<br />

simplifies them according to the rules it<br />

knows. This entails a longer computation<br />

time but with the advantage of eliminating all<br />

numerical coefficients. This compromise was<br />

found satisfactory.<br />

The only terms that would be useful to have<br />

simplified and that the program is unable to<br />

handle at this stage are those including sin(20)<br />

or cos(20) terms, encountered for example<br />

when there is a 180° pulse in the middle of a<br />

delay. This is not a serious limitation of the<br />

BASIC language, especially considering that<br />

even in Mathematica this simplification must<br />

be explicitly requested by the user.<br />

3 Conclusions<br />

EVOLVE has been applied to numerous<br />

complicated pulse sequences avoiding lengthy<br />

and monotonous calculations and providing<br />

the results in a way suitable for<br />

straightforward analysis.<br />

A copy of the program, including the source<br />

code, can be obtained by sending a 5.25" or<br />

3.5" diskette and return postage to the author.<br />

4 References<br />

[1] O. W. Stfrensen, G. W. Eich, M. H.<br />

Levitt, G. Bodenhausen, and R. R. Ernst,<br />

Prog. NMR Spectrosc 16,163 (1983).<br />

[2] See for example H. Kessler, M. Gehrke,<br />

and C. Griesinger, Angew. Chem. Int. Ed.<br />

Engl 27,490 (1988).<br />

[3] T. T. Nakashima and R. E. D. McClung, /.<br />

Magn. Reson. 70,187 (1986).<br />

[4] J. W. Shriver, /. Magn. Reson. 94, 612<br />

(1991).<br />

[5] A. Bax, R. H. Griffey, and B. L. Hawkins<br />

/. Magn. Reson. 55,301 (1983).


Vol. 14, No. 1-4 263<br />

SELECTIVE LONG-RANGE POLARI<br />

ZATION TRANSFER via DEPT.<br />

Introduction<br />

T. Parella, F. Sanchez-Ferrando* and A. Virgili.<br />

Departament de Quimica. Universitat Autonoma de Barcelona,<br />

08193 Bellaterra, Barcelona, Spain.<br />

The structural assignment of organic compounds<br />

containing quaternary carbons and/or heteroatoms is<br />

often greatly helped by measurements of long-range<br />

proton-carbon coupling constants, "J^, with particular<br />

emphasis on two-bond and three-bond couplings which<br />

usually show values in the range 3-10 Hz [1,2], depending<br />

on carbon hybridization, torsion angles, substttuent<br />

electronegativity and orientation, etc.<br />

Selective ID NMR methods, such as the selective<br />

INEPT method proposed by Bax [3], have long been<br />

used to reveal connectivities with a given proton.<br />

Thus, application of a selective INEPT (or, more<br />

frequently, refocussed INEPT) pulse sequence on a<br />

well resolved proton, with delays optimized for a longrange<br />

heteronuclear coupling (usually around 5-7<br />

Hz), results in a ID carbon spectrum displaying large<br />

intensity enhancements at the carbons coupled (at<br />

long range) with the perturbed proton. Furthermore,<br />

the intensity of these carbons shows a maximum when<br />

the true "iCH value is used for the delay optimization<br />

of the selective INEPT sequence. We have recently<br />

shown the use of this dependence to obtain a quick<br />

estimate of "J^, in several polycyclic derivatives [4].<br />

Selective 2D methods, however, are to be<br />

preferred because they can easily yield accurate values<br />

for the desired long-range coupling constants, provided<br />

the perturbed proton appears as a well resolved multiplet<br />

in the proton spectrum. Thus, the selective spin flip<br />

method first proposed by Bax and Freeman [5] has<br />

been widely used for the determination of "J^ values<br />

from a given proton.<br />

In recent years a number of methods have been<br />

suggested which combine the intensity enhancements<br />

obtained from selective polarization transfer with the<br />

easy measurement of "J^ characteristic of the selective<br />

spin flip method. Thus, Jippo et al. proposed a 2D<br />

selective INEPT method [6] which consists in a<br />

conventional 2D INEPT pulse train containing a<br />

selective spin echo sequence. We have modified this<br />

method by delivering all decoupler pulses in the<br />

selective mode [4], and in this way we have suppressed<br />

the artefact peaks otherwise observed with this method.<br />

Another combination of polarization transfer<br />

and selective spin flip was proposed by Uhrin et at. [7].<br />

In this method, a standard non-selective DEPT<br />

preparation period (optimized for I JCH) is followed by<br />

a conventional selective (or semiselective) spin flip<br />

sequence, yielding a J-resolved 2D spectrum displaying<br />

the desired long range couplings. More recently,<br />

Poppe and van Halbeek [8] have introduced inverse<br />

detection, either in ID or 2D methods, by combining<br />

selective polarization transfer with long range<br />

heteronuclear coupling measurements in lH-detected<br />

sequences.<br />

We now report two new u C-detected, 2D Jresolved<br />

sequences, based on a DEPT pulse train,<br />

which allow fast and accurate measurements of longrange<br />

heteronuclear coupling constants from well<br />

resolved protons. Both sequences are particularly<br />

useful for the determination of couplings to quaternary<br />

carbons, and their main feature is that all proton<br />

pulses are delivered as soft, selective pulses (20-30<br />

ms).


264 .<br />

Parametrization<br />

Starting from the 1D-SDEPT sequence (Scheme<br />

1), we have derived the new sequences 2D-SDEPT1<br />

(Scheme 2) and 2D-SDEPT2 (Scheme 3), in which<br />

SDEPT stands for Selective DEPT. Since ail proton<br />

pulsesareselective (yBjln=20-30 Hz), both sequences<br />

achieve a selective polarization transfer from the<br />

pulsed proton to long-range coupled carbons, provided<br />

that the fixed interpulse delay A is optimized for longrange<br />

couplings. Both sequences yield comparable<br />

results in similar experiment times.<br />

scheme 1<br />

scheme 2<br />

90° 180<br />

scheme 3<br />

180° 4>° 180°<br />

i i i i<br />

i A i A < Evolution •<br />

90" 180°<br />

180° 90° 180°<br />

In both sequences the incrementable evolution<br />

period t,, which brackets a selective spin echo moiety,<br />

allows only the evolution of heteronuclear long range<br />

couplings which will therefore be detected in Fl<br />

dimension in the final 2D J-resolved spectrum.<br />

Bulletin of Magnetic Resonance<br />

We have studied the dependence of carbon<br />

signal intensities on several experimental parameters,<br />

such as the delay A or the pulse angle of the last<br />

selective proton pulse in the DEPT part of the sequence,<br />

using camphor (Fig. 1) as a readily available, rigid<br />

model compound. These determinations have been<br />

most conveniently carried out using the ID selective<br />

DEPT sequence 1D-SDEPT.<br />

Fig.l<br />

The extent of polarization transfer is a function<br />

of both, the interpulse delay A and the pulse angle $.<br />

Fig. 2 shows the intensity of the quaternary C-l carbon<br />

signal of camphor when pulsing the H-4 methine<br />

proton, for 4>=45° and for =90°, as a function of<br />

interpulse delay A, using sequence 1D-SDEPT. In this<br />

case, the heteronuclear coupling constant involved is<br />

3 JH4^,= 4.4 Hz. As expected, the quaternary carbon C-<br />

1 intensity follows a typical sinus function, and maximum<br />

signal is obtained for (j>=90° and A=50-90 ms, a range<br />

which does not contain the theoretical value, (2* 3 JH4^<br />

C1)'—115 ms. Instead, this range of maximum signal is<br />

centered around the practical value (3* 3 JH4^1) l =75<br />

ms.<br />

1000-<br />

800-<br />

~ 600<br />

f<br />

£ 400<br />

200<br />

0<br />

0.02<br />

-i 1 « 1 1 1—r<br />

0.04 0.06 0.08 0.1<br />

Fig. 2<br />

Delay (s)


Vol. 14, No. 1-4<br />

However, when pulsing on a methyl proton the<br />

optimum delay has to be modified. Fig. 3 shows the<br />

variation of the quaternary O7 carbon signal of camphor<br />

when pulsing the methyl H-10 protons as a function of<br />

the pulse angle, for three A values, using also sequence<br />

1D-SDEPT. In this case, the heteronuclear coupling<br />

constant involved is 2 JH1(>C7=4.1 Hz (sign not determined).<br />

If the optimization is carried out as in a conventional<br />

DEPT, using A=(2J)S the intensities follow the typical<br />

sin*cos 2


266<br />

The results (Fig. 7) easily allow the assignment<br />

of the four quaternary carbons coupled to the H-5<br />

proton. The particular example shown was obtained<br />

using sequence 2D-SDEPT1, but the alternative sequence<br />

2D-SDEPT2 gave the same results. We also performed<br />

a comparison between 2D-SDEPT1 and our previous<br />

[4] sequence 2D-SINEPT. Both methods can yield the<br />

desired long-range heteronuclear coupling constants<br />

in less than one hour of accumulation (using a 400<br />

M Hz spectrometer), with excellent sensitivity even for<br />

quaternary carbons. However, the SDEPT method is<br />

to be preferred, because it is less sensitive to mismatch<br />

between the delay A and the long range coupling "J^.<br />

b)<br />

d)<br />

6'7'<br />

5'6 4' 87a' 7 4a 3a' 8a 5<br />

6 8 7<br />

'.58 128 1!!<br />

Fig. 7<br />

a) Decoupled "C spectrum; b) 1D-SDEPT spectrum after pulsing on<br />

H-5. Only long-range coupled carbons with it are presents. Then, The<br />

2D- J spectra (d) and its internal projection (c) obtained with the pulse<br />

sequence of the scheme 2.<br />

Acknowledgements.<br />

Financial support from DGICYT through project<br />

number PB89-0304 is gratefully acknowledged. We<br />

also thank the Servei de Ressonancia Magnetica Nuclear,<br />

UAB, for allocating instrument time to this project. A<br />

grant (to T.P.) from Universitat Autonoma de Barcelona<br />

is gratefully acknowledged.<br />

References<br />

Bulletin of Magnetic Resonance<br />

1.- P.E. Hansen, Progress in NMR Spectroscopy, 1980,<br />

14, 175.<br />

2.- J.L. Marshall, "Carbon-Carbon and Carbon-Proton<br />

NMR Couplings. Application to Organic<br />

Stereochemistry and Conformational Analysis", VCH<br />

Publishers, Deerfield Beach, Florida, 1982.<br />

3.- A. Bax, J. Magn. Reson., 1984, 57,314.<br />

4.- T. Parella, F. Sanchez-Ferrando and A. Virgili,<br />

Magn. Reson. Chem., 1992, in press.<br />

5.- A. Bax and R. Freeman, J. Am. Chem. Soc, 1982,<br />

104, 1099.<br />

6.- T. Jippo, O. Kamo and K. Nagayama, J. Magn.<br />

Reson., 1986,66, 344.<br />

7.- D. Uhrin, T. Liptaj, M. Hricovini and P. Capek, J.<br />

Magn. Reson., 1989,85,137.<br />

8.- L. Poppe and H. van Halbeek, J. Magn. Reson.,<br />

1991,92,636.


Vol. 14, No. 1-4 267<br />

Computer Simulations of High Resolution NMR Spectra<br />

1 Introduction<br />

Scott A. Smith, William E. Palke, and J. T. Gerig<br />

Department of Chemistry, University of California<br />

The evolution of a spin system during the application<br />

of an RF pulse is a central aspect of high<br />

resolution NMR spectroscopy. An appropriately<br />

applied field can lead to saturation or decoupling<br />

effects. When a field is used to maintain a spinlocked<br />

condition, both scalar and dipolar interactions<br />

come into play, leading to coherence transfers<br />

and nuclear Overhauser effects that form the<br />

basis of TOCSY, ROESY and related experiments.<br />

These experiments have important roles<br />

in structural studies of biological macromolecules<br />

in solution . The accompanying bad<br />

news is that the interpretation of these experiments<br />

is tricky; computer simulations can be<br />

very helpful in assisting the analysis of these<br />

spectra.<br />

The theoretical formalism of relaxation in<br />

the presence of an RF field can be written elegantly<br />

in superoperator notation , or in a more<br />

conventional notation laden with superscripts<br />

and subscripts. The latter approach displays<br />

more details and shows how the presence of the<br />

RF field introduces new combinations of frequencies<br />

into the relaxation expressions. Of<br />

course, it can be shown that both methods generate<br />

identical results. Details of the theoretical<br />

developments formulated in this lab are given in<br />

reference 7 and in a manuscript in preparation.<br />

We have used our theoretical results to<br />

extend the program GAMMA 8 to include the<br />

effects of relaxation in the presence of an RF<br />

field and present here the results of several calculations<br />

of double resonance or rotating frame<br />

experiments that illustrate the capabilities<br />

Santa Barbara, CA 93106, USA<br />

thereby developed. For the examples presented,<br />

only dipole-dipole relaxation was included, and<br />

isotropic diffusional tumbling was assumed with<br />

a correlation time of 1.0 nsec. In all cases, the<br />

spectrometer proton frequency was 500 MHz.<br />

2 Applications<br />

Presaturation - The first example is a presaturation<br />

experiment on a three spin system whose<br />

parameters were chosen to mimic three protons<br />

in the trans conformation of a glycine residue.<br />

Details are given in Table 1.<br />

Table 1: Spin System 1 Parameters<br />

R12 = 1.75A<br />

R13 = 3.00A<br />

R23 = 2.53A<br />

J12 = -16.3Hz<br />

J13=4.0Hz<br />

J23 = 12.0 Hz<br />

v = -700 Hz<br />

t> = -950Hz<br />

v = 950 Hz<br />

The pulse sequence consisted of a pre-irradiation<br />

period followed by an analyzing 90° pulse, as<br />

shown in Figure 1.<br />

Pre-saturation<br />

Figure 1. Pre-saturation Pulse Sequence<br />

The initial pulse length was chosen to be long<br />

enough to establish a steady state, and the RF frequency<br />

chosen to match the chemical shift of<br />

proton 2. The resulting spectra for several RF<br />

field strengths are shown in Figure 2. Notice the


268 Bulletin of Magnetic Resonance<br />

750<br />

960 950 940<br />

V3<br />

-690-700-710 -940-950-960<br />

V2<br />

Figure 2. Simulation of Presaturation. Shading is<br />

used to highlight NOE enhanced intensities.<br />

Units on both axes are Hz.<br />

non-symmetrical effects on spins 1 and 3.<br />

Decoupling - The second example illustrates the<br />

decoupling of a two spin proton system with a<br />

chemical shift difference of 1000 Hz and a spin<br />

coupling constant of 10 Hz. The protons were<br />

2A apart. After an initial 90° pulse, an RF field<br />

of a specified magnitude was applied during the<br />

collection of the FID. The pulse sequence is<br />

shown in Figure 3. Some spectra are shown in<br />

Figure 4. Notice the slight oscillation in the peak<br />

height as a function of the strength of the decoupling<br />

field.<br />

Figure 3. Decoupling Pulse Sequence<br />

ROESY/TOCSY Simulations - Two-dimensional<br />

rotating frame experiments may produce coherence<br />

transfer (TOCSY) and spin-spin cross<br />

relaxation (ROESY) effects simultaneously, and<br />

1125 1100<br />

v (Hertz)<br />

1075<br />

(Hertz)<br />

Figure 4. Simulation of Decoupling. Spectra at the<br />

first and last field strengths are shown to<br />

the upper left.<br />

ambiguities of interpretation arise when bothkinds<br />

of effects are present. Griesinger et al.<br />

have described a strategy for suppressing<br />

ROESY effects when doing TOCSY experiments<br />

9 . The interpretation of ROESY experiments<br />

is complicated because coherence transfer<br />

(COSY) and indirect interactions (HOHAHA)<br />

can lead to spectral effects that interfere with or<br />

imitate rotating frame Overhauser features 10 .<br />

Methods for suppressing these unwanted effects<br />

have been proposed , and it has been suggested<br />

that use of a train of short RF pulses to generate<br />

an effective spin-locking field leads to diminution<br />

of COSY-type cross peaks .<br />

Here we simulate several variations of<br />

ROESY experiments. First, consider the pulse<br />

sequence shown in Figure 5 in which a continuous<br />

spin-locking pulse is applied.<br />

JC/2 Spin Lock<br />

Figure 5. CW ROESY Pulse Sequence<br />

We apply this to an equilateral triangle of protons<br />

as described in Table 2. The 2D spectra that


Vol. 14, No. 1-4 269<br />

1<br />

1<br />

C-l. 0 sec<br />

M<br />

• 1 11<br />

150 100 50 0 -50 -100 -150 150 100 50 0 -50 -100 -150<br />

150 100 0 -50 -100 -150<br />

Figure 6. 2D ROESY spectra for differing spin-lock<br />

times.<br />

result from several different spin-lock times are<br />

shown in Figure 6. A point to notice here is that<br />

the ROESY peaks grow more slowly than the<br />

Table 2: Spin System 2 Parameters<br />

Rl2=V3A<br />

Ri3=V3A<br />

R23=V3A<br />

J12 = 0<br />

Jl3 = 0<br />

J23 = 4.0 Hz<br />

v = 140 Hz<br />

v = -90 Hz<br />

?? = -140Hz<br />

D -1.1 sec<br />

150 100 50 0 -50 -100 -150<br />

©<br />

.©<br />

-o<br />

o<br />

• o<br />

©<br />

1 V)<br />

-o<br />

COSY peaks, reaching their maximum intensity<br />

well after the latter have begun to subside.The-<br />

ROESY peaks also persist much longer and<br />

decay more smoothly than both the COSY and<br />

diagonal peaks for protons 2 and 3. The plots for<br />

1.0 and 1.1 sec. spin-lock times show oscillation<br />

in both the latter features while the ROESY<br />

peaks decay smoothly.<br />

Interesting contrasts to these results arise in<br />

changing to an isosceles triangle configuration.<br />

The spin system is given in Table 3. In this<br />

geometry, the ratio of the 1-3 distance to the 1-2<br />

©


270 Bulletin of Magnetic Resonance<br />

distance is \2. Thus, the direct 1-2 dipole-dipole<br />

interaction is 8 times stronger than the corresponding<br />

1-3 interaction. Keeping a fixed<br />

Table 3: Spin System 3 Parameters<br />

Ri2=V3A<br />

Ri3=V6A<br />

R23=V3A<br />

J12 = 0<br />

J13 = 0<br />

J23 = 0 -10 Hz<br />

v = 140 Hz<br />

v = -90 Hz<br />

v = -U0Uz<br />

geometry and a spin-lock time of 0.1 sec, the 2-<br />

3 spin-spin coupling is varied from 0 to 10 Hz as<br />

we repeat the pulse sequence of Figure 5. Slices<br />

through the resulting 2D spectra at 140 Hz are<br />

shown in Figure 7.<br />

J = 10 Hz<br />

J = 8Hz<br />

= 6Hz<br />

J = 4Hz<br />

J = 2Hz<br />

= 0Hz<br />

J)LJL_<br />

150 100 50 0 -50 -100 -150<br />

Figure 7. The effect of J23 variation on ROESY/<br />

TOCSY experiment.<br />

Indeed, the simulation with J23 = 0 shows an 8:1<br />

ratio of NOE peak intensities. However, variation<br />

of J23 generates a ROESY-like peak in the<br />

location of the 1-3 interaction that becomes even<br />

larger than the true 1-2 ROESY peak for some<br />

values of J23- This situation provides an excellent<br />

example of a spectrum that cannot directly<br />

yield reliable information about internuclear distances.<br />

In the next example, the frequency of the<br />

spin-locking RF field is varied. In this simulation<br />

spin system of Table 3 is used with J23 = 4 Hz. It<br />

is subjected to a 1.0 sec. spin-locking pulse as<br />

depicted in the ROESY pulse sequence in Figure<br />

5. The frequency of the RF field is indicated by<br />

the arrow on each of the stacked plots in Figure<br />

8. While the 1-2 ROESY peak is relatively unaffected<br />

by the frequency of the RF field, the 1-3<br />

peaks which are generated by indirect<br />

HOHAHA effects show extreme sensitivity both<br />

in magnitude and sign. As has been suggested in<br />

the literature, several experiments should be run<br />

with different RF frequencies to discriminate<br />

between these interactions.<br />

\<br />

1<br />

150 100 50 0 -50 -100 -150<br />

Figure 8. The effect of varying the applied RF frequency<br />

on ROESY/TOCSY experiment.<br />

i<br />

1 1<br />

As a final demonstration, we investigate the<br />

suggestion that dividing the spin-locking period<br />

into a sequence of numerous pulse-delay steps<br />

can emphasize the ROESY interactions. Again,<br />

the three spin proton system in Table 3 is used<br />

with J23 = 4 Hz. It was subjected to the ROESY<br />

pulse sequence utilizing a pulse train as shown<br />

in Figure 9. The length of the pulse and delay<br />

steps were varied keeping the average yBj con-<br />

I


Vol. 14, No. 1-4 271<br />

Figure 9. ROESY Pulse Sequence with pulse train<br />

150 100 50 0 -50 -100 -150<br />

Figure 10. Simulation of ROESY/TOCSY experiment<br />

using a spin-locking pulse train.<br />

o<br />

.©<br />

o<br />

-o<br />

150 100 50 0 -50 -100 -150<br />

Figure 11. Simulation of ROESY/TOCSY using continuous<br />

irradiation for spin-locking.<br />

o<br />

•o<br />

stant at 2000 Hz by adjusting the number of<br />

pulses per second. This simulation has been performed<br />

with pulse lengths that correspond to<br />

individual rotations of 180°, 30°, and 10°, and<br />

also with CW irradiation. The contour plot using<br />

30° pulses is presented in Figure 10. For comparison,<br />

the plot which was simulated using continuous<br />

irradiation during the spin lock is shown<br />

in Figure 11. Cross sections of these contour<br />

plots taken at -90 Hz are shown in Figure 12.<br />

= 10deg<br />

= 180deg<br />

150 100 50 0 -50 -100 -150<br />

Figure 12. Cross sections of ROESY/TOCSY simulations<br />

using a spin-locking pulse train.<br />

Shortening the pulse length indeed suppress the<br />

COSY peaks, but they remain even for the shortest<br />

pulses. Also, the limit of the sequence of<br />

shorter and more frequent pulses gives the same<br />

result as continuous irradiation. These two processes<br />

appear to become the same once the delay<br />

between pulses is sufficiently short that a negli-


272 Bulletin of Magnetic Resonance<br />

gible precession occurs before the next pulse.<br />

3 Conclusions<br />

It is clear that full computer simulation may be<br />

essential for the correct interpretation of ROESY<br />

experiments. Our initial results indicate that<br />

there may be additional information about structure<br />

and dynamics that can be extracted from<br />

classical one-dimensional multiple irradiation<br />

experiments with the aid of simulations. An<br />

appreciable expansion of this program in terms<br />

of the size of spin systems that can be handled is<br />

probably needed to make such applications practical;<br />

this along with other expansions of the program's<br />

capabilities is planned.<br />

4 References<br />

1 - L. Braunschweiler and R. R. Ernst, J. Magn.<br />

Reson. 53, 521 (1983).<br />

2 - A. Bax and D. G. Davis, /. Magn. Reson.<br />

63,207 (1985).<br />

3 - A. A. Bothner-By and R. Shukla, /. Magn.<br />

Reson. 77, 524 (1988).<br />

4 - M. Ranee and J. Cavanaugh, J. Magn.<br />

Reson. 87, 363 (1990).<br />

5 - S. J. Glaser and G. P. Drobny, Advan. Magn.<br />

Reson. 14, 35 (1990).<br />

6 - R. R. Ernst, G. Bodenhausen, and A. Wokaun,<br />

"Principles of Nuclear Magnetic Resonance<br />

in One and Two Dimensions",<br />

Clarendon Press, Oxford (1987); J. Jeneer,<br />

Advan. Magn. Reson. 10, 2 (1982); M.<br />

Ravikumar, R. Shukla, and A. A. Bothner-<br />

By, J. Chem, Phys. 95, 3092 (1991).<br />

7 - S. A. Smith, W. E. Palke, and J. T. Gerig, /.<br />

Magn. Reson. in press (1992).<br />

8- S.A. Smith, T. Levante, B.H. Meier, and<br />

R.R. Ernst, manuscript in preparation.<br />

9 - C. Griesinger, G. Otting, K. Wiithrich, and<br />

R.R. Ernst, J. Am. Chem Soc. 110, 7870<br />

(1988).<br />

10-D. Neuhaus and M. P. Williamson, "The<br />

Nuclear Overhauser Effect in Structural and<br />

Conformational Analysis". VCH, New York<br />

(1989) p. 312-327.<br />

11 - H. Kessler, C. Griesinger, R. Kessebaum, K.<br />

Wagner, and R.R. Ernst, J. Am. Chem Soc.<br />

109,607-609(1987).


Vol. 14, No. 1-4 273<br />

1. Introduction<br />

Variation of 13 C NMR Linewidths of<br />

Metallocenes as a Function of Magic<br />

Angle Sample Spinning Frequency<br />

In this paper we report high-resolution solid state 13 C<br />

NMR investigations of various metallocenes<br />

[(Ti5-C5H5)M(Ti5-C5H5); M = Fe, Ru, Ni], carried out as<br />

a function of magic angle sample spinning (MAS)<br />

frequency. Specifically, we focus on how the linewidth of<br />

the isotropic peak varies with MAS frequency at fixed<br />

temperature. Unexpectedly, it has been found that the<br />

linewidth increases as the MAS frequency is increased, and<br />

it is demonstrated that the underlying reason is that the *H<br />

decoupling becomes less efficient as the MAS frequency is<br />

increased. It is suggested that molecular motion within<br />

these solids is an important factor underlying this<br />

phenomenon.<br />

It is well known that, at room temperature, there is<br />

substantial molecular motion in crystalline metallocenes<br />

[1-3]. In ferrocene, for example, it has been shown [4-6]<br />

that there is rapid reorientation of the cyclopentadienyl<br />

(C5H5) rings via a five-fold jump mechanism, with<br />

correlation time xc = 5 X 10~ 12 s at 293 K. The<br />

correlation time for ring reorientation in nickelocene at<br />

room temperature is essentially the same as that for<br />

ferrocene, whereas the ring reorientation in ruthenocene is<br />

associated with a significantly longer correlation time (tc ~<br />

5 X 10- 10 s at 293 K) [4,5]. It has been suggested [2]<br />

that there may be some additional slower molecular<br />

motions in crystalline ferrocene at ambient temperature.<br />

All experiments reported in this paper were carried out at<br />

constant temperature, and hence the correlation time for<br />

molecular motion for a given metallocene can be assumed<br />

to be constant for the series of experiments discussed here.<br />

Before presenting our results, we discuss briefly relevant<br />

aspects of magic angle sample spinning (MAS) and high<br />

power *H decoupling in relation to the measurement of<br />

high-resolution 13 C NMR spectra for organic solids, with<br />

Ian J. Shannon, Kenneth D.M. Harris*, s* S. Arumugam<br />

Department of Chemistry<br />

University of St. Andrews<br />

St. Andrews<br />

Fife KY16 9ST<br />

Scotland<br />

particular emphasis on the application of the technique to<br />

systems in which there is substantial molecular motion.<br />

2. Theory<br />

* Author to whom all correspondence should be addressed.<br />

In solid state 13 C NMR of organic materials, the two<br />

major sources of line-broadening are chemical shift<br />

anisotropy (CSA) and direct ^C- 1 !! dipole-dipole<br />

interaction. Magic angle sample spinning (MAS) [7,8]<br />

will transform an NMR line broadened by these effects into<br />

a set of comparatively narrow, equally-spaced lines<br />

comprising an isotropic peak and spinning sidebands. The<br />

spacing between adjacent lines in this set is equal to the<br />

MAS frequency vr. The chemical shift of the isotropic<br />

peak is independent of vr, and only this line remains when<br />

the anisotropic interactions have been averaged completely<br />

(i.e. at sufficiently large vr).<br />

As discussed fully elsewhere [7-9], there is an important<br />

difference concerning the way in which CSA and direct<br />

dipole-dipole interaction are affected by MAS. For an<br />

NMR line broadened by CSA, relatively slow MAS is<br />

generally sufficient to transform this line into the set of<br />

narrow, equally-spaced lines discussed above, whereas a<br />

line broadened by direct dipole-dipole interaction will be<br />

narrowed significantly only if vr is in the region of (or<br />

greater than) the magnitude of the dipole-dipole interaction.<br />

The underlying reason for this difference [9] is that CSA<br />

gives rise to inhomogeneous broadening of the spectral<br />

line, whereas dipole-dipole interaction is a source of<br />

homogeneous broadening; the value of vr required to<br />

achieve effective line-narrowing is larger (relative to the<br />

linewidth of the static sample) in the case of homogeneous<br />

broadening. For most organic solids with natural isotopic<br />

abundance, the only appreciable dipole-dipole interaction<br />

directly affecting the 13 C NMR spectrum is that between


274<br />

13 C and 1 H. Since the magnitude of this interaction<br />

(typically ca. 30 kHz for rigid organic solids) is generally<br />

much larger than the MAS frequencies that can be obtained<br />

on conventional instruments, substantial averaging of this<br />

interaction cannot be achieved using MAS. For this<br />

reason, high power *H decoupling is generally applied (in<br />

addition to MAS) during acquisition of the 13 C spectrum<br />

in order to eliminate line-broadening due to direct ^C-^H<br />

dipole-dipole interaction.<br />

In 13 C NMR spectroscopy of many crystalline<br />

metallocenes (such as ferrocene and ruthenocene), GSA and<br />

direct ^C- l U dipole-dipole interaction are the important<br />

sources of line-broadening.<br />

A detailed publication [10] has considered, from both<br />

theoretical and experimental standpoints, the linewidth of<br />

the isotropic peak, recorded under MAS conditions, for a<br />

system that is subject to line-broadening by CSA. It was<br />

also shown that, at fixed temperature, the linewidth of the<br />

isotropic peak decreases as the' MAS frequency (vr) is<br />

increased, approaching a limiting value at sufficiently large<br />

vr. Another paper [11] has considered the linewidth of a<br />

spin system S, dipolar coupled to an unlike spin system /,<br />

under conditions of isotropic molecular motion and<br />

decoupling of the / spins. Using coi to denote the<br />

decoupler field strength, it was shown that, in the limit of<br />

long correlation time (i.e. (0itc » 1) the linewidth is<br />

proportional to (o>i)~ 2 , whereas in the limit of short<br />

correlation time (i.e. coixc « 1) the linewidth is<br />

independent of ca i. Thus, for a sample at fixed<br />

temperature (and hence fixed xc), the linewidth should<br />

either decrease as the decoupler field strength is increased<br />

(long correlation limit) or remain independent of the<br />

decoupler field strength (short correlation limit). The<br />

effects of anisotropic molecular motion on the measured<br />

spectrum were also discussed briefly in ref. 11.<br />

In the studies discussed in this paper, both MAS<br />

frequency and *H decoupler field strength are important in<br />

controlling the linewidth of the isotropic peak in the 13 C<br />

NMR spectrum. If the separate effects discussed above can<br />

be combined in a simple way, then it should be expected<br />

that: (a) at fixed temperature and fixed decoupler field<br />

strength, the linewidth should decrease with increasing<br />

MAS frequency (up to a limiting value, beyond which the<br />

linewidth should be effectively independent of the MAS<br />

frequency); and (b) at fixed temperature and fixed MAS<br />

frequency, the linewidth should either decrease or remain<br />

constant as the decoupler field strength is increased,<br />

depending on the motional regime (i.e. long or short<br />

correlation limit) of the sample at the temperature of<br />

interest.<br />

It is shown here that, for ferrocene and ruthenocene at<br />

room temperature, the effects of MAS frequency and *H<br />

decoupler field strength on the 13 C NMR linewidth cannot<br />

be combined in this simple way, since the effective<br />

decoupler field strength is modulated by altering the MAS<br />

frequency. Nickelocene is paramagnetic, and for this<br />

Bulletin of Magnetic Resonance<br />

reason it is not valid to consider nickelocene in the same<br />

way as ferrocene and ruthenocene in relation to the NMR<br />

properties discussed here.<br />

3. Experimental<br />

13 C NMR spectra were recorded at 125.758 MHz on a<br />

Bruker MSL500 spectrometer using a Bruker doublebearing<br />

magic angle spinning probe capable of MAS<br />

frequencies between ca. 1 kHz and 12 kHz with stability<br />

better than ca. ± 10 Hz. All spectra were recorded at room<br />

temperature (293 ± 2 K) with the samples contained in<br />

zirconia rotors (4 mm external diameter).<br />

The 13 C "single pulse" sequence was used to record the<br />

spectra, with high power *H decoupling applied during<br />

acquisition. Typical parameters were: 13 C 90° pulse<br />

length = 3.5 \xs; recycle delay = 10 s for ferrocene and 20 s<br />

for ruthenocene.<br />

The *H decoupler field was set on resonance for the *H<br />

of each metallocene. An accurate assessment of the<br />

decoupler field strength was made by measuring (for<br />

adamantane) the length of the *H 90° pulse (t9o( 1 H)) for<br />

the *H r.f. power level used in the experiments involving<br />

*H decoupling. For the discussion of results, tgo^H) has<br />

been converted to a decoupling frequency vi via:<br />

1<br />

vi is related to the decoupler field strength Hi and to the<br />

parameter coi used in ref. 11 by the equations:<br />

_ Y( 1 H) Hi _ ©!_<br />

L 71 Z %<br />

The linewidth of the isotropic peak in the 13 C NMR<br />

spectrum was measured as the full width at half maximum<br />

height, and the experimental error in the measured<br />

linewidth is estimated to be less than ca. ± 5 Hz.<br />

4. Results and Discussion<br />

Initially we present the main results for ferrocene [12], and<br />

then consider the other metallocenes studied. Highresolution<br />

13 C NMR spectra of ferrocene were recorded<br />

initially at two different decoupler field strengths (vi =<br />

64.9 kHz and 26.6 kHz), and at several MAS frequencies<br />

(vr) ranging from ca. 1 kHz to 11 kHz. The spectrum at<br />

vr


Vol. 14, No. 1-4<br />

increases as vr is increased. The relationship between A<br />

and vr is approximately linear, particularly at the higher<br />

decoupler field strength, and the gradient is greater at the<br />

lower decoupler field strength. At fixed vr, the linewidth<br />

A is smaller at higher decoupler field strength, as shown in<br />

Fig. 1.<br />

A/Hz<br />

400<br />

300<br />

a n<br />

vr / kHz<br />

Fig. 1 Linewidth (A) versus MAS frequency (vr) for the<br />

isotropic peak in the 13 C NMR spectrum of ferrocene.<br />

The spectra were recorded at 'H decoupler field<br />

strengths corresponding to Vi = 64.9 kHz (•) and Vj =<br />

26.6 kHz (A).<br />

Our observation that the linewidth of the isotropic peak<br />

increases as the MAS frequency is increased conflicts with<br />

the discussion in Section 2 that, under conventional<br />

conditions, A should decrease, or remain constant, as vr is<br />

increased. We propose that the observed increase in A<br />

with increasing vr for ferrocene arises as a result of MAS<br />

indirectly modulating the efficiency of the *H decoupling<br />

in such a way that the effective decoupler field strength is<br />

decreased, leading to line-broadening, as vr is increased.<br />

This conclusion is supported by the results of three further<br />

experiments.<br />

(1) l^C NMR spectra of ferrocene were recorded for a<br />

series of different decoupler field strengths (with vi in the<br />

range 20 kHz to 80 kHz) at fixed vr. As expected, A<br />

decreases as vi is increased at fixed vr (Fig. 2).<br />

Furthermore, in the limit of sufficiently high decoupler<br />

field strength, A becomes essentially independent of both<br />

vr and vi, and converges to a limiting value of ca. 100 Hz.<br />

From this it can be concluded that both values of vr (5.06<br />

kHz and 9.05 kHz) used to record the data shown in Fig. 2<br />

are sufficiently rapid to remove any Vr-dependent sources<br />

of line-broadening due to CSA; i.e. at sufficiently high<br />

decoupler field strength, A is essentially independent of vr<br />

at these values of vr. This is consistent with the view<br />

that, under the conditions of the experiments shown in Fig.<br />

1, the property that does depend on vr is the efficiency of<br />

the *H decoupling (and not the ability of MAS to remove<br />

275<br />

the line-broadening effects due to CSA). As discussed<br />

fully elsewhere [12], there is a linear relationship between<br />

A and (vi)~ 2 .<br />

5001<br />

400<br />

A/Hz 300-<br />

200<br />

o o<br />

20 SO 70 80<br />

/ kHz<br />

Fig. 2 Linewidth (A) versus Vj for the isotropic peak in the<br />

13 C NMR spectrum of ferrocene. The spectra were<br />

recorded at MAS frequencies vr = 5.06 kHz (O) and vr =<br />

9.05 kHz (•).<br />

(2)<br />

13 C NMR spectra of ferrocene were recorded as a<br />

function of vr, but with no *H decoupling field applied;<br />

the variation of A with vr in these experiments is shown<br />

in Fig. 3. Under these conditions, A decreases as vr is<br />

increased, as predicted from the discussion in Section 1 and<br />

from ref. 10. This relationship between A and vr reflects<br />

the direct effect of MAS on the linewidth of the isotropic<br />

peak for a system that is subject to line-broadening by<br />

CSA and by direct 13 C-*H dipole-dipole interaction. At<br />

the lowest value of vr studied (ca. 3 kHz), A is ca. 592 Hz,<br />

and it is clear that slow MAS alone can substantially<br />

average the dipole-dipole interaction (as well as CSA) in<br />

A/Hz<br />

6001<br />

400"<br />

300<br />

vr / kHz<br />

Fig. 3 Linewidth (A) versus MAS frequency (vr) for the<br />

isotropic peak in the 13 C NMR spectrum of ferrocene,<br />

with no 'H decoupling applied.<br />

12


276<br />

this system. Undoubtedly, this is a consequence of the<br />

fact that the direct ^C- 1 !! dipole-dipole interaction is<br />

already extensively averaged by molecular motion.<br />

(3)<br />

13 C NMR spectra were recorded (using the 13 C<br />

"single pulse" method with no *H decoupling field applied)<br />

for perdeuterated ferrocene (denoted ferrocene-dio) as a<br />

function of MAS frequency. Over the range vr« 1 kHz to<br />

12 kHz, the linewidth of the isotropic peak is independent<br />

of vr (Fig. 4), and the value (A = 117 Hz) is close to the<br />

limiting linewidth obtained in our experiments for<br />

undeuterated ferrocene. In 13 C NMR spectroscopy of<br />

ferrocene-dio, the principal source of line-broadening is<br />

CSA and, furthermore, the CSA should be substantially<br />

the same in ferrocene-dio and in undeuterated ferrocene.<br />

The fact that A is essentially independent of vr for<br />

ferrocene-dio thus strongly supports the view that the<br />

increase of A with vr for undeuterated ferrocene is not<br />

arising from CSA being modulated by MAS.<br />

A / Hz<br />

200l<br />

ISO-<br />

100<br />

50<br />

vr / kHz<br />

10 12<br />

Fig. 4 Linewidth (A) versus MAS frequency (vr) for the<br />

isotropic peak in the 13 C NMR spectrum of<br />

ferrocene-djo, with no J H decoupling applied.<br />

High-resolution 13 C NMR spectra of ruthenocene and<br />

nickelocene were also recorded as a function of MAS<br />

frequency in order to establish whether the behaviour<br />

observed for ferrocene is also exhibited by other<br />

structurally-related systems. In Fig. 5, the relationships<br />

between A and vr for ruthenocene and ferrocene are<br />

compared, with all spectra recorded at the same *H<br />

decoupler field strength (vi = 64.9 kHz).<br />

It is clear that ruthenocene exhibits the same general<br />

trend as ferrocene, with A increasing as vr is increased,<br />

although the relationship for ruthenocene is apparently less<br />

linear than that for ferrocene. At the higher values of vr<br />

studied, the gradient 3A/3vr is larger for ruthenocene.<br />

These small differences in behaviour between ferrocene and<br />

ruthenocene presumably reflect small differences in the<br />

Bulletin of Magnetic Resonance<br />

dynamic properties of these solids at 293 K (as suggested<br />

in ref. 5).<br />

2501<br />

200"<br />

A /Hz lso-<br />

50-<br />

c ! 1<br />

vr / kHz<br />

10 12 14<br />

Fig. 5 Linewidth (A) versus MAS frequency (yr) for the<br />

isotropic peak in the 13 C NMR spectra of ferrocene (•)<br />

and ruthenocene (A), at fixed *H decoupler field<br />

strength corresponding to Vj = 64.9 kHz.<br />

The linewidth in the 13 C NMR spectrum of nickelocene<br />

is substantially greater (A = 15.7 kHz at vr = 5 kHz) than<br />

that in spectra recorded for ferrocene and ruthenocene under<br />

the same conditions. The linewidth for nickelocene is not<br />

significantly affected by increasing the MAS frequency,<br />

although there is a measurable decrease to A = 15.4 kHz at<br />

vr = 11 kHz. The very large 13 C NMR linewidth<br />

observed for nickelocene, even under conditions of MAS<br />

and high power *H decoupling, is due to the paramagnetic<br />

properties of this molecule (which contains two unpaired<br />

electrons). For this reason, it is not expected that the<br />

NMR properties of nickelocene will be comparable, in any<br />

way, to those of ferrocene and ruthenocene.<br />

5. Conclusions<br />

The increase in linewidth of the isotropic peak in the 13 C<br />

NMR spectra of ferrocene and ruthenocene as the MAS<br />

frequency is increased (at fixed temperature and fixed *H<br />

decoupler field strength) is due to an indirect effect in which<br />

the effective *H decoupler field strength is decreased as vr<br />

is increased. Considering the results of high-resolution<br />

13 C NMR experiments for several crystalline organic<br />

solids carried out in our laboratory, it is clear that, for<br />

systems in which there is no appreciable molecular<br />

motion, A is essentially independent of vr. This fact<br />

tends to suggest that the molecular motions present within<br />

crystalline ferrocene and ruthenocene are important in<br />

relation to the observed increase in the linewidth of the<br />

isotropic peak with increasing MAS frequency.


Vol. 14, No. 1-4<br />

A similar anomalous relationship between isotropic 13 C<br />

NMR linewidth and temperature has been observed recently<br />

by Muller [13] in studies of thiourea inclusion compounds;<br />

in this case, linewidths for 13 C environments in the guest<br />

molecules (which undergo substantial molecular motion)<br />

have been observed to increase with increasing temperature,<br />

and again this effect has been attributed to an "interference"<br />

between the molecular motion and the efficiency of the *H<br />

decoupling.<br />

In view of the fact (see Section 1) that cyclopentadienyl<br />

ring reorientation occurs on a timescale of the order of ca.<br />

10~ 12 s for ferrocene and ca. 10~ 10 s for ruthenocene at<br />

room temperature, it is not clear whether ring reorientation<br />

is indeed the dynamic process responsible for influencing<br />

the value of A in our experiments; it might be expected<br />

that a motion occurring at a frequency comparable to vi<br />

and/or vr (and thus in the approximate frequency range 10 3<br />

Hz to 10 6 Hz) would be required. At present, we make no<br />

attempt to assign the dynamic process that is important in<br />

giving rise to the observed NMR phenomena for ferrocene<br />

and ruthenocene at room temperature. However, it may be<br />

that the slower motion in ferrocene alluded to in ref. 2 is<br />

influential in this regard. Future experiments, at different<br />

temperatures, will investigate the variation of A with vr at<br />

different values of the correlation time for this motion, and<br />

other experimental investigations of the dynamic properties<br />

of crystalline metallocenes are in progress. We are also<br />

currently investigating various theoretical implications of<br />

the results reported here.<br />

Finally, it is relevant to strike a cautionary note in regard<br />

to recording high-resolution solid state 13 C NMR spectra<br />

for systems that are subject to line-broadening by CSA and<br />

direct ^C-*H dipole-dipole interactions. In view of the<br />

results reported here, it is clear that there are circumstances<br />

under which optimum resolution can be obtained by<br />

recording the spectrum at high *H decoupler field strength<br />

and low MAS frequency, rather than at high *H decoupler<br />

field strength and high MAS frequency.<br />

Acknowledgements<br />

277<br />

We are grateful to the S.E.R.C. (studentship to US) and<br />

the Nuffield Foundation (research grant to KDMH) for<br />

financial support.<br />

References<br />

[I] D. Braga, Chem. Rev., 92 (1992) 633.<br />

[2] C.H. Holm and J.A. Ibers, /. Chem. Phys., 30<br />

(1959) 885.<br />

[3] B.T.M. Willis, Ada Cryst., 13 (1960) 1088.<br />

[4] A.B. Gardner, J. Howard, T.C.Waddington, R.M.<br />

Richardson and J. Tomkinson, Chem. Phys., 57<br />

(1981) 453.<br />

[5] A.J. Campbell, CA. Fyfe, D. Harold-Smith and<br />

K.R. Jeffrey, Mol. Cryst. Liq. Cryst., 36 (1976) 1.<br />

[6] A. Kubo, R. Ikeda and D. Nakamura, /. Chem. Soc,<br />

Faraday Trans. 2, 82 (1986) 1543.<br />

[7] E.R. Andrew, Int. Rev. Phys. Chem., 1 (1981) 195.<br />

[8] E.R. Andrew, Phil. Trans. Roy. Soc. (Lond.) A, 299<br />

(1981)505.<br />

[9] M.M. Maricq and J.S. Waugh, /. Chem. Phys., 70<br />

(1979) 3300.<br />

[10] D. Suwelack, W.P. Rothwell and J.S. Waugh,<br />

/. Chem. Phys., 73 (1980) 2559.<br />

[II] W.P. Rothwell and J.S. Waugh, /. Chem. Phys., 74<br />

(1981)2721.<br />

[12] I.J. Shannon, K.D.M. Harris and S. Arumugam,<br />

Chem. Phys. Lett., in press.<br />

[13] K. Muller,/. Phys. Chem., 96(1992) 5733.


278<br />

1 Introduction<br />

The structural role of water in silicate glasses:<br />

1H and 29si NMR evidence<br />

J. Kiimmerlen, T. Schaller, A. Sebald and H. Keppler<br />

Bulletin of Magnetic Resonance<br />

Bayerisches Geoinstitut, Universitat Bayreuth, Postfach 101251<br />

W-8580 Bayreuth, Germany<br />

Over the last years various spectroscopic<br />

methods [1-6] have been used to<br />

investigate hydrous silicate and aluminosilicate<br />

glasses and to clarify the structural<br />

role of water in such systems. Especially,<br />

the existence of molecular water and/or<br />

Si-OH/Al-OH species, i.e. the H2O-induced<br />

depolymerisation of the Si/Al or Si<br />

network in such glasses has been discussed<br />

controversially in the literature.<br />

We have chosen two hydrous Na2Si4O9<br />

glasses with different H2O-contents as an<br />

Al-free model system.<br />

The Al-free Na2Si4O9 system is particularly<br />

well suited for high-resolution<br />

solid-state NMR investigations as the<br />

various Q-species can easily be resolved in<br />

the respective 29 Si MAS and CP/MAS<br />

spectra.<br />

Various high resolution solid state NMR<br />

techniques were used: 2!? Si singie pulse<br />

MAS, *H -> 29 Si-CP/MAS and *H-CRAMPS.<br />

Only the combined use of all these methods<br />

provides an insight into the interactions<br />

between the Si-network and the proton<br />

system. Additional modifications of these<br />

standard techniques were then used to<br />

confirm and to refine the picture of the<br />

Na2Si4O


Vol. 14, No. 1-4 279<br />

selectively and to obtain separate<br />

information about the dipolar protoninteractions<br />

("dipolar dephasing").<br />

29 Si dipolar dephasing<br />

10% H2O<br />

I • .I . . . I . . . I . . .I...I<br />

-70 -90 -110 ppm<br />

dephasing time<br />

1 ms<br />

5 ms<br />

10 ms<br />

Fig. 1: 29 Si-CP/MAS dipolar dephasing<br />

spectra of Na2Si4O9 glass with<br />

9.1% H2O. The glass with the<br />

lower water content shows<br />

similar decays of the signal<br />

components.<br />

3 Results and Discussion<br />

The 29 Si-MAS and 29 Si-CP/MAS spectra<br />

of the hydrous glasses show the presence<br />

°f Q4» Q3 and Q2 silicon species. The higher<br />

water content corresponds to a higher<br />

relative intensity of the Q2 signal and to a<br />

lower relative intensity of the Q4 signal,<br />

respectively. Furthermore, the 29 Si-<br />

CP/MAS experiments show significant<br />

differences in the time dependence of the<br />

29 Si magnetization of the two glasses. For<br />

the glass with the higher water content<br />

the signal intensity as a function of<br />

contact time is describable by a<br />

biexponential curve according to the usual<br />

thermodynamic model used to describe I-S<br />

CP-dynamics. The cross polarization times<br />

Tis and the relaxation times Tip show a<br />

trend for (Q4) > (Q3) >(Q2). Both Tis and Tip<br />

are significantly longer for the glass with<br />

the lower water content. The 29 Si CP-data<br />

of the Q3 and Q4 species for the lower H2Ocontent<br />

material can only be described by<br />

double-biexponential magnetization<br />

curves. This leads to the following<br />

conclusions: i) in the glass with the<br />

higher water content a uniform *H spin<br />

lattice is established, ii) due to the dilution<br />

of the protons in the other glass at least<br />

two ^H spin baths are present and iii) in<br />

the lower H2O-content glass the average<br />

interatomic distances between 29 Si and *H<br />

for the two different reservoirs must be<br />

significantly different. In consequence,<br />

one can assume the presence of both Si-OH<br />

groups and molecular water in these<br />

hydrous glasses. In fact, quantitative<br />

determination of the Q2, Q3 and Q4 species<br />

in both hydrous glasses from 29 Si MAS<br />

spectra and comparison of these results<br />

with the respective stoichiometric<br />

requirements if fully in accord with the<br />

depolymerisation of the silicate network.<br />

Results of the 29 Si-CP/MAS experiment<br />

with interrupted decoupling ("dipolar<br />

dephasing") are illustrated in Fig. 1. The<br />

decay of the 29 Si magnetization is mainly<br />

determined by the strength of the dipolar<br />

interaction between the 29 Si nuclei and<br />

the nearest surrounding proton system.<br />

Obviously, for the loss of magnetization<br />

characterized by a decay time T2 the<br />

relation T2(Q2)


280<br />

'H dipolar dtphasing<br />

dtphasing tinw<br />

»%H20<br />

20 10 0 ppm 20 10 0 ppm<br />

Fig. 2: ^-dipolar dephasing CRAMPS<br />

spectra of both glasses (450<br />

scans). (The spectra on the left<br />

hand side show an experimental<br />

artefact at ca. 18ppm)<br />

To confirm these CP results various J H<br />

multiple pulse experiments were used. The<br />

^H-CRAMPS spectra clearly show two well<br />

resolved signals at ca. 4.7 ppm and 12.2<br />

ppm (with respect to TMS = 0.0 ppm). The<br />

chemical shifts and the results of the CP<br />

experiments allow tentative assignment of<br />

these signals as molecular water (4.7 ppm)<br />

and protons of Si-OH groups (12.2 ppm).<br />

This assignment was confirmed by the *H<br />

"dipolar dephasing" CRAMPS experiment<br />

(jc/2-T-rc-T-MREV8). As illustrated in Fig. 2,<br />

in both glasses the signal at 4.7 ppm decays<br />

much faster than the less shielded signal,<br />

i.e., the dipolar * H - 1 H interaction is<br />

significantly stronger. This experimental<br />

fact corroborates the model proposed<br />

above. Comparing the decay rates found<br />

for the two different glasses, differences<br />

in these rates were found (Fig. 3). Again,<br />

these differences are in good agreement<br />

with the results of the standard 29 s i-<br />

CP/MAS experiments.<br />

. »0-<br />

3<br />

C<br />

v<br />

100<br />

>*, 90<br />

V)<br />

C<br />

Fig. 3:<br />

Bulletin of Magnetic Resonance<br />

a water protons<br />

A Si-OH protons<br />

a D<br />

5 10 15 20 25<br />

dephasing time (/is)<br />

Signal intensities in the * H<br />

dipolar dephasing CRAMPS<br />

experiment for the<br />

glass with<br />

a) 4.9 % H2O (top)<br />

b) 9.1 % H2O (below)<br />

Additionally, the longitudinal relaxation<br />

times Ti of both *H signal components in<br />

both glasses were measured. The use of a<br />

JC-T-JC/2-T-MREV8 pulse sequence provides<br />

a uniform Ti of 0.70 ± 0.05 s for the glass<br />

with the higher water content. In contrast,<br />

the signal components of the other<br />

glass decay with Ti(SiOH) = 0.85 ± 0.05 s and<br />

Ti(H2O) = 0.97 ± 0.05 s, respectively. These<br />

data also confirm the interpretation of the<br />

CP experiments and provide an additional<br />

support for the following conclusions:


Vol. 14, No. 1-4 " 281<br />

(1)H2O does depolymerize the silicate<br />

network,<br />

(2) both Si-OH and molecular water are<br />

present, and<br />

(3) in the lower H2O-content glass the<br />

more dilute proton system, forming<br />

these components is best described as<br />

"separate" sub-systems.<br />

To answer further questions concerning<br />

the interactions between Si-OH and<br />

H2O protons and to address questions<br />

concerning the structural role of cations<br />

like e.g. Na + in such silicate networks, ID<br />

and 2D-spin diffusion experiments are in<br />

preparation.<br />

Acknowledgment: Support of this work<br />

by the Deutsche Forschungsgemeinschaft<br />

and the Alexander von Humboldt-<br />

Foundation is gratefully acknowledged.<br />

Literatur:<br />

[I] Bartholomew, R.F., Butler, B.L., Hoover,<br />

H.L., Wu, C.K.J., J. Am. Ceram. Soc. 1980,<br />

63, 481<br />

[2] Kohn, S.C., Dupree, R., Geochim.<br />

Cosmochim. Acta 1987, 53, 2925<br />

[3] McMillan, P., Holloway, J.R., Contrib.<br />

Mineral. Petrol. 1987, 97, 320<br />

[4] Mysen, B.O. Virgo, D., Chem. Geol. 1986,<br />

64, 2623<br />

[5] Stolper, E.M., Am. Mineral. 1989, 74,<br />

1247<br />

[6] Eckert, H., Yesinowski, J.P., Silver,<br />

L.A., Stolper, E.M., J. Phys. Chem. 1988,<br />

92, 2055<br />

[7] Kiimmerlen, J., Merwin, L.H., Sebald,<br />

A., Keppler, H., J. Phys. Chem. (in<br />

press)<br />

[8] Oppella, S.J., Frey, M.H., J. Am. Chem.<br />

Soc. 1979, 101, 5854<br />

[9] Bodenhausen, G., Stark, R.E., Ruben,<br />

D.J., Griffin, R.G., Phys. Lett. 1979, 67,<br />

424<br />

[10] Rhim, W.K., Elleman, D.D., Vaughan,<br />

R.W., j. Chem. Phys. 1973, 59, 3740<br />

[II] Bronnimann, C.E., Zeigler, R.C.,<br />

Maciel, G.E., J. Am. Chem. Soc. 1988,<br />

110, 2023


282<br />

High-Resolution Solid-State NMR Study<br />

of Microstructures in Layered Aluminosilicate<br />

Bulletin of Magnetic Resonance<br />

Shigenobu Hayashi, Takahiro Ueda, Kikuko Hayamizu, and Etsuo Akiba<br />

1. INTRODUCTION<br />

National Chemical Laboratory for Industry,<br />

Tsukuba, Ibaraki 305, Japan<br />

Kaolin ite, Al4Si4Qo(OH)8> is a layered<br />

aluminosilicate with a dioctah edr al 1:1 layer<br />

structure consisting of an octahedral aluminum<br />

hydroxide sheet and a tetrahedral silica sheet.<br />

Figure 1 shows the structure of kaolin ite. Ihe crystal<br />

structure is not fully understood because of the<br />

absence of a large single crystal.<br />

In the present paper, we have traced 2 ^Si, 27 A1,<br />

and !H NMR spectra of various kaolinites, using<br />

high-resolution solid-state techniques. Analyzing<br />

the spectra theoretically, correlations between the<br />

NMR data and the local structures are discussed<br />

quantitatively.<br />

2. EXPERIMENTAL<br />

Totally eight samples were used. Six samples<br />

were natural, which were Kanpaku kaolin (called<br />

Nl; Hinckley index 1.4), API No.9 standard kaolin ite<br />

specimen (N2; 1.4), Georgia kaolin (N3; 0.7), Hakone<br />

Cbwakudani kaolin (N4; 0.4), Kibushi clay (N5; 0.2),<br />

and Gairome clay (N6; 0.2). Two synthetic samples<br />

were used,wh ich were synth esized at 290°C(51; 0.9)<br />

and220°C(52;0.8).<br />

NMR spectra were traced at room temperature by<br />

aBruker MSL400 (a static m agn etic field of 9.4 T) and<br />

a JBXGSH200 (4.71). Th e lin e sh apes of th e spectra<br />

were analyzed using computer programs written by<br />

ourselves.<br />

3. RESULTS AND DISCUSSION<br />

3.1. 2 ^Si spectra<br />

Figur e 2 sh ows 29 Si O7 MAS NMR spectr a of th e<br />

sample Nl, which has the highest crystallinity and<br />

the lowest content of paramagnetic impurities<br />

Fig. 1. Projection of the structure of kaolin ite from<br />

the(100) direction.<br />

among the eight samples studied. The spectra have<br />

two signals at -90.8 and -91.4 ppm from<br />

tetramethylsilane, being ascribed to Q'(GAl). The<br />

line shapes do not depend on the contact time, and<br />

the two peaks have the same cross relaxation time<br />

between ^H and 29 Si, which is 2.0 HIS. Maximum<br />

intensities are obtained at the contact time of 8 ms.<br />

The field dependence experiments demonstrate<br />

clearly th at two in equivalent Si sites are present.<br />

The linewidth in the 29 Si CP/MAS spectra is<br />

originated from the dipole-dipole interaction with<br />

27 Al,the chemical shift dispersion due to structural<br />

disorders, and the anisotropic magnetic<br />

susceptibility due to paramagnetic impurities.<br />

For Kanpaku kaolin,the chemical shift dispersion<br />

is 0.39 ppm, while the contributions of the dipolar<br />

interaction are0.08and033ppm for the fields of 9.4<br />

and4.7T,respectively,being estimatedfrom th e field


Vol. 14, No. 1-4 283<br />

B<br />

-85 -90 -95<br />

ppm<br />

79.496MHz 104.263MHz<br />

39.683MHz<br />

Fig. 2. 29 Si O7 MAS NMR spectra of th e sample JV1,<br />

measured at (A) 79.496 MHz and (B) 39.683<br />

MHz. Chemical shifts are expressed with<br />

r esp ect to tetr am eth y lsilan e.<br />

dependence experiments. The effect of<br />

paramagnetic impurities is negligible. The<br />

contribution of the dipole-dipole interaction with<br />

27 Al spin sis calculated theoretically from the crystal<br />

structure. The estimated linewidths at 9.4Tare 0.14<br />

and 0.09 ppm for Si(l) and Si(2), respectively,<br />

whereas those at 4.7 Tare 055 and 0.35 ppm. These<br />

values are in excellent agreement with the values<br />

estimated experimentally.<br />

Qher kaolinites, with lower crystallinities and/or<br />

higher contents of paramagnetic impurities, have<br />

broader linewidths due to the structural disorders<br />

an d th e par am agn etic impurities.<br />

3.2. 27 Al spectra<br />

Figur e 3 sh ows 27 A1 DE^ MAS NMR spectr a of th e<br />

sampleiv"l,measuredatdifferent fields. The Al atom<br />

in the kaolinite structure is coordinated by six<br />

oxygen atoms,and they give a signal around 0 ppm<br />

with respect to 1M A1(NQ)3 aqueous solution. The<br />

spectrum at the lower field gives the broader signal<br />

atthelower frequency position,which suggests that<br />

the signal is the central transition, being broadened<br />

by the second-order quadrupole interaction. The<br />

B<br />

52.051MHz<br />

100 50 0 -50 -100<br />

ppm<br />

Fig. 3. 27 A1 Hy MAS NMR spectra of the sample M,<br />

measured at (A) 104.263 MHz and (B) 52.051<br />

MHz. Chemical shifts are expressed with<br />

r espect to 1 M A1(NC$)3 aqueous solution.<br />

line shapes are simulated by our computer<br />

programs. The observed spectra cannot be<br />

simulated by one component, but can be simulated<br />

much better by two components with equal<br />

intensities. The obtained parameters for Al(l) are a<br />

chemical shift of 7.8 ppm, a quadrupole coupling<br />

constant of 3.36 MHz, an dan asymmetry factor of the<br />

quadrupole interaction of 055, while they are 7.8<br />

ppm, 2.88 MHz, and 1.00 for Al(2). The crystal<br />

structure indicates the presence of two in equivalent<br />

Al sites with a population ratio of 1:1.<br />

The 27 A1 spectra are also recorded for .the other<br />

kaolin ites. The same quadrupole coupling<br />

parameters as those in the sample JV1 can well<br />

explain th e lin e sh apes of th e oth er kaolin ites.<br />

A small fraction of tetrahedral Al is observed at<br />

about 70 ppm for several samples, which might be<br />

ascribed to impurities.<br />

3.3. 1 H static spectra<br />

Figure 4Ashows an 1 H static NMR spectrum of<br />

the sample JV1. The spectrum consists of two<br />

components with different line shapes; a narrow<br />

Lorentzian line with a width of 1.4 kHz and a broad<br />

Gaussian with a28.9kHzwidth.


284 Bulletin of Magnetic Resonance<br />

40 -40<br />

20 10 -10<br />

ppa<br />

Fig.4. *H NMR spectra of the sample Nl, measured at<br />

400.136 MHz. (A) The ordinary single-pulse<br />

sequen ce is used for th e static sample. (B) Th e<br />

CRAMPS spectrum measured with the BR24<br />

pulse sequence in the quadrature phase<br />

detection mode. Chemical shifts are expressed<br />

with respect to tetramethylsilane.<br />

The narrow component can be ascribed to water<br />

molecules adsorbed on the outer surface. This<br />

component is easily diminished by evacuation, and<br />

they grow up gradually in the air atmosphere.<br />

Qi the other hand,the broad component can be<br />

ascribed to the hydroxyl groups in the kaolinite<br />

structure, whose second moment is 105 kHz 2 . The<br />

second moment estimated from the crystal structure<br />

is 92 kHz 2 , in which iH-iH and 1 H- 27 A1<br />

contributions are 67 and 25 kHz 2 , respectively. The<br />

calculated second moment agrees with the<br />

experimental value. These results demonstrate that<br />

the hydrogen atoms in the CH group is in a rigid<br />

lattice state at room temperature.<br />

The *H static NMR spectra of other kaolin ites also<br />

con sist of th e two compon en ts.<br />

3.4. 1 E CRAMPS<br />

The CRAMPS technique is successfully applied to<br />

the kaolinite samples. Figure 4Bshows l H CRAMPS<br />

spectra of the sample JV1. The BR24 pulse sequence is<br />

used in the quadrature phase detection mode.<br />

Considerably large spinning sidebands are observed<br />

on both sides of the central peak, which are caused<br />

by the strong dipole-dipole interaction between *H<br />

and 27 A1 spins. The linewidth in the static state, 29<br />

kHz, is reduced to about 600 Hz by the use of the<br />

CRAMPS technique, where the reduction factor is<br />

about 50. The chemical shift is 2.8 ppm from<br />

tetramethylsilane.<br />

The chemical sh ift of the hydroxyl groups in the<br />

kaolin ites changes slightly depending on the<br />

sample, which might reflect the strength of the<br />

hydrogen bonding or the acidity of the hydrogen .<br />

Acknowledgement<br />

The authors wish to express their thanks to Dr. R.<br />

Miyawaki at Government Industrial Research<br />

Institute, Nagoya and Dr. K. Kuroda at Waseda<br />

University for kindly supplying the kaolin ite<br />

samples.


Vol. 14, No. 1-4 285<br />

1 Introduction<br />

Broadline NMR of Structural Ceramics<br />

Magic angle spinning (MAS) has become the most widely<br />

used NMR technique for the study of inorganic solids.<br />

This popularity has come about because MAS removes<br />

the effects of chemical shift anisotropy, permitting acquisition<br />

of chemical shift spectra in these materials.<br />

However, there are two weaknesses to the MAS<br />

technique, which are accentuated in the study of structural<br />

ceramics.<br />

The first difficulty with MAS is that rotation at the<br />

magic angle does not remove the effects of second order<br />

quadrupolar broadening of the central (+1/2 «—» -1/2)<br />

transition [1]. This has limited the applicability of NMR<br />

for half-integral quadrupolar nuclei like 27 A1, U B, 17 O,<br />

and 91 Zr, which are important constituents of ceramics.<br />

While recent work has shown that the effects of second<br />

order quadrupolar broadening are reduced by working in<br />

larger magnetic fields [2], or by employing more sophisticated<br />

spinning techniques [3], these solutions can be<br />

difficult and expensive to implement.<br />

The other important limitation of MAS, also true of<br />

other sample spinning techniques, is that the physical<br />

form of the sample is restricted to fine powders, homogeneous<br />

cylinders, or chunks of material packed in an NMR<br />

inert powder of similar density. All of these forms present<br />

problems when working with ceramics. Many ceramics<br />

are challenging to machine or grind due to their extreme<br />

hardness. Even when grinding is possible, other<br />

requirements may limit the use of powders. For example,<br />

in determining the effect of long term or repeated heating<br />

of a ceramic, heating a powdered sample may provide<br />

unreliable data because of surface oxidation. Packing<br />

large chunks of sample in a powder with similar density<br />

becomes time consuming when a number of samples have<br />

to be studied, or the sample is being subjected to heating<br />

which changes its chemical or phase composition.<br />

We have used broadline NMR of static samples as an<br />

alternative to MAS for samples with large quadrupolar<br />

splittings. In favorable cases, the resulting powder<br />

pattern, produced by the first order quadrupolar coupling<br />

Chuck Connor<br />

Defence Research Establishment Pacific<br />

FMO Victoria, B.C., VOS 1B0, Canada<br />

[1], directly yields information about the electronic environment<br />

of the nuclei under investigation. Although the<br />

broadline technique for quadrupolar nuclei is not new,<br />

having been first applied over forty years ago [4], it is a<br />

simple, useful technique which is often overlooked. It is<br />

generally applicable only to nuclei with large quadrupolar<br />

splittings, which provide the most difficulty for MAS;<br />

accordingly the broadline technique can serve as a complement<br />

to MAS. The work reported here demonstrates<br />

that, for many ceramics, broadline spectra can provide<br />

useful structural information.<br />

2 Experimental<br />

All spectra were recorded on a Bruker MSL-300<br />

spectrometer, equipped with the BC-131 5 MHz 9 bit<br />

digitizer. The spectrometer operates at 96 MHz for U B,<br />

and 78 MHz for 27 A1. A standard Bruker multinuclear<br />

solenoid probe was used for the broadline spectra. Free<br />

induction decays were collected after a single 1 \is pulse.<br />

Although the excitation profile of this pulse drops to zero<br />

at 1 MHz, reasonable sensitivity is still maintained for<br />

satellites 600 to 700 kHz off resonance. The delay time<br />

before the start of data collection was typically 5 u.s.<br />

Usually the first 2 or 4 points of the FID were discarded<br />

because of pulse breakthrough. The relatively long<br />

deadtime usually prohibits observation of the broad<br />

pedestal portion of the powder pattern, but the cusps,<br />

which by themselves are sufficient to characterize the<br />

quadrupolar splitting, do not appear greatly affected. The<br />

width of the spectral window was 2.5 MHz (+1.25 MHz)<br />

for U B and 1.67 MHz (±0.833 MHz) for 27 A1. A filter<br />

bandwidth of 1 MHz, the largest available on the MSL-<br />

300 in quadrature mode, was used. Additional data collection<br />

parameters used for individual spectra are noted in<br />

the figure captions.<br />

Quadrupolar splittings were measured from a point on<br />

the outside edge of one satellite, at about 80% of the<br />

satellite peak height, to the corresponding point on the<br />

other satellite of the pair. The value 80% was obtained by


286 Bulletin of Magnetic Resonance<br />

comparing simulated powder patterns with and without<br />

Gaussian broadening. Measurements between the maxima<br />

of the satellites give values that are smaller than the true<br />

value, because dipolar broadening shifts the position of<br />

the maxima toward the central transition.<br />

3 Results and Discussion<br />

The 27 A1 (/ = 5/2) resonance in corundum (a-Al2O3) has<br />

been observed by a variety of NMR techniques [4], [5].<br />

For comparison, we show an 27 A1 powder pattern in<br />

Figure 1. For this sample, excellent sensitivity is obtained<br />

500000 0<br />

HERTZ<br />

-500000<br />

Fig. 1. 27 A1 powder pattern of a-Al2O3. 32 scans were coadded,<br />

using a recycle delay of eight seconds. The upper trace is<br />

a magnification of the lower trace by ten.<br />

with about four minutes of signal averaging. From the<br />

splitting of each pair of satellites, we find<br />

e 2 qQ/h = 2.42+0.02 MHz and r| = 0.0, in fair agreement<br />

with the values determined by Pound [4]. The slight<br />

reduction in precision of these numbers, as compared with<br />

Pound's work, is offset by the speed and ease with which<br />

the results can be obtained on a standard solids NMR<br />

spectrometer.<br />

One of the many applications for NMR of ceramics is<br />

to follow high temperature phase changes and reactions.<br />

We have studied the calcination of gibbsite (A1(OH)3) to<br />

form a-Al2O3, which is a complex and poorly understood<br />

process [6]. Results of preliminary work to determine the<br />

applicability of broadline NMR to this problem are shown<br />

in Figure 2. The ^Al spectrum of the unheated material<br />

(determined by X-ray diffraction to be at least 90%<br />

gibbsite) shows five pairs of satellite transitions, with<br />

splittings of 96 kHz, 397 kHz, 517 kHz, 608 kHz, and<br />

960 kHz. The peaks near ±700 kHz, which appear in<br />

many 27 A1 spectra, are probably artifacts. Assuming three<br />

sites are present, the splittings can be paired as follows:<br />

397 and 608 kHz, from a site with e 2 qQ/h = 2.1 MHz and<br />

500000<br />

0<br />

HERTZ<br />

-500000<br />

Fig. 2. 27 A1 powder patterns of gibbsite (A1(OH)3) before<br />

heating (lower trace), after heating at 700 C for 35 minutes<br />

(middle trace), and after further heating at 1070 C for 16 hours<br />

(upper trace). Each trace represents 1 to 2 hours of signal<br />

averaging, with a one second recycle delay.<br />

T] = 0.5; 517 and 960 kHz, from a site with<br />

e^qQ/h = 3.2 MHz and Y| = 0.2. The other possible pairing<br />

gives one site with e*qQlh = 1.8 MHz (r\ = 0.7) and one<br />

with e 2 qQlh = 3.3 MHz (r\ = 0.5). Only one pair of<br />

satellites from the remaining site is visible, with a splitting<br />

of 96 kHz. Two possibilities may explain this absence of<br />

a second pair. If e 2 qQ/h = 0.36 MHz and r\ ~ 1, the<br />

second pair of satellites will overlap the observed pair. If<br />

e*qQ/h = 0.32 MHz and r\ - 0, the second pair will have a<br />

splitting of about 48 kHz, and will not be resolved from<br />

the central transition. An accurate quantification of the<br />

relative population of the three sites from this data is<br />

difficult, but one can say the three sites are roughly equally<br />

populated. The nuclei with the larger couplings are<br />

probably in octahedral environments, as expected from the<br />

reported structure, in which all the aluminum nuclei are in<br />

octahedral environments [7]. The relatively large values<br />

of r| can be attributed to hydrogen bonding, which distorts<br />

the octahedral symmetry of the aluminum sites in gibbsite.<br />

The smaller quadrupolar coupling is about an order of<br />

magnitude less than that typically observed for octahedral<br />

sites (cf. a-Al2O3), so is probably due to a slightly<br />

distorted tetrahedral site. From a simulation of MAS<br />

results at 6.35 T [2], it appears that only two sites are<br />

observed by MAS, both of which are octahedral.<br />

Conversion of A1(OH)3 to CC-A12O3 involves several<br />

intermediate phases such as boehmite, X-A12O3, y-Al2O3,<br />

K-A12O3, 6-Al2O3, and 8-Al2O3, with formation of oc-<br />

A12O3 reportedly occurring at 1140 C [6]. The middle


Vol. 14, No. 1-4<br />

trace in Figure 2 shows the effect of heating gibbsite at<br />

700 C for 35 minutes. The satellites attributed to gibbsite<br />

have disappeared, and the central transition is flanked by<br />

broad features indicative of amorphous material. The<br />

width of the central transition, which may be attributed to<br />

the second order quadrupolar interaction, corresponds to<br />

values of e 2 qQ/h ranging up to 4.4 MHz. After further<br />

heating for 16 hours at 1070 C, the upper spectrum in<br />

Figure 2 was obtained. The sharp features indicate a<br />

substantial amount of amorphous material has converted<br />

to a-Al2O3. This contradicts the notion that conversion to<br />

a-Al2O3 occurs only at temperatures above 1140 C.<br />

However, the conversion is quite sluggish at 1070 C.<br />

These broadline spectra demonstrate conversion of<br />

gibbsite to a-Al2O3, but unfortunately information on the<br />

intermediate phases cannot be obtained from these spectra.<br />

Chemical shift spectra obtained with sample spinning will<br />

probably not be useful in determining which phases are<br />

present in the amorphous intermediate stage, considering<br />

the large quadrupolar couplings which appear to be<br />

present.<br />

Another problem of interest is the oxidation of<br />

ceramics at high temperatures. When boron nitride (BN)<br />

is heated in an oxidizing environment, conversion to<br />

boron oxide gradually takes place. Since both BN and<br />

B2O3 have large quadrupole couplings [8], [9], one can<br />

expect difficulty in resolving the signals from the two<br />

materials in MAS spectra. In fact, the second order<br />

quadrupolar broadening is on the order of 50 ppm, based<br />

on a Larmor frequency of 96 MHz, which is a significant<br />

fraction of the expected range of chemical shift for n B.<br />

To demonstrate that broadline NMR may prove more<br />

useful in monitoring the oxidation of BN, we prepared a<br />

mixture of roughly equal amounts of BN and B(OH)3.<br />

Boric acid was used instead of boron oxide because the<br />

fine powder required for MAS work is not hygroscopic, as<br />

is the B2O3 powder, and the boric acid was readily available.<br />

Boric acid has a quadrupolar splitting of 1282 kHz,<br />

similar to that of B2O3 (1308 kHz).<br />

U B (/ = 3/2) MAS and broadline spectra of the boron<br />

nitride/boric acid mixture were recorded, and are shown in<br />

Figures 3 and 4 respectively. From the MAS spectrum,<br />

one can see that the single peak observed is a composite of<br />

several resonance lines. However, one would have<br />

difficulty determining which boron species are present in<br />

the sample, or even how many species there are. This<br />

spectrum should be contrasted with the broadline<br />

spectrum shown in Figure 4. One immediately sees that<br />

two boron sites are present, with quadrupolar splittings of<br />

1462±10 kHz and 1256±10 kHz. By comparison with<br />

previous work [8], [9] the species are identified as boron<br />

nitride and boric acid respectively. Based on this model<br />

mixture, we expect to be able to monitor the oxidation of<br />

boron nitride.<br />

A variety of approaches, including double-rotation<br />

(DOR), dynamic-angle-spinning (DAS) [3], and MAS in<br />

larger static fields [2], have been used to improve the<br />

50000 0<br />

HERTZ<br />

-50000<br />

Fig. 3. MAS spectrum of n B in a mixture of boron nitride and<br />

boric acid, with a 5 kHz spinning rate. 2296 scans were<br />

collected, with a recycle delay of one second.<br />

resolution of chemical shift spectra beyond that attainable<br />

by conventional MAS. Broadline NMR may compare less<br />

favorably with these new techniques than it does with<br />

MAS. However, these techniques have not been widely<br />

exploited in materials research, and, like MAS, they suffer<br />

from the physical limitations imposed on the sample as<br />

discussed in the Introduction.<br />

Titanium boride (TiB^ has been presented as a candidate<br />

for lightweight armor [10]. The structure consists of<br />

alternating planar layers of boron and titanium, with each<br />

boron nucleus trigonally bound to three other boron nuclei<br />

[11]. To yield the 2:1 stoichiometry, each boron layer<br />

contains twice as many nuclei as a titanium layer. The<br />

crystal structure suggests only one type of boron site is<br />

present, and previous workers have reported only one site,<br />

with a quadrupolar splitting of 180*10 kHz [9]. However,<br />

500000 0<br />

HERTZ<br />

-500000<br />

AJL.<br />

Fig. 4. U B powder pattern of the mixture examined in Figure<br />

3. This spectrum was obtained after about 16 hours of signal<br />

averaging (56989 scans, one second recycle delay).<br />

287


288<br />

we see evidence for at least two sites in breadline U B<br />

NMR spectra (Figure 5), with quadrupolar splittings of<br />

500000<br />

0<br />

HERTZ<br />

-500000<br />

Fig. 5. n B powder pattern of titanium boride (TiBj), from 324<br />

co-added scans with a recycle delay of one second.<br />

177 kHz and 355 kHz. In addition, a weak, poorly<br />

resolved pair of satellites may be present with a splitting<br />

of 195 kHz. It appears that roughly four fifths of the<br />

boron nuclei are located in the 177 kHz site, one fifth in<br />

the 355 kHz sites, with a much smaller fraction in the<br />

195 kHz site.<br />

X-ray diffraction confirmed the sample was greater<br />

than 90% TiB2, with most of the remainder consisting of<br />

titanium oxides. The general features of the pattern<br />

matched that of A1B2, which also has planar layers. The<br />

diffraction pattern showed no similarity to patterns from<br />

compounds containing puckered boron layers, such as<br />

RhB2, TcB2, and Ru2B3. Thus it appears the two sites are<br />

in equivalent geometrical positions within the boron<br />

plane. The larger quadrupolar splitting could be due to<br />

decreased donation of electrons from the titanium atoms<br />

to the boron it orbitals in about one fifth of the boron<br />

nuclei [9]. This heterogeneity within the boron layers<br />

may explain why titanium boride does not fracture along<br />

the interface between layers, as does graphite and many<br />

other layered materials.<br />

4 Summary<br />

For many ceramics, the second order quadrupolar broadening<br />

in MAS spectra exceeds the dispersion provided by<br />

the chemical shift. The far greater dispersion from the<br />

first order quadrupolar interaction can be used to obtain<br />

resolved spectra from these samples. We have provided<br />

examples of several structural ceramics for which<br />

broadline spectra, obtained with a standard solids<br />

spectrometer, yield useful structural information. Thus<br />

the technique shows the potential to complement MAS in<br />

the study of inorganic materials.<br />

5 References<br />

Bulletin of Magnetic Resonance<br />

1. A. Abragam, Principles of Nuclear Magnetism<br />

(Clarendon Press, Oxford, 1961), Chapter VII.<br />

2. e.g. D.E. Woessner, Am. Mineral. 74, 203 (1989).<br />

3. B.F. Chmelka, K.T. Mueller, A. Pines, J. Stebbins,<br />

Y. Wu, and J.W. Zwanziger, Nature, 339, 42<br />

(1989).<br />

4. R.V. Pound, Phys. Rev. 79,685 (1950).<br />

5.a) D. Lee and P.J. Bray, /. Magn. Reson. 94, 51<br />

(1991).<br />

b) J. Haase and H. Pfeifer, /. Magn. Reson. 86, 217<br />

(1990).<br />

c) HJ. Jakobsen, J. Skibsted, H. Bilds0e, and N.C.<br />

Nielsen,/. Magn. Reson. 85,173 (1989).<br />

6. Engineered Materials Handbook, Vol. 4, Ceramics<br />

and Glasses (ASM International, 1991), p.112.<br />

7. A.F. Wells, Structural Inorganic Chemistry<br />

(Clarendon Press, Oxford, 1962), p. 552.<br />

8.a) A.H. Silver, J.Chem.Phys. 32, 959 (1960).<br />

b) C. Connor, J. Chang and A. Pines, J. Chem. Phys.<br />

93, 7639 (1990).<br />

9. PJ. Bray, AIP Conf. Proc. 140,142 (1986).<br />

10. D.J. Viechnicki, M.J. Slavin, and M.I. Kliman,<br />

Ceramic Bulletin, 70,1035 (1991).<br />

11. T. Lundstrom in Boron and Refractory Borides,<br />

edited by V.I. Matkovich (Springer-Verlag, Berlin,<br />

1977), p. 351.


Vol. 14, No. 1-4 289<br />

PERMEABILITY OF LIPOSOMAL MEMBRANES TO MOLECULES OF<br />

ENVIRONMENTAL INTEREST: RESULTS FROM NMR EXPERIMENTS<br />

EMPLOYING SHIFT AGENTS<br />

INTRODUCTION:<br />

by<br />

F.G. Herring*, W.R. Cullen, J.C. Nelson and P.S. Phillips,<br />

Department of Chemistry, University of British Columbia,<br />

Vancouver, B.C., Canada V6T 1Z1<br />

Dimethylarsinic acid (DMA) is a<br />

widely used pesticide and is an important<br />

intermediate in the marine bio-cycling of<br />

arsenic. Studies into the uptake mechanism<br />

of this organo-arsenical have shown that it<br />

enters cells by slow passive diffusion (1).<br />

The work presented in this article describes a<br />

NMR technique that has been developed to<br />

measure the rate of diffusion of compounds<br />

through a phospholipid bilayer. A similar<br />

method has been described by Prestegard<br />

et.al. (2) for the measurement of maleic acid<br />

diffusion constants.<br />

The diffusion rates of molecules<br />

across the membrane of liposomes have been<br />

measured using a variety of techniques, the<br />

most common of which is radio-labelling.<br />

The technique demonstrated in this study has<br />

many advantages over radio-labelling some<br />

of which are as follows; the cost and<br />

difficulties associated with working with<br />

radio-labelled compounds are eliminated, the<br />

method is readily automated, and sampling is<br />

eliminated. The NMR method is applicable<br />

to any water soluble compound which has a<br />

*H resonance signal which does not overlap<br />

*The abbreviation (DMA) does not<br />

distinguish between the protonated (DMAH)<br />

and the unprotonated (DMA") forms of the<br />

acid.<br />

with any other peaks or which can be shifted<br />

either upfield or downfield agent. It can be<br />

used on molelecules which have a<br />

permeability through the phospholipid bilayer<br />

of 10'^cm/s or less.<br />

We present here an investigation of<br />

DMA transport in a model membrane system<br />

by using this NMR method as applied to<br />

diffusion across the membranes of extruded<br />

large unilamella vesicles (LUVs) (3). The<br />

study contributes to our understanding of<br />

diffusive transport across bilayer membranes.<br />

It also illustrates the potential of NMR<br />

spectroscopy for membrane permeability<br />

studies in large unilamella vesicles.<br />

MATERIALS AND METHODS:<br />

Dry egg phosphatidyl-choline was<br />

hydrated with a buffered solution (in D2O)<br />

of DMA at the appropriate pH. The solution<br />

was then subjected to several freeze-thaw<br />

cycles using liquid nitrogen to enhance<br />

entrapment of DMA and to increase the<br />

degree of unilamellarity of the phospholipid<br />

bilayer. The multi-lamella suspension was<br />

extruded through polycarbonate filters with a<br />

200 nm pore diameter under high pressures<br />

to ensure that the vesicles used were all<br />

approximately the same size (3,4). The<br />

resultant solution of LUVs was then passed<br />

down a Sephadex column and eluted with<br />

buffer (in H2O) to remove most of the DMA


290<br />

that was not encapsulated; this procedure<br />

establishes the desired concentration gradient<br />

for diffusion studies. The eluted LUVs were<br />

added to a NMR tube which already<br />

contained Mn 2+ , HEPES, TSP, and glucose.<br />

NMR spectra were obtained at appropriate<br />

time intervals by using a Bruker AM<br />

4OO.The water signal was suppressed by presaturation.<br />

The FIDs were processed with a<br />

line-broadening of 10 Hz. The DMA and<br />

HEPES peaks were integrated and the ratio<br />

of the integrals taken as a measure of the<br />

efflux of DMA, this procedure was adopted<br />

to account for instrument variability.<br />

THEORY:<br />

-dnin/dt = dnout/dt = k (nin/Vin-n0Ut/V0Ut)<br />

Inside peak:<br />

nt. 11 =neq. +(no. 1 .neq. II \e-(l+f)kt/V.<br />

e<br />

in " in v in m> in<br />

Outside peak:<br />

P = k/A<br />

°out<br />

(1)<br />

(4)<br />

Equation (1) is the basic rate<br />

equation for the flux of particles across the<br />

bilayer assuming that the rate of appearance<br />

of the particles on the outside is equal to the<br />

rate of disappearance of the particle on the<br />

inside. Equations (2) and (3) are the<br />

equations which relate the number of<br />

particles on either side of the membrane as a<br />

function of time. Equation (4) relates the<br />

internal volume to the external volume.<br />

Equation (5) relates the rate constant to the<br />

permeability coefficient which is the standard<br />

measure of the rate of diffusion of a<br />

compound through a membrane.<br />

For the integral of the composite<br />

methyl resonance (DMAH and DMA") we<br />

have:<br />

(5)<br />

Bulletin of Magnetic Resonance<br />

where kobs = a ICDMAH Vjn, is the<br />

observed rate constant of diffusion for<br />

DMAH and a = Ka / (Ka + [H+]).<br />

Equation (6) is valid if it is assumed that<br />

^DMA"


Vol. 14, No. 1-4 291<br />

0.25<br />

both sides of the membrane. The membrane<br />

is impermeable to these three additives over<br />

the time scale of an experiment. It should be<br />

noted that DMA is a weak acid which is<br />

about 80% dissociated at pH = 7 (pKa =<br />

6.28).<br />

The decrease of the NMR peak for<br />

those molecules effusing and a corresponding<br />

increase in the peak for DMA outside is<br />

displayed in Fig 1. At the equilibrium position<br />

(the last spectrum shown) the ratio of<br />

the integrals of these peaks correspond to<br />

the ratio of the volume inside to outside (eq<br />

(4))-<br />

The amplitude ratios (total integral of<br />

DMAH and DMA" to the integral of the<br />

HEPES buffer) of both the inside and outside<br />

peaks for the 25 spectra which make up a<br />

single experimental run are shown as a function<br />

of time in Fig. 2. An iterative fit of<br />

these curves using a spreadsheet program<br />

(QPRO) permits the estimation of the rate<br />

constant.<br />

10 15<br />

TIME(SECS.)<br />

(Thousands)<br />

Fig. 2<br />

20 25<br />

In order to demonstrate that the<br />

transport is dominated by the neutral species<br />

(DMAH) as suggested above, we performed<br />

experiments at different pH. The rate of<br />

change of the integral ratio will depend upon<br />

the concentration of DMAH. The integral<br />

ratio will decrease faster due to the increased<br />

fraction (a) of DMAH contributing to the<br />

single methyl resonance observed for both<br />

DMA" and DMAH. The table illustrates the<br />

results we obtained.<br />

Table<br />

The pH dependence of the<br />

observed rate constant<br />

pH<br />

7.00<br />

7.15<br />

7.40<br />

7.73<br />

7.97<br />

(/cm 3 s-l*10" 5 )<br />

14.50<br />

8.89<br />

5.65<br />

2.39<br />

1.56


292<br />

Equation (6) gives the true rate<br />

constant for DMAH from the measured<br />

observed rate constant. The true rate<br />

constant and permeability coefficient are<br />

1.08.x 10" 3 cm 3 /s and 3 x 10' 8 cm/s<br />

respectively. Similar studies for monomethyl<br />

arsonic acid (MMAH) show that the<br />

permeability is 2 x 10"*" cm/s. This<br />

difference is consistent with the general rule<br />

that replacing a hydroxyl group with a<br />

methyl group will increase the permeability<br />

of the molecule by approximately two to<br />

three orders of magnitude(6).<br />

ACKNOWLEDGMENTS:<br />

This work is supported by operating<br />

grants from the NSERC of Canada. The<br />

exceptional technical assistance of Alice Ho<br />

and Anna Mason is appreciated. Alice and<br />

Bulletin of Magnetic Resonance<br />

Anna also received Summer '92 Studentships<br />

from NSERC.<br />

REFERENCES:<br />

(1) W.R. Cullen, B.C. McBride, A.W.<br />

Pickett, Appl. Organomet. Chem. 4,<br />

119,(1990).<br />

(2) J.H. Prestegard, J.A. Cramer and<br />

D.B. Viscio, Biophys. J. (Biophysical<br />

Society) 26, 575, (1979).<br />

(3) M.J. Hope, M.B. Bally, G. Webb and<br />

P.R. Cullis, Biochim. Biophys. Acta<br />

872. 55-65, (1985).<br />

(4) L.D. Mayer, M.J. Hope and<br />

P.R.Cullis, Biochim. Biophys. Acta<br />

858, 181,(19861<br />

(5) W.C. Stein, Channels Carriers and<br />

Pumps, Academic Press, New York,<br />

(1990).


Vol. 14, No. 1-4<br />

NUCLEAR MAGNETIC RESONANCE PARTITIONING STUDIES OF<br />

SOLUTE ACTION IN LIPID MEMBRANES<br />

Lan Ma, Theodore F. Taraschi, and Nathan Janes<br />

Department of Pathology and Cell Biology,<br />

Thomas Jefferson University, 1020 Locust St., Philadelphia, PA 19107<br />

INTRODUCTION: Lipid theories of<br />

anesthesia implicate perturbation of membrane<br />

lipids as the locus for acute anesthetic<br />

action. [1] Chronic exposure to alcohols and<br />

anesthetics induces an adaptive response in<br />

membrane phospholipids that confers resistance<br />

to many of the acute actions of alcohols<br />

and anesthetics. [2]<br />

We have proposed a colligative thermodynamic<br />

reformulation of the Meyer-Overton<br />

hypothesis for anesthetic action. [3,4] This<br />

reformulation implicates configurational<br />

entropy (Scf), the entropy imparted by a solute<br />

upon a membrane structure in the partitioning<br />

process, as the driving force of solute action<br />

on cooperative membrane equilibria. Solute<br />

potency is determined by the competing contributions<br />

of configurational and thermal<br />

entropy (ASt). Equilibria most susceptible to<br />

solute action (where dilute concentrations of<br />

solutes induce a perturbation equivalent to a<br />

large change in temperature) involve large<br />

changes in configurational entropy and small<br />

changes in thermal entropy according to the<br />

following relation. [3]<br />

AT/Tm = AScf/ASt (1)<br />

AT is the perturbation of the midpoint temperature,<br />

Tm, from its value in the absence of<br />

solute. The thermal entropy of an equilibrium<br />

is deduced from calorimetry and is approximately<br />

constant for solute levels of biological<br />

relevance. The remaining unknowns are the<br />

configurational entropy, which is determined<br />

from the partitioning of the solute, and the<br />

perturbation of the equilibrium midpoint.<br />

The colligative thermodynamic framework<br />

implicates solute partitioning as the energetic<br />

force that drives perturbations of cooperative<br />

membrane equilibria by altering the relative<br />

free energies of membrane states. Tests of<br />

the framework require simultaneous measures<br />

of solute partitiomng and membrane structure<br />

over a range of solute concentrations and<br />

temperatures.<br />

293<br />

Spin label* partitioning protocols have often<br />

been used in ESR studies of membrane structure.<br />

[5] Such studies are designed so that the<br />

spin label partitioning probe is a<br />

nonperturbing reporter of membrane structure.<br />

To study solute action on membranes,<br />

however, a partitioning probe should serve the<br />

multifarious role of membrane perturbant,<br />

reporter of perturbations, and reporter of<br />

solute partitioning. Since NMR methods are<br />

not limited to dilute solute levels, such flexibility<br />

is offered. Furthermore, complementary<br />

structural information is available from<br />

simultaneous wideline X H [6], 2 H [7], or 31 P<br />

[8] studies.<br />

PARTITIONING APPROACH TO SOLUTE<br />

ACTION: In this abstract, we describe a H<br />

NMR partitioning approach based on the<br />

uncharged local anesthetic alcohol, benzyl<br />

alcohol. Benzyl alcohol is a clinically used<br />

topical bacteriostatic agent. A variety of<br />

commercial pharmaceutical agents prepared<br />

for injection contain benzyl alcohol for its<br />

preservative properties and for pain relief.<br />

The partitioning approach is based on (/) the<br />

sensitivity of the ring proton chemical shift to<br />

the polarity of its environment and («) the<br />

sensitivity of the ring proton linewidth to<br />

membrane binding. The chemical shift of the<br />

ring resonances in a hydrophobic environment<br />

are shielded and resolved from the ring resonances<br />

of the aqueous alcohol. The sensitivity<br />

of the ring proton resonance to its environment<br />

provides a means of discriminating the<br />

aqueous alcohol resonance from the partitioned<br />

alcohol resonance. The dependence of<br />

the three chemically distinct ring proton<br />

chemical shifts on their environment is shown<br />

in Table 1 for benzyl alcohol (5 mole fraction<br />

%) in a variety of bulk solvents at 22°C. The<br />

resonance exhibits a diamagnetic shift in<br />

hydrophobic solvents. A modest correlation<br />

between the chemical shift and Hildebrandt's<br />

solubility parameter ($*) for the solvent is<br />

evident.


294<br />

SOLVENT<br />

Water<br />

Methanol<br />

1-Propanol<br />

1-Butanol<br />

1-Octanol<br />

1-Decanol<br />

Acetone<br />

Methylene Chloride<br />

Chloroform-dj<br />

Carbon Tetrachloride<br />

Hexane<br />

I<br />

7.8<br />

I<br />

7.6<br />

I<br />

7.4<br />

7.2<br />

PPM<br />

5*<br />

23.4<br />

14.5<br />

11.9<br />

11.4<br />

10.3<br />

9.9<br />

9.8<br />

9.2<br />

8.6<br />

7.3<br />

7.0 6.8<br />

7.41<br />

7.32<br />

7.29<br />

7.29<br />

7.27<br />

7.27<br />

7.34<br />

7.29<br />

7.32<br />

7.19<br />

7.20<br />

Figure 1: The ring proton resonances of benzyl alcohol<br />

are shifted upfield and broadened upon binding to<br />

lecithin membranes.<br />

Further discrimination stems from the<br />

motional restrictions imparted by the membrane<br />

environment that is reflected in the<br />

spin-spin relaxation. The ring resonances<br />

corresponding to the free and bound drug are<br />

shown in Figure 1 for a lecithin model membrane<br />

in the L_ state. The resonance of the<br />

bound agent is oroadened due to immobilization<br />

in the membrane. The T2 of the bound<br />

agent is approximately 6 msec, while the T2 of<br />

the free agent is more than three orders of<br />

magnitude greater (11 sec). The different<br />

TABLE 1<br />

(ppm from TMS)<br />

7.41<br />

7.32<br />

7.25<br />

7.24<br />

7.21<br />

7.21<br />

7.30<br />

7.29<br />

7.32<br />

7.16<br />

7.20<br />

7.41<br />

7.23<br />

7.17<br />

7.16<br />

7.12<br />

7.12<br />

7.21<br />

7.29<br />

7.32<br />

7.16<br />

7.13<br />

Bulletin of Magnetic Resonance<br />

S (ppm) avg<br />

7.41<br />

7.29<br />

7.24<br />

7.23<br />

7.20<br />

7.20<br />

7.28<br />

7.29<br />

7.32<br />

7.17<br />

7.18<br />

relaxation properties allow for spectral editing<br />

based on T2 using spin echoes.<br />

The partitioning of benzyl alcohol into<br />

membranes is modest. Consequently, the<br />

lipid to water ratios of the sample must be<br />

large to obtain accurate simulations of the<br />

broad bound resonance, while a sample size<br />

and geometry consistent with high Zeeman<br />

field homogeneity must be maintained. In<br />

practice, to reduce sample demands, an internal<br />

acetate standard was used to determine<br />

the aqueous alcohol concentration in a dilute<br />

membrane suspension. Since the lipid concentration<br />

is known, the intramembrane concentration<br />

is obtained by difference to yield<br />

the partition coefficient. To ensure that the<br />

integrated aqueous resonance is not contaminated<br />

by the broad bound resonance, a<br />

CPMG sequence is used to delay acquisition<br />

by 25 msec in order to filter the broad bounddrug<br />

resonance. This filtering method also<br />

removes the dipolar broadened lipid resonance<br />

to improve baseline definition.<br />

Since the colligative thermodynamic framework<br />

equates the action of solute and temperature<br />

through entropy, precise temperature<br />

regulation is required. The aqueous benzyl<br />

alcohol resonance exhibits a temperature<br />

dependent chemical shift. In order to maintain<br />

a consistently reproducible temperature,<br />

we have taken advantage of this temperature<br />

dependence. The chemical shift differences<br />

between the HOD and free benzyl alcohol<br />

resonances as a function of temperature is<br />

shown in Figure 2.


Vol. 14, No. 1-4<br />

Q.<br />

0.<br />

UI<br />

o<br />

cc<br />

UI<br />

u.<br />

u.<br />

a<br />

u.<br />

X<br />

in<br />

2.9<br />

2.8<br />

2.7<br />

2.6<br />

2.5<br />

2.4<br />

2 ^<br />

-<br />

-<br />

-<br />

c<br />

i 1<br />

1 , 1<br />

) 10 20<br />

Figure 2: The chemical shift difference between t<br />

benzyl alcohol ring proton resonance and the HOD resonance<br />

shows a temperature dependence.<br />

1<br />

1<br />

30<br />

TEMPERATURE (C)<br />

-<br />

_<br />

-<br />

_<br />

1<br />

40 5<br />

ANALYSIS OF PARTITIONING: The<br />

degree of anesthetic partitioning into a membrane<br />

system is sensitive to and characteristic<br />

of the state of the lipid assembly. The equilibrium<br />

constant, K , is deduced from the<br />

partitioning changes characteristic of the<br />

interchange between membrane states. The<br />

temperature dependence of the partitioning<br />

exhibits the following functional form for a<br />

state change between two membrane structures.<br />

[3]<br />

KP =<br />

C =<br />

expC(T-Tm)<br />

AHvH<br />

„<br />

RTT m<br />

(2)<br />

(3)<br />

The partition coefficient for the membrane<br />

states a and ft are Kp« and Kp", respectively.<br />

These partition coefficients are not necessarily<br />

constant and may be altered to include a<br />

temperature dependence. The total partition<br />

coefficient is Kp. The midpoint temperature<br />

is Tm. A fit of the experimental data to this<br />

function yields partition coefficients for each<br />

phase, the midpoint temperature, and the<br />

van't Hoff enthalpy (AHvH), a measure of the<br />

cooperativity of the equilibrium.<br />

295<br />

TEMPERATURE (K)<br />

Figure 3: The molal partition coefficient of benzyl alcohol<br />

into multilamellar DPPC membranes is shown as a<br />

function of temperature for two concentrations of benzyl<br />

alcohol. The fit corresponds to the theoretical multiparameter<br />

least-squares analysis described in the text. The<br />

derivative of the fit to the data is shown offset below.<br />

The percent mole fraction intramembrane benzyl alcohol<br />

concentrations at the L<br />

p,<br />

p p equilibrium<br />

midpoint are as follows: Panel A: Lp , = 0.23%, Pp ' =<br />

1.1%; Panel B: 2.2%, 16.8%; The mole fraction benzyl<br />

alcohol concentrations at the P^/ -* La equilibrium<br />

midpoint are as follows: Panel A: Pp, = 0.84%, La=<br />

2.1%; Panel B: 11.7%, 24.2%; For comparative purposes,<br />

general anesthetic intramembrane concentrations<br />

are considered less than 5 mole fraction percent.<br />

The analytical framework presented is not<br />

specific to the partitioning analysis. It is<br />

broadly applicable to any technique in which<br />

the observable is characteristic of each state.<br />

ALCOHOL ACTION IN MODEL<br />

MEMBRANES: The lecithin membrane,<br />

DPPC (l,2-dipalmitoyl-.stt-glycero-3-phosphocholine),<br />

adopts three well-studied structures<br />

or phases, a gel-structure (L«,), a ripplestructure<br />

(Pp,), and a fluid bilayer-structure<br />

(La). [9] Since the interchange between these<br />

three membrane structures is driven by


296<br />

entropy, changes in solute, temperature and<br />

pressure alter the energetic balance to favor a<br />

given structure. The Lp, •* P^/ equilibrium<br />

(pretransition) exhibits an equilibrium midpoint<br />

temperature determined by calorimetry<br />

as 34.8°C. [9] This change in state is accompanied<br />

by a small change in thermal entropy<br />

(12.5 J mol" 1 K- 1 ). The P^, -* La (main transition)<br />

exhibits an equilibrium midpoint temperature<br />

determined by calorimetry as 41.0°C.<br />

This change in state is accompanied by a relatively<br />

large change in thermal entropy (85.6 J<br />

mol' 1 K*i). [9] The large difference between<br />

the thermal entropy changes associated with<br />

these two equilibria provides a simple system<br />

in which to test the predictions of the colligative<br />

thermodynamic framework, that solute<br />

action occurs through entropy and that equilibria<br />

characterized by a small thermal<br />

entropy change should be most susceptible to<br />

perturbation.<br />

The temperature dependence of benzyl<br />

alcohol partitioning at two substantially different<br />

alcohol concentrations is shown in<br />

Figure 3. Figure 3A corresponds to benzyl<br />

alcohol concentrations below that required for<br />

general anesthesia; whereas, the concentration<br />

in Figure 3B is near that required for<br />

local anesthesia. The partition coefficients<br />

obtained by the NMR method are in excellent<br />

agreement with direct radiolabel measures.<br />

[10] Two discontinuities correspond to the<br />

two membrane equilibria. It is these changes<br />

in partitioning that provide the configurational<br />

entropy by which solutes perturb equilibria.<br />

The low entropy Le, -+ Ppr equilibrium<br />

exhibits greater sensitivity to the alcohol<br />

than the high entropy P^, -*• La, as qualitatively<br />

predicted by the thermodynamic model.<br />

The quantitative test for the model is<br />

shown in Figure 4. The partitioning method<br />

provides intramembrane solute concentrations,<br />

which, in turn, provide the magnitude of<br />

the configurational entropy imparted to each<br />

membrane structure. This contribution<br />

lowers the free energy of each state according<br />

to the magnitude of the partitioning, and<br />

thereby alters the difference in free energy<br />

and shifts the equilibrium. The experimental<br />

points are in good agreement with the theoretical<br />

predictions (represented by the lines)<br />

at dilute alcohol concentrations for which the<br />

thermodynamic treatment is derived and<br />

which corresponds to pharmacological levels<br />

Bulletin of Magnetic Resonance<br />

1 10<br />

INTRAMEMBRANE HOLE FRACTION U) DIFFERENCE<br />

Figure 4: The dependence of the equilibrium midpoint<br />

temperature of DPPC on the presence of benzyl alcohol.<br />

The benzyl alcohol intramembrane concentration difference<br />

between the initial and final states at the equilibrium<br />

midpoint is shown. Data are presented for the low<br />

entropy Lp, -+ Fg, (pretransition; filled circles) and<br />

the high entropy Fp, -*• La (main transition; open circles)<br />

equilibria. The coUigative thermodynamic predictions<br />

(eq. 1) are represented by the lines. The solid<br />

portions of the lines designate the average<br />

intramembrane concentrations at the midpoint which<br />

correspond to the range of pharmacological relevance<br />

for general anesthesia.<br />

Particularly striking is the dramatic contrast<br />

in benzyl alcohol sensitivity exhibited by these<br />

two equilibria. At average intramembrane<br />

concentrations of 5 m.f.%, the low entropy<br />

equilibrmm is perturbed by approximately<br />

12°C, whereas the high entropy equilibrium is<br />

perturbed by approximately 1°C. Not only<br />

does this observation support the predictions<br />

of the thermodynamic model, but it demonstrates<br />

that remarkably low intramembrane<br />

concentrations of nonspecific solutes can precipitate<br />

quite substantial effects upon membrane<br />

structure.<br />

ALCOHOL ACTION IN LIPOSOMES<br />

MADE FROM RATS CHRONICALLY<br />

EXPOSED TO ANESTHETICS: Rat liver<br />

microsomes obtained from rats exposed to<br />

nitrous oxide or fed ethanol were isolated and<br />

liposcmes formed from the extracted phospholipids.<br />

Shown in Figure 5 are representative<br />

benzyl alcohol partitioning traces for the<br />

liposomes prepared from the ethanol-fed and<br />

control animals. The partition coefficient into


Vol. 14, No. 1-4<br />

UJ<br />

FFIC]<br />

UJ<br />

o<br />

t i<br />

TION (<br />

IRTI<br />

«*.<br />

Q.<br />

40LAL<br />

40<br />

35<br />

30<br />

25<br />

• •! 1<br />

• •<br />

-i—' 1 ' r~ —i 1 1 r—<br />

\mr •<br />

^ T<br />

1 , 1 , 1<br />

i . \<br />

20 30 40 50 60 70<br />

TEMPERATURE<br />

Figure 5: Benzyl alcohol partitioning traces are shown<br />

for liposomes made from rat-liver microsomal-phospholipids.<br />

The samples represented in the lower trace<br />

(filled triangles) were prepared from chronically<br />

ethanol-fed rats. The samples represented in the upper<br />

trace (filled circles) were prepared from their pair-fed<br />

littermates.<br />

the control samples is larger than the treated<br />

samples. This difference in partitioning is<br />

characteristic of 'membrane tolerance'. [2] A<br />

structural equilibrium is apparent in the control<br />

samples near 37°C that is lacking in the<br />

samples obtained from treated animals. Similar<br />

results are obtained from the chronic<br />

nitrous oxide paradigm. These results evidence<br />

an adaptive response to the chronic<br />

presence of anesthetic agents that results in<br />

altered domain structure in the reconstituted<br />

system. Similarly, structural lipid domains are<br />

predicted at the anesthetic locus in our thermodynamic<br />

reformulation of the Meyer-<br />

Overton hypothesis.<br />

CONCLUSIONS: Alcohols and anesthetics<br />

act through the entropy imparted by partitioning<br />

to modify membrane architecture.<br />

Analysis of anesthetic action requires simultaneous<br />

measures of solute partitioning and<br />

membrane structure over a wide range of<br />

-<br />

-<br />

297<br />

solute concentrations. NMR partitioning<br />

methods offer unique advantages in such<br />

inquiry since the solute can serve the multifaceted<br />

role of perturbant, reporter of membrane<br />

perturbations, and reporter of solute<br />

partitioning.<br />

METHODS: Spectra were obtained on a Bruker 8.5T<br />

AM spectrometer operating at 360 MHz with deuterium<br />

lock. Lipids were dried under N2, evacuated (


7. Smith, R.L. and Oldfield, E. Science 225,<br />

280-288 (1984).<br />

8. Taraschi, T.F., Lee, Y.-C, Janes, N., and<br />

Rubin, E. Annals New York Acad. Sci.<br />

625,698-706(1991).<br />

9. Chen, S.C. and Sturtevant, J.M.<br />

Biochemistry 20. 713-718 (1981V<br />

10. Colley, CM. and Metcalfe, J.C. FEBS<br />

Letters 24,241-246 (1972).<br />

11. Ellingson, J.S., Janes, R, Taraschi, T.F.,<br />

and Rubin, E. Biochim. Biophys. Acta<br />

1062,199-205 (1991).<br />

Bulletin of Magnetic Resonance


Vol. 14, No. 1-4<br />

ABSTRACT<br />

WEAK MOLECULAR INTERACTIONS: NMR SPECTROSCOPY<br />

OF ORIENTED MOLECULES<br />

C.L. Khetrapal<br />

National Institutes of Health, Bethesda, MI)-20892, USA<br />

"Indian Institute of Science, Bangalore 560 012, India<br />

NMR spectroscopy of oriented<br />

molecules is employed to study<br />

weak molecular interactions. The<br />

information is obtained from the<br />

changes in the molecular order<br />

and the structure as a result of<br />

the complexation.<br />

Specific results on the pi<br />

and charge transfer complexes<br />

formed by iodine, chloroform, and<br />

silver nitrate with nitrogen<br />

heterocycles, aromatic systems<br />

and acetonitrile are reported. A<br />

method for the determination of<br />

the order parameter and the<br />

structure of the 'complexed'<br />

species is presented and its<br />

utility demonstrated. The use of<br />

mixture of liquid crystals of<br />

opposite diamagnetic anisotropies<br />

to investigate the extraordinary<br />

symmetry of Buckminster<br />

Fullerene , Cgo> 1S illustrated.<br />

INTRODUCTION<br />

Application of NMR to<br />

weak molecular interactions<br />

study<br />

from<br />

changes in chemical shifts,<br />

indirect spin-spin coupling<br />

constants and line width is quite<br />

well known and has been employed<br />

since the early days of NMR.<br />

However, the use of direct<br />

dipolar couplings/order parameters<br />

derived from the spectra<br />

of molecules oriented in thermotropic<br />

liquid crystals is relatively<br />

new (1) and provides more<br />

quantitative results since<br />

changes in molecular structure if<br />

any as a result of complexation<br />

* Address for Correspondence<br />

can also be determined directly<br />

and used for such a purpose.<br />

Results on the pi and the<br />

charge transfer complexes formed<br />

by chloroform, iodine, silver<br />

nitrate with aromatic systems,<br />

nitrogen heterocycles and acetonitrile<br />

are described in this<br />

communication. A study of<br />

Buckminster Fullerene (C, )<br />

oriented in mixture of liquid<br />

crystals of opposite diamagnetic<br />

anisotropies is also reported in<br />

order to establish the extraordinary<br />

symmetry of the molecule<br />

indicating, thereby, negligible<br />

solvent-solute interactions.<br />

2. METHODOLODY<br />

Information on weak molecular<br />

complexes has been obtained<br />

from the changes in the degree of<br />

order as well as the molecular<br />

structure produced as a result of<br />

the complex formation.<br />

The degree of order of a<br />

molecule dissolved in a "hematic<br />

liquid crystal usually decreases<br />

with the increase of temperature<br />

or concentration. Any abnormal<br />

change in the degree of order is<br />

attributed to the formation of<br />

the complex(es). Recently, we<br />

have investigated metal ion<br />

ligand interactions between monovalent<br />

metal ions such as Li+ and<br />

Ag+ in LiCIO , LiBF and AgNC<br />

with ligands like aoetonitri1e?<br />

dimethyl sulphoxide, pyridine and<br />

acetone in thermotropic liquid<br />

crystals (2-4). Interactions of<br />

iodine (and bromine) with<br />

299


300 Bulletin of Magnetic Resonance<br />

pyridine (4-6), pyrimidine (7)<br />

and quinazoline (8) have also<br />

been investigated. Work on chloroform-benzene<br />

pi complexes has<br />

also been undertaken. We have<br />

also employed the use of mixture<br />

of liquid crystals of opposite<br />

diamagnetic anisotropies in order<br />

to find out if any detectable<br />

distortions in the spherical symmetry<br />

are present in Buckminster<br />

Fullerene , C 59, as result of<br />

solvent-solute interactions (9).<br />

Some such results are reported<br />

below briefly.<br />

LiCIO •-! igand<br />

1igands used in interactions:<br />

(a)<br />

The 1 igands "'used in such studies<br />

are dimethyIsulphoxide, acetonitrile<br />

and pyridine. The liquid<br />

crystals employed are Schiff<br />

bases such as N-(p-methoxy (or<br />

ethoxy) benzylidene)-p-n-butyl<br />

aniline (MBBA or EBBA) and their<br />

deuterated (-d2) analogue with<br />

deuterium substituted at position<br />

ortho to the -N= group in the<br />

butyl aniline ring. The results<br />

indicate ' the formation of two<br />

types of complexes in these cases<br />

with one having "isotropic-1ike"<br />

structure. The "isotropic-1ike"<br />

complex may be of the type ML4<br />

where M is the metal ion and L is<br />

the 1igand and the other could be<br />

the one containing different<br />

ratios<br />

gands.<br />

of the metal ions to li-<br />

(b) LiBF -Ligand solutions: The<br />

1igand used in such studies was<br />

acetonitrile. The liquid crystals<br />

employed were trans, trans-4-npropyl(1,1'-bicyclohexyl)-4'carbonitrile<br />

(ZLI-1184) and<br />

trans-4-pentyl-1-(4-cyanophenyl)<br />

cyclohexane (S-1114). The results<br />

indicate the formation of tetrahedral<br />

1ithium-acetonitrile complexes<br />

and the coordination of<br />

lithium from LiBF^ and LiClC^<br />

with the cyano group of ZLI-1184.<br />

No evidence was found to support<br />

either lithium ion complexation<br />

with S-1114 or the tetrafluorobo-<br />

rate moiety in these systems.<br />

(c) AgJNO3-ligand complexes: In<br />

this case AgNC^-CI^CN (CD3CN)<br />

complexes have been investigated<br />

in two different liquid crystals,<br />

namely, ZLI-1167 (a ternary eutectic<br />

mixture of propyl, pentyl<br />

and heptyl bicyclohexyl carbonitrile)<br />

and S-1114 from both the<br />

proton and the deuteron NMR<br />

studies. The results indicate the<br />

formation of different types of<br />

AgNO -CHoCN complexes, i.e., ML4,<br />

ML 3.?..etc.<br />

(d) Benzene-chloroform complex:<br />

It has been studied from the<br />

reduction of molecular order of<br />

benzene upon addition of iodine.<br />

From the shape considerations,<br />

the order parameter of such a<br />

complex should be either opposite<br />

in sign to or smaller in magnitude<br />

than that of benzene under<br />

comparable conditions. The observed<br />

spectrum of benzene in<br />

ZLI-1167 containing iodine which<br />

arises from the average orientation<br />

of the "complexed" and the<br />

"free" benzene, therefore, has an<br />

order parameter which is smaller<br />

in magnitude than that of the'<br />

"uncomplexed" benzene. The change<br />

of the chemical shift of the<br />

benzene carbon is also consistent<br />

with this observation.<br />

(e) Iodine(bromine)-pvridine<br />

charge transfer complex: Such<br />

complexes using NMR spectroscopy<br />

of oriented systems were first<br />

studied in 1973 and 1983 (5, 6)<br />

from the drastic change of the<br />

order parameter of the C2~axis of<br />

symmetry of pyridine. We have<br />

however, detected the formation<br />

of the "inner" complexes<br />

[(PYlfl"] as well as those of<br />

the type PY.I2 unambiguously. The<br />

order parameters of the<br />

"complexed" species have also<br />

been determined (4).


Vol. 14, No. 1-4<br />

(f) Extraordi nary symmetry<br />

Buckmi nster Fullerene. C,_<br />

of<br />

in<br />

REFERENCES<br />

nematic solutions: The "^ C-NMR<br />

spectrum of C6o was studied in a<br />

mixture of nematic liquid<br />

crystals of opposite diamagnetic 1.<br />

anisotropies. The method makes<br />

use of the change in the anisotropic<br />

parameters resulting from<br />

the switching of the order parameters<br />

at the critical point in<br />

the mixture; the change is by a 2.<br />

factor of 2 or -1/2 depending<br />

upon the direction of approach of<br />

the critical point (10). By an<br />

appropriate adjustment of the<br />

concentration and temperature, it<br />

is possible to observe both the 3.<br />

types of orientations to coexist.<br />

Even in tetrahedral molecules<br />

such as methane and tetramethysilane,<br />

the coexistence of 4.<br />

the two spectra has been observed<br />

(11). On the other hand, the<br />

observation of a single line at<br />

the same position as in the isotropic<br />

phase at the critical 5.<br />

point where the coexistence of<br />

the two spectra in methane or<br />

tetramethylsi 1ane is observed,<br />

indicates the absence of any 6.<br />

detectable distortions in the<br />

molecule. This has actually been<br />

observed in a solution of C6Q in<br />

a 1:1 mixture of liquid crystals<br />

ZLI-1167 and S-1114 containing 7.<br />

tetramethylsilane, at 332.4 K.<br />

The spectrum clearly shows the<br />

coexistence of two spectra for<br />

tetramethylsi 1ane whereas for the 8.<br />

C60 only a single line at its<br />

isotropic position (140.37 ppm<br />

with respect to TMS) is observed.<br />

The- results, therefore, establish 9.<br />

that there are no detectable<br />

distortions from spherical symmetry<br />

in C in the nematic solvents.<br />

TcP® our knowledge this is<br />

the first molecule where no detectable<br />

distortation from spherical<br />

symmetry have been observed.<br />

It should, therefore, serve<br />

as an ideal reference for the 10<br />

study of the chemical shift<br />

sotropy.ani-<br />

301<br />

P. Diehl, H.R. Wasser, G.A.<br />

Nagana Gowda, N. Suryaprakash<br />

and C.L. Khetrapal,<br />

Chem. Phys. Lett. 159. 199<br />

(1989)<br />

G.A. Nagana Gowda, N. Suryaprakash,<br />

R.G. Weiss, C.L.<br />

Khetrapal and P. Diehl,<br />

Magn. Reson. Chem. .28, 642<br />

(1990)<br />

G.A. Nagana Gowda, R.G.<br />

Weiss and C.L. Khetrapal,<br />

Liq. Cryst. 10, 659 (1991)<br />

C.L. Khetrapal, G.A. Nagana<br />

Gowda and N. Suryaprakash,<br />

Spect. Chim. Acta. (in the<br />

press)<br />

C.A. Veracini, M. Longeri<br />

and P.L. Barili, Chem. Phys.<br />

Lett. 19., 592 (1973)<br />

D. Catalano, C.A. Veracini,<br />

P.L. Barili and M. Longeri,<br />

J. Chem. Soc. Perk Trans II,<br />

171, 1983<br />

N. Suryaprakash, R. Ugolini<br />

and P. Diehl, Magn. Reson.<br />

Chem. 29., 1024 (1991 )<br />

N. Suryaprakash, C.L.<br />

Khetrapal and P. Diehl (To<br />

be published)<br />

C.N.R. Rao, T. Pradeep, R.<br />

Seshadri, R. Nagarajan, V.N.<br />

Murthy, G.N. Subbanna, F.<br />

D'Souza, V. Krishnan, G.A.<br />

Nagana Gowda, N. Suryaparakash,<br />

C.L. Khetrapal and<br />

S.V. Bhat, Ind. J. Chem. 31.<br />

F5 (1992).<br />

C.L. Khetrapal and A.C.<br />

Kunwar, Chem. Phys. Lett.<br />

.82, 170 ( 1981 ) .


302 Bulletin of Magnetic Resonance<br />

11. C.L. Khetrapal, A.C. Kunwar<br />

and M.R. Lakshminarayana,<br />

Mol. Cryst. Liq. Cryst. 111.<br />

189 (1984).


Vol. 14, No. 1-4 303<br />

Introduction:<br />

Structural Studies of Collagen by Solid State NMR.<br />

Richard J. Wittebort and Anne Marie Clark<br />

University of Louisville, Department of Chemistry,<br />

Louisville, KY 40292 USA<br />

Collagen is one of the most abundant<br />

proteins nature and has the same structure and<br />

virtually the same composition for diverse<br />

species. It provides the organic matrix for teeth<br />

and bones, and gives tendons and blood<br />

vessels their strength. Collagen has three<br />

polypeptide chains, two al chains and one a2<br />

chain, coiled together in a 3-1 helix. Each<br />

chain is formed from a repeating (Gly-X-Y)<br />

triad with Gly always in the first position. The<br />

other two positions can have any amino acid<br />

but frequently (about 30%) have proline (Pro)<br />

and hydroxyproline (Hyp). Hyp always<br />

occurs in position 3 while Pro usually occurs in<br />

position 2 and only rarely in position 3 in adult<br />

animals. The principle difference between the<br />

types of chains is their direction of orientation.<br />

Several models have been developed to<br />

explain the three dimensional structure of<br />

collagen. From steric considerations, it is<br />

thought that the collagen triple helix has Gly<br />

alpha hydrogens pointing inside. Despite<br />

numerous attempts to solve the X-ray structure,<br />

collagen's lack of long range order has made it<br />

impossible to determine a unique structure from<br />

diffraction methods. 1 " 5<br />

Collagen is a uniaxially ordered fiber<br />

with helical symmetry about the fiber axis. In<br />

order to do structural studies by solid state<br />

NMR, the fiber axis must be aligned with the<br />

magnetic field. Opella used solid state<br />

techniques to determine the three dimensional<br />

structure of bactreiophage fd coat protein. 6<br />

This experiment requires at least one direction<br />

of orientation and the molecular sites of interest<br />

must be immobilized and uniformly oriented<br />

with respect to the magnetic field. In single<br />

crystals, any arbitrary sample orientation can be<br />

studied; however, for uniaxially oriented -<br />

samples the direction of orientation must be<br />

along the applied field to obtain single crystal-<br />

like spectra. The bacteriophages are well suited<br />

for this type of experiment. They<br />

spontaneously align with the magnetic field and<br />

their large size and rod-like shape immobilize


304<br />

the protein subunits. The alpha helical<br />

structure determined agrees well with X-ray<br />

structures of alpha helical proteins.<br />

The three dimensional structure of<br />

polypeptides can be described as a series of<br />

connected peptide planes. By selectively<br />

labelling specific sites and determining their<br />

orientation, structural constraints will be added<br />

to the three dimensional structure of collagen.<br />

The frequency of resonance lines in<br />

solid state NMR spectra depend on the<br />

orientation of the local molecular environment<br />

relative to the magnetic field Ho. The observed<br />

splitting can be used to determine the angle a<br />

bond makes with the magnetic field by the<br />

following equation, assuming an axially<br />

symmetric interaction:<br />

Av = vn(3cos 2 6-1) Eqn. 1<br />

Av is the observed splitting, 0 is the angle<br />

between the bond and the applied field and vn<br />

is the quadrupolar coupling constant. It has<br />

been found that in order to determine the<br />

orientation of a peptide plane at least two labels<br />

per plane must be examined.<br />

Experimental:<br />

The solenoidal coil geometry typically<br />

used in solid state NMR probes is<br />

Bulletin of Magnetic Resonance<br />

unsatisfactory since a bundle of fibers is only<br />

conveniently placed in the coil in the<br />

perpendicular orientation. Attempts were made<br />

to wrap tendons around several small slides<br />

and stack them in the sample holder but<br />

satisfactory alignment was not obtained. A<br />

different probe design is called for by this<br />

experiment. One probe design is based on a<br />

small flat coil geometry suggested by Opella. 7<br />

Small, flat frames were made for us upon<br />

which to wind the coil. A series of rat-tail<br />

tendons were tied together and wrapped around<br />

a 1 x 1 cm polycarbonate card. The card was<br />

then placed in the coil frame and the coil wound<br />

around it. This arrangement consistently gives<br />

a 90 degree pulse width of 2.7 JIS at 1 kW<br />

power. This design has the advantage of<br />

allowing the fiber axis to be placed at any angle<br />

relative to Ho. Proper turn spacing for good<br />

RF field homogeneity and consistent<br />

inductance is insured by using a frame with set<br />

holes. This also allows the sample to be kept<br />

in the center of the coil, away from the edges<br />

where the field homogeneity is poor. We<br />

obtained satisfactory results with this probe.<br />

We were, however, limited to a small sample


Vol. 14, No. 1-4 305<br />

size due to the polycarbonate cards and<br />

resultant poor filling factor. We had to retain<br />

the cards to keep the fibers aligned; otherwise,<br />

they shrink during changes in humidity and<br />

temperature.<br />

We are principally interested in<br />

structural information obtainable from samples<br />

with the fiber axis along Ho, that is, the parallel<br />

orientation. From that premise, we designed a<br />

probe using a Helmhotz coil similar to liquid<br />

NMR probes. The probe was constructed<br />

based on our circuit for a double resonance<br />

design with a 0.5 x 1 cm saddle. The probe<br />

has a 90 pulse width of 3.2 microseconds with<br />

400 watts power at the 2 H frequency<br />

(38.8 MHz). This probe is capable of holding<br />

ten times the sample as the flat probe and has<br />

provisions for stretching the sample.<br />

We have labelled selected sites in the<br />

collagen molecule by incorporating deuterated<br />

amino acid into the rat tail tendon. Torchia<br />

incorporated labelled glycine into 1/3 of the Gly<br />

positions in rat tail tendon by injecting it into<br />

rats. 8 We slightly modified his injection<br />

scheme in order to label selected amino acids in<br />

rat tail tendon. Given the low natural<br />

abundance of deuterium, we were confident<br />

that the only signal we would see would be<br />

from our covalently bound label.<br />

The first covalently labelled position we<br />

tried was alpha-d-Pro. The solutions for<br />

injection were 1.3 M in the amino acid and<br />

were neutralized with NaOH. Ten rat pups<br />

were injected once a day for 21 days and then<br />

sacrificed. The rat tail tendon was extracted for<br />

use in our experiments. The degree of<br />

incorporation of d-Pro in the tendon was<br />

determined by GC/MS and found to be 5.4%<br />

which agrees well with our NMR results. No<br />

incorporation of deuterium in any amino acids<br />

other than Pro and Hyp was seen. The<br />

oriented fibers were run on the Helmholtz coil<br />

probe. A splitting of Av=l 17 kHz was<br />

observed and the angle of the C-D bond relative<br />

to the magnetic field determined.<br />

The glycine amides were labelled by<br />

exchange. Samples were heated at 40°C and at<br />

constant humidity to completely exchange the<br />

labile H's. At 78% relative humidity, exchange<br />

reaches a constant level after 24 hours, at 38%<br />

relative humidity three days are required to<br />

reach a constant level of exchange. The<br />

samples were cooled to room temperature at<br />

constant humidity for an hour and then back


306<br />

exchanged in liquid H2O. Samples that were<br />

not heated to label the amide positions showed<br />

an extremely rapid loss of the solid echo peak.<br />

The heated sample, after back-exchange, had<br />

10% of the solid echo remaining, indicating the<br />

exchanged sample has a significant amount of<br />

bound water. The doublet splitting of 155 kHz<br />

shows that the hydrogen bonded amides are in<br />

fact nearly perpendicular to the fiber axis in<br />

agreement with the predictions from the X-ray<br />

structure. Recent packing studies predict the<br />

collagen molecule is tilted 4 to 5 degrees off the<br />

fiber axis.<br />

Results and Conclusions:<br />

Bulletin of Magnetic Resonance<br />

The asymmetric lineshape of the back<br />

exchanged sample indicates a distribution of<br />

orientation about 6=89 degrees. The spectral<br />

simulation for a Gaussian distribution of<br />

orientations P(6) a exp(-6 2 /2a) with 6=90 and<br />

o=17 matches our experimental data quite well.<br />

We have successfully labelled selected<br />

positions and determined their orientation<br />

relative to the fiber axis. By carefully choosing<br />

our labels, we will be able to determine the<br />

orientation of various peptide planes in collagen<br />

and gradually build the three-dimensional<br />

structure.<br />

Table I. Angle between X-D bond and the fiber axis comparison of solid state NMR results and<br />

various models derived from X-ray crystallography data.<br />

Deuteron Experiment Fraser's 4 Ramachandran's 1 Ramachandran's 2 Yonath's 5<br />

Rich-Crick bridging water two bonded<br />

ProCa-D 90<br />

GlyN-D 89<br />

References:<br />

81.3<br />

89.4<br />

1 Ramachandran, G. N. and Kartha, G.,<br />

Nature, 1955, 776, 593.<br />

2 Ramachandran, G. N. and<br />

Chandrasekaran, R., Biopolymers, 1968,<br />

6,1649.<br />

3 Rich, A. and Crick, F. H. C, Nature,<br />

1955,276,915.<br />

4 Fraser, R. D. B., /. Mol Biol., 1979,<br />

129, 463.<br />

73.8<br />

78.7<br />

73.1<br />

78.4 83.1<br />

5 Yonath, A. and Traub, W., J.Mol. Biol,<br />

1969,43,461.<br />

6 Opella, S. J., Quart. Rev. Biophys.,<br />

1987,79,7.<br />

7 Bechinger, B. and Opella, S. J., /. Magn.<br />

Reson., 1991, 95, 585.<br />

8 Jelinski, L. W. and Torchia, D. A., /.<br />

Mol. Biol, 1979,133, 45.


Vol. 14, No. 1-4 307<br />

Calender of Forthcoming<br />

Conferences in Magnetic<br />

Resonance<br />

March 8-14, 1993<br />

1993 Keystone Symposia on Molecular & Cellular<br />

Biology, Taos, New Mexico, USA<br />

Frontiers of NMR in Molecular Biology - III;<br />

Organizers: Thomas L. James, Stephen W. Fesik<br />

and Peter E. Wright.<br />

Information:<br />

Keystone Symposia<br />

Drawer 1630<br />

Silverthorne, CO 80498<br />

Phone: 303-262-1230<br />

March 14-18, 1993<br />

Experimental Nuclear Magnetic Resonance Conference,<br />

The Adam's Mark Hotel, St. Louis, Missouri,<br />

USA<br />

There will be sessions covering high resolution<br />

and multi-dimensional NMR in liquids, materials<br />

imaging, biological imaging, hardware, NMR<br />

in solids, calculations and data processing, and that<br />

old standby, miscellaneous. The deadline for submitting<br />

poster abstracts will be December 30,<br />

1992.<br />

Contact:<br />

ENC<br />

815 Don Gaspar<br />

Santa Fe, New Mexico 87501<br />

Phone: 505-989-4573<br />

FAX: 505-989-5073<br />

April 1993<br />

High Resolution NMR Spectroscopy (a residential<br />

school), University of Sheffield, England<br />

For information contact:<br />

Ms. L. Hart<br />

The Royal Society of Chemistry<br />

Burlington House<br />

Piccadilly, London W1V 0BN<br />

England<br />

Tel: 071-437-8656<br />

September 6-10, 1993<br />

Second International Conference on Magnetic<br />

Resonance Microscopy, Heidelberg, Germany<br />

The program includes plenary lectures, as well<br />

as oral and poster contributions selected from the<br />

submitted papers. The papers will be judged solely<br />

on the basis of an abstract. Preregistration deadline:<br />

April 30, 1993.<br />

Organization and Program: Winfried Kuhn<br />

(St. Ingbert), Bernhard Blumich (Mainz).<br />

International Advisory Board: J. L. Ackerman<br />

(Charlestown), L. J. Berliner (Columbus),<br />

P. T. Callaghan (Palmerston), W. Edelstein<br />

(Schenectady), A. N. Garroway (Washington),<br />

A. Haase (Wiirzburg), W. E. Hull (Heidelberg),<br />

L. W. Jelinski (Ithaca), J. L. Koenig (Cleveland),<br />

G. A. Johnson (Durham), P. Jonson (London),<br />

P. C. Lauterbur (Urbana), P. Mansfield (Nottingham),<br />

B. Maraviglia (Rome), G. D. Mateescu<br />

(Cleveland), J. M. Pope (Kensington), V. Sarafis<br />

(Richmond), S. Sarkar (King of Prussia)<br />

For further information please write to:<br />

Dr. Winfried Kuhn<br />

Fraunhofer Institute<br />

Ensheimer Str. 48<br />

DW-6670 St. Ingbert, Germany<br />

phone: +49-6894-89738<br />

FAX: +49-6894-89750<br />

or<br />

Dr. Bernhard Blumich<br />

Max-Planck Institute for Polymer Research<br />

Postfach 3148<br />

D-6500 Mainz, Germany<br />

phone: +49-6131-379125<br />

FAX: +49-6131-379100<br />

The editor would be pleased to receive<br />

notices of future meetings in the field of<br />

magnetic resonance so that they could be<br />

recorded in this column.


308<br />

Recent Magnetic Resonance Books<br />

1 Magnetic Resonance Spectroscopy in Biology<br />

and Medicine (1992). Edited by J. D. De Certaines,<br />

W. M. M. J. Bovee and F. Podo. Contents: Presents<br />

the experimental and basic aspects of functional and<br />

pathological tissue characterization of MRS. A balance<br />

is drawn between the basic science, practical<br />

technologies and biomedical applications. Covers<br />

recent developments in the field: localization, 2D<br />

NMR, spectroscopic imaging, data quantification<br />

and quality assessment, as well as the basic principles<br />

of magnetic resonance spectroscopy. Pergamon<br />

Press, ISBN 0-08-0410170 (flexicover) $70.00; ISBN<br />

0-08-0410189 (hardcover) $170.00.<br />

1 In Vivo Magnetic Resonance Spectroscopy I.<br />

Probeheads and Radiofrequency Pulses, Spectrum<br />

Analysis (1992). Edited by M. Rudin, Springer, 345<br />

pp. ISBN 0-387-54547-6 (hardcover) $119.00.<br />

l In Vivo Magnetic Resonance Spectroscopy II.<br />

Localization and Spectral Editing (1992). Edited by<br />

M. Rudin and J. Seelig, Springer, 368 pp. ISBN<br />

0-387-55022-4 (hardcover) $119.00.<br />

^In Vivo Magnetic Resonance Spectroscopy III.<br />

In Vivo MR Spectroscopy: Potential and Limitations<br />

(1992). Edited by M. Rudin and J. Seelig,<br />

Springer, 293 pp. ISBN 0-387-55029-1 (hardcover)<br />

$98.00.<br />

1 Annual Reports on NMR Spectroscopy. Volume<br />

24 (1992). Contents: Developments in solid state<br />

NMR. Solid state NMR imaging. NMR studies of<br />

interfacial phenomena. NMR measurements of intracellular<br />

ions in living systems. 199Hg NMR parameters.<br />

Applications of NMR methods in coal<br />

research.<br />

1 Progress in Nuclear Magnetic Resonance Spectroscopy.<br />

Volume 24 No. 3(1992). Contents: Structural<br />

characterization of noncrystalline solids and<br />

glasses using solid state NMR.<br />

1 Progress in Nuclear Magnetic Resonance Spectroscopy.<br />

Volume 24 No. 2 (1992). Contents: 129 Xe<br />

J New additions to the list.<br />

Bulletin of Magnetic Resonance<br />

NMR as a probe for the study of microporous solids:<br />

A critical review. Simulation of 2D NMR spectra for<br />

determination of solution conformations of nucleic<br />

acids.<br />

Progress in Biophysics & Molecular Biology.<br />

Volume 57 No. 1 (1992). Contents: ENDOR and<br />

EPR of metalloproteins. Free energy transduction<br />

in polypeptides and proteins based on inverse temperature<br />

transitions.<br />

Progress in Nuclear Magnetic Resonance Spectroscopy.<br />

Volume 24 No. 1 (1992). Contents: 13 C<br />

NMR spectroscopy of oleanane triterpenoids.<br />

NMR at Very High Field (1991). Guest editor:<br />

J. B. Robert, Springer, 168 pp. ISBN 0-387-52946-2<br />

(hardcover) $79.00.<br />

1 Transition Metal Nuclear Magnetic Resonance<br />

(1991). Edited by P. S. Pregosin. Contents: The<br />

book contains a collection of review articles concerned<br />

with measuring, understanding and using the<br />

nuclear magnetic resonance spectra of the metals of<br />

Groups 3-12. The reader is provided with a view<br />

on how these nuclei are currently being approached,<br />

and what information can be obtained. The authors<br />

have liberally reproduced spectra as well as correlations<br />

relating metal-NMR data to different physical<br />

characteristics of their molecules. 364 pp. ISBN<br />

0-444-88176-X $169.00.<br />

Chemical Reviews. Volume 91 No. 7 (1991).<br />

Contents: Low-temperature solid-state NMR of proteins.<br />

Structure and dynamics of solid polymers<br />

from 2D- and 3D-NMR. NMR under high gas pressure.<br />

Nuclear magnetic resonance at high temperature.<br />

Gas-phase NMR spectroscopy. Selective<br />

excitation in high-resolution NMR. Application of<br />

the linear prediction method to NMR spectroscopy.<br />

High-resolution fluorine-19 magnetic resonance of<br />

solids. NMR determination of enantiomeric purity.<br />

Solid-state NMR studies of molecular sieve<br />

catalysis. Pulsed electron-nuclear double resonance<br />

methodology. Multidimensional NMR and data processing.<br />

One- and two-dimensional high-resolution<br />

solid-state NMR studies of zeolite lattice structures.<br />

Solid-state NMR studies of DNA structure and dynamics.<br />

Spin-lattice relaxation of coupled nuclear


Vol. 14, No. 1-4 309<br />

spins with applications to molecular motion in liquids.<br />

Progress in Nuclear Magnetic Resonance Spectroscopy.<br />

Volume 23 No. 2 (1991). Contents: Solvent<br />

signal suppression in NMR.<br />

Progress in Nuclear Magnetic Resonance Spectroscopy.<br />

Volume 23 No. 3 (1991). Contents: Modern<br />

methods of NMR data processing and data evaluation.<br />

X H NMR magic angle spinning (MAS) studies<br />

of heterogenous catalysis.<br />

NMR - Basic Principles and Progress. Volume<br />

23: Deuterium and Shift Calculation (1991).<br />

Eds.: P. Diehl, E. Fluck, H. Giinther, R. Kosfeld,<br />

J. Seelig. Contents: M.L. Martin, G.J. Martin,<br />

Nantes, France: Deuterium NMR in the Study of<br />

Site-Specific Natural Isotope Fractionation (SNIF-<br />

NMR); H.-H. Limbach, Freiburg, FRG: Dynamic<br />

NMR Spectroscopy in the Presence of Kinetic Hydrogen/Deuterium<br />

Isotope Effects; W. Kutzelnigg,<br />

U. Fleischer, M. Schindler, Bochum, FRG: The<br />

IGLO-Method: Ab-initio Calculation and Interpretation<br />

of NMR Chemical Shifts and Magnetic Susceptibilities.<br />

Approx. 270 pp. 92 figs. 45 tabs.<br />

ISBN 3-540-52949-7.<br />

NMR - Basic Principles and Progress. Volume<br />

24: High Pressure NMR (1991). Eds.: P. Diehl,<br />

E. Fluck, H. Giinther, R. Kosfeld, J. Seelig, J.<br />

Jonas, University of Illinois, Urbana, IL (Guest-<br />

Ed.). Contents: D. Brinkmann, Zurich, Switzerland:<br />

Solid-State NMR Studies at High Pressure;<br />

K.O. Prins, Amsterdam, The Netherlands: High<br />

Pressure NMR Investigations of Motion and Phase<br />

Transitions in Molecular Systems; J. Jonas, Urbana,<br />

IL: High Pressure NMR Studies of the Dynamics<br />

in Liquids and Complex Systems; E.W. Lang, H.-<br />

D. Liidemann, Regensburg, FRG: High Pressure<br />

NMR Studies on Water and Aqueous Solutions;<br />

J.W. Akitt, A.E. Merbach, Lausanne, Switzerland:<br />

High Resolution Variable Pressure NMR for Chemical<br />

Kinetics; H. Yamada, Kobe, Japan: Glass Cell<br />

Method for High-Pressure, High-Resolution NMR<br />

Measurements. Applications to the Studies of Pressure<br />

Effects on Molecular Conformation and Structure.<br />

Approx. 270 pp. 148 figs. 28 tabs. ISBN<br />

3-540-52938-1.<br />

NMR - Basic Principles and Progress. Volume<br />

25: NMR at Very High Field (1991). Eds.: P. Diehl,<br />

E. Fluck, H. Giinther, R. Kosfeld, J. Seelig, J.B.<br />

Robert, CNRS, Grenoble,-France (Guest-Ed.). Contents:<br />

R. Freeman, Cambridge, UK, J.B. Robert,<br />

Grenoble, France: A Brief History of High Resolution<br />

NMR; E.W. Bastiaan, C. MacLean, Amsterdam,<br />

The Netherlands: Molecular Orientation<br />

in High-Field High-Resolution NMR; D. Canet,<br />

Vandoeuvre-les-Nancy, France, J.B. Robert, Grenoble,<br />

France: Behaviour of the NMR Relaxation Parameters<br />

at High Fields; D. Marion, Orl ans, France:<br />

Structural Studies of Biomolecules at High Field; U.<br />

Haeberlen, Heidelberg, FRG: Solid State NMR in<br />

High and Very High Magnetic Fields. Approx. 175<br />

pp. 44 figs. 10 tabs. ISBN 3-540-52946-2.<br />

Modern NMR Techniques and Their Application<br />

in Chemistry (Practical Spectroscopy Series Volume<br />

11). Edited by Alexander I. Popov and Klaas Hallenga,<br />

Marcel Dekker, Inc. (1991). ISBN 0-8247-<br />

8332-8<br />

Annual Reports of NMR Spectroscopy. Volume<br />

23 (1991). Contents: NMR studies of isolated spin<br />

pairs in the solid state. The oxidation-state dependence<br />

of transition-metal shieldings. The Cinderella<br />

nuclei. Permutation symmetry in NMR relaxation<br />

and exchange. Nuclear spin relaxation in organic<br />

systems and solutions of macromolecules and aggregations.<br />

NMR of coals and coal products.<br />

Progress in Nuclear Magnetic Resonance Spectroscopy.<br />

Volume 23 No. 1 (1991). Contents:<br />

Nuclear magnetic resonance imaging in the solid<br />

state. Applications of three-and four-dimensional<br />

heteronuclear NMR spectroscopy to protein structure<br />

determination. Angiography and perfusion<br />

measurements by NMR.<br />

EPR Imaging and in vivo EPR (1991). Edited<br />

by Gareth R. Eaton, Sandra S. Eaton, and Keiichi<br />

Ohno, CRC Press, Boca Raton, FL. 320 pages,<br />

$89.95, ISBN: 0-8493-4923-0.<br />

Basic One-and Two-dimensional NMR Spectroscopy<br />

by Horst Friebolin (1991). VCH, New York.<br />

344 pages.


310<br />

Advances in Magnetic and Optical Resonance<br />

Volume 16 (1991). Contents: Laser excitation and<br />

detection of magnetic resonance. Deuterium relaxation<br />

in molecular solids. On the growth of multiple<br />

spin coherences in NMR of solids.<br />

1 Progress in Biophysics & Molecular Biology<br />

Volume 56 No. 1 (1991). Contents: An evaluation<br />

of computational strategies for use in the determination<br />

of protein structure from distance constraints<br />

obtained by nuclear magnetic resonance.<br />

1 Radiospectroscopy of Natural Substances by B.<br />

F. Alekseev, Y. V. Bogachev, V. Z. Drapkin, A. S.<br />

Serdjuk, N. B. Strakhov and S. G. Fedin, Engl. Tr.<br />

Norell Pr., New Jersey, 1991.<br />

1 Electron Paramagnetic Resonance of Exchange<br />

Coupled Systems by A. Bencini and D. Gatteschi,<br />

Springer Verlag, Berlin, 1990.<br />

Modern Pulsed and Continuous Wave Electron<br />

Spin Resonance by L. Kevan and M. K. Bowman<br />

(1990). Wiley, New York.<br />

1 Transition Ion Electron Paramagnetic Resonance<br />

by J. R. Pilbrow, Clarendon Press, Oxford,<br />

1990.<br />

1 Electron Paramagnetic Resonance of Exchange<br />

Coupled Systems by A. Bencini and D. Gattechi<br />

(1990). Springer, 287 pp. ISBN 0-387-50944-5<br />

(hardcover) $83.00.<br />

1 Isotope Effects in NMR Spectroscopy by S.<br />

Berger, J. M. Risley, N. M. Sergeyev and<br />

R. L. Van Etten (1990). Springer, 173 pp. ISBN<br />

0-387-51286-1 (hardcover) $83.00.<br />

17 O NMR Spectroscopy in Organic Chemistry<br />

(1990). Edited by David W. Boykin. This book provides<br />

a comprehensive review of the application of<br />

17 O NMR spectroscopy to organic chemistry. Topics<br />

include the theoretical aspects of chemical shift,<br />

quadrupolar and J coupling; 17 O enrichment; the<br />

effect of steric interactions on 17 O chemical shifts of<br />

functional groups in flexible and rigid systems; the<br />

additions to the list.<br />

Bulletin of Magnetic Resonance<br />

application of 17 O NMR spectroscopy to hydrogen<br />

bonding investigations; mechanistic problems in organic<br />

and bioorganic chemistry; and 17 O NMR spectroscopy<br />

of oxygen monocoordinated to carbon in<br />

alcohols, ethers, and derivatives. CRC Press, Inc.,<br />

Florida. ISBN: 0-8493-4867-6.<br />

Advances in Magnetic and Optical Resonance.<br />

Volume 15 (1990). Contents: Iterative methods in<br />

the design of pulse sequences for NMR excitation.<br />

Electron-nuclear polarization transfer in the nuclear<br />

rotating frame. Multipole NMR. Solid state and solution<br />

NMR of nonclassical transition metal polyhydrides.<br />

Low-frequency magnetic resonance with<br />

a dc SQUID.<br />

Advances in Biophysical Chemistry. Volume<br />

1 (1990). Contents: Stable-isotope-assisted protein<br />

NMR spectroscopy in solution.<br />

31 P and<br />

1<br />

H two-dimensional NMR and NOESY-distance<br />

restrained molecular dynamics methodologies for<br />

defining sequence-specific variations in duplex<br />

oligonucleotides: A comparison of NOESY two-spin<br />

approximation and the relaxation matrix analyses.<br />

NMR study of B- and Z-DNA hairpins of d[(CG)3]<br />

in solution. Molecular dynamics simulations of carbohydrate<br />

molecules. Diversity in the structure of<br />

hemes.<br />

Biological Magnetic Resonance. Volume 9<br />

(1990). Contents: Phosphorus NMR of membranes.<br />

Investigation of ribosomal 5S ribonucleic acid solution<br />

structure and dynamics by means of highresolution<br />

nuclear magnetic resonance spectroscopy.<br />

Structure determination via complete relaxation<br />

matrix analysis (CORMA) of two-dimensional nuclear<br />

overhauser effect spectra: DNA fragments.<br />

Methods of proton resonance assignment for proteins.<br />

Solid-state NMR spectroscopy of proteins.<br />

Methods for suppression of the H2O signal in proton<br />

FT/NMR spectroscopy: A review.<br />

Progress in Nuclear Magnetic Resonance Spectroscopy.<br />

Vol. 25 pt. 5 (1990). Contents: Solid<br />

state NMR techniques for the study of surface phenomena.<br />

A primer on isotopic labeling in NMR investigations<br />

of biopolymers. Vanadium-51 NMR.


Vol. 14, No. 1-4 311<br />

One-dimensional and Two-dimensional NMR<br />

Spectra by Modern Pulse Techniques. Koji Nakanishi.<br />

(1990). University Science Books, Mill Valley,<br />

CA. 234 p.<br />

Annual Reports on NMR Spectroscopy. Volume<br />

22 (1990). Contents: Metal-ion NMR studies of ion<br />

binding. NMR studies of ligand-macromolecule interactions.<br />

Applications of NMR in the analysis of<br />

agrochemicals and pesticides. NMR nuclear shielding<br />

and the electronic structures of macromolecules.<br />

207 Pb-NMR parameters. Nuclear spin relaxation in<br />

diamagnetic fluids part 1. General aspects and inorganic<br />

applications.<br />

Fourier Transforms in NMR, Optical, and Mass<br />

Spectrometry: A User's Handbook. By A. G. Marshall<br />

and F. R. Verdun (Ohio State University). Elsevier:<br />

Amsterdam and New York. 1990. xvi + 450<br />

pp. $107.25. ISBN 0-444-87360-0.<br />

Progress in Nuclear Magnetic Resonance Spectroscopy.<br />

Volume 22, pt. 1 (1990). Contents: Scaling<br />

in one and two dimensional NMR spectroscopy<br />

in liquids. Oligosaccharide conformations: Application<br />

of NMR and energy calculations. Relaxation<br />

matrix analysis of 2D NMR data.<br />

Progress in Magnetic Resonance Spectroscopy.<br />

Volume 22, Part 3 (1990). Contents: NMR parameters<br />

of alkynes. Improved methods for quantitative<br />

spectral analysis of NMR data.<br />

Advances in Magnetic Resonance. Volume 14<br />

(1990). Contents: Measurement of dipole-dipole<br />

cross correlation by triple-quantum filtered twodimensional<br />

exchange spectroscopy. Assessment<br />

and optimization of pulse sequences for homonuclear<br />

isotropic mixing. Spin-1/2 description of spins 3/2.<br />

Optical pumping measurements of nuclear cross relaxation<br />

and electrix doublets.<br />

Quarterly Review of Biophysics. Volume 23<br />

(Number 1) February 1990. Contents: Biosynthetic<br />

incorporation of 15 N and 13 C for assignment and interpretation<br />

of nuclear magnetic resonance spectra<br />

of proteins. Heteronuclear niters in two-dimensional<br />

[1H, 1H]- NMR spectroscopy: combined use with<br />

isotope labelling for studies of macromolecular con-<br />

formation and intermolecular interactions.<br />

Quarterly Reviews of Biophysics. Volume 23<br />

(Number 2) May 1990. Contents: Heteronuclear<br />

three-dimensional NMR spectroscopy of isotopically<br />

labelled bioilogical macromolecules. Deuterium labelling<br />

in NMR structural analysis of larger proteins.<br />

Use of deuterium labelling in NMR studies of<br />

antibody combining site structure.<br />

Principles of Nuclear Magnetic Resonance in<br />

One and Two Dimensions. Richard R. Ernst and<br />

Geoffrey Bodenhausen. Oxford University Press.<br />

1990. 640 pp. paper $39.95<br />

A Dictionary of Concepts in NMR. S.W.<br />

Homans. Oxford University Press. 1990. 352 pp."<br />

$80.00<br />

Nuclear Magnetic Resonance: Principles and<br />

Theory. Ryozo Kitamaru. Elsevier, New York,<br />

1990.<br />

Quantum Description of High-Resolution NMR<br />

in Liquids. Maurice Goldman. Oxford University<br />

Press. 1990. 288 pp. $65.00<br />

Modern Pulsed and Continuous-wave Electron<br />

Spin Resonance. Edited by Larry Kevan and<br />

Michael K. Bowman. Wiley, New York, 1990. 440<br />

p.<br />

Principles of Magnetic Resonance, Second Ed.<br />

by C. P. Slichter, Springer, New York, 1990. 655 p.<br />

Soviet Scientific Reviews Section B: Chemistry<br />

Reviews. Volume 14, Part 2 (1990). Contents:<br />

Pulsed NMR study of molecular motion in solids.<br />

Progress in Nuclear Magnetic Resonance Spectroscopy,<br />

Volume 22 No. 6 1990. Contents: Fieldcycling<br />

relaxometry of protein solutions and tissue.<br />

Implications for MRI. Solid state NMR studies of<br />

local motions in polymers.<br />

Nuclear Magnetic Resonance, Volume 20 1989/<br />

1990. Contents: NMR books and reviews. Theoretical<br />

and physical aspects of nuclear shielding. Applications<br />

of nuclear shielding. Theoretical aspects


312<br />

of spin-spin couplings. Applications of spin-spin<br />

couplings. Nuclear spin relaxation in liquids and<br />

gases. Solid state NMR Multiple pulse NMR Natural<br />

macromolecules. Synthetic macromolecules.<br />

Conformational analysis. Nuclear magnetic resonance<br />

spectroscopy of living systems. Nuclear magnetic<br />

resonance imaging of living systems. NMR of<br />

paramagnetic species. NMR of liquid crystals and<br />

micellar solutions.<br />

1 Spin Labeling: Theory and Applications.<br />

Edited by L. J. Berliner and J. Reuben, Academic<br />

Press, New York, Vol. 3, 1989.<br />

1 Advanced EPR: Applications in Biology and<br />

Biochemistry. Edited by A. J. Hoff, Elsevier, Amsterdam,<br />

1989.<br />

1 Pulsed EPR: A New Field of Applications.<br />

Edited by C. P. Keijzers, E. J. Reijerse and J.<br />

Schmidt, North Holland, Amsterdam, 1989.<br />

1 Electron Spin Resonance, Specialist Periodical<br />

Report, Vol. 1 IB, Royal Chemical Society, London,<br />

1989.<br />

Nuclear Magnetic Resonance: Structure and<br />

Mechanism. Edited by Norman J. Oppenheimer and<br />

Thomas L. James, Academic Press, New York, 1989.<br />

507 p. (Methods in Enzymology).<br />

NMR Spectroscopy and Polymer Micro structure.<br />

The Conformation Connection. Alan E. Tonelli.<br />

VCH, New York, 1989. x 252 pp., illus. $69.50.<br />

Methods in Stereochemical Analysis.<br />

Annual Reports on NMR Spectroscopy, Vol. 21.<br />

Edited by G. A. Webb, Academic Press, London,<br />

1989. ISBN: 0-12-505321-5.<br />

EPR of Exchange-Coupled Systems. Alessandro<br />

Bencini and Dante Gatteschi. Springer-Verlag,<br />

Berlin, 1989. 287 pages. ISBN: 0-387-50944-5.<br />

Nuclear Magnetic Resonance, Vol. 18, Specialist<br />

Periodical Reports, G. A. Webb, Senior Reporter,<br />

Royal Society of Chemistry, London, 1989. 511<br />

x New additions to the list.<br />

pages. ISBN: 0-85186-412-0.<br />

Bulletin of Magnetic Resonance<br />

Advances in Magnetic Resonance Imaging.<br />

Edited by Ephraim Feig, IBM Research Division,<br />

Thomas J. Watson Research Center. Ablex Publishing<br />

Corporation. 1989. 272 pp. $55.00<br />

Advances in Magnetic Resonance. Volume 13<br />

(1989). Contents: Single crystal nuclear magnetic<br />

resonance studies of high temperature superconductors.<br />

Deuterium nuclear magnetic resonance and<br />

molecular dynamics in alkane/urea inclusion compounds.<br />

1 H nuclear magnetic resonance imaging of<br />

solids with magic-angle spinning. Two-dimensional<br />

nuclear magnetic resonance experiments for studying<br />

molecular order and dynamic in static and rotating<br />

solids. Electrophoretic nuclear magnetic resonance<br />

experiments for studying molecular order<br />

and dynamics in static and rotating solids. Electrophoretic<br />

nuclear magnetic resonance. Ultraslow<br />

atomic motion by site-selective excitation of highly<br />

resolved nuclear magnetic resonance lines in dilute<br />

spin systems. Two-dimensional hybrid experiments<br />

for the measurement of small anisotropies in magicangle<br />

spinning nuclear magnetic resonance.<br />

Analytical NMR. Edited by L. D. Field and S.<br />

Sternhell. Wiley, New York, 1989, 250 p.<br />

Modern NMR spectroscopy. A workbook of<br />

chemical problems by Jeremy K.M. Sanders, Edwin<br />

C. Constable and Brian K. Hunter. Oxford University<br />

Press, New York, 1989, 1.18 p.<br />

Introduction to Pulse NMR Spectroscopy, Second<br />

Edition., Thomas C. Farrar. The Farragut<br />

Press, Madison, Wisconsin 53705, 1989. 211 pages.<br />

$39.95 (hard cover); $24.95 (paperback).<br />

Nuclear Magnetic Resonance. Volume 19<br />

(1988/1989) Contents: Theoretical and physical aspects<br />

of nuclear shielding. Applications of nuclear<br />

shielding. Theoretical aspects of spin-spin couplings.<br />

Applications of spin-spin coupling. Nuclear<br />

spin relaxation in liquids. Solid state NMR Multiple<br />

pulse NMR Conformational analysis. Nuclear<br />

magnetic resonance of living systems. Oriented molecules.<br />

Heterogeneous systems.


Vol. 14, No. 1-4 313<br />

1 Electron Nuclear Double Resonance Spectroscopy<br />

of Radicals in Solution, by H. Kurreck, B.<br />

Kirste and W. Lubitz. VCH Publishers, New York,<br />

1988.<br />

Quantum Description of High-Resolution NMR<br />

in Liquids. Maurice Goldman. Clarendon Press,<br />

Oxford University Press, New York, 1988.<br />

Electron Spin Resonance. Volume 11B (1988).<br />

Contents: In vivo detection of free radical metabolites<br />

by spin trapping. Theoretical aspects of ESR<br />

Transition metal ions. Recent developments of EN-<br />

DOR spectroscopy in the study of defects in solids.<br />

Inorganic and organometallic radicals and clusters<br />

prepared in a rotating cryostat by metal vapour<br />

techniques. Inorganic and organometallic radicals.<br />

Metalloproteins. Complexes of paramagnetic metals<br />

with paramagnetic ligands.<br />

Nuclear Magnetic Resonance<br />

Spectroscopy. Frank A. Bovey, Lynn Jelinski and<br />

Peter A. Mirau. Academic, New York, 1988. 653 p.<br />

Coherence and NMR, Michael Munowitz. Wiley,<br />

New York, 1988. 289 pages. $39.95.<br />

Biomedical Magnetic Resonance Imaging, Principles,<br />

Methodology, and Applications. Edited by<br />

Felix W. Wehrli, Derek Shaw, and J. Bruce Kneeland.<br />

VCH Publishers, New York, 1988. 601 pages.<br />

$95.00.<br />

Interpretation of Carbon-13 NMR Spectra, F. W.<br />

Wehrli, A. P. Marchand, and S. Wehrli. Wiley, New<br />

York, 1988. 484 pages. $89.95. ISBN 0-471-91742-<br />

7.<br />

Introduction to NMR Spectroscopy. By R. J.<br />

Abraham (University of Liverpool) et al. John Wiley<br />

and Sons: Chicester and New York. 1988. xiii<br />

+ 271 pages. $44.95. ISBN 0-471-91893-8.<br />

Nuclear Magnetic Resonance. Volume 18<br />

(1987/1988) Contents: Theoretical and physical<br />

aspects of nuclear shielding. Theoretical aspects<br />

of spin-spin couplings. Applications of spin-spin<br />

1 New additions to the list.<br />

coupling. Nuclear spin relaxation in liquids and<br />

gases. Solid state NMR Multiple pulse NMR Natural<br />

macromolecules. Synthesis macromolecules.<br />

Conformational analysis. Nuclear magnetic resonance<br />

of living systems. NMR of paramagnetic<br />

species. NMR of liquid crystals and micellar solutions.<br />

The Nuclear Overhauser Effect in Stereochemical<br />

and Conformational Analysis. David Neuhaus<br />

and Michael Williamson. Contents: Integrated account<br />

of the theory, experimental practice and applications<br />

of the nuclear Overhauser effect (NOE). Describes<br />

such recent developments as NOE difference<br />

spectroscopy, NOESY, heteronuclear NOE, and rotating<br />

frame NOE. Covers experimental design in<br />

depth and the underlying theory of NOE. Assumes<br />

a graduate level knowledge of NMR spectroscopy.<br />

496 pp. 1987. $95.00. .<br />

Modern NMR Spectroscopy: A Guide for<br />

Chemists. Jeremy K.M. Sanders and Brian K.<br />

Hunter. Oxford University Press. 1987.<br />

Stereochemical Analysis of Alicyclic Compounds<br />

by C-13 NMR Spectroscopy, James K. Whitesell and<br />

M. A. Minton. Chapman & Hall, London, 1987. 231<br />

pages. $60.00.<br />

Frequency Synthesizers, Theory and Design, 3rd<br />

Edition, Vadim Manassewitsch. Wiley, New York,<br />

1987. 608 pages. $52.95.'<br />

Carbon-13 NMR Spectroscopy. High-Resolution<br />

Methods and Applications in Organic Chemistry<br />

and Biochemistry, 3rd Edition, Eberhard Breitmaier<br />

and Wolfang Voelter. VCH Publishers,<br />

New York/Weinheim, Germany, 1987. 515 pages.<br />

$135.00. ISBN 0-89573-493-1 (3-527-26466-3 at<br />

Weinheim).<br />

Two-dimensional NMR Spectroscopy: Applications<br />

for Chemists and Biochemists. Edited by<br />

William R. Croasmun and Robert M. K. Carlson.<br />

VCH Publishers: New York and Weinheim. 1987<br />

xx + 511 pages. $95.00 ISBN 0-89573-308-0.


314<br />

Instructions for Authors<br />

Because of the ever increasing difficulty of keeping<br />

up with the literature there is a growing need for<br />

critical, balanced reviews covering well-defined areas<br />

of magnetic resonance. To be useful these must<br />

be written at a level that can be comprehended by<br />

workers in related fields, although it is not the intention<br />

thereby to restrict the depth of the review.<br />

In order to reduce the amount of time authors must<br />

spend in writing we will encourage short, concise<br />

reviews, the main object of which is to inform nonexperts<br />

about recent developments in interesting aspects<br />

of magnetic resonance.<br />

The editor and members of the editorial board<br />

invite reviews from authorities on subjects of current<br />

interest. Unsolicited reviews may also be accepted,<br />

but prospective authors are requested to<br />

contact the editor prior to writing in order to avoid<br />

duplication of effort. Reviews will be subject to critical<br />

scrutiny by experts in the field and must be<br />

submitted in English. Manuscripts should be sent<br />

to the editor, Dr. David G. Gorenstein, Chemistry<br />

Department, Purdue University, West Lafayette, Indiana<br />

47906, USA/(317) 494 7851. Fax No. 317 494<br />

0239.<br />

MANUSCRIPTS must be submitted in triplicate<br />

(one copy should be the original), on approximately<br />

22x28 cm paper, type-written on one side<br />

of the paper, and double spaced throughout. If the<br />

manuscript cannot be submitted on computer tapes,<br />

floppy disks, or electronically (see below), please<br />

type with a carbon ribbon using either courier 10<br />

or 12, gothic 12, or prestige elite type face with 10<br />

or 12 pitch. All pages are to be numbered consecutively,<br />

including references, tables, and captions to<br />

figures, which are to be placed at the end of the<br />

review.<br />

ARRANGEMENT: Considerable thought<br />

should be given to a logical ordering of the subject<br />

matter and the review should be divided into<br />

appropriate major sections, and subsections, using<br />

Roman numerals, capital letters, and Arabic numerals<br />

respectively. A table of contents should be included.<br />

TABLES: These are to be numbered consecutively<br />

in the text with Arabic numerals. Their place<br />

of insertion should be mentioned in the text, but<br />

they are to be placed in order at the end of the<br />

Bulletin of Magnetic Resonance<br />

paper, each typed on a separate sheet. Each table<br />

should be supplied with a title. Footnotes to tables<br />

should be placed consecutively, using lower case letters<br />

as superscripts.<br />

FIGURES are also to be numbered consecutively<br />

using Arabic numerals and the place of insertion<br />

mentioned in the manuscript. The figures are<br />

to be grouped in order at the end of the text and<br />

should be clearly marked along the edge or on the<br />

back with figure number and authors' names. Each<br />

figure should bear a caption, and these should be<br />

arranged in order and placed at the end of the text.<br />

Figures should be carefully prepared in black ink<br />

to draftsman's standards with proper care to lettering<br />

(typewritten or freehand lettering is not acceptable).<br />

Graphs should include numerical scales<br />

and units on both axes, and all figures and lettering<br />

should be large enough to be legible after reduction<br />

by 50-60%. Figures should be generally placed on<br />

sheets of the same size as the typescript and larger<br />

originals may be handled by supplying high-contrast<br />

photographic reductions. One set of original figures<br />

must be supplied; reproduction cannot be made<br />

from photocopies. Two additional copies of each figure<br />

are required. Complex molecular formula should<br />

be supplied as ink drawings.<br />

. REFERENCES to the literature should be<br />

cited in order of appearance in the text by numbers<br />

on the line, in parentheses. The reference list is to be<br />

placed on separate sheets in numerical order at the<br />

end of the paper. References to journals should follow<br />

the order: author's (or authors') initials, name,<br />

name of journal, volume number, page, and year of<br />

publication. The abbreviation of the journal follows<br />

that used in the Chemical Abstracts Service Source<br />

Index. References to books should include in order:<br />

author's (or authors') initials, name, title of book,<br />

volume, edition if other than the first, publisher,<br />

address, date of publication, and pages.<br />

FOOTNOTES should be used sparingly and<br />

only in those cases in which the insertion of the<br />

information in the text would break the train of<br />

thought. Their position in the text should be<br />

marked with a superscript Arabic numeral and the<br />

footnotes should be typed at the bottom of the relevant<br />

page in the text, separated from the latter by<br />

a line.<br />

SYMBOLS AND ABBREVIATIONS:<br />

Mathematical symbols should be typewritten wher-


Vol. 14, No. 1-4 315<br />

ever possible. Greek letters should be identified in<br />

pencil in the margin. In reviews containing a number<br />

of mathematical equations and symbols, the author<br />

is urged to supply a list of these on a separate<br />

sheet for the assistance of the printer; this will not<br />

appear in print. Standard abbreviations will follow<br />

the American Chemical Society's HANDBOOK<br />

FOR AUTHORS names and symbols for units.<br />

PERMISSIONS: It is the responsibility of the<br />

author to obtain all permissions concerned with the<br />

reproduction of figures, tables, etc, from copyrighted<br />

publications. Written permission must be obtained<br />

from the publisher (not the author or editor) of the<br />

journal or book. The publication from which the<br />

figure or table is taken must be referred to in the<br />

reference list and due acknowledgement made, e.g.<br />

reprinted by permission from ref. (00).<br />

REPRINTS: Thirty reprints of a review will<br />

be supplied free to its senior author and additional<br />

reprints may be purchased in lots of 100.<br />

INSTRUCTIONS FOR SUBMITTING<br />

MANUSCRIPTS ON FLOPPY DISKS<br />

OR ELECTRONICALLY: If you have used a<br />

word processor to type your manuscript, please forward<br />

your manuscript after review and revision, in a<br />

computer readable form. Tape should be unlabeled,<br />

in a standard IBM format, 1600 BPI, 80x80 blocks<br />

and ASCII image. Floppy disks or tapes readable on<br />

IBM, DEC, or Macintosh computers are also acceptable.<br />

For direct submission over the phone lines use<br />

300 - 2400 baud ASCII transmission. Submission<br />

via INTERNET (address David@chem.purdue.edu)<br />

is also recommended. Additional details will be provided<br />

for the appropriate hand-shake requirements.<br />

Please supply us with the code for interpreting superscripts,<br />

greeks, etc. on your word processor.


INTERNATIONAL SOCIETY OF MAGNETIC RESONANCE<br />

(<strong>ISMAR</strong>)<br />

<strong>ISMAR</strong> MEMBERSHIP APPLICATION<br />

Name and Address (Telephone & Fax numbers, e-mail) of <strong>ISMAR</strong> Member<br />

Membership Fee for 1991, $20.00 $.<br />

Membership Fee for 1992, $20.00 $<br />

Membership Fee for 1993, $20.00 $<br />

Membership Fee for 1994, $20.00 $<br />

If this is a new membership application, please check Q<br />

If you wish to be registered as a member of the Division of Biology and Medicine,<br />

in addition to <strong>ISMAR</strong> itself, at no extra charge, please check box. D<br />

BULLETIN OF MAGNETIC RESONANCE<br />

The Quarterly Review Journal of the<br />

International Society of Magnetic Resonance<br />

Telephone: (317)494-7851 DAVID G. GORENSTEIN, Editor<br />

FAX: (317)494-0239 Department of Chemistry<br />

Internet: david@adam.chem. Purdue University<br />

purdue.edu W. Lafayette, IN, U.SA.47907<br />

Subscription to Bulletin of Magnetic Resonance: <strong>ISMAR</strong> Member Rate<br />

Volume 13 (1991), $25.00 $<br />

Volume 14 (1992), $25.00 $<br />

Volume 15 (1993), $25.00 $<br />

Volume 16 (1994), $25.00 $<br />

Subscription to Bulletin of Magnetic Resonance: Library & NON-<strong>ISMAR</strong> Member Rate<br />

Volume 13 (1991), $80.00 $<br />

Volume 14 (1992), $80.00 $<br />

Volume 15 (1993), $80.00 $<br />

Volume 16 (1994), $80.00 $<br />

TOTAL $<br />

Send this completed form with your check made payable to:<br />

International Society of Magnetic Resonance<br />

Professor Regitze R. Void, Treasurer<br />

Department of Chemistry, 0342<br />

University of California, San Diego<br />

9500 Gilman Drive<br />

La Jolla,CA 92093-0342<br />

(619) 534-0200; Fax (619) 534-7042<br />

Bitnet: rrvold@ucsd.bitnet

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!