Crystal Ucle I C Rot Eins
Crystal Ucle I C Rot Eins
SERIES EDITOR
B. D. HAMES
Department of Biochemistry and Molecular Biology
University of Leeds, Leeds LS2 9JT, UK
Edited by
ARNAUD DUCRUIX
Laboratoire de Cristallographie et
RMN Biologiques, Faculte de Pharmacie
Universite de Paris V, Paris
and
RICHARD GIEGE
Institut de Biologie Moleculaire
et Cellulaire du CNRS, Strasbourg
OXFORD
UNIVERSITY PRESS
OXFORD
UNIVERSITY PRESS
Great Clarendon Street, Oxford OX2 6DP
Oxford University Press is a department of the University of Oxford
and furthers the University's aim of excellence in research, scholarship,
and education by publishing worldwide in
Oxford New York
Athens Auckland Bangkok Bogota Buenos Aires Calcutta
Cape Town Chennai Dar es Salaam Delhi Florence Hong Kong Istanbul
Karachi Kuala Lumpur Madrid Melbourne Mexico City Mumbai
Nairobi Paris Sao Paulo Singapore Taipei Tokyo Toronto Warsaw
and associated companies in Berlin Ibadan
Oxford is a registered trade mark of Oxford University Press
Published in the United States
by Oxford University Press Inc., New York
© Oxford University Press, 1999
All rights reserved. No part of this publication may be reproduced,
stored in a retrieval system, or transmitted, in any form or by any means,
without the prior permission in writing of Oxford University Press.
Within the UK, exceptions are allowed in respect of any fair dealing for the
purpose of research or private study, or criticism or review, as permitted
under the Copyright, Designs and Patents Act, 1988, or in the case
of reprographic reproduction in accordance with the terms of licenses
issued by the Copyright Licensing Agency. Enquiries concerning
reproduction outside those terms and in other countries should be
sent to the Rights Department, Oxford University Press,
at the address above.
This book is sold subject to the condition that it shall not, by way
of trade or otherwise, be lent, re-sold, hired out, or otherwise circulated
without the publisher's prior consent in any form of binding or cover
other than that in which it is published and without a similar condition
including this condition being imposed on the subsequent purchaser
Users of books in the Practical Approach Series are advised that prudent
laboratory safety procedures should be followed at all times. Oxford
University Press makes no representation, express or implied, in respect of
the accuracy of the material set forth in books in this series and cannot
accept any legal responsibility or liability for any errors or omissions
that may be made.
A catalogue record for this book is available from the British Library
Library of Congress Cataloging in Publication Data
Crystallization of nucleic acids and proteins : a practical approach /
edited by Arnaud Ducruix, Richard Giege — [2nd ed.]
(The practical approach series : 210)
Includes bibliographical references and index.
1. Proteins—Analysis. 2. Nucleic Acids—analysis. I. Ducruix.
A. (Arnaud) II. Giege, R. (Richard) III. Series.
QD431.25.A53C79 1999 547.7'5046—dc21 99-15222
ISBN 0-19-963679-6 (Hbk)
0-19-963678-8 (Pbk)
Typeset by Footnote Graphics,
Warminster, Wilts
Printed in Great Britain by Information Press, Ltd,
Eynsham, Oxon.
Preface
With the development of genomics and proteomics, and their applications in
basic biological research and the biotechnologies, there is an increasing need
of three-dimensional structural knowledge of proteins, nucleic acids, and
multi-macromolecular assemblies by X-ray methods. To achieve this aim,
crystals diffracting at high resolution are needed. The major aim of the second
edition of this book in the Practical Approach series is to present an update of
the methods employed to produce crystals of biological macromolecules and
to outline the newest trends that have entered the field. Since the first edition,
which appeared in 1992, the science of crystallogenesis, which was then in its
infancy, has grown rapidly balancing between the physics of crystal growth and
blind-screen crystallizations. The advances can be appreciated from the Pro-
ceedings of the different International Conferences on the Crystallization of
Biological Macromolecules (ICCBM1 to 7) which appear every two years, the
latest covering ICCBM-7, held in June 1998 in Granada, and published in
J. Crystal Growth, Vol. 196, January 1999.
As usual in the series, the emphasis of the present book is to give detailed
laboratory protocols throughout the chapters. However, we have not given
protocols just as 'recipes', but instead we have intended to always present the
methods with reference to the theoretical concepts and principles underlying
them. In fact one of the aims of this book was to fight against the fallacious
idea according to which crystal growth of biological macromolecules is more
an 'art' than a science. Although this is probably sometimes true from a prac-
tical point of view, it is certainly incorrect in its principle. Therefore emphasis
has been given to the physical parameters involved in crystallization and on
the large knowledge on the crystal growth of small molecules, as well as to the
particular biochemical and physico-chemical properties of biological macro-
molecules.
This book is intended to be read by a wide range of scientists. First, by the
crystallographers who have to solve three-dimensional structures of macro-
molecules. Secondly, by all molecular biologists who have access to macro-
molecules but often do not know how to handle them for crystallization, and
who may consider crystallization as an esoteric undertaking, because of lack of
basic knowledge about the crystallization process. Thirdly, by the physico-
chemists and physicists who for other reasons consider biology as an esoteric
science. It is our wish that this book will contribute to a better understanding
of crystallogenesis by these scientists and to the improved perception of the
biological requirements that have to be taken into account for physical studies.
Finally, the book should be a laboratory guide for all students and beginners,
helping them to avoid making mistakes when entering the field of crystal
preparation.
Preface
Chapter 1 is an introduction to crystallogenesis of biological macro-
molecules. It includes a brief historical survey of the subject and introduces the
general principles and major achievements of this new discipline. The prepara-
tion of biological macromolecules and the concept of 'crystallography-grade
purity' are developed in Chapter 2. Chapter 3 is new and introduces the use of
molecular biology methods to 'customize' domains for structural biology. It
also includes the preparation of protein crystals made of protein molecules
containing selenomethionine residues and outlines why they can be used for
the multiple anomalous dispersion (MAD) method. Screen-like methods are
now widespread but do not provide suggestions when they fail. The answer
may then come from statistical methods presented in Chapter 4 which explains
their theory and gives practical advice (and a computer program) for protocol
design. One of the goals of this book is to give to crystal growers of biomacro-
molecules the conceptual and methodological tools needed to control crystal-
lization. This is examined in Chapter 5 which includes a description of the
classical crystallization methods together with workshop examples. Crystal-
lization in gels is described with theoretical and practical considerations in
Chapter 6. This chapter includes a novel section describing the gel acupunc-
ture method; it also contains information on crystal growth under microgravity
and hypergravity conditions. Because it is sometimes difficult to reproduce
appropriate nucleation conditions, Chapter 7 is devoted to seeding procedures
with preformed crystalline material, including micro-, macro-, and cross-
seeding with numerous examples.
The special cases of nucleic acids (and their complexes with proteins and
nucleoprotein assemblies) and membrane proteins are covered in two indi-
vidual chapters (8 and 9). Their crystallization is still challenging, but the
novel developments in the field have led to a number of recent breakthroughs,
that are encouraging for experimenters entering the field. For nucleic acids,
Chapter 8 gives emphasis to the strategies for the design and preparation of
appropriate DNA or RNA fragments and to the specific features character-
izing their crystallization, either free or in complexes with proteins. For
membrane proteins, the already published genomes show that a good third of
the expressed proteins belong to this category and it is expected that the inter-
est for membrane proteins will expand quickly, especially among structural
biologists. It is our wish that the methods described in Chapter 9 will help them
to reach this goal.
The link between protein solubility and the physico-chemical parameters
governing crystal growth is presented in Chapter 10 with a strong emphasis on
the practical issues. Chapter 11 deals with physical methods and gives an intro-
duction to the physics of crystal growth. In particular, the use of light scattering
methods to monitor early nucleation events is advocated with examples and a
description of the material used.
Chapter 12 is new and covers the expanding field of the two-dimensional
crystallization of soluble proteins on planar lipid films. It presents many proto-
Preface
cols which may be readily used. Soaking of crystals of biological macro-
molecules is of great interest for crystallographers, either for resolving a
structure (heavy-atom derivatives), or for diffusing inhibitors, activators, or
cofactors (eventually photoactivable). As in all previous chapters, practical
aspects were the driving force and Chapter 13 is illustrated by a variety of pro-
tocols. The editors thought that an introduction to X-ray crystallography
should be included in this book. This is done in Chapter 14, that is geared
toward biochemists wanting to characterize crystals by themselves rather than
explaining how to solve a structure.
It is a great pleasure to acknowledge our gratitude to a number of friends
and colleagues. First, our colleagues from Gif-sur-Yvette/Paris and Strasbourg
deserve particular thanks for having participated over the years in the develop-
ment of our studies on crystallogenesis; their enthusiasm was essential and
gave us the impetus for the preparation of a book covering this field and for
updating it in its second edition. However, without the invaluable help of
many friends from both sides of the Atlantic who agreed to cover specialized
topics, this venture would not have been possible. We would like to warmly
thank all of them. The French Centre National de la Recherche Scientifique
(CNRS) and Centre National d'Etudes Spatiales (CNES) are acknowledged
for their permanent support in developing biological crystallogenesis and their
interest for the physico-chemical aspects of the field.
While preparing this second edition, our dear colleague Roland Boistelle
closed his eyes. He was a source of inspiration for all of us and one of the first
who geared the biology oriented scientists to the physics of crystal growth. His
contribution to the field of macromolecules crystallogenesis was essential and
we would like to dedicate this book to his memory.
Paris and Strasbourg A. D.
June 1999 R. G.
IX
This page intentionally left blank
Contents
List of Contributors
Abbreviations
2. Seeding 178
Supersaturation and nucleation 178
Crystal growth 179
Seeding techniques 179
3. Crystallization procedures 180
Pre-seeding: sitting drop vapour diffusion 180
Analytical seeding 185
4. Production seeding methods 188
Microseeding 188
Macroseeding 191
5. Heterogeneous seeding 196
Cross-seeding 197
Epitaxial nucleation 200
6. Crystallization of complexes 202
Considerations in the crystallization of complexes 202
Use of streak seeding in protein complex crystallization 204
Analytical techniques for determination of crystal content 204
7. Concluding remarks 206
Acknowledgements 207
References 207
XIX
This page intentionally left blank
Contributors
W. BERGSMA-SCHUTTER
Department of Chemistry / Biophysical Chemistry, University of Groningen,
Nijenborgh 4, NL-9747 AG Groningen, The Netherlands,
P. F. BERNE
Rhone Poulenc Rorer, 13, quai Jules Guesde, 94403 Vitry sur Seine Cedex,
France.
R. BOISTELLE (+ DECEASED 1998)
Centre de Recherche sur les Me'canismes de la Croissance Cristalline du
CNRS, Campus de Luminy, Case 913, 13288 Marseille Cedex, France.
A. BRISSON
Department of Chemistry / Biophysical Chemistry, University of Groningen,
Nijenborgh 4, NL-9747 AG Groningen, The Netherlands.
C.W. CARTER JR
Department of Biochemistry, CB 7260, University of North Carolina at
Chapel Hill, Chapel Hill, NC 27599-7260, USA.
A.-C. DOCK-BREGEON
Institut de Genetique et de Biologie Moleculaire et Cellulaire, 1, rue Leon
Fries, Pare d'Innovation, BP163, F-67404 Illkirch Cedex, France.
S. DOUBLIE
Department of Microbiology and Molecular Genetics, The Markey Center for
Molecular Genetics, University of Vermont, Burlington, VT 05045, USA.
A. DUCRUIX
Laboratoire de Cristallographie et RMN Biologiques, Faculte de Pharmacie,
Universite de Paris V, 4, Avenue de l'Observatoire, 75270 Paris Cedex 06,
France.
J. M. GARCIA-RUIZ
Institute Andaluz de Ciencias de la Tierra, CSIC-Universidad de Granada,
Facultad de Ciencias, 18002-Granada, Spain.
R. GIEGE
Institut de Biologie Moleculaire et Cellulaire du CNRS, 15 rue Rene
Descartes, F-67084 Strasbourg Cedex, France.
T. GLEICHMANN
Anorganisch-Chemisches Institut, Westfalische Wilhelms-Universitat,
Wilhelm-Klemm Str. 8, D-48149 Munster, Germany.
Contributors
O. LAMBERT
Department of Chemistry/Biophysical Chemistry, University of Groningen,
Nijenborgh 4, NL-9747 AG Groningen, The Netherlands.
B. LORBER
Institut de Biologic Moleculaire et Cellulaire du CNRS, 15 rue Rene
Descartes, F-67084 Strasbourg Cedex, France.
D. MORAS
Institut de Genetique et de Biologie Moleculaire et Cellulaire, 1 rue Leon
Fries, Pare d'Innovation, BP 163. F-67404 Illkirch Cedex, France.
F. OTALORA
Institute Andaluz de Ciencias de la Tierra, CSIC-Universidad de Granada,
Facultad de Ciencias, 18002-Granada, Spain.
D. PICOT
Institut de Biologie Physico-Chimique, 13 rue P. et M. Curie, F-75005 Paris.
F. REISS-HUSSON
Centre de Genetique Moleculaire du CNRS, Avenue de la Terrasse, F-91190
Gif-sur-Yvette Cedex, France.
M. RIES-KAUTT
Laboratoire de Cristallographie et RMN Biologiques, Faculte de Pharmacie,
Universite de Paris V, 4, Avenue de 1'Observatoire, 75270 Paris Cedex 06,
France.
M.-C. ROBERT
Laboratoire de Mineralogie-Cristallographie, Universites Pierre et Marie
Curie, 4 place Jussieu, F-75252 Paris Cedex 05, France.
L. SAWYER
Structural Biochemistry Group, The University of Edinburgh, Swann Build-
ing, King's Buildings, Mayfield Road, Edinburgh EH9 3JR, UK.
E. A. STURA
Dept. d'Ingenierie et d'Etudes des Proteines, Bat. 152, CEA/Saclay, 91191
Gif-sur-Yvette Cedex, France.
M. A. TURNER
X-ray Structure Laboratory, Department of Biochemistry, Hospital for Sick
Children, 555 University Avenue, Toronto, Ontario, Canada.
S. VEESLER
Centre de Recherche sur les Mecanismes de la Croissance Cristalline du
CNRS, Campus de Luminy, Case 913, F-13288 Marseille Cedex, France.
O. VIDAL
Laboratoire de Mineralogie-Cristallographie, Universites Pierre et Marie
Curie, 4 place Jussieu, F-75252 Paris Cedex 05, France.
xxii
Abbreviations
BPTI bovine pancreatic trypsin inhibitor
BTP bis Tris propane
CMC critical micellar concentration
DEPC diethylpyrocarbonate
DiFP diisopropylfluorophosphate
DLS dynamic light scattering
DMSO dimethyl sulfoxide
DOPC dioleoylphosphatidylcholine
DTT dithiothreitol
EDTA ethylenediaminetetraacetic acid
ESI electronspray ionization mass spectra
GST glutathione-S-transferase
HEW hen egg white
HEWL hen egg white lysozyme
HIC hydrophobic interaction chromatography
IEF isoelectric focusing
LS light scattering
MAD multiple wavelength anomalous dispersion
MALDI matrix-assisted desorption/ionization
MeTEOS methyltriethoxysilane
MIR multiple isomorphous replacement
MPD 2-methyl-2,4-pentane diol
MR molecular replacement
OD optical density
OP osmotic pressure
PBC periodic bond chain
PC phosphatidylcholine
PCR polymerase chain reaction
PEG polyethylene glycol
pTS para-toluenesulfonate
RPC reverse-phase chromatography
SANS small angle neutron scattering
SAXS small angle X-ray scattering
SIR single isomorphous replacement
SLS static light scattering
TEOS tetraethoxysilane
TLC thin-layer chromatography
TMOS tetramethoxysilane
This page intentionally left blank
1
An introduction to the
crystallogenesis of biological
macromolecules
R. GIEGE and A. DUCRUIX
'II y a la des mysteres, qui preparent a 1'avenir d'immenses travaux et appellent des
aujourd'hui les plus serieuses meditations de la science'
Pasteur, 1860, in Leqons de Chimie.
1. Introduction
The word 'crystal' is derived from the Greek root 'krustallos' meaning 'clear
ice'. Like ice, crystals are chemically well defined, and many among of them
are of transparent and glittering appearance, like quartz, which was for a long
time the archetype. Often they are beautiful geometrical solids with regular
faces and sharp edges, which probably explains why crystallinity, even in the
figurative meaning, is taken as a symbol of perfection and purity. From the
physical point of view, crystals are regular three-dimensional arrays of atoms,
ions, molecules, or molecular assemblies. Ideal crystals can be imagined as
infinite and perfect arrays in which the building blocks (the asymmetric units)
are arranged according to well-defined symmetries (forming the 230 space
groups) into unit cells that are repeated in the three-dimensions by trans-
lations. Experimental crystals, however, have finite dimensions. An implicit
consequence is that a macroscopic fragment from a crystal is still a crystal,
because the orderly arrangement of molecules within such a fragment still
extends at long distances. The practical consequence is that crystal fragments
can be used as seeds (Chapter 7). In laboratory-grown crystals the periodicity
is never perfect, due to different kinds of local disorders or long-range
imperfections like dislocations. Also, these crystals are often of polycrystal-
line nature. The external forms of crystals are always manifestations of their
internal structures and symmetries, even if in some cases these symmetries
may be hidden at the macroscopic level, due to differential growth kinetics of
the crystal faces. Periodicity in crystal architecture is also reflected in their
macroscopic physical properties. The most straightforward example is given
R. Giege and A. Ducruix
by the ability of crystals to diffract X-rays, neutrons, or electrons, the
phenomenon underlying structural chemistry and biology (for introductory
texts see refs 1 and 2), and the major aim of this book is to present the
methods employed to produce three-dimensional crystals of biological macro-
molecules, but also two-dimensional crystals (Chapter 12), needed for
diffraction studies. Other properties of invaluable practical applications
should not be overlooked either, as is the case of optical and electronic
properties which are at the basis of non-linear optics and modern electronics
(for an introduction to physical properties of molecular crystals see ref. 3).
Crystals furnish one of the most beautiful examples of order and symmetry in
nature and it is not surprising that their study fascinates scientists (4).
What characterizes biological macromolecular crystals from small molecule
crystals? In terms of morphology, one finds with macromolecular crystals the
same diversity as for small molecule crystals (Figure 1). In terms of crystal
size, however, macromolecular crystals are rather small, with volumes rarely
exceeding 10 mm3, and thus they have to be examined under a binocular
microscope. Except for special usages, such as neutron diffraction, this is not
too severe a limitation. Among the most striking differences between the two
families of crystals are the poor mechanical properties and the high content of
solvent of macromolecular crystals. These crystals are always extremely
fragile and are sensitive to external conditions. This property can be used as a
preliminary identification test: protein crystals are brittle or will crush when
touched with the tip of a needle, while salt crystals that can sometimes
develop in macromolecule crystallization experiments will resist this treat-
ment. This fragility is a consequence both of the weak interactions between
macromolecules within crystal lattices and of the high solvent content (from
20% to more than 80%) in these crystals (Chapter 14). For that reason,
macromolecular crystals have to be kept in a solvent-saturated environment,
otherwise dehydration will lead to crystal cracking and destruction. The high
solvent content, however, has useful consequences because solvent channels
permit diffusion of small molecules, a property used for the preparation of
isomorphous heavy-atom derivatives needed to solve the structures (Chapters
13 and 14). Further, crystal structures can be considered as native structures,
as is indeed directly verified in some cases by the occurrence of enzymatic re-
actions within crystal lattices upon diffusion of the appropriate ligands (5, 6).
Other characteristic properties of macromolecular crystals are their rather
weak optical birefringence under polarized light: colours may be intense for
large crystals but less bright than for salt crystals (isotropic cubic crystals or
amorphous material will not be birefringent). Also, because the building blocks
composing macromolecules are enantiomers (L-amino acids in proteins—
except in the case of some natural peptides—and D-sugars in nucleic acids)
macromolecules will not crystallize in space groups with inversion sym-
metries. Accordingly, out of the 230 possible space groups, macromolecules
do only crystallize in the 65 space groups without such inversions (7). While
1: An introduction to the crystnllogaiwsis
1- PURIFICATION OF MACROMOLECULES
from wild-type, engineered or overproducing organisms
(possibility of in vitro synthesis for nucleic acids and small peptides)
2- CRYSTALLIZATION
by de novo crystallization or seeding techniques
3- DATA MEASUREMENTS
characterization of space group and diffraction resolution;
measurements of diffraction intensities on an electronic area detector
(possible use of neutron and frequently of tunable X-ray synchrotron radiation);
frequent data acquisition by cryo-crystallographic methods.
4- PHASE DETERMINATION
using methods based on isomorphous replacement (preparation of heavy atom derivatives),
anomalous scattering, molecular replacement and non-crystallographic symmetry,
or direct calculations (e.g. from maximum entropy)
6- MODEL REFINEMENT
least-square refinements; restrained refinements;...
Figure 2. Steps involved in the resolution of the 3D structure of a biological macro-
molecule (for more details see Chapter 14).
3. General principles
3.1 A multiparametric process
Biocrystallization, like any crystallization, is a multiparametric process involv-
ing the three classical steps of nucleation, growth, and cessation of growth.
What makes crystal growth of biological macromolecules different is, first, the
much larger number of parameters than those involved in small molecule
crystal growth (Table 1) and, secondly, the peculiar physico-chemical pro-
perties of the compounds. For instance, their optimal stability in aqueous
media is restricted to a rather narrow temperature and pH range. But the
main difference from small molecule crystal growth is the conformational
flexibility and chemical versatility of macromolecules, and their consequent
greater sensitivity to external conditions. This complexity is the main reason
why systematic investigations were not undertaken earlier. Furthermore, the
importance of some parameters, such as the geometry of crystallization
vessels or the biological origin of macromolecules, had not been recognized. It
is only recently that the hierarchy of parameters has been perceived. A prac-
tical consequence of this new perception was the development of statistical
methods to screen crystallization conditions (Chapter 3). For a rational design
of growth conditions, however, physical and biological parameters have to be
controlled. One of the aims of this book is to give to crystal growers of
biological macromolecules the conceptual and methodological tools needed
to achieve such control.
3.2 Purity
Because macromolecules are extracted from complex biological mixtures,
purification plays an extremely important role in crystallogenesis (Chapter 2).
Purity, however, is not an absolute requirement since crystals of macro-
molecules can sometimes be obtained from mixtures. But such crystals are
mostly small or grow as polycrystalline masses, are not well shaped, and are of
bad diffraction quality, and thus cannot be used for diffraction studies. How-
ever, crystallization of macromolecules from mixtures may be used as a tool
for purification (47), especially in industry (48). For the purpose of X-ray
R. Giege and A, Ducruix
Biological parameters
• Rarity of most biological macromolecules
• Biological sources and physiological state of organisms or cells (e.g. thermophiles
versus halophiles or mesophiles, growing versus stationary phase)
• Bacterial contaminants
Purity of macromolecules
• Macromolecular contaminants (odd macromolecules or small molecules)
• Sequence (micro) heterogeneities (e.g. fragmentation by proteases or nucleases—
fragmented macromolecules may better crystallize —, partial or heterogeneous post-
translational modifications)
• Conformational (micro) heterogeneities (e.g. flexible domains, oligomer and conformer
equilibria, aggregation, denaturation)
• Batch effects (two batches are not identical)
aAlthough all these parameters have not been screened systematically, especially for the crystal-
lization of a given macromolecule, all of them have been evaluated individually in isolated cases.
3.5 Packing
With biological macromolecules, crystal quality may be correlated with the
packing of the molecules within the crystalline lattices, and external crystal
morphology with internal structure. As shown by the periodic bond chain
(PBC) method, direct protein-protein contacts are essential in determining
packing and morphology (Chapter 11). Forces involved in packing of macro-
molecules may be considered as weak as compared to those maintaining the
cohesion of small molecule crystals. They involve salt bridges, hydrogen bonds,
Van der Waals, dipole-dipole, and stacking interactions (58-60). It must also
be borne in mind that the weak cohesion of macromolecular crystals results
from the fact that only a small part of macromolecular surfaces participate in
intermolecular contacts (61), the remaining being in contact with the solvent
(exceptions may be found for small proteins). This explains the commonly
observed polymorphism of biological macromolecular crystals.
12
1: An introduction to the crystallogenesis
5. Is it a fusion protein? If yes, which protease is used to cleave the
protein?
6. How many mg/litre of culture can you produce?
7. How many mg of protein can you obtain per standard purification?
B. Biochemistry
1. How long does it take to purify one batch of protein?
2. How do you assess the purity of the protein?
(a) Electrophoresis (native/denaturating conditions).
(b) HPLC (which phase).
(c) Ion spray (mass spectrometry).
(d) Checking of N-terminus.
(e) Activity assay.
2. What are the principal characteristics of the protein?
(a) Molecular weight.
(b) Isoelectric point (calculated or measured).
(c) Glycosylation (yes/no).
(d) Number of free cysteine(s).
(e) Number of disulfide bridges.
(f) Hydrophobicity or hydrophilicity.
(g) What are the ligands?
3. What are the friendly (or unfriendly) solvents?
4. Is the protein monomeric of oligomeric? How did you check it
(chromatography, light scattering, others)? Does your protein has a
tendency to aggregate?
5. What is the stability of the protein versus time, temperature, or pH?
The better these questions can be answered, the easier will be the design of
a crystallization strategy. Good knowledge of the characteristics of the pro-
tein, of its availability, will guide the experimenter. Quite often, it helps
people to be aware that the premise of crystallization may be as important as
the crystallization itself.
References
1. Pickworth Glusker, J. and Trueblood, K. N. (1985). Crystal structure analysis, a
primer. Oxford University Press, New York.
13
R. Giege and A. Ducruix
2. Drenth, J. (1995). Principles of protein X-ray crystallography. Springer-Verlag,
Berlin.
3. Wright, J. D. (1987). Molecular crystals. Cambridge University Press, Cambridge.
4. Lima-de-Faria, J. (ed.) (1990). Historical atlas of crystallography. Kluwer,
Dordrecht.
5. Hajdu, J., Acharaya, K. R., Stuart, D. I., Barford, D., and Johnson, L. N. (1988).
Trends Biochem. Sci., 13, 104.
6. Mozzarelli, A. and Rossi, G. L. (1996). Annu. Rev. Biophys. Biomol. Struct., 25,
3430.
7. Blundell, T. L. and Johnson, L. M. (1976). Protein crystallography. Academic
Press, New York.
8. Wukovitz, W. and Yeates, T. O. (1995). Nature Struct. Biol., 2, 1060.
9. Usha, R., Johnson, J. E., Moras, D., Thierry, J.-C, Fourme, R., and Kahn, R.
(1984). J. Appl. Crystallogr., 17, 147.
10. Kay, L. L. (1986). ISIS/J. Hist. Sci. Soc., 77, 450.
11. Feigelson, R. S. (1988). J. Cryst. Growth, 90, 1.
12. Boistelle, R. and Astier, J.-P. (1988). J. Cryst. Growth, 90, 14.
13. Rosenberger, F., Vekilov, P. G., Muschol, M., and Thomas, B. R. (1996). J. Cryst.
Growth, 168, 1.
14. Lehman, C. G. (1853). Lehrbuch der physiologische Chemie. Leipzig.
15. Reichert, E. T. and Brown, A. P. (1909). The differentiation and specificity of
corresponding proteins and other vital substances in relation to biological classifi-
cation and evolution: the crystallography of hemoglobins. Carnegie Institution,
Washington DC.
16. Debru, C. (1983). L'esprit des proteines: histoire et philosophic biochimiques.
Hermann, Paris.
17. McPherson, A. (1990). J. Cryst. Growth, 110, 1.
18. Sumner, J. B. (1926). J. Biol. Chem., 69, 435.
19. Northrop, J. H., Kunitz, M., and Herriot, R. M. (1948). Crystalline enzymes.
Columbia University Press, New York.
20. Jabri, E., Carr, M. B., Hausinger, R. P., and Karplus, P. A. (1995). Science, 268,
998.
21. Dounce, A. L. and Allen, P. Z. (1988). Trends Biochem. Sci., 13, 317.
22. Bernal, J. D. and Crowfoot, D. (1934). Nature, 133, 794.
23. Jones, C., Mulloy, B., and Sanderson, M. R. (ed.) (1996). In Methods in molecular
biology. Crystallographic methods and protocols, Vol. 114, pp. 1-394. Humana
Press, Totowa, NJ, USA.
24. McPherson, A. (1982). Preparation and analysis of protein crystals. Wiley, New
York.
25. Kim, S. H. and Rich, A. (1968). Science, 162, 1381.
26. Dock, A.-C, Lorber, B., Moras, D., Pixa, G., Thierry, J.-C., and Giege, R. (1984).
Biochimie, 66, 179.
27. Michel, H. (1982). J. MoL Biol, 158, 567.
28. Kuhlbrandt, W. (1988). Q. Rev. Biophys., 21, 429.
29. Arnoux, B., Ducruix, A., Reiss-Husson, F., Lutz, M., Norris, J., Schiffer, M., et al.
(1989). FEES Lett, 258, 47.
30. Michel, H. (ed.) (1991). Crystallization of membrane proteins. CRC Press, Boca
Raton, FL, USA.
14
1: An introduction to the crystallogenesis
31. Jeruzalmi, D. and Steitz, T. A. (1997). J. Mol. Biol, 274, 748.
32. Giege, R., Dock, A.-C., Kern, D., Lorber, B., Thierry, J.-C, and Moras, D. (1986).
J. Cryst. Growth, 76, 554.
33. Yonath, A., Frolow, F., Shoham, M., Mtissig, J., Makowski, I., Glotz, C, et al.
(1988). J. Cryst. Growth, 90, 231.
34. Trakhanov, S., Yusupov, M., Shirikov, V., Garber, M., Mitschler, A., Ruff, M., et
al. (1989). J. Mol. Biol, 209, 327.
35. Feigelson, R. S. (ed.) (1986). Proc. 1st Int. Conf. Protein Crystal Growth,
Stanford, CA, USA, 1985. J. Cryst. Growth, 76, 529.
36. Giege, R., Ducruix, A., Fontecilla-Camps, J., Feigelson, R. S., Kern, R., and
McPherson, A. (ed.) (1988). Proc. 2nd Int. Conf. Crystal Growth of Biological
Macromolecules, Bischenberg, France, 1987. J. Cryst. Growth, 90, 1.
37. Carter, C. W., Jr. (ed.) (1990). Methods: a companion to methods in enzymology,
Vol. 1, pp. 1-127.
38. Ward, K. and Gilliland, G. (ed.) (1990). Proc. 3rd Int. Conf. Crystal Growth of
Biological Macromolecules, Washington, DC, USA, 1989. J. Cryst. Growth, 110, 1.
39. Stezowsky, J. J. and Littke, W. (ed.) (1992). Proc. 4th Int. Conf. Crystal Growth of
Biological Macromolecules, Freiburg, Germany, 1991. J. Cryst. Growth, 122, 1.
40. Glusker, J. P. (ed.) (1994). Proc. 5th Int. Conf. Crystal Growth of Biological
Macromolecules, San Diego, CA, USA, 1993. Acta Cryst., D50, 337.
41. Miki, K., Ataka, M., Fukuyama, K., Higuchi, Y., and Miyashita, T. (ed.) (1996).
Proc. 6th Int. Conf. Crystal Growth of Biological Macromolecules, Hiroshima,
Japan, 1995. J. Cryst. Growth, 168, 1.
42. Drenth, J., and Garcia-Ruiz, J. M. (ed.) (1999). Proc. 7th Int. Conf. Crystal
Growth of Biological Macromolecules, Granada, Spain, 1998. J. Cryst. Growth, 196,
pp. 185-720.
43. Wood, S. P. (1990). In Protein purification applications: a practical approach (ed.
E. L. V. Harris and S. Angal), pp. 45-58. IRL Press, Oxford.
44. Weber, P. C. (1991). Adv. Protein Chem., 41, 1.
45. McPherson, A., Malkin, A. J., and Kuznetsov, Y. G. (1995). Structure, 3, 759.
46. Durbin, S. D. and Feher, G. (1996). Annu. Rev. Phys. Chem., 47,171.
47. Jakoby, W. B. (1971). In Methods in enzymology (ed. W. B. Jakoby), Vol. 22,
pp. 248-52. Academic Press, London.
48. Judge, R. A., Johns, M. R., and White, E. T. (1995). Biotechnol. Bioeng., 48, 316.
49. Lorber, B., Jenner, G., and Giege, R. (1996). J. Cryst. Growth, 103, 117.
50. Von Hippel, P. H. and Schleich, T. (1969). In Structure and stability of biological
macromolecules (ed. S. N. Timasheff and G. D. Fasman), Vol. 2, pp. 417-574.
Dekker.
51. Mikol, V. and Giege, R. (1989). J. Cryst. Growth, 97, 324.
52. Ries-Kautt, M. and Ducruix, A. (1989). J. Biol. Chem., 264, 745.
53. Cacioppo, E., Munson, S., and Pusey, M. L. (1991). J. Cryst. Growth, 110, 66.
54. Gilliland, G. L., Tung, M., Blakeslee, D. M., and Ladner, J. E. (1994). Acta Cryst.,
D50, 408.
55. Thaller, C., Weaver, L. H., Eichele, G., Wilson, E., Karlson, R., and Jansonius, J.
N. (1981). J. Mol. Biol., 147, 465.
56. Giege, R., Drenth, J., Ducruix, A., McPherson, A., and Saenger, W. (1995). Prog.
Cryst. Growth Charact., 30, 237.
57. McPherson, A. (1997). Trends Biotechnol. 15, 197.
15
R. Giege and A. Ducruix
58. Bergdoll, M. and Moras, D. (1988). J. Cryst. Growth, 90, 283.
59. Salemme, F. R., Genieser, L., Finzel, B. C., Hilmer, R. M., and Wendolosky, J. J.
(1988). J. Cryst. Growth, 90, 273.
60. Wang, A. H. J. and Teng, M. K. (1988). J. Cryst. Growth, 90, 295.
61. Carugo, O. and Argos, P. (1997). Protein Set., 6, 2261.
62. Chayen, N. E., Boggon, T. J., Cassetta, A., Deacon, A., Gleichmann, T., Habash,
J., etal. (1996). Q. Rev. Biophys., 29, 227.
63. Seeman, N. C. (1991). Curr. Opin. Struct. BioL, 1, 653.
16
2
1. Introduction
The quality and quantity of the macromolecular samples are important pre-
requisites for successful crystallizations. Proteins and nucleic acids extracted
from living cells or synthesized in vitro differ from small molecules by
additional properties intrinsic to their chemical nature and their larger size.
They are frequently difficult to prepare at a high degree of purity and
homogeneity. Besides traces of impurities, harsh treatments may decrease their
stability and activity through different kinds of alterations. Consequently, the
quality of biomacromolecules depends on the way they are prepared and
handled. As a general rule purity and homogeneity are regarded as conditions
sine qua non. Accordingly, purification, stabilization, storage, and handling of
macromolecules are essential steps prior to crystallization attempts. Other
difficulties in crystal growth may come from the source of the biological
material. It is advisable to have at disposal a few milligrams of material when
starting first crystallization trials although structures were solved with sub-
milligram quantities of protein (1). Once crystals suitable for X-ray analysis
can be produced, additional material is often needed to improve their quality
and size and to prepare heavy-atom derivatives. It is thus essential that
isolation procedures are able to supply enough fresh material of reproducible
quality. Similar situations are encountered with multi-macromolecular
assemblies (e.g. viruses, nucleosomes, ribosomal particles, or their subunits).
This chapter discusses biochemical methods used to prepare and character-
ize macromolecules intended for crystallization assays. Practical aspects
concerning manipulation and qualitative analyses of soluble proteins will be
emphasized. The cases of nucleic acids and membrane proteins are described
in more detail in Chapters 8 and 9. Peculiar aspects of molecular biology that
are important for crystallogenesis are presented in Chapter 3. They include
the design of engineered macromolecules with new physical properties or
B. Lorber and R. Giege
modified to simplify purification or crystallographic analysis. Finally, methods
for identification of macromolecular content of crystals and measurements of
their density are presented as well.
• Cell culture
Fermentors, culture flasks and plates, thermostated cabinets
High capacity centrifuges or filtration devices for cell recovery
• Cell disruption
Mechanical disruption devices (grinders, glass bead mills, French press)
Chemical treatments (e.g. phenolic extraction of small RNAs)
Biochemical treatments (e.g. cell lysis by enzymes)
Others (e.g. sonication, freezing/thawing)
• Centrifugation
Low speed centrifuge (to remove cell debris or recover precipitates)
High speed centrifuge (to fractionate subcellular components)
• Dialysis and ultrafiltration
Dialysis tubing (hollow fibres or membranes of various porosities and sizes)
Concentrators (from 0.5 ml to a few litres with high flow rate low macromolecule-
binding membranes of various cut-offs)
• Chromatography (prefer metal-free systems)
Low pressure equipment for fast separation (FPLC, hyperdiffusion, perfusion)
High pressure equipment (HPLC)
Columns of various capacities filled with various matrices (particle size 10-30 mm)
Pumps, programmer, on-line absorbance detector, fraction collector, recorder
• Preparative electrophoresis and isoelectric focusing
Electrophoresis apparatus for large rod or slab gels
Preparative liquid IEF apparatus (column or horizontal cells)
Power supplies
• Detection, characterization, and quantitation
Spectrophotometer, fluorimeter
pH meter, conductimeter, refractometer (to monitor chromatographic elution)
Liquid scintillation counter (for radioactivity detection)
Analytical gel electrophoresis, capillary electrophoresis, and IEF equipment
20
2: Biochemical aspects and handling of macromolecular solutions
The preparation of a cellular extract and fractionation of its components
are the two stages common to most purification protocols (except for macro-
molecules secreted in culture media). Intracellular macromolecules are
released using physical, chemical, or biological disruption methods and
extracts are clarified by centrifugation or ultrafiltration. Membrane proteins
and proteins with hydrophobic surfaces are solubilized with detergents or
sulfobetaines (26) (Chapter 9). Extracellular compounds and macromolecules
synthesized in vitro may be recovered either by ultrafiltration, centrifugation,
flocculation, or liquid-liquid partitioning.
3.1.2 Proteins
Gross fractionation includes one or several precipitations induced either by
addition of salts (e.g. ammonium sulfate), organic solvents (e.g. acetone), or
organic polymers (e.g. PEG). Temperature or pH variation are applied to
decrease solubility or stability of unwanted macromolecules. Fractionation
between two liquid phases and selective precipitation (e.g. of nucleic acids by
protamine) are additional methods. The next steps involve more resolutive
methods, generally a combination of column chromatographies. These are
based on separation by charge (adsorption, anion or cation exchange,
chromatofocusing), hydrophobicity (hydrophobic interaction (HIC) or
reverse-phase (RFC) chromatographies), size (exclusion chromatography),
peculiar structural features (e.g. affinity for heparin, antibodies, metal ions, or
thiol groups), or activity (affinity for catalytic sites, receptors, or biomimetic
compounds). HPLC yields higher resolution than standard techniques
because of the monodispersity and small size of the spherical matrix particles
(27) and new matrices take advantage of hyperdiffusion (Beckman) or per-
fusion (PerSeptive Biosystems) to accelerate elution. Preparative IEF is
carried on in gels with free or immobilized ampholytes (Immobiline®,
Pharmacia) (28) in rotating cells divided in compartments by permeable nylon
grids (Rotofor®, Bio-Rad) or in multichamber units holding fixed pH mem-
branes (Isoprime™, Pharmacia). Differential centrifugation and free flow
electrophoresis (29) are other methods. Monitoring of specific activities
during the purification procedure helps to identify unsatisfactory steps in
which macromolecules are lost or inactivated. Guidelines for effective protein
purification may be summarized as follows:
• work in the cold room (i.e. at 4°C) with chilled equipment and solutions if
the protein is unstable at higher temperature
• use precipitation steps to speed up fractionation
• limit the number of chromatographies (to three or four)
• prefer quick assays to characterize macromolecules
• use short and efficient non-denaturing intermediary treatments (e.g.
repeated dialysis)
• add stabilizing agents and protease inhibitors (see Section 5.5).
21
B. Lorber and R. Giege
In summary, success in crystallization is often dependent on rapid purifi-
cation. To reach this aim advanced equipment and chromatographic systems
enabling high flow rate (e.g. advanced HPLC and perfusion chromatography)
are recommended.
3.1.3 Nucleic acids
Purification of nucleic acids requires specific methods. For tRNAs, phenol
extraction precedes counter-current distribution or chromatography on
benzoylated DEAE-cellulose (also on Sepharose® or other matrices).
Further purification is based on anion exchange, adsorption, reverse-phase,
mixed-mode, hydrophobic interaction, perfusion, or affinity chromato-
graphies. Intermediary treatments include precipitation by ethanol, dialysis,
and concentration by evaporation under vacuum. HPLC on coated silica
substituted by short aliphatic chains (C4) gives separations with good
resolution. Oligo-DNA or RNA are synthesized chemically on solid phase
supports or enzymatically in vitro. Abortive sequences are eliminated by
denaturing gel electrophoresis (Prepcell™, Bio-Rad) or HPLC (30-32). For
further details see Chapter 8.
3.4 Ageing
Properties of macromolecular samples change with time as explicitly illus-
trated for lysozyme crystallization (46). Ageing results from the action of
contaminants present or introduced in samples or of modifications generated
by oxidants. In the example of lysozyme, changes in crystallizability are due to
the presence of fungi that multiply in the stored protein solution (46). Water
molecules or metal ions induce slow hydrolysis in RNAs (Chapter 8).
Self-cleaving macromolecules, like certain proteases and ribozymes, pose
specific problems. For protease crystallization the problem can be solve by
storing and co-crystallizing the protease with an active site inhibitor. For
instance, the three-dimensional structure of a human protease of the ICE type
participating in apoptosis could be solved because it was crystallized as a
covalent complex with a tetrapeptide inhibitor (47). For hammerhead
ribozyme, self-cleavage was prevented by introducing modified bases into the
molecule (48).
Method
1. Prepare a stock solution to make gels having a total acrylamide con-
centration T = 8% (w/v), a cross-linker concentration C = 5% (w/w),
and containing 8 M urea by mixing:
• 20 ml acrylamide-bisacrylamide solution
• 5 ml of 20 x Tris-borate buffer
• 50 g urea
• distilled water up to 100 ml
2. Filter the above stock solution on a 0.45 um pore size membrane.
24
2: Biochemical aspects and handling of macro-molecular solutions
3. To 10 ml solution add 10 ul TEMED and 100 ul fresh 5% (w/v)
ammonium peroxodisulfate solution. Mix and pour in the mould.
Polymerization occurs in about 30 min.
4. Denature nucleic acids in a solution containing 8 M urea, 20% (w/v)
saccharose (or 20% (v/v) glycerol), and 0.025% (w/v) bromophenol
blue. Load samples onto the gel and run electrophoresis under
appropriate voltage.
5. Stain nucleic acids by soaking gels in the dark in a solution containing
30 mg 'Stains all', 100 ml dimethylformamide, and distilled water up to
1 litre. Destain in the light. Another technique employs electrophoresis
buffer containing 0.5 ug/ml ethidium bromide. Wait 10 min and view
the gel in UV light (254 nm). Silver stain techniques for proteins are
also suitable to visualize nucleic acids (49).
where 5690 and 1280 are the rnolar absorption coefficients at 280 nm of
tryptophan and tyrosine, and nx and ny the numbers of tryptophan and
tyrosine residues, respectively (57). Hence, protein concentrations are
obtained from:
aNon-soluble species.
bSee Chapter 5, Protocol 1.
28
2: Biochemical aspects and handling of macromolecular solutions
31
B. Lorbor and R. Giege
titration shows the mobility of individual proteins as a function of pl I (92).
The latter method can also suggest the type of chromatography (i.e. anion or
cation exchange, chromatofocusing) suitable for further purification or guide
toward other chromalographies (adsorption. size exclusion, hydrophobic
interaction, or affinity). Capillary electrophoresis is well adapted for purity
analysis (52). Ammo acid composition and sequencing of N- and C-termim
verify in part the integrity of primary structure (54), ESI and MA1.DI mass
spectrometries are powerful tools in recombinant protein chemistry (93).
Homogeneity of nucleic acids is probed by electrophoresis in gels containing
urea (50, 89). Radioactive end-labelling enables detection of low levels of
cleavage in ribose-phosphate chains (84). NMR detects small size con-
taminants and gives structural information on biomolecules (94). Useful
methods for detecting conformational heterogeneity are given in Table 6.
Figure 1. (Left) Comparison of the resolution of ion exchange HPLC and IEF. Aspartyl-
tRNA synthetase from yeast (250 ug, pure according to standard ion exchange
chromatography) was fractionated by anion exchange HPLC on a Mono Q column (i.d.
5 mm x length 5 cm, v 1 ml, Pharmacia) in 50 mM Tris-HCI buffer pH 7.5, and was
eluted at 0.5 ml/min with increasing NaCI concentration. IEF was performed on aliquots
(3 (j.g protein) of the fractions. The polyacrylamide gel was 10 x 10 cm2 (thickness
0.5 mm) and contained 2% (w/v) ampholytes (pH range 4-7). Staining with Coomassie
Blue R-250 reveals several protein populations differing by charge. (Right) Batch-
dependent variation in the microheterogeneity of pure aspartyl-tRNA synthetase. Six
batches of protein purified according to a standard procedure and having the same
specific activity, were compared by IEF under native conditions (samples of 5 ug protein,
a dimer with a subunit Mr of 60000, were analysed. Differences in charge result from
uncontrolled proteolysis between positions 14 and 33 in the polypeptide chain.
32
2: Biochemical aspects and handling of macromolecular solutions
material, change the sequence of events (by inverting chromatographic steps)
or the steps themselves (by using other chromatographic matrices). To avoid
cross-contamination never mix batches of pure macromolecules even when
they look apparently identical. A small shift in the elution from a chroma-
tography column or a preparation done on the same columns but at another
scale or temperature can introduce other contaminants in active fractions.
Such variability can sometimes be detected by IEF (Figure 1). Clean and
sterilize by filtration (e.g. over 0.22 um porosity membranes) all solutions in
contact with pure macromolecules. Use chemically inert and autoclavable
chromatography matrices which do not release molecules (e.g. Trisacryl, IBF
Biotechnics, or TSK gels, Merck).
Macromolecules can be rendered more homogeneous in various ways.
Addition of protease inhibitors is generally effective (Table 4) (95, 96). Assays
to detect proteases by solubilization of clotted protein or by degradation of
labelled peptides are commercially available (e.g. Peptag™, Boehringer). A
cocktail of inhibitors should contain at least one specific for each protease
class; an example is given in Protocol 2, On a small scale, chromatography
over a column of immobilized inhibitors (e.g. a2-macroglobulin) or substrate
analogues (like arginine or benzamidine) may trap proteases. The major
drawback of inhibitors lies in their possible binding to or inactivating of the
proteins they should protect. Over-production in strains deprived of harmful
proteases is a common solution to proteolysis (97).
Reagents
• Buffer solutions containing 10% (v/v) • 10-3 M stock solutions of peptidic inhib-
glycerol and 10-3 M EDTA itorsa (pepstatin, bestatin, and E-64 from
. 0.1 M stock solution of DIPF (Sigma) pre- Sigma) in ethanol:water (50:50)
pared by diluting a 1 g commercial sample • Reducing agents (2-mercaptoethanol and
(about 1 ml) in 50 ml cold anhydrous isopro- DTE or DTT) as stock solutions at 10-1 M
panol; always keep this solution at -20°C
Method
1. Add DIPF, peptidic inhibitors, and 2-mercaptoethanol (DTE or DTT)b in
buffer solutions just prior to use (final concentrations 5 x 10-4 M, 5 x
10-6 M, and 5 x 10-3 M, respectively).
33
B. Lorber and R. Giege
Protocol 2. Continued
2. Add inhibitors afresh before cell disruption and at each step of the
isolation procedure.
a
Experimenters should be aware of low solubility, limited stability, affinity, and reversibility of
inhibitors.
b
Final concentration 10-3 to 10-4 M.
Dissolved crystals
Spectrophotometry/fluorimetry Characterization and quantitation of
molecules
Gel or capillary electrophoresis Characterization of macromolecules, size
Gel or capillary IEF (for proteins only) Characterization and charge
Column chromatography (microscale) Characterization and quantitation (115)
Activity assays Biological activity
Mass spectrometry Molecular mass of components, detection
of microheterogeneities, of counterions,
sequencing (52-54, 93)
Sequencing (protein, nucleic acid) Integrity of primary structure
'Other methods may be employed in particular cases, e.g. dichroism (116), analytical ultracentrifuga-
tion(117).
b
c
For proteins, most stains used in light microscopy. For nucleic acids, see Protocol 1.
d
See Chapter 14.
See Section 6.
for the analysis of crystals. Some are applicable to the crystalline material
itself whereas others require solubilized molecules. Since crystals contain only
micrograms of macromolecules most methods must be scaled down. In all
cases, the aims are:
(a) To verify that crystals contain the desired macromolecules and in the
right stoichiometry in the case of co-crystals.
(b) To ensure that macromolecules are in an active conformation within
crystals (for enzymes, this can be asserted by in situ catalytic assays pro-
vided active sites are accessible and ligands can diffuse harmlessly within
the crystalline lattice).
(c) To compare the macromolecules in the crystalline state with those in
solution (e.g. by laser Raman spectroscopy).
35
B. Lorber and R. Giege
Prior to any analysis, uncrystallized macromolecules or amorphous material
(present within the mother liquor or deposited onto the crystal faces) must be
washed away. This is done by transferring crystals several times in large
volumes of mother liquor.
where N is Avogadro's number (6.02 X 1023), rs the density of free and bound
solvent (g/cm3), M the protomer molar weight (g/mol), and vp the partial
specific volume of the dry protein (cm3/g). Hence, the solvent fraction:
In the case of proteins, the mass per asymmetric unit Mp is obtained with good
approximation with the simplified formula (120):
assuming the density of the solvent equals that of water (rw = 1.0) and vp =
0.74 cm3/g. In practice, the volume of the unit cell is calculated from cell
parameters measured on X-ray diffraction patterns (Chapter 14). The molar
weight of the macromolecule is determined by biochemical or biophysical
methods (Table 6). Partial specific volumes of proteins (vp) are either assumed
to be equal to the inverse of their density, computed from partial specific
volumes of individual amino acids (121) and the amino acid composition, or
approximated as above. Partial specific volumes of nucleic acids can be
approximated to 0.54 cm3/g (122). The densities of solvent and crystals are
determined experimentally. The resulting number of protomers is rounded to
the nearest integer.
36
2: Biochemical aspects and handling of macromolecular solutions
Two common methods for the measurement of crystal densities use either
organic solvents or solutions of large polymers in which the sedimentation of
the crystals is compared with internal references. Both are applicable to
proteins, nucleic acids, and viral or ribosomal particles. They have their
advantages and limitations. Other methods are described separately below.
The densities of most protein crystals range from 1.10-1.60 g/cm3 (123).
6.2.1 Organic solvent and Ficoll™ methods
The first method, adapted from the small molecule field, uses mixtures of
water saturated carbon tetrachloride or chloroform and xylene. Experimental
details are given in Protocol 3, part A. Measurements are sensitive to the
presence of mother liquor around the crystals and difficulties may arise when
trying to remove the excess solution. Readers are referred to ref. 123 for
further details.
B. Ficoll™ method
1. Prepare a series of solutions (from 30-60%, w/w) by mixing appropriate
37
B. Lorber and R. Giege
Protocol 3. Continued
amounts of Ficoll™ powder and water. Heat at 55°C to dissolve.
Cooled solutions are yellowish and viscous.
2. Work at constant temperature. Prepare gradients in transparent glass
tubes of 0.7-1 mm inner diameter. Use a gradient maker or proceed by
layers of decreasing density. Centrifuge after each layer (5 min, 1000 g)
and end with a longer centrifugation (e.g. 1 h at 3000 g) to smooth the
gradient.
3. Calibrate gradients with drops of carbon tetrachloride (or chloroform)
and toluene mixtures (all saturated with water). The densities of these
organic solutions are measured as in part A.
4. Deposit one or two crystals and a droplet of mother liquor onto the
gradient with a glass micropipette. Centrifuge at 1000 g for 1-60 min
and compare the displacement of the crystals with that of calibrated
drops.
5. Take X-ray diffraction pictures to verify that unit cell parameters are
unchanged.
and the unit cell parameters, the fraction volume occupied by the solvent can
be estimated and an approximate density can be extrapolated. More accurate
results are expected when molar absorption coefficients are high. For crystals
containing more than one type of macromolecule, protein and nucleic acid or
a macromolecule and a smaller molecule (e.g. a protein and a ligand), their
stoichiometry can be calculated when molecules differ by their spectroscopic
properties.
References
1. Wierenga, R. K.? Lalk, K. H., and Hol, W. G. J. (1987). J. Mol. Biol, 198, 109.
2. Matthews, B. V. (1968). J. Mol. Biol., 33, 491.
3. Garber, M., Davydova, N., Eliseikina, I., Fomenkova, N., Grysaznova, O.,
Gryshkovskaya, I., et al. (1996). J. Cryst. Growth, 168, 301.
4. Thygesen, J., Krumholz, S., Levin, I., Zaytzev-Bashan, A., Harms, J., Bartels, H.,
et al. (1996). J. Cryst. Growth, 168, 308.
5. Zaccai, G. and Eisenberg, H. (1990). Trends Biochem. Sci., 15, 333.
6. Kern, D., Giege, R., Robbe-Saul, S., Boulanger, Y., and Ebel, J.-P. (1975).
Biochimie, 57, 1167.
7. Adelman, R. C. and Dekker, E. E. (ed.) (1985). Modification of proteins during
aging. Alan R. Liss, New York.
8. Lorber, B. and DeLucas L, J. (1990). FEBS Lett., 241, 14.
9. Sambrook, J., Fritsch, E. F., and Maniatis, T. (ed.) (1989). Molecular cloning: a
laboratory manual, 3 Vols. Cold Spring Harbor Laboratory Press.
10. Goeddel, D. V. (ed.) (1991). Methods in enzymology, Vol. 185, pp. 1-599.
Academic Press, London.
11. Kurland, C. G. and Dong, H. (1996). Mol. MicrobioL, 21, 1.
12. Schein, C. H. (1990). Biotechnology, 8, 308.
13. Miroux, B. and Walker, J. E. (1996). J. Mol. Biol., 260, 289.
14. Zahn, K. and Landy, A. (1996). Mol. Microbiol, 21, 69.
15. Calderone, T. L., Stevens, R. D., and Oas, T. G. (1996). J. Mol. Biol, 262, 407.
16. Carter, C. W. (1988). J. Cryst. Growth, 90, 168.
17. Horwich, A. L., Neupert, W., and Hartl, F. U. (1990). Trends Biotech., 8, 126.
18. Richardson, C. D. (ed.) (1995). Methods in molecular biology, Vol. 39, pp. 1-420.
Humana Press, Totowa, NJ.
19. Marston, F. A. O. (1986). Biochem. J., 240, 1.
20. Spirin, A. S., Baranov, V. L, Ryabova, L. A., Ovodov, S. Y., and Alakhov, Y. B.
(1988). Science, 242, 1162.
21. Perona, J. J., Swanson, R., Steitz, T. A., and Soil, D. (1988). J. Mol. Biol., 202, 121.
22. McPherson, M. J. (ed.) (1991). Directed mutagenesis: a practical approach,
pp. 1-288. IRL Press, Oxford.
23. Oubridge, C., Io, N., Teo, C. H., Fearnley, I., and Nagai, K. (1995). J. Mol. Biol.,
249, 409.
24. Harris, E. L. V. and Angal, S. (ed.) (1990). Protein purification applications: a
practical approach, pp. 1-192. IRL Press, Oxford.
25. Yin, Y. and Carter, C. W., Jr. (1996). Nucleic Acids Res., 24, 1279.
39
B. Lorber and R. Giege
26. Vuillard, L., Baalbaki, B., Lehmann, M., Norager, S., and Legrand, P. (1996). /.
Cryst. Growth, 168, 150.
27. Oliver, R. W. A. (ed.) (1989). HPLC of macromolecules: a practical approach,
pp. 1-252. IRL Press, Oxford.
28. Righetti, P. G. (1983). Isoelectric focusing: theory, methodology and application.
Elsevier, Amsterdam.
29. Wagner, H. (1989). Nature, 341, 669.
30. Ott, G., Dorfler, S., Sprinzl, M., Muller, U., and Heinemann, U. (1996). Acta
Cryst., D52, 871.
31. Wahl, M. C., Ramakrishnan, B., Ban, C, Chen, X., and Sundaralingam, M. (1996).
Acta Cryst., D52, 668.
32. Gate, J. H., Gooding, A. R., Podell, E., Zhou, K., Golden, B. L., Kundrot, C. E., et
al (1996). Science, 273, 1678.
33. Aoyama, K. (1996). J. Cryst. Growth, 168, 198.
34. Gulewicz, K., Adamiak, D., and Sprinzl, M. (1985). FEBS Lett., 189, 179.
35. Giege, R., Dock, A.-C., Kern, D., Lorber, B., Thierry, J.-C., and Moras, D. (1986).
J. Cryst. Growth, 76, 554.
36. Stoll.V. S. and Blanchard, J. S. (1990). In Methods in enzymology (ed. M. P.
Deutscher), Vol. 182, pp. 24-38. Academic Press, London.
37. Yang, A. S. and Honig, B. (1993). J. Mol. Biol., 231, 459.
38. Darby, N. J. and Creighton, T. E. (ed.) (1989). Protein structure: a practical
approach, pp. 1-120. IRL Press, Oxford.
39. Norde, W. (1995). Cells Mater., 5, 97.
40. Baldwin, R. L. (1996). Biophys. J., 71, 2056.
41. Strambini, G. B. and Gabellieri, E. (1996). Biophys. J., 70, 971.
42. Sousa, R. (1995). Acta Cryst., D51, 271.
43. Crowe, L. M., Reid, D. S., and Crowe, J. H. (1996). Biophys. J., 71, 2087.
44. Jeruzalmi, D. and Steitz, T. A. (1997). J Mol. Biol., 274, 748.
45. Buttler, S. L. and Falke, J. J. (1996). Biochemsitry, 35, 10595.
46. Chayen, N. E., Radcliffe, J. W., and Blow, D. M. (1993). Protein Sci., 2, 113.
47. Mittl, P. R. E., Di Marco, S., Krebs, J. F., Bai, X., Karanewsky, D. S., Priestle, J. P.,
etal. (1997). J. Biol. Chem., 272, 6539.
48. Scott, W. G., Finch, J. T., Grenfell, R., Fogg, J., Smith, T., Gait, M. J., et al. (1995).
J. Mol. Biol., 250, 327.
49. Hames, B. D. and Rickwood, D. (ed.) (1990). Gel electrophoresis of proteins: a
practical approach, pp. 1-250. IRL Press, Oxford.
50. Rickwood, D. and Hames, B. D. (ed.) (1990). Gel electrophoresis of nucleic acids:
a practical approach, pp. 1-336. IRL Press, Oxford.
51. Righetti, P. G., Gianazza, E., and Gelfi, C. (1988). Trends Biochem. Sci., 13, 333.
52. Karger, B. L. and Hancock, W. S. (ed.) (1996). Methods in enzymology, Vol. 270,
pp. 1-611 and Vol. 271, pp. 1-543. Academic Press, London.
53. Muddiman, D. C., Bakhtiar, R., Hofstaler, S. A., and Smith, R. C. (1997). J. Chem.
Edu.,74, 1288.
54. Ries-Kautt, M., Ducruix, A., and Van Dorsselaer, A. (1994). Acta Cryst., D50, 366.
55. Matsudaira, P. (ed.) (1993). A practical guide to protein and peptide purification
and microsequencing. Academic Press, Inc., San Francisco.
56. Stoscheck, C. M. (1990). In Methods in enzymology (ed. M. P. Deutscher), Vol.
182, pp. 50-68. Academic Press, London.
40
2: Biochemical aspects and handling of macro molecular solutions
57. Pace, C. N., Vajdos, F., Fee, L., Grimsley, G., and Gray, T. (1995). Protein Sci., 4,
2411.
58. Warburg, O. and Christian, W. (1941). Biochem. Z., 310, 384.
59. Ehresmann, B., Imbault, P., and Weil, J.-H. (1973). Anal. Biochem., 54, 454.
60. Fersht, A. (1977). Enzyme structure and mechanism. Freeman & Company,
Reading.
61. Gueron, M. and Leroy, J.-L. (1978). Anal. Biochem., 91, 691.
62. Ames, B. N. and Dubin, D. T. (1960). /. Biol. Chem., 235, 769.
63. Holley, R. W., Apgar, J., and Merrill, S. H. (1961). /. Biol. Chem., 236, PC42.
64. Judge, R. A., Johns, M. R., and White, E. T. (1995). Biotech. Bioeng., 48, 316.
65. Vekilov, P. G. and Rosenberger, F. (1996). J. Cryst. Growth, 158,540.
66. Skouri, M., Lorber, B., Giege, R., Munch, J.-P., and Candau, S. J. (1996). J. Cryst.
Growth, 152, 209.
67. McPherson, A., Malkin, A. J., Kuznetsov, Y. G., and Koszelak, S. (1996). J Cryst.
Growth, 168, 74.
68. Provost, K. and Robert, M.-C. (1995). J. Cryst. Growth, 156, 112.
69. Hirschler, J., Charon, M.-H., and Fontecilla-Camps, J. C. (1995). Protein Sci., 4, 2573.
70. Doudna, J. A., Grosshans, C., Gooding, A., and Kundrot, C. E. (1993). Proc. Natl.
Acad. Sci. USA, 90, 7829.
71. Lawson, D. M., Artymiuk, P. J., Yewdall, S. J., Smith, J. M. A., Livingstone, J. C.,
Treffry, A., et al. (1991). Nature, 349, 541.
72. Ray, W. J. and Puvathingal, J. (1985). Anal. Biochem., 146, 307.
73. Jurnak, F. (1986). J. Cryst. Growth, 76, 577.
74. Bello, J. and Nowoswiat, E. F. (1965). Biochim. Biophys. Acta, 105, 325.
75. Kuhlbrandt, W. (1988). Q. Rev. Biophys., 21, 429.
76. Lorber, B., Bishop, J. B., and DeLucas, L. J. (1990). Biochim. Biophys. Acta, 1023,
254.
77. Achstetter, T. and Wolf, D. H. (1985). Yeast, 1, 139.
78. Barrett, A. J. and Salvesen, G. (ed.) (1986). Research monographs in cell and
tissue physiology, Vol. 12, pp. 1-661. Elsevier, Amsterdam.
79. Calling, M. J. (ed.) (1986). Plant proteolytic enzymes, 2 Vol. CRC Press, Boca
Raton, FL.
80. Bond, J. S. and Butler, P. E. (1987). Annu. Rev. Biochem., 56, 333.
81. Arfin, S. M. and Bradshaw, R. A. (1988). Biochemistry, 27, 7979.
82. Wandersman, C. (1989). Mol. Microbiol, 3, 1825.
83. Howe, C. J. and Ward, E. S. (ed.) (1989). Nucleic acids sequencing: a practical
approach, pp. 1-254. IRL Press, Oxford.
84. Moras, D., Dock, A.-C., Dumas, P., Westhof, E., Romby, P., Ebel, J.-P., et al.
(1985). J. Biomol. Struct. Dyn., 3, 479.
85. Wold, F. and Moldave, K. (ed.) (1984). Methods in enzymology, Vol. 106,
pp. 1-574 and Vol. 107, pp. 1-688. Academic Press, London.
86. Kendall, R. L., Yamada, R., and Bradshaw, R. A. (1990). In Methods in
enzymology (ed. D. V. Goeddel), Vol. 185, pp. 398-08. Academic Press, London.
87. Rudd, P. M. and Dwek, R. A. (1997). Crit. Rev. Biochem. Mol. Biol., 32, 1.
88. Grosjean, H. and Benne, R. (ed.) (1998). Modification and editing of RNA. ASM
Press, Washington DC.
89. Lorber, B., Giege, R., Ebel, J.-P., Berthet, C., Thierry, J.-C, and Moras, D. (1983).
J. Biol. Chem., 258, 8429.
41
B. Lorber and R. Giege
90. Van der Laan, J. M., Swarte, M. B. A., Groendijk, H., Hol, W. G. J., and Drenth,
J. (1989). Eur. J. Biochem., 179, 715.
91. Rhodes, D. G. and Laue, T. M. (1990). In Methods in enzymology (ed. M. P.
Deutscher), Vol. 182, pp. 555-66. Academic Press, London.
92. Lorber, B., Kern, D., Mejdoub, H., Boulanger, Y., Reinbolt, J., and Giege, R.
(1987). Eur. J. Biochem., 165, 409.
93. Andersen, J. S., Svensson, B., and Roepstorff, P. (1996). Nature Biotech., 14, 449.
94. Wuthrich, K. (1995). Acta Cryst, D51, 249.
95. Beynon, R. J. and Bond, J. S. (ed.) (1989). Proteolytic enzymes: a practical
approach, pp. 1-278. IRL Press, Oxford.
96. Beynon, R. J. (1989). In Protein purification methods: a practical approach (ed. E.
L. V. Harris and S. Angal), pp. 40-51. IRL Press, Oxford.
97. Gottesman, S. (1990). In Methods in enzymology (ed. D. V. Goeddel), Vol. 185,
pp. 119-29. Academic Press, London.
98. Berger, S. L. and Birkenmeier, C. S. (1979). Biochemistry, 18, 5143.
99. Scheele, G. and Blackburn, P. (1979). Proc. Natl. Acad. Sci. USA, 76, 4898.
100. Thompson, S. T., Cass, K. H., and Stellwagen, E. (1975). Proc. Natl. Acad. Sci.
USA, 72, 669.
101. Wierenga, K., Huizinga, J. D., Gaastra, W., Welling, G. W., and Beintema, J. J.
(1973). FEBS Lett., 31, 181.
102. Bergfors, T., Rouvinen, J., Lehtovaara, P., Caldentey, X., Tomme, P.,
Claeyssens, M., et al. (1989). J. Mol. Biol, 209, 167.
103. Kalisz, H. M., Hecht, H. J., Schomburg, D., and Schmid, R. D. (1990). J. Mol
Biol, 213, 207.
104. Stura, E. A., Chen, P., Wilmot, C. M., Arevalo, J. H., and Wilson, I. A. (1992).
Proteins, 12, 24.
105. Baker, H. M., Day, C. L., Norris, G. E., and Baker, E. N. (1996). Acta Cryst.,
DSO, 380.
106. Nitanai, Y., Satow, Y., Adachi, H., and Tsujimoto, M. (1996). J. Cryst. Growth,
168, 280.
107. Anderson, W. F., Boodhoo, A., and Mol, C. D. (1988). J. Cryst. Growth, 90, 153.
108. Bentley, G. A., Boulot, G., Riottot, M. M., and Poljak, A. (1988). J. Cryst.
Growth, 90, 213.
109. Spangler, B. D. and Westbrook, E. M. (1989). Biochemistry, 28, 1333.
110. Moreau, H., Abergel, C., Carriere, R, Ferrato, F., Fontecilla-Camps, J. C.,
Cambillau, C., et al. (1992). J. Mol. Biol, 225, 147.
111. Weber, W., Wenisch, E., Gunther, N., Marnitz, U., Betzel, C., and Righetti, P. G.
(1994). J. Chromatogr., A679, 181.
112. Kirsten, H. and Christen, P. (1983). Biochem. J., 211, 427.
113. Chen, M., Lord, R., Giege, R., and Rich, A. (1975). Biochemistry, 14, 4385.
114. Mozzarelli, A. and Rossi, G. L. (1996). Annu. Rev. Biophys. Biomol. Struct., 25,
3430.
115. Lorber, B. and Giege, R. (1985). Anal. Biochem., 146, 402.
116. Ford, R. C., Picot, D., and Garavito, R. M. (1987). EMBO J., 6, 1581.
117. Schuster, T. M. and Toedt, J. M. (1996). Curr. Opin. Struct. BioL, 6, 650.
118. Palmer, K. J., Ballantyne, M., and Galvin, J. A. (1948). J. Am. Chem. Soc., 70, 906.
119. Salemme, F. R., Genieser, L., Finzel, B. C., Hilmer, R. M., and Wendoloski, J. J.
(1988). J. Cryst. Growth, 90, 273.
42
2: Biochemical aspects and handling of macromolecular solutions
120. Bode, W. and Schirmer, T. (1985). Biol. Chem. Hoppe-Seyler, 366, 287.
121. Cohn, E. J. and Edsall, J. T. (ed.) (1943). In Proteins, amino acids and peptides,
p. 370. Van Nostrand Reinhold, Princeton, NJ.
122. Cohen, G. and Eisenberg, H. (1968). Biopolymers, 6, 1077.
123. Matthews, B. W. (1985). In Methods in enzymology (ed. H. W. Wyckoff, C. H.
W. Hirs, and S. N. Timasheff), Vol. 114, pp. 176-87. Academic Press, London.
124. Westbrook, E. M. (1985). In Methods in enzymology (ed. H. W. Wyckoff, C. H.
W. Hirs, and S. N. Timasheff), Vol. 114, pp. 187-96. Academic Press, London.
125. Sewell, B. T., Bouloukos, C., and Von Holt, C. (1984). J Microsc., 136, 103.
126. Mikol, V. and Gieg6, R. (1992). In Crystallization of nucleic acids and proteins: a
practical approach (ed. A. Ducruix and R. Giege), 1st edn, pp. 219-40. IRL Press,
Oxford.
43
This page intentionally left blank
3
1. Introduction
The number of published 3D structures has increased exponentially in the last
decade and the resulting mass of structural data has contributed significantly
to the understanding of mechanisms underlying the biology of living cells.
However, these mechanisms are so complex that structural biologists face still
greater challenges, such as the study of higher-order functional complexes. As
an example, we can mention the protein complexes that assemble around
activated growth factor receptors to allow the transduction of extracellular
signals through the membrane and inside the cell (1).
Because of their diverse intrinsic properties, proteins exhibit variable diffi-
culty for structural biology studies. Before the rise of recombinant expression
methods, only a minority of protein structures were determined, representing
mainly favourable cases: proteins of high abundance in their natural source
which could be purified and crystallized, in contrast to rare proteins that
were often refractory to crystallization. The advent of methods for recom-
binant protein overexpression was a breakthrough in this area. It was
followed by an increasing number of publications describing the crystal-
lization of proteins, not under their native form, but in modified versions after
sequence engineering.
First we will consider the classical use of molecular biology applied to
optimize the expression system for a recombinant protein for structural
biology, without modification of its sequence. In the second part, we will deal
with molecular biology procedures aimed at engineering the properties of a
protein through sequence modifications in order to make its crystallization
possible. In the last part we will give an example where molecular biology can
help solve a crystallographic problem, namely that of phase determination by
introducing anomalous scatterers (e.g. selenium atoms) into the protein of
interest.
P. F. Berne et al.
Method
1. Resolubilize the insoluble protein in 6-8 M guanidinium hydrochloride.
2. Dilute it to ~ 100 ug/ml in a buffer with ~ 1.5 M NaCI, 4.5 M guani-
dinium hydrochloride, 10% glycerol, 10 mM DTT, 5% ethylene glycol,
and other components necessary for the activity of the protein. Use a
weakly buffered solution compatible with activity (i.e. ~ 40 mM
acetate).
3. Apply the mixture to a hipropyl hydrophobic interaction column pre-
equilibrated with the same buffer used in step 2, plus 2 M NaCI, 10-20
mM DTT, additional components as before, 10% glycerol, and 5%
ethylene glycol. For a 7.75 x 100 mm2 semi-preparative column, load
-1.0 mg of diluted protein. For a preparative 2.2 x 10 cm2 column,
load ~ 5.0 mg.
4. Elute with a gradient of the starting buffer mixed progressively with
the same buffer without NaCI.
49
P. F. Berne et al.
Alternatives to renaturation trials, when a protein is produced in the
bacteria as a totally insoluble form, are to turn to other expression hosts,
to try to express the protein as a fusion protein, or to have it secreted by E.
coli, which will provide a different environment for the folding of the protein.
This is done by fusing the protein to an N-terminal signal peptide. One
commercially available vector is suitable for this purpose (Table 1).
The response surface methodology described in Chapter 4 offers a rational,
coherent way to evaluate and optimize the variety of factors influencing the
proportion of soluble protein.
51
P. F, Berne et al.
Method
1. Perform a standard PCR reaction (25 cycles: 30 sec at 94°C, 1 min at
60°C, 1 min per kb amplified at 72°C) with the selected primers using
1 ng of a plasmid carrying your gene as template. Include control
reactions with individual primers.
2. Isolate the PCR product using a phenol:chloroform extraction,
followed by ethanol precipitation or using one of the commercially
available kits.
3. Digest the acceptor vector (1 ug) and your PCR product with the
selected restriction enzymes. Isolate the fragments of interest using
agarose gel electrophoresis. Under long UV illumination (365 nm), cut
out gel pieces containing the digested vector and PCR insert. Use one
of the commercial kits for the extraction of DNA out of the gel.
4. Perform a standard ligation reaction in the presence of about 1 nM of
the linearized vector and 2 nM of the PCR insert (concentration is not
critical). Include a ligation of the vector alone as control.
5. Transform competent bacteria (e.g. TG1) strain with 2 ul of ligation.
6. Prepare a small amount of DNA (commercial miniprep kit) of a few or
several clones and check for the presence of the insert using
restriction enzymes.
7. Sequence a few positive clones to make sure that no error has been
introduced by the PCR reaction. The complete region amplified by PCR
has to be checked.
Method
1. Prepare, as substrate for the assays, a stock solution of your protein
(2-10 mg/ml) in a buffer in which it is stable.
2. Select proteases exhibiting broad specificity (with several potential
sites inside proteins, e.g. trypsin, chymotrypsin, subtilisin, thermo-
lysin, papain, endoproteinase Glu-C). Prepare stock solutions (2-3
mg/ml) and store them as frozen aliquots in appropriate buffers (see
Table 3) containing 20% glycerol.a
58
3: Molecular biology for structural biology
3. Prepare on ice serial dilutions (0.1 mg/ml down to 10 ng/ml) of the
proteases in their specific incubation buffers. In a first series of
experiments, serial tenfold dilutions might be done and the conditions
may then be refined in subsequent experiments.
4. Mix a small volume of the concentrated protein solution (2-5 ul) with
the diluted protease solution (40 ul for example) in order to reach a
final concentration of about 0.2 mg/ml.
5. Incubate for 1 h at room temperature or 37°C. Stop the reaction by
addition of a specific inhibitor. We recommend PMSF (1 mM final
concentration) for all the mentioned serine proteases and E-64 for
papain (cysteine protease) (see also Chapter 2).
6. Analyse the reaction products by migration on SDS-PAGE. Try to
identify the products of proteolysis at preferred sites. These products
correspond to a unique cleavage site at low protease concentration,
whereas a higher protease concentration leads to non-specific
cleavage at multiple sites.
7. For each protease, optimize the protease concentration that leads to
the limited cleavage at the preferred site. Once these conditions have
been established, perform a larger scale reaction in order to isolate a
larger amount of the proteolysis products.
8. Isolate the gel band of interest using PAGE followed by transfer on a
polyvinylidene difluoride (PVDF) membrane (Millipore). Give the
samples to a specialized service for N-terminal sequence analysis.6
9. The sequence information originating from various proteases may
pinpoint areas in the protein that are especially sensitive to proteo-
lysis. These positions probably correspond to connections between
various domains in the protein.
'Appropriate storage conditions have to be investigated for other proteases. Most importantly,
the activity has to be strictly reproducible from one experiment to another, and this is the
reason for using frozen aliquots (see also Chapter 2).
b
Alternatively, the protein species might be analysed by mass spectrometry, as in ref. 27.
Method
1. Design two mutagenic primers, a sense primer (primer a), encoding
the target sequence, and an antisense primer (primer b), strictly com-
plementary to primer a. The sense primer could include one or a few
modified bases, an insertion, or a deletion, provided that these modi-
fications are surrounded by two 13 base segments strictly com-
plementary to the starting sequence. These segments are necessary to
ensure that the primers will recognize the template in spite of the
sequence changes.
2. In order to design two external primers (20- to 22-mers with 50-60%
GC content), identify two unique restriction sites located each side of
the mutation, separated by 200-1500 base pairs. Choose a sense
primer (primer c) overlapping with or immediately upstream from the
upstream site and an antisense primer (primer d) overlapping with or
immediately downstream from the downstream site.
3. Perform two independent PCR reactions using the starting vectora as
template (10 ng) and standard PCR conditions. The first reaction uses
primers a and d, the second reaction uses primers b and c. In this way,
the PCR products will be two fragments overlapping over the segment
encoded by the mutagenic primers.
4. Isolate the two PCR products using agarose gel electrophoresis.
5. Perform a third PCR reaction using the two external primers (c and d)
and, as a template, a mixture of the two previous PCR products (~ 10
ng each). Using gel electrophoresis, check an aliquot for the presence
of the expected final PCR product.b
6. Follow Protocol 2, steps 3-7.
a
If available, use as template a vector containing your gene but different from the final vector.
That will facilitate the identification of the mutated versus wild-type clones. The template
vector should, of course, contain all the region between the two external primers and be
recognized by these primers.
b
During the hybridization step, some hybrids will appear carrying one strand of each fragment
annealed through their overlapping segment. These hybrids will be complemented to double-
strand DNA by the polymerase and will constitute an efficient template for further
amplification by the external primers.
60
3: Molecular biology for structural biology
The overlap extension method is described in Protocol 4, adapted from ref.
35. Note that there are simpler strategies for mutagenesis in some particular
cases (e.g. mutations close to an existing restriction cleavage site). The
method proposed here has the advantage of being general. It is cheap and
relatively easy to perform, and one can envisage constructing several directed
mutations in parallel, if required. In this case, the external primers are
common to all mutations and only two oligonucleotides are required for
each specific mutant. Additionally, it is often possible to combine directed
mutagenesis with subcloning in another expression vector, as described in
Protocol 2, by designing appropriately the external primers.
Method
1. Choose the part of the protein DNA sequence that you want to
mutagenize randomly. It should be included between two unique
restriction sites. It may be necessary to first introduce these restriction
sites by directed mutagenesis.
61
P. F. Berne et al.
Protocol 5. Continued
2. Design oligonucleotides as primers for a PCR amplification. Choose a
sense primer immediately upstream from the first restriction site and
an antisense primer immediately downstream from the second site.
3. Perform the PCR reaction under specific conditions in order to favour
the misincorporation of deoxyribonucleotides by Taq DNA polymer-
ase. The PCR reaction should include, in addition to the usual buffer,
0.5 mM MnCI2, 7 mM MgCI2, and modified concentrations of dNTPs
(0.2 mM of dATP and dGTP, 1 mM of dTTP and dCTP). Use a very low
initial concentration of the plasmid template (e.g. 1 pg in a reaction
volume of 50 ul), as the error rate of the PCR will increase if the
amplification factor is increased.a
4. Follow Protocol 2, steps 4-7.
5. Sequence several clones to determine the average number of muta-
tions per clone. An average rate of one per clone seems reasonable for
most applications. It might be necessary to sequence a series of clones
coming from PCR reactions starting with various amounts of DNA
template in order to identify the conditions that generate the
appropriate rate of mutation.b
6. Using the ligation mixture selected according to step 5, transform an
appropriate expression strain, e.g. BL21.
7. Grow small volume cultures of several clones to characterize potential
protein variants.
a
It might be necessary to adjust the amount of polymerase, the annealing temperature, the
length of elongation, the number of cycles, in order to get a good yield of amplification,
depending on the length of the amplified fragment and on the choice of the primers.
b
The rate of various mutation types might be adjusted by varying the nucleotide
concentration, as described by Fromant et al. (63).
Method
1. Add 100 ul of a bacterial growth medium, supplemented with the
appropriate antibiotics, to each well of a microtitre plate.
2. With a tip, pick up colonies out of a Petri dish originating from a
transformation made the day before. Inoculate each well with a unique
colony and strike the tip on a new dish at an identified position to keep
a replicate of each clone.
3. Let the cultures grow by incubating the plate for 6 h at 25°C. In our
case, the gene was cloned into vector pET3®, which exhibited a
sufficient level of expression without IPTG under these conditions.
When working with a vector with a more tightly regulated promoter, it
could be necessary to induce the expression in a more specific way.
63
P. F. Berne et al.
Protocol 6. Continued
4. Collect the cells by centrifuging the plate for 10 min at 2000 g (Sigma
centrifuges, for instance, have rotors that accept microtitre plates).
Suck in carefully the medium. The cells can be frozen at this stage for
future use.
5. This step and the following are done easily using a multichannel
pipette. Lyse the cells by resuspending the pellets in 100 ul of a buffer
containing 50 mM Tris, 20 mM NaCI, 0.1% Triton X-100, 0.5 mM EDTA,
10 ug/ml lysozyme pH 8.4. Incubate for 20 min at room temperature.
6. Add 10 ul of a DNase solution (10 ug/ml in the same buffer as above
plus 20 mM MgCI2) and homogenize each well. Incubate for 20 min at
room temperature.
7. Centrifuge the plate for 20 min at 2000 g. Carefully transfer the
supernatant to another plate.
8. Perform a purification step, which could be an incubation with an
affinity resin. In our case, the protein could be precipitated selectively
by the addition of 0.6 M ammonium sulfate and recovered after
centrifugation and elimination of the supernatant.
9. Analyse the protein by native PAGE electrophoresis. This gives a good
idea of the state of the protein under native conditions. Ideally, the
protein is expected to migrate as a unique and sharp band.
Ingredients Concentration
1. Minimal mediuma Minimal medium Ab or M9 (18), with carbon
source 5 g/litre
2. All amino acids except methionine" 40 mg/litre
3. Selenomethioninec 20-60 mg/litre
4. LB (12) x%(v/v) to be determinedd
a
One can also use LeMaster's medium (60) instead of 1 and 2.
b
Ausubel, F. M., Brent, R., Kingston, R. E., Moore, D. D., Seidman, J. G., Smith, J. A., and Struhl, K.
(ed.) (1987). Current protocols in molecular biology. Greene Publishing Associates and Wiley
Interscience, New York, NY.
c
L-selenomethionine can be purchased from Fisher/Acros or Sigma.
d
As an example, Yang et al. used a 100 ml starter medium containing 5% (v/v) LB for a 20 litre
fermenter (43).
Ingredients Concentration
1. Minimal mediuma Minimal medium Ab or M9 (12), with carbon
source 5 g/litre
2. All amino acids except metnioninea 40 mg/litre
3. Selenomethionine 20-60 mg/litre
4. Thiamine 2 mg/litre
5. Biotin (if needed) 2 mg/litre
a
One can also use LeMaster's medium (60) instead of 1 and 2.
b
Ausubel, F. M., Brent, R., Kingston, R. E., Moore, D. D., Seidman, J. G., Smith, J. A., and Struhl, K.
(ed.) (1987). Current protocols in molecular biology. Greene Publishing Associates and Wiley
Interscience, New York, NY.
residual methionine in the purified protein, the dilution factor for the final
inoculation must be adjusted according to the amount of LB used in the
starter inoculum. This amount will be a compromise between a better growth
rate and complete selenomethionine substitution. For each particular strain,
one will have to determine the amount of rich medium in the starter culture,
as well as the optimal concentration of selenomethionine throughout the
fermentation (see Protocol 7).
Method
1. Isolate single colonies by streaking an LB plate (18) supplemented with
antibiotics with strain of interest. Incubate at 37°C.
66
3: Molecular biology for structural biology
2. Ferment the cells in medium containing all appropriate antibiotics.
(a) Inoculate 100 ml of starter medium with a single colony. Shake at
37°C.
(b) Inoculate a 10-20 litre fermentera containing pre-warmed ferment-
ation medium with the 100 ml starter inoculum in mid-log phase.
(c) Monitor cell growth in order to identify times for induction and
harvest.
3. Induce, if necessary.
4. Harvest the cells in mid- to late log phase.b Resuspend the harvested
cells in the appropriate lysis buffer and quick-freeze in dry ice or liquid
nitrogen. Store at-80°C.
a
Cell growth should be done in a fermenter because regulated temperature and pH improve
the yield.
b
We have noticed cell lysis shortly after cells reached late log phase. Care must be taken to
harvest cells as quickly as possible.
Ingredients Concentration
1. Minimal medium Minimal medium Aa or M9 (18), with carbon
source 5 g/litre
2. Thiamine 2 mg/litre
3. Biotin (optional) 2 mg/litre
4. Lysine, phenylalanine, and threonine 100 mg/litre
Isoleucine, leucine, and valine 50 mg/litre
5. Selenomethionineb 60 mg/litre
a
Ausubel, F. M., Brent, R., Kingston, R. E., Moore, D. D., Seidman, J. G., Smith, J. A., and Struhl, K. (ed.)
(1987). Current protocols in molecular biology. Greene Publishing Associates and Wiley Interscience,
New York, NY.
b
L-selenomethionine can be purchased from Fisher/Acros or Sigma.
67
P. F. Berne et al.
involving a met- strain and should be applicable to any prokaryotic strain (see
Protocol 8).
Method
1. Isolate single colonies by streaking an LB plate (18) supplemented with
antibiotics with strain of interest. Incubate at 37°C.
2. Inoculate 1 ml of LB medium with a single colony. Grow overnight.
3. Spin down cells (2 min at 1300 g in a microcentrifuge) and resuspend
in 1 ml of supplemented minimum medium (items 1, 2, and 3 in Table
6). Inoculate 1 litre of the same, pre-warmed medium.
4. Add all seven amino acids at mid-log phase.
5. Induction is done 15 min after addition of the amino acids, if
necessary.
6. Harvest the cells in mid- to late log phase. Resuspend the harvested
cells in the appropriate lysis buffer and quick-freeze in dry ice or liquid
nitrogen. Store at -80°C.
4.3 Eukaryotes
Animal organisms are naturally auxotrophic for methionine and cells can
grow in a selenomethionine containing medium with good incorporation of
the modified amino acid. There are to date only a few examples of seleno-
methionyl protein production in eukaryotic cells. Selenomethionyl human
chorionic gonadotropin (hCG) was produced in insect cells (Sf9) (50, 51) as
well as in Chinese hamster ovary (CHO) cells (52, 53). The selenomethionyl
variant of a functional fragment of sialoadhesin was also produced in CHO
cells (54). The substitution rates reported for these three proteins range from
84-92%.
4.4 Purification
Introduction of selenomethionine into proteins has two consequences that
impact on purification. The altered chemistry of selenium makes substituted
proteins more sensitive to oxidation than natural proteins. Moreover, if
selenium atoms are solvent exposed, they can alter protein solubility and
behaviour on chromatography resins. These properties require the following
modifications to the normal purification as shown in Protocol 9.
68
3: Molecular biology for structural biology
Method
1. Purify as quickly as possible, with modifications to avoid oxidation.
(a) Degas all buffers by boiling or evacuation.
(b) Include a reducing agent such as DTT and a chelator such as EDTA
to remove traces of metals that could catalyse oxidation (55) (see
also Chapter 2). Use 0.2-1 mM EDTA and 5-20 mM DTT.
2. Expect selenomethionyl proteins to be slightly less soluble than their
natural counterparts.
(a) Anticipate lower optimal ammonium sulfate concentrations in
trituration protocols.
(b) Anticipate increased retention in some chromatography procedures.
3. Store purified protein in an oxygen-free environment, at -80°C in the
presence of glycerol, or if possible, as frozen droplets at -180°C.
4. Mass spectroscopy is the most accurate method to quantitate seleno-
methionine incorporation. If this technique is not available, one can
undertake an amino acid analysis to check the percentage of subs-
titution. Selenomethionine is destroyed under the acid hydrolysis
conditions used in amino acid analysis, so that it is the disappearance
of methionine that is monitored.
4.5 Crystallization
Experience to date suggests that selenomethionyl proteins crystallize in con-
ditions that are very similar to those used with native proteins (42, 45, 48, 49).
As a consequence of the lowered solubility of selenomethionyl proteins,
either the protein or the precipitant concentration should be slightly reduced
to achieve comparable degrees of supersaturation. It is often the case that
growth of selenomethionyl protein crystals require microseeding with a
crushed wild-type protein crystal. Selenomethionine oxidation can lead to
aberrant X-ray fluorescence spectra in which the position and shape of the K-
edge are altered and the white line intensity decreased (37). This can be
avoided by maintaining the crystals in a solution containing DTT and EDTA.
Crystals should be stored in an oxygen-free environment such as an anaerobic
chamber if possible (see Chapter 5 for a crystallization method suitable for
oxygen-sensitive proteins). They should be irradiated as soon as possible or
69
P. F. Berne et al.
flash-frozen and stored in liquid nitrogen while they await data collection.
Selenomethionine incorporation does not appear to alter diffraction limits
and selenomethionyl protein crystals are generally isomorphous with native
crystals. However, they can be more sensitive to radiation damage.
There are now several large protein structures (> 90 kDa) that have been
solved by the MAD method using solely the anomalous signal of selenium
(48, 56; J. L. Smith, personal communication). It is also clear that large
numbers of selenium sites (15 or more) can be readily located with direct
methods programs. This realization should increase even further the wide-
spread use of selenomethionyl proteins for phase determination. There are
also encouraging results regarding the incorporation of telluromethionine
into proteins (57, 58). Even though tellurium cannot be used as an anomalous
scatterer (its K-edge (0.389 A) and L-edges (> 2.5 A) correspond to wave-
lengths not usually reachable at synchrotron facilities), its 36 e~ difference
with sulfur has been successfully used in conventional isomorphous replace-
ment methods (58). This procedure will reach its true potential when telluro-
methionine becomes commercially available.
4.6 Warning
Selenium is an essential element for most animal and bacterial life (59), but
it is also a very toxic compound because of its ability to replace sulfur. In
mammals (e.g. protein crystallographers), ingested methionine is a source of
sulfur. As a result, selenomethionine can be harmful or even fatal if inhaled,
swallowed, or absorbed through the skin. Experiments should always be done
in a hood and the experimenter should be sure to wear gloves. Experimenters
should contact the Health and Safety office at their institution and inquire
about proper disposal of selenomethionine containing media.
5. Conclusion
Structural studies require homogeneous concentrated solutions of the protein
of interest. Some proteins cannot be obtained in this form and optimizing the
expression system can sometimes solve the problem. However, even an
apparently perfect protein solution might be reluctant to crystallize, perhaps
because of surface residues that could form destabilizing intermolecular con-
tacts. In some cases, such a protein can be stabilized by the addition of specific
agents like glycerol, detergents, zwitterions, or amino acids (61, 62) (see also
Chapter 2). An alternative approach consists in identifying residues respons-
ible for the instability of the protein and mutating them. Chances are that
these residues, being located at the surface, will not be crucial for keeping the
overall 3D folding of the protein. This chapter gives an overview of the genetic
engineering tools that one can exploit to optimize expression systems and to
design variants by systematic or random mutagenesis. These modifications
70
3: Molecular biology for structural biology
can lead to improved crystallization or to novel physical properties useful for
structure determinations by X-ray diffraction studies. Such macromolecular
engineering strategies may find routine use in modern structural biology.
References
1. Pawson, T. (1995). Nature, 373, 573.
2. Gruenwald, M. D. and Heitz, M. S. (1993). In Baculovirus expression vector
system: procedures and methods. A manual published by Pharmingen.
3. Miegel, A., Sano, K. I., Yamamoto, K., Maeda, K., Maeda, Y., Taniguchi, H., et al
(1996). FEBS Lett., 394, 201.
4. Marchot, P., Ravelli, R. B. G., Raves, M. L., Bourne, Y., Vellom, D. C., Kanter, J.,
et al. (1996). Protein Sci., 5, 672.
5. Cregg, J. M., Vedvick, T. S., and Raschke, W. C. (1993). Biotechnology, 11, 905.
6. Kern, D., Dietrich, A., Fasiolo, F., Renaud, M., Giege, R., and Ebel, J.-P. (1977).
Biochimie, 59, 453.
7. Weijland, A., Williams, J. C., Neubauer, G., Courtneidge, S. A., Wierenga, R. K.,
and Superti-Furga, G. (1997). Proc. Natl. Acad. Sci. USA, 94, 3590.
8. Friehs, K. and Reardon, K. F. (1993). Adv. Biochem. Eng. Biotech., 48, 53.
9. Olins, P. O. and Lee S. C. (1993). Curr. Opin. Biotech., 4, 520.
10. Dale, G. E., Broger, C., Langen, H., D'Arcy, A., and Stuber, D. (1994). Protein
Eng., 7, 933.
11. Kim, K. K., Yokota, H., Santoso, S., Lerner, D., Kim, R., and Kim, S. H. (1998).
J. Struct. Biol., 121, 76.
12. Kozak, M. (1997). EMBO J., 16, 2482.
13. Schein, C. H. (1989). Biotechnology, 7, 1141.
14. Blackwell, J. R. and Horgan, R. (1991). FEBS Lett, 295, 10.
15. Caspers, P., Stieger, M., and Burn, P. (1994). Cell. Mol. Biol, 40, 635.
16. Yasukawa, T., Kanei-Ishii, C., Maekawa, T., Fujimoto, J., Yamamoto, T., and
Ishii, S. (1995). J. Biol. Chem., 270, 25328.
17. Mayer, M. P., Prestwich, G. D., Dolence, J. M., Bond, P. D., Wu, H., and Poulter,
CD. (1993). Gene, 132, 41.
18. Maniatis, T., Fritsch, E. F., and Sambrook, J. (ed.) (1989). Molecular cloning: a
laboratory manual, 2nd edn. Cold Spring Harbor Press, Cold Spring Harbor, NY.
19. Geng, X. and Chang, X. (1992). J. Chromatogr., 599, 185.
20. Ford, C. F., Suominen, I., and Glatz, C. E. (1991). Protein Expression Purification,
2, 95.
21. Stoll, V. S., Manohar, A. V., Gillon, W., Macfarlane, E. L. A., Hynes, R. C., and
Pai, E. F. (1998). Protein Sci., 7, 1147.
22. Zhang, F., Robbins, D. J., Cobb, M. H., and Goldsmith, E. J. (1993). J. Mol. Biol.,
233, 550.
23. Fields, B. A., Ysern, X., Poljak, R., Shao, X., Ward, E. S., and Mariuzza, R. A.
(1994). J. Mol. Biol., 239, 339.
24. Campbell, I. and Duffus, J. H. (ed.) (1989). Yeast: a practical approach. IRL Press,
Oxford.
25. Jenkins, T. M., Hickman, A. B., Dyda, F., Ghirlando, R., Davies, D. R., and
Craigie, R. (1995). Proc. Natl. Acad. Sci. USA, 92, 6057.
71
P. F. Berne et al.
26. Woodcock, S., Moraon, J.-P., and Henrissat, B. (1992). Protein Eng., 5, 629.
27. Wei, L., Hubbard, S. R., Hendrickson, W. A., and Ellis, L. (1995). J. Biol. Chem.,
270, 8122.
28. Dickinson, C. D., Gay, D. A., Parello, J., Ruoslahti, E., and Ely, K. R. (1994).
J. Mol. Biol., 238, 123.
29. Martinez-Hackert, E., Harlocker, S., Inouye, M., Berman, H. M., and Stock, A. M.
(1996). Protein Sci., 5, 1429.
30. Windsor, W. T., Walter, L. J., Syto, R., Fossetta, J., Cook, W. J., Nagabhushan,
T. L., et al. (1996). Proteins: Structure, Function, Genetics, 26, 108.
31. Zhang, G., Liu, Y., Qin, J., Vo, B., Tang, W. J., Ruoho, A. E., et al. (1997). Protein
Sci., 6, 903.
32. Bergfors, T., Rouvinen, J., Lehtovaara, P., Caldentey, X., Tomme, P., Claeyssens,
M., et al. (1989). J. Mol. Biol, 209, 167.
33. Cohen, S. L., Ferre-D'Amare, A. R., Burley, S. K., and Chait, B. (1995). Protein
Sci., 4, 1088.
34. Pfuetzner, R. A., Bochkarev, A., Frappier, L., and Edwards, A. M. (1997). J. Biol.
Chem., 272, 430.
35. Ho, S. N., Hunt, H. D., Horton, R. M., Pullen, J. K., and Pease, L. R. (1989). Gene,
77, 51.
36. Cadwell, C. and Joyce, G. F. (1992). PCR Methods Applications, 2, 28.
37. Smith, J. L. and Thompson, A. (1998). Structure, 6, 815.
38. Hendrickson, W. A. (1988). In Crystallographic computing, Vol. 4 (ed. N. W.
Isaacs and M. R. Taylor), pp. 97-108. Oxford University Press, Oxford, UK.
39. Hendrickson, W. A. (1985). Trans. Am. Crystallogr. Assoc., 21, 11.
40. Hendrickson, W. A., Pahler, A., Smith, J. L., Satow, Y., Merrit, E. A., and
Phizackerley, R. P. (1989). Proc. Natl. Acad. Sci. USA, 86, 2190.
41. Yang, W., Hendrickson, W. A., Crouch, R. J., and Satow, Y. (1990). Science, 249,
1398.
42. Hendrickson, W. A., Horton, J. R., and LeMaster, D. M. (1990). EMBO J., 9, 1665.
43. Cowie, D. B. and Cohen G. N. (1957). Biochim. Biophys. Acta, 26, 252.
44. Doublie, S. (1997). In Methods in enzymology (eds C. W. Carter and R. Sweet),
Academic Press, London. Vol. 276, p. 523.
45. Yang, W., Hendrickson, W. A., Kalman, E. T., and Crouch, R. J. (1990). J. Biol.
Chem., 265, 13553.
46. Van Duyne, G. D., Standaert, R. F., Karplus, P. A., Schreiber, S. L., and Clardy, J.
(1993). J. Mol. Biol., 229, 105.
47. Benson, T. E., Filman, D. J., Walsh, C. T., and Hogle, J. M. (1995). Nature Struct.
Biol., 2, 644.
48. Doublie, S., Tabor, S., Long, A. M., Richardson, C. C., and Ellenberger, T. (1998).
Nature, 391, 251.
49. Lau, A. Y., Scharer, O. D., Samson, L., Verdine, G. L., and Ellenberger, T. (1998).
Cell, 95, 249.
50. Chen, W. Y. and Bahl, O. P. (1991). J. Biol. Chem., 266, 8192.
51. Chen, W. Y. and Bahl, O. P. (1991). J. Biol. Chem., 266, 9355.
52. Wu, H., Lustbader, J. W., Liu, R. E., Canfield, W., and Hendrickson, W. A. (1994).
Structure, 2, 545.
53. Lustbader, J. W., Wu, H., Birken, S., Pollak, S., Gawinowicz Kolks, M. A., Pound,
A. M., et al. (1995). Endocrinology, 136, 640.
72
3: Molecular biology for structural biology
54. May, A. P., Robinson, R. C., Aplin, R. T., Bradfield, P., Crocker, P. R., and Jones,
E. Y. (1997). Protein Sci., 6, 717.
55. Scopes, R. K. (1982). Protein purification. Springer-Verlag, New York, NY.
56. Turner, M. A., Yuan, C. S., Borchardt, R. T., Hersfield, M. S., Smith, G. D., and
Howell, P. L. (1998). Nature Struct. Biol., 5, 369.
57. Boles, J. O., Lewinski, K., Kunke, M., Odom, J. D., Dunlap, B. R., Lebioda, L., et
al. (1994). Nature Struct Biol., 1, 283.
58. Budisa, N., Karnbrock, W., Steinbacher, S., Humm, A., Prade, L., Neuefeind, T.,
et al. (1997). J. Mol Biol., 270, 616.
59. Stadman, T. C. (1980). Annu. Rev. Biochem., 49, 93.
60. LeMaster, D. M. and Richards, F. M. (1985). Biochemistry, 24, 7263.
61. Sousa, R., Lafer, E. M., and Wang, B. C. (1991). J. Cryst. Growth, 110, 237.
62. Jeruzalmi, D. and Steitz, T. A. (1997). J. Mol. Biol., 274, 748.
63. Fromant, M., Blanquet, S., and Plateau, P. (1997). In Genetic engineering with
PCR (ed. R. M. Horton and R. C. Tait), p. 39. Horizon Scientific Press,
Wymondham, UK.
73
This page intentionally left blank
4
1. Introduction
This chapter is about practical uses of mathematical models to simplify the
task of finding the best conditions under which to crystallize a macromolecule.
The models describe a system's response to changes in the independent
variables under experimental control. Such a mathematical description is a
surface, whose two-dimensional projections can be plotted, so it is usually
called a 'response surface'.
Various methods have been described for navigating an unknown surface.
They share important characteristics: experiments performed at different
levels of the independent variables are scored quantitatively, and fitted im-
plicitly or explicitly, to some model for system behaviour. Initially, one
examines behaviour on a coarse grid, seeking approximate indications for
multiple crystal forms and identifying important experimental variables.
Later, individual locations on the surface are mapped in greater detail to
optimize conditions. Finding 'winning combinations' for crystal growth can be
approached successively with increasingly well-defined protocols and with
greater confidence. Whether it is used explicitly or more intuitively, the idea
of a response surface underlies the experimental investigation of all multi-
variate processes, like crystal growth, where one hopes to find a 'best' set of
conditions. The optimization process is illustrated schematically in Figure 1.
In general, there are three stages to this quantitative approach:
(a) Design. One must first induce variation in some desired experimental result
by changing the experimental conditions. Experiments are performed
according to a plan or design. Decisions must be made concerning the
experimental variables and how to sample them.
(b) Experiments and scores. Each experiment provides an estimate for how
the system behaves at the corresponding point in the experimental space.
C. W. Carter Jr
Figure 2. Mathematical models provide a two-way link between crystal properties and
experimental effects. (a) The mathematical relationships linking response-surface para-
meters and experimental scores via the design matrix, F[k. (scorek) is a row vector of the
N experimental scores, (B i ) T is a column vector of the model coefficients. (b) Schematic
presentation of the reciprocity between parameters and scores described in the text.
Figure 3. Example of an incomplete factorial design matrix, showing how scores and
contrasts (model parameters) are connected to each other via the experimental matrix
(shaded). Here, six experiments are used to define a model with four parameters, which
include the constant term and the three main effect contrasts,
79
C. W. Carter Jr
model states that the expected score for the kth experiment, Scorecalc, k, is given
by the average observed score, (30 = 19.33, plus or minus a constant value, pi,
from each of the three columns:
where (30 is the average score. Thus, for the first experiment, Scorecalc,1 = 19.33
+ [(-1 X 4.0) + (1 X -1.75) + (-1 X 7.75)] = 5.8. Similarly, the three
coefficients of the model, {ft}, are obtained from the average sum of the
products of each score and the matrix element from the appropriate column:
where the sums over experiments at the two levels are kept separately, in case
there are not the same numbers of experiments at each level. Equation 2 is
also called the contrast between the high and low levels for the ith factor.
Other contrasts can be calculated, and may be important. Most important
are those for the two-way interactions between pairs of factors. Consider the
data for the full-factorial design in Table 1, and in which the incomplete
factorial sample from Figure 3 is indicated by bold face. The effects of three
variables are each tested at two levels. Contrasts for all possible n-factor
interactions are shown, in addition to those for the main effects. There are
three two-factor interactions and one three-factor interaction. Entries for
each interaction are the products of the entries for the respective main effects.
Contrasts for higher-order interactions are generally smaller than those for
main effects.
In summary, the following aspects of the entries in Table 1 should be noted.
(a) All possible combinations for the three variables are tested at two levels
by the 8 (~ 23) experiments in the full design. This means that the
Table 1. Extended matrix and contrasts for a three-factor, two-level full-factorial design
80
4: Experimental design and quantitative analysis
experiments are necessarily uniformly distributed among the different
possible combinations.
(b) Each column is a different linear combination of the eight scores. This is
true both for the full-factorial design and the 6-experiment sample.
Designs for which any two Fij columns are the same are said to involve
confounding or aliasing of the effects denoted by identical columns. It is
impossible to distinguish which of the confounded columns is responsible
for the contrast in the experimental scores without additional experiments
specifically designed to distinguish between the multiple possibilities (19).
(c) The seven columns all have equal numbers of 1s and -1s. This is true for
the complete factorial and for all but the temp x [Prt] column of the 6-
experiment sample. Both the full design and the sampled design are
therefore balanced with respect to the main effects.
2.3.3 Balance
If each level is tested by the same number of experiments, the design is said to
be balanced. Balance is important. The standard deviation of an average
value, <x>, is given by
Their use has been described previously (15-18). Factor levels are chosen
randomly and then balanced to achieve nearly uniform sampling (Figure 4).
Levels for each main effect are sampled the same number of times, and all
two-factor interactions are sampled as uniformly as possible. This strategy
preserves the ability to detect large main effects and two-factor interactions
with minimal confounding (18). This process leads to a very flexible sampled
factorial design which has given superior performance in a wide variety of
contexts (15,21-24).
The number of experiments in a design should be chosen relative to the
size, N, of the full-factorial design, which, in turn, depends on the number of
variables, and their levels according to Equation 1. Incomplete factorial de-
signs with a sampling density as coarse as roughly VM2 can be found that are
quite evenly balanced with respect to main effects and two-factor interactions,
and entirely free of explicit confounding. This rule of thumb is more useful for
83
C. W. Carter Jr
larger factorial designs; designs with smaller numbers of factors are less
efficient, as they require more experiments for adequate signal-to-noise.
Table 2. Hardin-Sloane minimum integrated variance design matrix for four factors, 20
experiments*
aThis design was prepared specifically for use in the experiments reported here by N. J. A. Sloane,
using GOSSET (39). Matrix entries should be interpreted as: 0 = the centre, -1 = the low end, and 1 =
the high end of the variable range. The same design has been used repeatedly in different contexts, by
assigning the matrix entries to different parameters and/or ranges.
been described (40). INFAC designs have optimal coverage and minimal
aliasing of main effects with each other and with two-way interactions. The
program is entirely interactive, prompting for all necessary information.
The GOSSET interface is less intuitive and requires explicit description of
variable ranges and the type of response-surface function to be fitted. Much
can be obtained from the examples in the users manual, but the overview in
Protocol 1 of how to generate a design like that in Table 2 should be helpful.
Method
86
4: Experimental design and quantitative analysis
those that will take on discrete values, e.g. 20 discrete T 4 14 21 can be
used to specify three discrete values for temperature.
3. Enter the model for which the design will be used, e.g. 30 model
(1+x+y+z+T) A 2 will specify a quadratic polynomial model in four
variables, x, y, z, and T. Such a model has (4+1)*(4+2)/2 = 15
parameters, so the design must have at least that number of
experiments.
4. Compile the above 'program' using: compile (from this point, no line
numbers are entered!).
5. Compute the matrix of 'experimental moments' using: moments
n=1000000. These moments are used internally to represent the
impacts of experimental points on the predicted variance (39).
6. Ask for the design. The command: 'design' will generate a design with
the minimum required number of experiments for the specified
model. Various modifications include:
(a) design runs=24 n=20. This forces the program to generate 24
experiments and to find the best design from 20 different starting
points.
(b) design type=I extra=5. This forces the program to generate an I-
optimal design (the default choice, which optimizes the Integrated
prediction variance) with 5 more than the minimum number of
experiments. The resulting design will have 20 experiments in this
case.
7. Generate a formatted file with the design: interp >20expt_xyzT.design.
This converts the output file into a formatted table in the file
20expt_xyzT.design.
Table 3. Hardin-Sloane minimum integrated variance design matrix for five factors, 30
experiments a
aThis five-variable design was prepared using the GOSSET program (39). It was designed to
compensate for the failure of one of the experiments to produce a score, and is called J-optimal. It
contains 30 experiments, which are nine more than the 21 experiments required to estimate
parameters for a five-variable quadratic model. Although variable names have been suggested, this
matrix can be used in any desired context.
88
4: Experimental design and quantitative analysis
Figure 5. Crystal growth variable space. Variables are organized hierarchically, from
those directly under explicit experimental control on the left, to those that dictate crystal
growth behaviour on the right. To challenge one experimental design in screening and/or
optimization of crystal growth is to find an efficient parameterization of variables in this
space that will permit post-hoc quantitative analysis.
Method
1. Gather information about the protein to be crystallized from all in-
volved with expression and purification; from previous screening
experiments; and from databases (Biological Macromolecule Crystal-
lization Database): (http://ibm4.carb.nist.gov:4400/bmcd/
bmcd.html) (30, 58).
2. List factors that might influence crystal growth.
(a) Estimate the variation of net charge with pH from the amino acid
composition (Chapter 10). Verify the pi by isoelectric focusing. Use
a pH range with values above and below the pi.
(b) Factors required for stability and monodispersity (59, 60).
(c) Ligands and other factors likely to influence the conformation of
the macromolecule.
4. Choose from this list the factors to be screened in the current
experiment and the levels to test.
5. Choose the number of tests. In general, this number should be
somewhat more than the number of factors to be screened. There is
no hard and fast rule; experience from diverse sources (20, 61)
suggests using ~ VN/2 tests if the full-factorial design requires N
tests.
6. Compile the experimental matrix itself (the computer program INFAC
http://russell.med.unc.edu/~carter/designs will do steps
a-e interactively).
(a) Choose factor levels at random, working down each column, and
from column to column.
(b) Balance each column, readjusting levels to equilibrate the
numbers of tests at each level.
(c) Balance each two-factor interaction by compensating readjust-
ments to two columns. In this case, it is adequate to ensure that
each combination is represented by at least one test.
(d) Verify the balance of all columns and two-factor interactions.
(e) Examine the experimental treatments for possible confounding.
Confounding is indicated whenever two different effects have
90
4: Experimental design and quantitative analysis
identical patterns of level assignments. If a confounded effect
turns out to be large, this knowledge is useful in further
experimentation to distinguish between the two possibilities.
(Designs generated by INFAC are selected to minimize
confounding.)
Exp't A B C D E F G H I J
1 2 2 4 1 1 3 2 3 1 1
2 4 5 5 1 2 3 1 1 1 2
3 5 1 4 2 1 1 3 1 2 2
4 1 2 3 1 2 2 3 3 2 1
5 3 5 2 3 2 1 1 2 1 2
6 4 3 1 1 1 3 1 3 2 1
7 5 4 1 3 3 1 1 2 2 2
8 1 4 3 3 1 2 2 2 2 2
9 2 1 2 1 3 2 2 3 1 1
92
4: Experimental design and quantitative analysis
Table 4. continued
Exp't A B C D E F G H I J
10 3 3 5 3 3 3 3 3 1 2
11 1 1 5 2 3 2 1 1 2 1
12 2 1 1 2 1 1 3 2 1 1
13 3 1 2 2 2 3 2 2 2 1
14 5 1 4 3 2 2 3 1 1 1
15 4 5 3 2 3 1 2 1 1 2
16 5 2 2 3 1 1 3 1 2 2
17 4 4 2 2 2 2 3 2 1 2
18 2 5 1 3 3 1 2 3 2 1
19 3 5 4 2 1 2 2 2 1 1
20 1 4 3 2 3 2 1 2 2 2
21 5 5 5 1 1 3 1 3 2 2
22 2 4 5 3 2 1 1 1 1 2
23 1 3 4 3 1 3 2 1 2 1
24 3 4 1 2 2 3 3 3 1 1
25 4 2 1 1 3 3 3 3 1 2
26 5 3 3 1 2 3 2 3 2 2
27 4 1 3 3 3 3 1 2 2 1
28 3 2 5 2 1 1 3 2 1 1
29 2 3 5 2 3 1 1 1 1 1
30 1 2 3 1 1 2 2 1 2 2
31 5 4 4 1 3 2 1 3 2 1
32 3 3 1 3 2 2 3 1 2 2
33 2 2 4 3 1 3 1 1 1 2
34 1 5 2 1 2 1 2 2 1 2
35 4 3 2 1 1 2 3 1 1 1
36 5 2 5 1 2 1 1 2 2 1
37 3 4 2 2 1 1 3 3 1 1
38 4 5 2 3 3 1 2 2 2 1
39 2 1 5 3 2 3 2 3 2 2
40 1 3 3 3 3 3 3 2 1 2
41 4 2 2 1 3 2 2 1 2 2
42 2 5 5 3 2 2 1 3 1 1
43 5 1 1 2 2 3 3 1 2 1
44 3 2 1 1 2 2 2 3 1 2
45 1 4 5 1 3 3 2 1 1 2
46 1 2 4 2 3 1 3 2 1 1
47 5 4 1 1 3 1 1 2 1
48 4 3 2 2 2 1 2 2 2 2
49 3 3 4 3 1 1 1 2 1 2
50 2 4 3 2 3 2 3 3 2 2
51 3 1 2 3 2 1 3 2 2 1
52 5 5 4 2 1 2 1 3 1 1
53 2 3 1 3 3 1 1 1 1 2
54 1 5 1 2 3 3 2 3 2 1
55 4 1 4 1 3 2 1 2 2 2
56 3 5 3 3 1 2 3 3 2 1
57 4 4 1 1 2 2 2 2 1 1
58 2 2 3 1 1 1 2 1 1 1
59 5 3 5 2 1 3 3 1 2 2
60 1 1 3 2 2 3 1 3 1 2
93
C. W. Carter Jr
motivated by that goal. It provides a balanced incomplete factorial screen
covering a similar sample of the conditions from Figure 5 to that represented
in the Hampton kit(s) (25), and is informed by the underlying physical
chemistry. It can be used either with a standard set of pHs or centred on a
known or estimated pi value.
Method
Identify columns in Table 4 with factors such as the following:
A. pH: five levels, 4.5, 5.5, 6.5, 7.5,8.5, cover the range normally observed
for crystallization of proteins. Alternately, if the pi is known five levels
can be centred on the pI.
B. Use five ionic crystallizing agents, with the following anions: sulfate,
phosphate, acetate, chloride, nitrate from the Hofmeister series.
C. Organic crystallizing agent: five levels, isopropanol, methylpentane-
diol, PEG 4000, PEG 8000, PEG 20000.
D. Protein concentration: three levels, Max, Max/2, Max/3. Three levels are
chosen because this variable may not be linear, and may have a maxi-
mum. Some effort is placed on estimating the curvature of its behaviour.
E. Temperature: three levels, for instance 4°C, 14°C, 22°C.
F. 'Moment' of non-polar reagent, M: 0%, 35%, 70%, as defined in
Section 4.4.2.
G. Divalent cation: three levels, none, Mg2+, and either Ca2+, Cd2+, or
Mn2+. Salts with acetate, nitrate, sulfate, and chloride are sufficiently
soluble. All are problematical with phosphate. This is a constant
weakness of all such plans. Phosphate experiment(s) calling for level
three of this variable can use Mn2+.
H. Additive: three levels, none, arginine, (JOG. These suggested choices
are based on the selection of additives that either stabilize proteins
(arginine ~ 50-100 mM) (59, 60) or destabilize weak, non-specific
contacts (00G 0.1-0.3%, w/v).
I. Glycerol: two levels, 0%, 10% (v/v). This variable samples additional
osmotic pressure exerted by the solution over and above that
produced by the crystallizing agent(s) themselves.
J. Ligand: two levels, presence or absence of a substrate or inhibitor.
5. Quantitative scoring
Previous quantitation and statistical analysis of crystal growth experiments (1,
2, 17, 73) are sufficiently compelling to justify a more intensive investigation
of practical requirements and procedures for using quantitative analysis of
crystal growth more routinely. We presume throughout this chapter that
experiments can be quantitatively scored in a variety of different ways. This
section addresses problems associated with scoring crystallization experi-
ments, and suggests protocols for solving those problems.
Method
1. Determine and score how many experiments in the design have some
kind of phase separation.
(a) If a majority are still clear, adjust the experiments to increase
supersaturation. Regression of this first score against the experi-
mental variables can reveal which changes to make in experi-
mental conditions to increase the proportion of experiments
producing a solid phase.
(b) If a majority have a solid phase, proceed to step 2.
2. Determine and score, according to Table 5, how many experiments in
the design have crystals.
(a) If most experiments have crystals, how many different morpho-
logies are represented? If there is only a single crystal habit,
proceed to step 4. Otherwise, proceed to step 5.
(b) If most experiments have precipitates, oils, and spherullites,
streak seeding (Chapter 7) can be helpful.
3. Examine a 'scatterplot matrix' showing how each score depends on
each independent variable for obvious trends and/or non-linearities.
4. Measure crystal sizes and shapes.
(a) Two dimensions, and occasionally a third can usually be measured
using a microscopic ruler. Enter dimensions for representative
samples from each experiment.
(b) Calculate sizes (surface area or volume) and shapes (aspect ratio,
or width/length).
98
4: Experimental design and quantitative analysis
(c) Determine the maximum, minimum, mean value, for each score.
Note significant variations.
5. Count the numbers of crystals with the same habit in each experiment.
Enter separate scores for each distinct morphology. These scores can
help resolve polymorphs (Section 8).
Result Score, Q
Cloudy/amorphous precipitates 1.0
Gelatinous/particulate precipitates 2.0
Oils 3.0
Spherulites 4.0
Needles 5.0
Plates 6.0
Prisms 7.0
Method
1. Prepare a file with the experimental matrix and all scores. Most
statistics programs provide separate modules for data entry, statistical
model building and analysis, and presentation graphics. For factors
whose levels are attributes (e.g. different ions), create a separate
column for each attribute, giving values of 0 or 1 for its presence
or absence. It may also be useful to approximate monotonic behaviour
like that of the ions of the Hofmeister series in a separate column.
2. Determine the maximum, minimum, mean value, for each score (this
may be programmed in EXCEL; ,in SYSTAT, use the Stats/Stats
command). Identify scores with significant variation.
3. Examine the data graphically. A matrix of two-way scatterplots for
each score versus each experimental variable is a useful, visual
presentation of the data, showing obvious correlations.
4. Try to express such trends as linear models explaining the variation in
scores (Protocol 6).
101
C. W. Carter Jr
italicized phrase suggests, the simple contrast sum calculations presented in
Section 2.3.2, Equation 2, can be misleading, and should be verified by a full
analysis of variance whenever possible.
Objective Design/model
103
C. W. Carter Jr
quadratic functions. These are the simplest functions that assume maxima,
minima, and saddle points, within the range of independent variables. Once a
model has been fitted, its stationary points can be determined analytically by
partial differentiation with respect to all the independent variables and
equating the gradient to zero. Stationary point coordinates provide estimates
for the factor levels giving the best result. This is the basis of the response-
surface method (19).
Method
1. Examine the mean value, range, and standard deviation statistics for
each score. Identify scores with significant variations.
2. Examine scatterplot matrices of each score for each independent
variable.
(a) Look for and note any obvious trends.
(b) Identify any categorical variables that might usefully be re-
ordered. These include cases where scores are clustered within a
category, but where their mean values show no pattern. Re-order
numerical assignments for categories, if doing so would create a
physically sensible monotonic or parabolic series.
(c) Rationalize patterns created in step 2(b) in terms of physically
reasonable effects.
3. Decide whether or not the variation in mean scores (step 1) and
suggested patterns (step 2) are sufficient to support meaningful
regression against the experimental variables.
4. Build and test trial models as described in Protocol 7, testing all main
effects first, then including two-way interactions and quadratic terms.
104
4: Experimental design and quantitative analysis
The analysis involves two interdependent tasks: model selection and
parameter estimation. The terms of a model represent the calculable
effects of the experimental factors, pjFj, their interactions, PyFiFj, and their
squares, 3HF2i, plus the intrinsic variation or noise, e, associated with the
experimental set-up and scoring. It is essential to discard terms that do not
contribute significant information about the response, and use the extra
degrees of freedom to improve the estimate for the residual error, thereby
reducing the variances of parameter estimates. Incorrect fitting of question-
able parameters can lead to model bias. Similarly, finding the best 'inter-
pretation' for a response-surface experiment can be haphazard because co-
efficients and their statistical significance change when the model itself
changes. Choosing which coefficients should be retained in the response-
surface model is therefore a challenging task (19, 82). Once the best set of
predictors has been identified, their coefficients are estimated by multiple
regression least squares.
These tasks are the job of a full statistics program. We have used two such
programs. SYSTAT (80) (now SPSS/SYSTAT) and JMP (80). Both provide a
powerful multiple regression module, with the appropriate statistical calcula-
tions, and graphing tools. The following illustrations present output from the
SYSTAT MGLH (Multiple regression, General Linear Hypothesis) module.
This module cannot evaluate partial derivatives or solve for stationary points,
but it is easy to use, intuitive, and fast, and the former tools are available in
Mathematica (81). The JMP user interface is well-developed, and JMP may be
easier to use.
Method
1. Define (on a command line or in a dialog box) a linear model for a
single dependent variable (the score) as a function of a set of
independent variables.
2. Identify the best subset of terms by using 'stepwise multiple re-
gression'. As the name suggests, this algorithm is an automated
procedure for choosing the best subset of coefficients. Generally, start
with a complete model and gradually eliminate insignificant contrib-
utors. It is sometimes useful to work forwards, finding the most
significant terms first. Stepping can be carried out with a variety of
different tolerance and threshold criteria for including or eliminating
terms, as appropriate, until the model appears stable and sensible.
Depending on the algorithm, terms may be recycled if they appear to
regain or lose significance as the stepping proceeds.
105
C. W. Carter Jr
Protocol 7. Continued
3. Think about what the model is saying; add/delete coefficients that
might/might not make sense.
4. Compare different models according to three types of criteria:
(a) The squared multiple correlation coefficient, R2. This indicates the
percentage of the variation that can be 'explained' by the model. It
should be as high as possible.
(b) The probability of the F-ratio test under the null hypothesis that
the model has no predictive value. This value should be very
small.
(c) Individual Student t-test probabilities for each coefficient. These
should be as small as possible.
5. Verify coefficients indicating two-way interactions; calculate and
examine the average scores for all four of the combinations (—, +-,
-+, ++) in the two-way matrix.
6. Expect that useful models of optimum behaviour will have positive
linear and negative quadratic coefficients. Look for models with these
characteristics.
7. Plot two-dimensional projections of the model surface. Super-
imposing these plots onto the observed scores is a good way to get a
feel for what the model has to tell about the system.
8. Verify that plots actually reflect the data. Discrepancies usually mean
errors in entering data, scores, and/or model coefficients, but can point
to unexpected effects.
9. Generally, the residuals, ([Qobs,i - Qcalcj]), can be saved and analysed in
the same ways used for the scores (Protocols 5-8). Examine them for
clues about where the model may be deficient.
Table 7. Multiple regression and analysis of variance for the illustration in Figure 3
A second kind of statistic, the 'P values', give the probability of obtaining
equally good models by random processes, i.e. under the 'null hypothesis' that
the variation is uncorrelated with changes in the experimental variables. This
information is available for the Student t-tests of individual coefficients and
the overall F-ratio. The Student t value is the ratio of a coefficient to its
standard error, so it is a statistic about the signal-to-noise of an estimated
coefficient. The overall F-ratio is the squared distance between calculated and
average scores for all experiments divided by the squared distance between
calculated and observed scores. It can be considered an estimate of the signal-
107
C. W. Carter Jr
to-noise of the model. Useful models can have P-values as high as 10-3, our
best models are better than 10-11. Protocol 8 and the associated statistical
criteria apply equally to the investigation of any model.
Try models using main effects one or two at a time, to see how much of the
variation is explained (multiple R2), and at what cost in terms of the F-ratio
probability. How much better does the model get by adding another factor?
Since this procedure is the most difficult, it requires some intuitive feel, which
comes from practice. This approach is illustrated for data from Table 1 in
Table 7, which illustrates the trade-off that necessarily accompanies sampling.
All models have quite high R2 values, and satisfy the criterion of predictive
power. In other respects, however, the quality of the models depends on the
amount of available data. The complete ensemble of eight experiments
affords the best model by all criteria: it supports a model with more significant
parameters, while at the same time giving the best F-ratio and t-test prob-
abilities. The incomplete factorial subset performs nearly as well, giving
nearly the same parameter estimates. Their statistical significance, however, is
degraded by about two orders of magnitude. Deleting additional data, as with
the two different orthogonal arrays, degrades the models well beyond the point
where they are useful. None of the parameters is statistically significant.
Usually, the default stepwise regression will produce a reasonable repre-
sentation of what is in the data. That model can occasionally be improved
using different tricks. Check individual t-tests and try deleting the worst one
(with the highest probability). Omit the constant term only when its t-test is
poor (has a high probability). It is recommended to retain the main effect in
any model that involves a higher-order interaction, even if this makes the
model worse. There are exceptions to all of these guidelines. A trade-off must
always be made between the decreased variance of the model parameters,
achieved by reducing their number, and the potential loss of real information
about the response surface that occurs when a 'true' parameter with a large
variance is deleted from the model. A more detailed description, with
examples, is provided in ref 83). The ultimate test is a model's usefulness.
Figure 8. (a) Ridges observed in the temperature x concentration level surface for
response surfaces determined for two different TrpRS crystal polymorphs (1). The
ridgeline represents combinations of temperature and [protein] which produce the same
supersaturation level, suggesting that supersaturation is a natural search variable, (b)
Level surfaces for a third TrpRS polymorph determined using as one of the search
variables an approximation to supersaturation (87). Here, the three level surfaces
involving [protein] all show optima.
7. Optimization
The term 'optimization' is used frequently in discussions of crystal growth.
Usually, it refers to variation of some experimental variables with the aim of
empirically finding 'better' crystals from among the conditions tested. Given
sufficient time and materials, and a fortunate choice of experimental variables
to explore, any search strategy can lead to better crystals. Often, however,
both time and materials are limited. In such cases, there are two ways in which
the search for optimal conditions can be made more systematic and efficient.
One uses either a line search or a more elaborate variant called 'simplex'
optimization (86), the other uses response surfaces. The two approaches are
complementary; the former may actually be more appropriate if existing
conditions are far from an optimum.
Method
1. A least four variables should be sampled simultaneously to use
quadratic polynomial models.
2. Two variables should always represent the solubility diagram,
irrespective of the other variables.
(a) Protein concentration and supersaturation are more nearly ortho-
gonal than protein concentration and precipitant concentration,
facilitating sampling of the nucleation zone (Figure 9).
(b) A product, either [protein] x [crystallizing agent] (87) or In
[protein] + [crystallizing agent] are useful approximations to super-
saturation or ln[supersaturation] when the latter are unknown.
3. Choose additional variables based on any available prior information.
Regression analyses of an incomplete factorial screening experiment
often provide indications of the most significant main effects (24).
4. Centre the experiment close to the best known set of conditions.
(a) Exploitation of response-surface experiments is most successful
when quantitative and reliable scores have been obtained for all
or nearly all experiments.
(b) Designs not centred on conditions known to produce crystals are
effectively screening experiments, and should be carried out using
qualitatively different experimental matrices.
5. Choose sufficiently large ranges for each variable to induce significant
variation in the score without losing the response itself. Follow the
guidelines in Section 4.3.
7.2.2 Replication
The variances of replicate experiments done at a random combination of
parameters near a stationary point should vary inversely with the distance of
that point from the stationary point (Figure 7). Our experience suggests that
this is indeed the case. An important consequence is that the estimates of the
variance obtained from an ensemble of unrcplicated experiments may fail to
capture the information about stationary points that can be obtained from
replicated sets of experiments, which would provide an approximate map of
the variances over the experimental space. For this reason, it is useful to carry
out each experiment in a design twice.
Figure 9. Orthogonalization of the nucleation zone for sampling purposes, (a) The
nucleation zone is shown as the curved figure bounded by the metastable and
precipitation zones, (b) The product [macromolecule] x [crystallizing agent] is constant
along the rectangular hyperbolae in (a). Thus, the nucleation zone becomes a rectangle
which can be sampled evenly on a regular grid (87).
114
4: Experimental design and quantitative analysis
7.2.3 An example
A response-surface experiment for monoclinic tryptophanyl-tRNA synthetase
(TrpRS) crystals is presented step-by-step to illustrate the process of building
and analysing the model. This crystal form had always previously given long
and thin crystals that had to be grown bigger by repeated reseeding (84, 85).
Intense efforts to improve the size of the initial crystals by systematic
variation of protein concentration had led only to modest improvements.
After some initial successes with other forms of TrpRS, we proceeded as
follows with the monoclinic form (1):
(a) We selected [protein], {[prot] X [ppnt]}, temperature, and the con-
centration of the additive, PEG 400, as the independent variables. A
range of values (+\- 9-23%) was chosen to surround the best conditions
we had previously achieved. Values of these four variables were assigned
to 20 experiments of the Hardin-Sloane matrix in Table 2. These experi-
ments were each done twice and scored using three different criteria:
volume, the ratio of the smallest dimension to the largest, and a subjective
assessment of their uniformity and freedom from satellite crystals. These
scores are more objective and quantitative than the subjective scale we
used previously to score screening experiments and should be easier to
use. They were input to a SYSTAT data file together with the H-S matrix.
(b) Using the MGLH (Multiple Regression, General Linear Hypothesis)
module in the statistics program SYSTAT for the Macintosh (82), tests
were carried out first using a model containing all 15 terms of equation
(HI) (a constant, four each of the main effects and the quadratic terms,
and the six two-factor interactions). The initial and subsequent models
were evaluated according to criteria described in Section 6.2.1. This
complete model had an F-ratio probability of10-4.R2 was 0.87, indicating
that all but about 13% of the variation in observed scores could be
attributed to the model. Nevertheless, in several respects the model
needed adjustment. In particular, three coefficients had t-test probabili-
ties > 0.05, indicating that they were without significance and should be
removed.
(c) The full model was pruned by backward stepwise regression, eliminating
three of the 15 parameters. The final model was obviously very significant;
its F-ratio probability, P, was 10-11, R2 was 0.95; and t-test probabilities of
the 11 coefficients of the model were nearly all below the 5% confidence
limit (1). Two were around 0.1 and of questionable significance. However,
removing them resulted in a serious deterioration, causing P to decrease
by an order of magnitude, and four additional factors with significant
t-tests in the best model became completely insignificant.
Prediction and verification. Partial derivatives of the calculated score with
respect to all variables were evaluated from the model expression. Equating
115
C. W. Carter Jr
the tour derivatives to zero and solving for the coordinates ([protein]opt.
[ p r o _ p t j l j p l , [PEG]l)pt, and Temp0|,,) of the optimum predicted that crystals
grown at these values would he better any of those observed in the H-S
experiment. Crystals grown at the optimum point were two orders of
magnitude larger and of sufficient volume for diffraction experiments. Similar
analysis for two other scores (volume and uniformity) showed that the three
optima were essentially in the same place, simultaneously optimizing all three
scores (1).
8. Resolution of polymorphs
A recurring problem with screening experiments is that they sample con-
ditions that may he far from optimal for a particular crystal form. Attempts to
optimize such cases using Hardin-Sloane designs can give rise to surfaces like
that in Figure JO, where instead of an optimum, the surface represents a
saddle. Sometimes, this problem is confounded by the appearance of multiple
crystal forms in the same experiments. Indeed, the appearance of saddle points
is often diagnostic of the superposition of overlapping response surfaces for
different poymorphs. Since different crystal polymorphs rarely, if ever, have
the same dependence on all experimental variables, finding optimal stationary
points for different polymorphs can be a useful way to 'purity' them away
from one another.
Figure II shows how a line search helped relocate the centre for a Hardin-
Figure 10. Response surface determined for E. colicytidine deaminase at the neighbour-
hood of a 'hit' from a kit (73) showing that the conditions are nearly at a saddle point and
hence far from an optimum. The dot indicates experiment number 28 of Jancarik and Kim
125).
UK
4: Experimental design and quantitative analysis
Figure 11. Use of steepest ascent (line search) to resolve polymorphs of E. colicytidine
deamlnase crystals, (a) The proportion of the desired form (Form I) fitted to the ridge
function shown above the figure. Tests selected to fall on the line as described in the text
showed that this proportion increased along that search, until at the X, only Form I was
observed, (b) A Hardin-Sloane design was then centred on the X, giving the optimum
shown (73).
Sloane design away from the saddle point in Figure 10, ultimately locating
optimal conditions for a new polymorph of E. coli cytidine deaminase (73).
Experiments at higher temperature in the original Hardin-Sloane design had
variable amounts of a second polymorph. The proportion of this second
polymorph was used as a score and fitted to the ridge surface in Figure 11a.
The initial conditions from the Hampton screen lay near the bottom of a
steeply sloped ridge in two variables, temperature and [Na Acetate]. The
gradient at that point indicated that the proportion of this form increased by
0.114 for every degree the temperature was raised, and by 2.07 as the con-
centration of acetate was reduced by 1 mole. Combining the two indications
gives that for each degree of temperature increase, the acetate concentration
should be reduced by 0.114/2.07 = 0.055 M. Experiments stepped out along
this gradient showed the expected increase in the fraction of the desired
crystal form. When the proportion was equal to 100%, a new Hardin-Sloane
design was performed, giving the surface in Figure 11b.
The first step in separating out the polymorphs was to identify the variables
critical to the different polymorphs. This was inferred from how the poly-
morphs behaved with respect to changes in all variables in the first Hardin-
Sloane design. A subset was identified by model fitting and analysis, and the
surface used to identify the gradient for a steepest ascent search to resolve the
different polymorphs from one another. Subsequent, re-centred response-
surface experiments can locate the stationary points associated with each
117
C. W. Carter Jr
References
1. Carter, C. W., Jr and Yin, Y. (1994). Acta Cryst., D50,572.
2. Carter, C. W., Jr., Doublie, S., and Coleman, D. E. (1994). J. Mol. Biol., 238,346.
3. Schutt, C. E., Lindberg, U., Myslik, J., and Strauss, N. (1989). 7. Mol. Biol., 209,
735.
4. Anderson, C. M., Zucker, F. H., and Steitz, T. A. (1979). Science, 204,375.
5. Sack, J. S., Saper, M. A., and Quiocho, F. (1989). J. Mol. Biol., 206,171.
6. Bullough, P. A., Hughson, F. M., Skehel, J. J., and Wiley, D. C. (1994). Nature,
371, 37.
7. Colombo, M. F., Rau, D. C., and Parsegian, A. V. (1992). Science, 256, 665.
8. Rand, R. P., Fuller, N. L., Butko, P., Francis, G., and Nicholls, P. (1993).
Biochemistry, 32, 5925.
9. Rand, R. P. (1992). Science, 256,618.
10. Reid, C. and Rand, R. P. (1997). Biophys. J., 72,1022.
11. Parsegian, V. A., Rand, R. P., and Rau, D. C. (1995). In Methods in enzymology
(eds M. L. Johnson and G. Ackers), Academic Press, Inc., London. Vol. 259,
p. 43.
12. Leikin, S. and Parsegian, V. A. (1994). Proteins: Structure, Function, Genetics, 19,73.
13. Royer, W. E., Pardnamani, A., Gibson, Q. H., Peterson, E. S., and Friedman, J. M.
(1996). Proc. Natl. Acad. Sci. USA, 93,14526.
14. Ladner, J. E., Finch, J. T., King, A., and Clark, B. F.C. (1972). J. Mol. Biol., 72,99.
15. Carter, C. W., Jr. and Carter, C. W. (1979). J. Biol. Chem., 254,12219.
16. Carter, C. W., Jr., Baldwin, E. T., and Frick, L. (1988). 7. Cryst. Growth, 90, 60.
17. Carter, C. W., Jr. (1990). Methods: a companion to methods in enzymology, 1,12.
18. Carter, C. W., Jr. (1992). In Crystallization of proteins and nucleic acids: a practical
approach (ed. R. Giege and A. Ducruix), p. 47. IRL Press, Oxford University
Press, Oxford, UK.
19. Box, G. E. P., Hunter, W. G., and Hunter, J. S. (1978). Statistics for experimenters.
Wiley Interscience, New York.
20. Carter, C. W. (1995). Qual. Eng., 8,181.
21. Bell, J. B., Jones, M. E., and Carter, C. W., Jr. (1991). Proteins: Structure, Function,
Genetics, 9,143.
22. Belts, L., Frick, L., Wolfenden, R., and Carter, C. W., Jr. (1989). J. Biol. Chem.,
264, 6737.
23. Doublie, S., Xiang, S., Bricogne, G., Gilmore, C. J., and Carter, C. W. J. (1994).
Acta Cryst A50,164.
24. Yin, Y. and Carter, C. W., Jr. (1996). Nucleic Acids Res., 24,1279.
25. Jancarik, J. and Kim, S.-H. (1991). J. Appl. Cryst., 24, 409.
26. Doudna, J. A., Grosshans, C., Gooding, A., and Kundrot, C. (1993). Proc. Natl.
Acad. Sci. USA, 90,7829.
27. Song, L. and Gouaux, E. (1997). In Methods in enzymology, ibid ref. 54. Vol. 276,
p. 60.
118
4: Experimental design and quantitative analysis
28. Kundrot, C. E. (1997). In Methods in enzymology, Vol. 276, p. 143.
29. Scott, W. G., et al. (1995). J. Mol. Biol, 250, 327.
30. Gilliland, G. L., Tung, M., Blakeslee, D. M., and Ladner, J. E. (1994). Acta Cryst.,
D50,408.
31. Rosenberg, J. M., et al. (1997). In Annual Meeting 56 (American Crystallographic
Association, St. Louis).
32. Shaanan, B. (1983). J. Mol. Biol., 175,159.
33. Smith, F. R., Lattman, E. E., and Carter, C. W., Jr. (1991). Proteins: Structure,
Function, Genetics, 10,81.
34. Janin, J. and Wodak, S. J. (1993). Proteins, 15,1.
35. Srinivasan, R. and Rose, G. D. (1994). Proc. Natl. Acad. Sci. USA, 91,11113.
36. Smith, F. R. and Simmons, K. C. (1994). Proteins: Structure, Function, Genetics,
18, 295.
37. Plackett, R. L. and Burman, J. P. (1946). Biometrika, 33,305.
38. Kingston, R. L., Baker, H. M., and Baker, E. N. (1994). Acta Cryst., D50,429.
39. Hardin, R. H. and Sloane, N. J. A. (1993). J. Stat. Plan. Inference, 37, 339.
40. Audic, S., Lopez, F., Claverie, J.-M., Poirot, O., and Abergel, C. (1997). Proteins:
Structure, Function, Genetics, 29,252.
41. Ferrone, F., Hofrichter, J., and Eaton, W. A. (1985). /. Mol. Biol., 183,611.
42. Ducruix, A. F. and Ries-Kautt, M. (1990). Methods: a companion to methods in
enzymology, 1,25.
43. Ries-Kautt, M. and Ducruix, A. (1992). In Crystallization of nucleic acids and
proteins: a practical approach (ed. A. Ducruix and R. Giege), p. 195. IRL Press,
Oxford, UK.
44. Carbonnaux, C., Ries-Kautt, M., and Ducruix, A. (1995). Protein Sci., 4,2123.
45. Ries-Kautt, M. and Ducruix, A. (1997). In Methods in enzymology ibid ref. 54.
Vol. 276, p. 23.
46. Ataka, M. and Tanaka, S. (1986). Biopolymers, 25,337.
47. Ataka, M. and Michihiko, A. (1988). J. Cryst. Growth, 90,86.
48. Rosenberger, F. (1996). J. Cryst. Growth, 166, 40.
49. Muschol, M. and Rosenberger, F. (1997). J. Chem. Phys., 107,1953.
50. Rosenberger, F., Muschol, M., Thomas, B. R., and Vekilov, P. G. (1996). J. Cryst.
Growth, 168,1.
51. Retailleau, P. (1997). Thesis Ph. D. Universite Paris XI Orsay.
52. Retailleau, P., Ries-Kautt, M., and Ducruix, A. (1997). Biophys. J., 73,2156.
53. Luft, J. R, et al. (1994). J. Appl. Cryst., 27,443.
54. Luft, J. R. and DeTitta, G. T. (1997). In Methods in enzymology (eds C. W. Carter
and R. Sweet), Academic Press, Inc., London. Vol. 276, p. 110.
55. Rosenberger, F., Vekilov, P. G., Lin, H., and Alexander, J. I.D. (1997).
Microgravity science and technology 10,29.
56. Vekilov, P. G., Thomas, B. R., and Rosenberger, F. (1998). J. Phys. Chem., 2, 5208.
57. Vekilov, P. G. and Rosenberger, F. (1998). J. Cryst. Growth, 186,251.
58. Gilliland, G. and Bickham, D. M. (1990). Methods: a companion to methods in
enzymology, 1,6.
59. Guilloteau, J. P., et al. (1996). Proteins: Structure, Function, Genetics, 25,112.
60. Timasheff, S. N. (1992). In Pharmaceutical biotechnology (ed. T. J. Ahera and
M. C. Manning), p. 265. Plenum, New York.
61. Bricogne, G. (1993). Acta Cryst., D49, 37.
119
C. W. Carter Jr
62. Rosenberger, F. (1986). J. Cryst. Growth, 76, 618.
63. Rosenberger, F. and Meehan, E. J. (1988). J. Cryst. Growth, 90,74.
64. Stura, E. and Wilson, I. (1990). Methods: a companion to methods in enzymology,
1,38.
65. Stura, E. A., Satterthwait, A. C., Calvo, J. C., Kaslow, D. C., and Wilson, I. A.
(1994). Acta Cryst., D50,448.
66. Stura, E. (1998). In Crystallization of nucleic acids and proteins: a practical
approach (ed. A. Ducruix and R. Giege). IRL Press, Oxford, UK.
67. Hofrichter, J., Ross, P. D., and Eaton, W. A. (1974). Proc. Natl. Acad. Sci. USA,
71,4864.
68. Hofrichter, J., Ross, P. D., and Eaton, W. A. (1976). Proc. Natl. Acad. Sci. USA,
73,3035.
69. Collins, K. D. and Washabaugh, M. W. (1985). Q. Rev. Biophys., 18,323.
70. Collins, K. D. (1997). Biophys. J., 72,65.
71. Sousa, R. (1996). In Methods in enzymology Vol. 276, p. 131.
72. Scarborough, G. (1994). Acta Cryst., D50,643.
73. Kuyper, L. and Carter, C. W., Jr. (1996). J. Cryst. Growth, 168,155.
74. Edwards, A. L. (1979). Multiple regression and the analysis of variance and
covariance. W. H. Freeman and Company, San Francisco.
75. Milliken, G. A. and Johnson, D. E. (1989). Analysis of messy data: Volume 2,
nonreplicated experiments, pp. 1-199. Chapman and Hall, New York.
76. Milliken, G. A. and Johnson, D. E. (1992). Analysis of messy data: Volume 1,
designed experiments, pp. 1-173. Chapman and Hall, New York.
77. Neter, J. and Wasserman, W. (1974). Applied linear statistical models, pp. 1-842.
Richard D. Irwin, Homewood, IL.
78. Morrison, D. F. (1990). Multivariate statistical methods. McGraw Hill, New York.
79. Mardia, K. V., Kent, J. T., and Bibby, J. M. (1979). Multivariate analysis.
Academic Press, London.
80. SAS Institute, I. (1989). SAS Institute, Inc., Gary, NC.
81. Wolfram, S. (1994). Mathematica, the student book. Addision Wesley, Reading,
MA.
82. Wilkinson, L. (1987). SYSTAT, Inc., Evanston, IL 60601.
83. Carter, C. W., Jr. (1997). In Methods in enzymology Vol. 276, p. 74.
84. Thaller, C., et al. (1981). J. Mol. Biol., 147, 465.
85. Thaller, C., et al. (1985). In Methods in enzymology Vol. 114, p. 132.
86. Wilson, L. (1998). In SpaceBound 1997 (ed. J. Sygusch). Canadian Space Agency,
Montreal, Quebec, Canada.
87. Carter, C. W., Jr. (1996). Acta Cryst., 53,647
120
5
Methods of crystallization
A. DUCRUIX and R. GIEGE
1. Introduction
There are many methods to crystallize biological macromolecules (for reviews
see refs 1-3), all of which aim at bringing the solution of macromolecules to a
supersaturation state (see Chapters 10 and 11). Although vapour phase equilib-
rium and dialysis techniques are the two most favoured by crystallographers
and biochemists, batch and interface diffusion methods will also be described.
Many chemical and physical parameters influence nucleation and crystal
growth of macromolecules (see Chapter 1, Table 1). Nucleation and crystal
growth will in addition be affected by the method used. Thus it may be wise to
try different methods, keeping in mind that protocols should be adapted (see
Chapter 4). As solubility is dependent on temperature (it could increase or
decrease depending on the protein), it is strongly recommended to work at
constant temperature (unless temperature variation is part of the experi-
ment), using commercially thermoregulated incubators. Refrigerators can be
used, but if the door is often open, temperature will vary, impeding repro-
ducibility. Also, vibrations due to the refrigerating compressor can interfere
with crystal growth. This drawback can be overcome by dissociating the
refrigerator from the compressor. In this chapter, crystallization will be
described and correlated with solubility diagrams as described in Chapter 10.
Observation is an important step during a crystallization experiment. If you
have a large number of samples to examine, then this will be time-consuming,
and a zoom lens would be an asset. The use of a binocular generally means the
presence of a lamp; use of a cold lamp avoids warming the crystals (which
could dissolve them). If crystals are made at 4°C and observation is made at
room temperature, observation time should be minimized.
2. Sample preparation
2.1 Solutions of chemicals
2.1.1 Common rules
Preparation of the solutions of all chemicals used for the crystallization of
biological macromolecules should follow some common rules:
• when possible, use a hood (such as laminar flux hood) to avoid dust
A. Ducruix and R. Giege
• all chemicals must be of purest chemical grade (ACS grade)
• stock solutions are prepared as concentrated as possible with double
distilled water.
Solubility of most chemicals are given in Merck Index. Filter solutions with
0.22 um minifilter. If you use a syringe, do not press too hard as it will enlarge
the pores of the filter. Filters of 0.4 um will retain large particles whereas
0.22 um filters are supposed to sterilize the solution. Label all solutions
(concentration, date of preparation, initials) and store at 4°C. Characterize
them by refractive index from standard calibrated solutions. Use molar units
(mole per litre) in preference to percentage. This avoids confusion between
weight to weight (w/w), weight to volume (w/v), and volume to volume (v/v).
Quite often crystallization articles refer to percentage without any inform-
ation, making the results difficult to reproduce. As an example, a 20% (w/v)
stock solution twice diluted will give a 10% solution whereas this would not
be the case if starting from a 20% (w/w) solution.
2.1.2 Buffer
The chemical nature of the buffer is an important parameter for protein crystal
growth. It must be kept in mind that the pH of buffers is often temperature-
dependent; this is particularly significant for Tris buffers. Buffers, which must
be used within one unit from their pK value, are well described in standard
text books (4).
122
5: Methods of crystallization
Method
1. Pour a column (2.5 x 10 cm2) with a mixed-bed strong ion exchange
resin in the H+-OH- form (e.g. Bio-Rad AG501X8). Wash with 300 ml
methanol:water (3:7, v/v) then with 500 ml water.
2. Dissolve 200 g PEG in water (500 ml final volume). Measure the
refractive index of the solution. Degas for 30 min under vacuum (water
aspirator) with gentle magnetic stirring. Add 1.24 g Na2S2O4.5H20 and
let stand for 1 h.
3. Pass the solution through the column at a flow rate of 1 ml/min.
Discard the first 30 ml and collect the following eluate.
4. Check the concentration of PEG by refractometry and store frozen in
small aliquots at -20°C.
5. Before use, an antioxidant can be added (para-hydroxyanisole, stock
solution in isopropanol, 1.3 mg/ml; add 1 ul per ml PEG stock
solution).
Method
1. Boil the tubing for 30 min in solution A. Avoid puncture at this stage
when mixing with glass rods or magnetic stirrers.
2. Rinse several times with distilled water.
3. Store in 50% (v/v) ethanol solution.
4. Check each tubing integrity for possible puncture.
5. Prior to crystallization, rinse membranes several times with distilled
water then with buffer.
2.2.2 Concentration
Whatever the crystallization method used, it requires high concentrations of
biological macromolecules as compared to normal biochemistry conditions.
Before starting a crystallization experiment, a concentration step is generally
needed. Keep pH and ionic strength at desired values, since pH may vary
when the concentration of the macromolecule increases. Also, low ionic
strength could lead to early precipitation (see Chapter 2 for further practical
advice). It could be very frustrating when the macromolecule precipitates
irreversibly or adsorbs on concentration apparatus membrane and/or support.
Many commercial devices are available; they are based on different principles
and operate:
124
5: Methods of crystallization
(a) Under nitrogen pressure.
(b) By centrifugation (e.g. Centricon).
(c) By lyophilization (because it may denature some proteins, test first on
a small amount). Non-volatile salts are also lyophilized and will
accumulate.
Choice of the method of concentration depends on the quantity of macro-
molecule available. Dialysis against high molecular PEG proved to be
successful in our hands. We use a dialysis chamber (volume 50-500 ul), the
top of which is covered by a glass coverslip which is sealed to the plastic
chamber with grease. Figure 1 describes the apparatus. This allows for an easy
access from the top of the dialysis chamber.
Figure 2. Dialysis button. (a) Diameter of the buttons (A) generally varies between 10-20
mm and volume of the biological macromolecule chamber is 5-350 ul. (b) To install a
dialysis membrane, use a pipetter tip of diameter adapted to the concave shape of the
dialysis button.
occur when crystals stick to the wall of the chamber. In this case a whisker can
be used to gently free the crystal.
iii. Microcap dialysis
The technique, described in Figure 3 and adapted from ref. 13, was useful for
membrane proteins (see Chapter 9). Although it is more difficult to observe
crystal growth with this method it is very convenient for storing, and micro-
caps are disposable. The method is quite easy to use (Protocol 3) and you can
play with the ratio of the diameter versus height of the microcap to influence
the kinetics of crystallization. It should be noted that when the macro-
molecule does not entirely fill the microcap chamber, the presence of air
(which is compressible) allows osmotic pressure to develop, and thus modifies
the macromolecule concentration.
Method
1. Commercial microcaps are cut with a glass saw (for instance a 50 ul
cap is cut in three parts to fit in an Eppendorf tube of 1.5 ml).
2. Wrap a piece of dialysis membrane around one end (the one which is
smooth) and secure with a piece of tubing of diameter 1.3 mm.
127
A. Ducruix and R. Giege
Protocol 3. Continued
3. Load the biological macromolecule with a Hamilton syringe.
4. Shake the assembly to bring the biological macromolecule solution in
contact with the membrane.
5. Seal the free microcap end with wax molten by soldering bit.
6. Split a second ring of tubing of diameter 3 mm and superpose it to the
first one. The aim is to prevent the membrane from touching the
bottom of the Eppendorf tube which would limit the exchange with the
reservoir.
7. Insert in an Eppendorf tube (volume 1.5 ml) containing 1 ml of the
crystallizing solution.
8. Close cap and wrap top of Eppendorf tube with Parafilm (American
Can Company).
Figure 3. Crystallization by microcap dialysis. (a) Place the dialysis membrane on the
microcap; (b) secure with Tygon ring; (c) load the protein; (d) close the extremity with
wax; (e) fill up a 1.5 ml Eppendorf tube with crystallizing agent and insert microcap.
128
5: Methods of crystallization
Figure 4. Double dialysis set-up (adapted from ref. 14). Macromolecule is contained in a
conventional dialysis button placed in a second dialysis set-up. The equilibration rate
depends upon the volumes of buffers in the different compartments.
Method
1. Prepare the dialysis burton as in Section 3.2.2 with a solution of
crystallizing agent at a concentration in which the biological macro-
molecule is undersaturated. This is called the 'inner compartment'.
2. Insert the conventional dialysis burton in a vial (about 10 ml) called the
'middle compartment' containing a solution of crystallizing agent at
a concentration in which the biological macromolecule is super-
saturated.
3. Cover with a dialysis membrane maintained by an O-ring.
129
A. Ducruix and R. Giege
Protocol 4. Continued
4. Place in a larger vial (for instance a beaker of 50 ml) containing the
solution of crystallizing agent at a concentration in which the bio-
logical macromolecule will precipitate completely. This is the 'outer
compartment'.
5. Cover with Parafilm or a stopper.
4.1 Principle
The principle of vapour diffusion crystallization is indicated in Figure 5. It is
very well suited for small volumes (down to 2 ul or less). A droplet containing
the macromolecule to crystallize with buffer, crystallizing agent, and addi-
tives, is equilibrated against a reservoir containing a solution of crystallizing
agent at a higher concentration than the droplet. Equilibration proceeds by
diffusion of the volatile species (water or organic solvent) until vapour
pressure in the droplet equals the one of the reservoir. If equilibration occurs
by water exchange from the drop to the reservoir, it leads to a droplet volume
decrease. Consequently, the concentration of all constituents in the drop will
increase. For species with a vapour pressure higher than water, the exchange
occurs from the reservoir to the drop. In such a 'reverse' system, the drop
volume will increase as well as the concentration of the drop constituents.
This last solution, less widely used, has led to the crystallization of tRNAAsp
(16) and of several proteins (17, 18). The same principle applies for hanging
drops, sitting drops, and sandwich drops.
Glass vessels in contact with macromolecular solutions should be treated
Figure 5. Schematic representation of hanging drop, sitting drop, and sandwich drop.
130
5: Methods of crystallization
A. Silanization
1. Place coverslips in a bath of toluene containing 1% dimethyl-
dichlorosilane at 60°C for 10 min.
2. Coverslips are then washed with soap solution and rinsed with
distilled water and ethanol.
3. The same procedure is used for Pyrex plates.
4. Dry overnight at 120°C to sterilize vessels.
B. Siliconization
1. This can be achieved with commercially available reagent solutions
(e.g. Sigmacoat). Coverslips are washed in the solution, and dried
overnight at 120°C.
"All operations can be performed under vacuum when dealing with narrow vessels like
capillaries.
Figure 6, A device for treating coverslips (Perspex or Teflon made). The set-up displayed
(in Teflon) can be manufactured in the laboratory workshop. Coverslips are held in the
threading of the two bottom axis; a smaller unthreaded axes secures the coverslips. The
set-up shown is about 20 cm long and permits handling of about 60 coverslips.
131
A. Ducruix and R. Giege
Figure 7, A Linbro box for vapour diffusion crystallization in hanging drops. The photo-
graph shows the box with its cover which is held by Plasticine in the corners. For a better
display, two drops were prepared with dyes (at the left).
Drops are set up following Protocol 6. Most of the people use a 'magic'
ratio of two between the concentration of the crystallizing agent in the
reservoir (well) and in the droplet. This is conveniently achieved by mixing a
droplet of protein at twice the desired final concentration with an equal
volume of the reservoir at the proper concentration (other ratios can be used
as well). Avoiding local over-concentration can be achieved by placing the
two drops (protein and reservoir) on each side of an Eppendorf tube and
vortexing it quickly.
132
5; Methods of crystallization
Method
1. Grease rims with silicone grease.a
2. Fill up reservoir with 1 ml of filtered (0.22 um) crystallizing agent.
3. Spray glass coversllp with antidust.
4. Mix a 2-10 ul drop of filtered (0.22 um) biological macromolecule
solution with an equivalent volume of reservoir.
5. Layer the drop on the 22 mm diameter coverslip (do not touch the
coverslip with the extremity of the tip of the pipettor or it will spread)
so that a nearly hemispherical drop is formed.b
6. Return the coverslip with a pair of brussel (or fingers). First train
yourself with water!
7. Set on the grease rim and gently press to seal the well with the
grease. Do not press too firmly, otherwise the coverslip will break.
8. Check the sealing by inspecting the rim in an azimuthal way. If sealing
is not properly done, the drop will concentrate as well as the
reservoir. Crystallization will occur eventually, but will be very
difficult to reproduce.
9. Adjust glass coverslips tangent to each other otherwise they overlap.
10. Put Plasticine in the corners to avoid the contact between the cover
and grease, otherwise slips get stuck at the cover.
a
To dispense the grease, fill up a syringe and replace the needle by an Eppendorf yellow tip.
b
You may layer several microdrops on a coverslip.
i. Problems
For membrane proteins (see Chapter 9), the presence of detergent tends to
spread out the drops and lower the surface tension. In all cases gravity will
tend to sink the drops containing the macromolecule in the reservoir for
volumes exceeding 25 ul. Shaking Linbro boxes will give the same results.
Boxes must be transported horizontally and carefully. It is always painful for
beginners to ruin an experiment when transporting a box. If you prepare
133
A. Ducruix and R. Giege
boxes at room temperature and transfer them to 4^°C, condensation will
occur on the surface of the coverslip. Water droplet will surround the
macromolecule drop. If it mixes, the protein will dilute and probably stay in
an undersaturated state. To avoid this problem, set up boxes at the final
experiment temperature and cover boxes with polystyrene sheets.
It is possible to recrystallize macromolecules using hanging drops as
described in Protocol 7.
A. For purification
1. Remove mother liquor.
2. Wash crystals with fresh buffer.
3. Redissolve in fresh crystallizing agent solution.
4. Recrystallize.
B. For crystal growth
1. Redissolve crystals by replacing crystallizing agent in the reservoir by
buffer. Depending on the biological macromolecule, it may take a few
hours or a few days. The drop volume will increase; so control the
process, otherwise the drop will sink.
2. Filter the drop in a minifilter (e.g. Costar, Millipore) by centrifugation.
Warning: the dead volume is at least 5 ul.
3. Place the drop on a clean glass coverslip.
4. Add some grease to the rim.
5. Fill the reservoir with crystallizing agent and recrystallize.
Figure 8. ACA CrystalPlate® (courtesy of ICN Flow). This is a versatile system for vapour
diffusion crystallization allowing individual experiments on sitting, hanging, or
sandwiched drops under classical or automated conditions.
tank in Figure 8 — may be removed with pliers and a rubber septum inserted
so that the reservoir solution may be changed with an hypodermic syringe.
A crystallization set-up based on the same versatile concept as the
CrystalPlates® is the so-called Q-Plate™ (supplied by Hampton Research).
i. Advantages
Because of the glass windows, when looking at drops under polarized binocular,
birefringence of crystals can be observed. Drops with macromolecules are no
longer above the reservoir thus eliminating sinking. Large drops can be
prepared with the sandwich method.
Warning: if you use a rubber septum, plates stick when you translate them
during observation, eventually provoking disasters.
Method
1. Prepare the plate by filling up the upper and lower troughs of each
well with ordinary hydrocarbon vacuum pump oil or grease.
2. To dispense oil or grease, fill up a syringe and replace the needle by
an Eppendorf yellow tip.
3. Put 0.5 ml of crystallizing agent into each reservoir.
4. Position one of the 14 x 14 mm2 glass coverslips over the hole in each
well. The coverslips should seal quickly if there is enough oil or grease.
(a) For hanging or sitting drops use 14 x 14 x 0.2 mm3 glass
coverslips.
135
A. Ducruix and R. Giege,
Protocol 8. Continued
(b) For sandwich drops (5-25 ul) use 14 x 14 x 1 mm3 glass cover-
slips.
(c) For sandwich drops (15-75 ul) use 14 x 14 X 1.5 mm3 glass cover-
slips.
5. Put a drop of the biological macromolecule solution in the centre of
the lower (for sitting or sandwich drops) or upper (for hanging drops)
glass coverslips and then set one of the 24 x 30 x 1 mm3 glass
coverslip in position on the upper trough.
136
5: Methods of crystallization
to use the HANGMAN framework. Here the protein droplets are first in-
stalled on the tape which is than inverted in a second step over the plate (22).
A variety of other set-ups have been designed in many laboratories, allow-
ing for instance Linbro or VDX boxes to be used for sitting drop experiments
(with the depression on a small plastic bridge (23) or on a glass rod as on the
Oxford or Perpetual Systems Corporation set-ups, respectively); or doing
vapour phase equilibration in capillaries (24, 25), or even directly in X-ray
capillaries as was described for ribosome crystallizations (26) or in the gel
acupuncture method. For extremely fragile crystals, when transfer from
crystallization cells to X-ray capillaries (see Chapter 14) can lead to internal
damage and mechanical cracks of crystals, this last method may be well
adapted.
Figure 10. Water evaporation kinetics in the presence of ammonium sulfate (AS), MPD,
PEG, and NaCI as dehydrating agents. (a) Measurements done in drops set in Linbro
boxes with AS, MPD, and PEG. The data also show the influence of protein (in AS
experiment), and of initial drop volume (in MPD experiment) on final drop volume. V0 is
the initial volume of the drop; experiments were conducted with a concentration of
crystallizing agent in the reservoir twice that in the drop, time at zero. Adapted from ref.
32 (b) Measurements done on hanging drops (24 ul) set over test-tubes with distances
between the drop and reservoir varying from 7.6-78.3 mm. Experiments were conducted
at 22.9°C with 1.0 M NaCI as initial concentration in the drop and 2.0 M NaCI in the
reservoirs. Adapted from ref. 35.
only present to maintain constant vapour pressure. Because one starts from
supersaturation, nucleation tends to be too large. However, in some cases
fairly large crystals can be obtained when working close to the metastable
region. This is illustrated in Chapter 10. If supersaturation is too high, a
precipitate may develop in a batch crystallization vessel; do not discard such
experiments because crystals can eventually grow from the precipitate by
Ostwald ripening (41). Notice, however, that the growth kinetics under such
circumstances are decreased.
An automated system for microbatch macromolecule crystallizations and
screening has been described (42) allowing the set up of samples of less than
2 ul. Reproducibility of experiments is guaranteed because samples are
139
A. Ducruix and R. Giege
dispensed and incubated under paraffin oil, thus preventing evaporation and
uncontrolled concentration changes of the components in the micro-droplets.
Using silicone oils that are slightly soluble in water, or appropriate mixtures
of paraffin and silicone oils, results in gradual protein concentration in the
droplets like in vapour diffusion experiments (43).
A variation of classical batch crystallization is the sequential extraction
procedure of Jakoby (44), based on the property that many proteins (not all)
are more soluble in concentrated salt (e.g. ammonium sulfatc) when lowering
the temperature. The method can be adapted for microassays and was
successfully applied for the crystallization of a proteolytic fragment of
methionyl-tRNA synthetase from Escherichia coli (45).
Figure 11. Crystallization in floating drops. Droplets (40 ul) of four different crystallizing
agent solutions placed at the interface of two silicone fluids are displayed. The two
silicone fluids at 20°C (800 ul each with low density fluid PS037 layered over high density
PS181 from Huls America, Inc.) [silicone oils may be provided by Hampton Research) are
placed in square glass or polystyrene spectrophotometer cuvettes. Adapted from ref. 48.
140
5: Methods of crystallization
Several proteins and a spherical plant virus were crystallized in the
temperature range 4-20°C using this method (48). Its main advantage is to
reduce the nucleation rate. Thus crystallization in floating drops provides a
means to obtain a small number of larger crystals in an homogeneous liquid
medium. Because drops are not in contact with air, the method may be
convenient to crystallize proteins sensitive to oxidation. Further, when im-
plemented in a thermostated device, the method provides a simple and
convenient way for kinetic measurements of macromolecule crystal growth
(48).
Other advanced methods useful to prepare crystals for diffraction studies
(e.g. in gelified media, under microgravity, and the gel acupuncture method)
are described in Chapter 6.
7.1 Dialysis
In the case of dialysis buttons, if one considers that stretching of the
membrane is negligible, the macromolecule concentration will remain con-
stant. However, if the macromolecule solution does not fill the chamber
entirely, leaving room for air, it is no longer exactly true since the macro-
molecule concentration may vary (increase or decrease depending on the
situation). The initial concentration of the crystallizing agent in the reservoir
(this could be buffer) leaves the macromolecule in an undersaturated state.
141
A. Ducruix and R. Giege
Figure 12. Schematic solubility diagram and correlation between macromolecule and
crystallizing agent concentrations in a crystallization experiment using a dialysis set-up.
Ci is the initial concentration of crystallizing agent and C the constant protein con-
centration. The area between the precipitation and solubility curves is the supersaturated
region where crystallization can occur. Precipitation and solubility curves can be
determined experimentally, although for the latter one crystals should be obtained first.
For more details see Chapter 10.
Figure 13. Schematic solubility diagram and correlation between macromolecule and
crystallizing agent concentrations in crystallization experiments using vapour diffusion
set-ups. Situation without (a) and with (b) crystallization. See legend to Figure 12 and text
for further explanations.
Figure 14. Schematic solubility diagram and correlation between macromolecule and
crystallizing agent concentrations in crystallization experiments using a batch method (in
closed vessels). See legend to Figure 12 and text for further explanations.
curves (point B). In that case the arrow describes the variation of the
remaining concentration of protein in solution. In the last case (point C) the
protein will precipitate immediately because supersaturation is too high. In
some cases, however, crystals may grow from the precipitates (41).
8. Practising crystallization
It is a good exercise to train oneself with a cheap easy accessible protein.
Lysozyme, thaumatin, thermolysin, and BPTI are good candidates which are
commercially available from various manufacturers and which crystallize
readily. Examples of crystallization using various methods in hanging drops
are given in Protocols 9 and 10.
143
A. Ducruix and R. Giege
Method
1. Prepare stock solutions of 3 M NaCI and 40 mg/ml (2.74 mM)
lysozyme in 50 mM acetate pH 4.5 and buffer stock solution (50 mM
sodium acetate pH 4.5). Filter all solutions with a 0.22 um microfilter.
2. Prepare a Linbro box as described in Section 4.2.1.
3. Fill up reservoirs of row A with solutions of NaCI ranging from
0.5-1.5 M in steps of 0.2 M.
4. On a coverslip, mix 4 ul of protein stock solution with 4 ul of
reservoir. Flip it and set it on the greased rim.
5. Fill up reservoirs of row B with solutions of NaCI ranging from
0.8-1.8 M in steps of 0.2 M. Repeat the experiment on row B after
diluting the protein stock solution by a factor two to obtain a new one
of 20 mg/ml (1.37 mM).
6. Fill up reservoirs of row C with solutions of NaCI ranging from 1.5-2.5
M in steps of 0.2 M. Repeat the experiment after diluting the protein
stock solution by a factor two.
7. Use row D for duplicate or testing particular parameters (e.g. volume
of drops to see the influence of kinetic effects on growth).
8. Store the experiments at 18°C.
9. Observe the experiments once a day for a week.
10. Train yourself in mounting crystals (see Chapter 14).
Method
1. Prepare stock solutions as in Protocol 9.
2. Fill the protein chamber of a dialysis button with protein stock solution
144
5: Methods of crystallization
diluted to 10 mg/ml. The solution must form a dome above the entry.
Install the dialysis membrane of appropriate cut-off as described in
Section 3.2.2.
3. Fill up with 1 ml of reservoir solution the reservoir of a Linbro box and
drop the button in it with the aperture of the protein chamber on the
top.
4. Store the experiments at 18°C.
5. Increase the concentration in the reservoir by 200 mM of crystallizing
agent every day.
6. Observe the next day.
7. Repeat steps 5 and 6 until crystallization occurs.
9. Concluding remarks
In this chapter we have described the most common crystallization methods;
all of them have advantages and drawbacks. Most of crystallographers favour
vapour phase diffusion which provides an easy way to practise crystallization.
It is also the method of choice for robotics (51). Dialysis presents the
advantage that macromolecular concentrations remain constant, so that only
one parameter varies at a time and nature of buffer or crystallizing agent can
be changed easily. It differs with a classical vapour phase equilibrium crystal-
lization experiment where all constituents in the drop are concentrated.
Beside the classical and advanced crystallization methods described in this
chapter, less standard methods may be useful in particular cases. These can be
methods based on old ideas not yet well explored, as pulse-diffusion of pre-
cipitant combining dialysis and free diffusion in capillaries (52), or combina-
tions of dialysis and electrophoresis (53). Also, crystallization in particular
environments should be considered, such as under levitation (54), in centri-
fuges (55, 56), or in magnetic (57, 58) or electric (59) fields. Particular
attention should be given to crystallization methods where convection is
reduced; e.g. the gel acupuncture method, crystallization in gels which is
becoming popular in the macromolecule field, and crystallization in micro-
gravity (Chapter 6). Methods based on temperature diffusion, which are
widely used in material sciences (60), may be adapted under certain con-
ditions for macromolecule crystallization (61, 62). Finally, the use of novel
types of crystallization cells may represent an interesting alternative for
growing better crystals. In particular, cells based on principles developed for
microgravity experiments may be appropriate (see refs 63-65). It is our hope
that the methods and ideas discussed in this chapter will help readers, not only
to solve their crystallization problems, but also to improve existing methods,
and even to develop new crystallization methodologies.
145
A. Ducruix and R. Giege
References
1. McPherson, A. (1985). In Methods in enzymology, (eds Wickoff, H. W., Hirs,
C. H., and Timasheff, S. N.), Academic Press, London. Vol. 114, p. 112.
2. Giege, R. (1987). In Crystallography in molecular biology (ed. D. Moras,
J. Drenth, B. Strandberg, D. Suck, and K. Wilson), NATO ASI Series, Vol. 126,
pp. 15-26. Plenum Press, New York and London.
3. McPherson, A. (1990). Eur. J. Biochem., 189, 1.
4. Perrin, D. D. and Dempsey, B. (1974). Buffer for pH and metal ion control.
Chapman and Hall Ltd., London and New York.
5. Jurnak, F. (1986). J. Cryst. Growth, 76, 577.
6. Ray, W. J. and Puvathingal, J. (1985). Anal. Biochem., 146, 307.
7. Chayen, N., Akins, J., Campbell-Smith, S., and Blow, D. M. (1988). /. Cryst.
Growth, 90, 112.
8. Winterbourne, D. J. (1986). Biochem. Soc. Trans., 14, 1179.
9. Mikol, V. and Giege, R. (1989). J. Cryst. Growth, 97, 324.
10. Bradford, M. M. (1976). Anal. Biochem., 72, 248.
11. Zeppenzauer, M. (1971). Methods in enzymology, ibid ref. 1. Vol. 22, p. 253.
12. Weber, B. H. and Goodkin, P. E. (1970). Arch. Biochem. Biophys., 141, 489.
13. Pronk, S. E., Hofstra, H., Groendijk, H., Kingma, J., Swarte, M. B. A., Dorner, F.,
et al. (1985). J. Biol. Chem., 260, 13580.
14. Thomas, D. H., Rob, A., and Rice, D. W. (1989). Protein Eng., 2, 489.
15. Hampel, A., Labananskas, M., Conners, P. G., Kirkegard, L., RajBhandary, U. L.,
Sigler, P. B., et al. (1968). Science, 162, 1384.
16. Giege, R., Moras, D., and Thierry, J.-C. (1977). J. Mol. Biol., 115, 91.
17. Richard, B., Bonnete, F., Dym. O., and Zaccai, G. (1995). In Archaea, a laboratory
manual, pp. 149-54. Cold Spring Harbor Laboratory Press.
18. Jeruzalmi, D. and Steitz, T. A. (1997). J. Mol. Biol., 274, 748.
19. Zuk, W. M. and Ward, K. B. (1991). J. Cryst. Growth, 110, 148.
20. Kim, S.-H. and Quigley, G. (1979). In Methods in enzymology, Vol. 59, p. 3.
21. Dock, A.-C., Lorber, B., Moras, D., Pixa, G., Thierry, J.-C., and Giege, R. (1984).
Biochimie, 66, 179.
22. Luft, J. R., Cody, V., and DeTitta, G. T. (1992). J. Cryst. Growth, 122, 181.
23. Harlos, K. (1992). J. Appl. Crystallogr., 25, 536.
24. Phillips, G. N., Jr. (1985). Methods in enzymology, ibid ref. 1. Vol. 104, p. 128.
25. Luft, J. R. and Cody, V. (1989). J. Appl. Crystallogr., 22, 386.
26. Yonath, A., Mussig, J., and Wittmann, H. G. (1982). J. Cell. Biochem., 19, 145.
27. McPherson, A. (1985). In Methods in enzymology, Vol. 114, p. 125.
28. McPherson, A. (1995). J. Appl. Crystallogr., 28, 362.
29. Mikol, V., Rodeau, J.-L., and Giege, R. (1989). /. Appl. Crystallogr., 22, 155.
30. McPherson, A. and Shlichta, P. (1988). Science, 239, 385.
31. Chayen, N. E., Lloyd, L. F., Collyer, C. A., and Blow, D. M. (1989). J. Cryst.
Growth, 97, 367.
32. Mikol, V., Rodeau, J.-L., and Giege, R. (1990). Anal. Biochem., 186, 332.
33. Luft, J. R., Arakali, S. V., Kirisits, M. J., Kalenik, J., Wawzzak, I., Cody, V., et al.
(1994). /. Appl. Crystallogr., 27, 443.
34. Arakali, S. V, Easley, S., Luft, J. R., and DeTitta, G. T. (1994). Acta Cryst., D50, 472.
146
5: Methods of crystallization
35. Luft, J. R., Albrigth, D. T., Baird, J. K., and DeTitta, G. T. (1996). Acta Cryst.,
D52, 1098.
36. Fowlis, W. W., DeLucas, L. J., Twigg, P. J., Howard, S. B., Meehan, E. J., and
Baird, J. K. (1988). J. Cryst. Growth, 90, 117.
37. Chayen, N. E. (1997). J. Appl. Crystallogr., 30, 198.
38. McPherson, A. (1976). J. Biol. Chem., 251, 6300.
39. Gernert, K. M., Smith, R., and Carter, D. C. (1988). Anal. Biochem., 168, 141.
40. Przybylska, M. (1989). J. Appl. Crystallogr., 22, 115.
41. Ng, J., Lorber, B., Witz, J., Theobald-Dietrich, A., Kern, D., and Giege, R. (1996).
J. Cryst. Growth, 168, 50.
42. Chayen, N. E., Shaw Stewart, P. D., Maeder, D. L., and Blow, D. M. (1990).
J. Appl. Crystallogr., 23, 297.
43. d'Arcy, A., Elmore, C., Stihle, M., and Johnston, J. E. (1996). J. Cryst. Growth,
168, 175.
44. Jakoby, W. B. (1971). In Methods in enzymology, Vol. 22, p. 248.
45. Waller, J. P., Risler, J.-L., Monteilhet, C., and Zelwer, C. (1971). FEBS Lett., 16,
186.
46. Visuri, K., Kaipainen, E., Kivimaki, J., Niemi, H., Leissla, M., and Palosaari, S.
(1990). Biotechnology, 547.
47. Lorber, B., Jenner, G., and Giege, R. (1996). J. Cryst. Growth, 158, 103.
48. Lorber, B. and Giege, R. (1996). J. Cryst. Growth, 168, 204.
49. Chayen, N. E. (1996). Protein Eng., 9, 927.
50. Salemne, F. R. (1972). Arch. Biochem. Biophys., 151, 533.
51. Zuk, W. M. and Ward, K. B. (1991). J. Cryst. Growth, 110, 148.
52. Koeppe, R. E., Stroud, R. M., Pena, V. A., and Santi, D. V. (1975). J. Mol. Biol.,
98, 155.
53. Chin, C.-C., Dence, J. B., and Warren, J. C. (1976). J. Biol. Chem., 251, 3700.
54. Rhim, W.-K. and Chung, S. K. (1990). Methods: a companion to methods in
enzymology, 1, 118.
55. Barynin, V. V. and Melik-Adamyan, V. R. (1982). Sov. Phys. Crystallogr., 27, 588.
56. Lenhoff, A. M., Pjura, P. E., Dilmore, J. G., and Godlewski, T. S., Jr. (1997).
J. Cryst. Growth, 180, 113.
57. Sazaki, G., Yoshida, E., Komatsu, H., Nakada, T., Miyashita, S., and Watanabe,
K. (1997). J. Cryst. Growth, 173, 231.
58. Ataka, M., Katoh, E., and Wakayama, N. I. (1997). J. Cryst. Growth, 173, 592.
59. Taleb, M., Didierjean, C., Jelsch, C., Mangeot, J.-P., Capelle, B., and Aubry, A.
(1998). J. Cryst. Growth,
60. Feigelson, R. S. (1988). J. Cryst. Growth, 90, 1.
61. Lorber, B. and Giege, R. (1992). /. Cryst. Growth, 122, 168.
62. DeMattei, R. C. and Feigelson, R. S. (1993). /. Cryst. Growth, 128, 1225.
63. Stoddard, B. L., Strong, R. K., and Farber, G. K. (1988). J. Cryst. Growth, 110,
312.
64. Bosch, R., Lautenschlager, P., Potthast, L., and Stapelmann, J. (1992). J. Cryst.
Growth, 122, 310.
65. McPherson, A. (1996). Crystallogr. Rev., 6, 157.
147
This page intentionally left blank
6
1. Introduction
From the first studies showing the feasibility of macromolecular crystal
growth in gels (1), an increasing attention has been paid to applications of gel
techniques to the domain of biological macromolecules. Confidence in these
techniques is such that kits of crystallization in gels are now commercially
available (Hampton Research, Laguna Hills, CA, USA).
Basically, the protein crystallization process consists of two consecutive
steps:
• first, the transport of growth units towards the surface of the crystals
• second, the incorporation of the growth units into a crystal surface position
of high bond strength.
The whole growth process is dominated by the slowest of these two steps
and is either transport controlled or surface controlled. Avoiding convection
in the growth environment will increase the possibility of growing the crystal
under slow diffusive mass transport providing that the surface interaction
kinetics are faster than the characteristic diffusive flow of macromolecules (in
the range of 10-6 cm2/sec for proteins). The ratio between transport to surface
kinetics, which can be tuned by either enhancing or reducing transport pro-
cesses in the solution, has been shown (2) to control the amplitude of growth
rate fluctuations (which is thought to reduce crystal quality). These are the
main reasons why gels (as well as capillaries and microgravity conducted
experiments), if correctly designed, are expected to enhance the quality of
crystals. This quality enhancement (3), as well as the possibility of getting
crystals when conventional solution techniques failed (4), have been experi-
mentally demonstrated. However, up to now, gel methods have been used on
a rather empirical basis, as a simple transposition of solution techniques, and
recent fundamental studies of nucleation and growth in gels show that the
situation is not as simple as first expected (5, 6).
After summarizing the main characteristics of crystal growth in gels, we will
M.-C. Robert et al.
examine what are the best conditions using a gel method. Recipes for the
preparation of different gel growth experiments will be given. Considering gel
growth as a possible simulation of experiments under reduced gravity, recent
results of space experiments will be reviewed. Mention will also be made to
growth under hypergravity conditions.
2. General considerations
Gels used for crystal growth are hydrogels with a growth solution soaking a
polymeric network. For physical gels like gelatin or agarose, sol-gel transition
is obtained by decreasing the temperature (physical parameter variation).
Polymerization corresponds to the formation of weak bonds and this process
is reversible with some hysteresis (~ 50°C for agarose). For chemical gels,
such as polyacrylamide, polymerization corresponds to strong bonding and is
not reversible. Although formation of silica gels also results from a chemical
reaction, it rather corresponds to an intermediate case between chemical and
physical gels: as a matter of fact, the chemical reaction leads to the formation
of dense beads which further aggregate by weak bonding (7). As far as we
know, only agarose and silica gels have been successfully used for macro-
molecule crystal growth.
150
6: Crystallization in gels and related methods
Figure 1. (a) Basic repeat unit of agarose chains: the agarobiose. (b) Schematic structure
of the agarose gel showing the disordered junction zones.
The waste products of the reactions are, in the first case a sodium salt and in
the second case an alcohol (if necessary, the waste products can be washed
off, after gel setting).
The silicic acid tends to polymerize according to the reaction:
Figure 2. (a) The polymerization of the silicic acid gives rise to colloidal silica beads
whose internal structure is due to siloxane bonding, leaving silanol groups on the
surface, (b) The growth of beads depends on the chemical conditions: (c) network
obtained at high pH, (d) network obtained at low pH.
Figure 4. Rocking curve and topography (inset) acquired from a tetragonal HEW lysozyrne
crystal grown in high concentration gel. Although the gel network is embedded into the
crystal lattice, the crystal quality is preserved.
High resolution diffraction has been observed for crystals grown from
rather firm agarose gels (12). However the use of light gels is advisable, except
for the special cases discussed above, first to make removal of crystals out of
the gel matrix easier, and secondly to minimize the gel contamination.
2.2.3 Nucleation inside the gel
Although seeding can be used, it appears that most of the gel-grown crystals
arc obtained by spontaneous nucleatkm inside a macroscopically homo-
geneous gel. When the gel well adheres to the walls of the container (without
intercalated liquid film), no nucleation occurs on the cell walls, neither on dust
or fibres which have been embedded in the gel. So heterogeneous nucleation
is strongly reduced, if not suppressed. Another type of nucleation, namely
secondary nucleation, is due to attrition of a previous crystal by the solution
flux. It is quite clear that, in gels, this type of nucleation is prevented. One can
154
6: Crystallization in gels and related methods
case in gel-free solutions. One generally observes that, in silica gels, the
number of nucleated crystals decreases by increasing the gel content. This has
been shown with a large variety of macromoleeules (14). In agarose gels, on
the contrary, the number of crystals increases by increasing the gel content
(15.16).
The influence of the gel media on the nucleation rate cannot be explained
by changes of solubility values as far as a solubility curve relates equilibrium
between a crystalline phase and a solution. However, supersaturation can be
lowered as observed in silica gels where protein molecules ean adsorb on the
gel surface. Small angle neutron scattering (SANS) spectra of proteins,
differing markedly from those corresponding to the gel-free solution, account
for this effect. So, with HEW lysozyme solutions at pH 4,5, part of the protein
molecules are adsorbed on the gel through electrostatic and H-bond inter-
actions, which reduces the content of free molecules remaining in solution.
The concentration of protein adsorbed increases by increasing the total
protein content (until binding sites on the gel surface are covered) and by
increasing the supersaturation (through protein-protein interaction) (see e.g.
Figure 5 in ref. 6). The latter process is reversible. Thus, reduction of the
nucleation rates is simply explained by a reduction of the actual super-
saturation. One can counterbalance this effect by using mixed silica gels
(TMOS and MeTEOS) (Section 3.1.1), which increases the nucleation rate.
In agarose gels, the initial free protein concentration is the same as in gel-
free solution. Differences in the SANS signals of gelled and gel-free HEW
156
6: Crystallization in gels and related methods
lysozyme solution are only visible in the very small scattering vector range.
These signals are related to the presence of aggregates (in the 100 nm range):
when the protein solution is trapped in agarose gel the signal is enhanced (5).
For macromolecular crystals, it is assumed that nucleation could occur via
restructuring of amorphous aggregates (17, 18). Here the concomitant observ-
ations of enhanced aggregation and enhanced nucleation, both increasing
with the gel content, supports this hypothesis. As large aggregates cannot
sediment in gel, they are maintained in the whole bulk as shown by refractive
index measurements as a function of time during the nucleation process (19).
An opposite effect is visible in solution under normal gravity or hypergravity
conditions (Section 5.2).
From a practical point of view, use of either silica or agarose gels is inter-
esting because situations exist where, in solution, nucleation is either too
abundant or too scarce.
2.2.4 Parameters influenced by the presence of a gel
Taking into account the above considerations, one can select from the list of
parameters influencing biological macromolecule crystal growth (see Chapter 1,
Table 7) those which are influenced by the presence of a gel structure. They
are parameters either related to the supply of reactants or related to the
mechanical behaviour of the solid or liquid phases.
A priori, gels are not expected to improve crystal growth with regards to
biochemical parameters like purity or degree of denaturation of the macro-
molecule. However, this assertion is refuted by a comparative study (in gel
and in solution) on the effect of contamination by a parent molecule (20, 21).
Indeed, it was shown that crystals of good quality can be obtained, even when
contamination levels are much higher in the gel than in free solution. In
current crystallization conditions, impurities are either rejected or incorpor-
ated in the crystal. Consequently, the impurity concentration nearby the
interface increases (or decreases) with respect to the concentration in the
bulk. In free solution, on earth, convections provoke fluctuations of the im-
purity content at the interface resulting in time-dependent incorporation of
impurity in the crystal (growth striations). In gels, due to the supply of solute
by diffusion, such fluctuations are damped; furthermore, with a slow growth,
foreign molecules or molecules having a distorted conformation can be
rejected from the growth interface instead of being buried in the crystal
network. In silica gels, the adsorption-desorption process could also act as a
purification process, assuming that ill-folded molecules are more strongly
bound to the gel surface.
3. Practical consideration
Gel growth is a particular case of solution growth so that it must always be
considered downstream with respect to classical solution growth. It results
157
M.-C. Robert et al.
that the same chemical components would be chosen among those which have
given the best results in solution. Then using gels, one tries to offset the
drawbacks encountered during these first trials. A first possibility is to
nucleate and grow the crystals inside the gel. This implies that the protein
solution is either gelified or brought into a gel previously set. The different
procedures are detailed in Section 3.2.1. A second possibility is to use the gel
as a diffusion medium to monitor the supply of reactants, the crystal growing
outside the gel. This technique, known as gel acupuncture method, will be
described in Section 3.3.1.
A. Hydrolysis of siloxane
1. Tetramethoxysilane or tetraethoxysilane are liquids very soluble in
alcohol but not much in water. Add drop by drop 1 ml TMOS to 20 ml
buffer solution (e.g. 0.1 M acetate buffer solution). Dissolution occurs
through a vigorous stirring of siloxane droplets in water (Figure 7a);
this emulsifying provides a large contact surface, which allows the
dissolution of a small amount of siloxane. The reaction proceeds
according to Equation 2 (Section 2.1.2) and methanol is progressively
released which makes easy the complete siloxane dissolution. This
process consumes 12.5 ml water and releases 54.5 ml methanol per
litre of solution. This must be taken into account to know the final
growth solution composition.
2. The homogenization step must be achieved as quickly as possible
because it competes with the polymerization reaction step, which
begins as soon as monomers are available in the medium. It is
possible to delay the polymerization step by keeping the mixture in a
water/ice bath (Figure 7b).
3. The lower the pH, the more rapid is homogenization.
4. The mixture first looks like an oily emulsion in water, then, when no
parasitic reaction occurs, it becomes clear and homogeneous.
158
6: Crystallization in gels and related methods
5. One can reduce the interactions between protein and silica gel by
adding some amount (~ 30% in weight) of methyltriethoxysilane,
MeTEOS, (C2H5O)3Si-CH3 (for which the hydrolysis reaction is similar)
to the TMOS so that Si-CH3 groups are substituted to Si-OH groups on
the gel surface (6).
Siloxanes are corrosive liquids. Careful protection of skin and eyes are
recommended. All vessels in contact with them must be thoroughly
rinsed with alcohol prior to water cleaning.
ii. Polymerization
After having added an aliquot of the silica sol to the solution at the required
composition, the preparation is thoroughly mixed (Figure 7c). It must remain
homogeneous (no flocculation). The mixture is poured in clean, dried
crystallization containers and allowed to gelify without mechanical disturb-
ances. The gel must stick to the container cell walls. It can look somewhat
opalescent, but without macroscopic heterogeneities such as fissures.
Dehydration of the gel surfaces must be avoided, either by sealing them with
a minimum air volume enclosed or by closing them in a vessel containing a
reservoir of solution giving the suitable vapour pressure (Figure 7d).
Gelation time depends on many parameters such as concentration of gelling
agent, nature and concentration of species in solution, pH, and temperature.
So, at room temperature it can vary from a few minutes at pH 7.0 to several
hours at pH 4.0. With thermally stable solutions, one can shorten this time by
increasing temperature (typically 40°C for further use at room temperature).
The gel can be considered as set when it resists pouring, though it undergoes
some further evolution as shown by light scattering techniques.
3.1.2 Agarosegel
Commercially available agaroses are powders which can be dissolved in water
as described in Protocol 2. Homogeneous preparations at concentrations
higher than 2% (w/v) are difficult to achieve.
159
M.-C. Rnhnrt et al.
Figure 7. Two-step procedure for crystal growth in a CG silica gel. After setting the silica
gel at T - 40°C, the protein solution diffuses into the gel. Then, the crystal growth occurs
under the final concentrations of protein CP, salt Cs, and buffer CB/2.
160
6: Crystallization in gels and related methods
Method
1. Add progressively 0.1 g agarose to 10 ml water at room temperature
(not the reverse which could leads to agglomerates) with a slow
stirring (Figure 8a).
2. Keep stirring for a couple of hours at ambient temperature.
3. Raise the temperature to 100°C (possibly in a water-bath to avoid
overheating) and maintain the slow stirring. The mixture must rapidly
become as clear and transparent as water (Figure 8b).
4. Keep it on the hot plate (80°C) until use.
161
M.-C. Robert et al.
Protocol 3. Continued
Method
1. The silica sol is prepared as explained in Protocol 1 in a buffer at
concentration CB. It is kept in a water/ice bath to avoid a premature
gelation. The different growth settings are presented Figure 7.
2. Prepare a growth solution with a protein concentrations CP, salt
concentration Cs, gel concentration CG.
3. Mix equal volumes of stock salt solution (at concentration 2 Cs) and
silica sol (at concentration 2 CG) (Figure 7c).
4. Suck a few microlitres of this preparation in capillaries (or prepare
droplets of this preparation on coverslips). Place the capillaries
opened at one end (or the coverslip) in a closed vessel containing a
reservoir of salt at concentration Cs to avoid dehydration during
gelling. Gelling can be accelerated by setting the closed vessel in an
incubator at 40°C (Figure 7d).
5. (a) For hanging drops. When the gel is set, pour carefully a volume V
of protein solution at concentration Cp on the gel surface taking
care not to touch it with the pipette (Figure 7e).
(b) For capillaries. Set again the capillaries (or droplets) in the closed
vessel. Diffusion of protein in the gel matrix and dehydration of
the liquid droplet occurs simultaneously so that the final result is a
gelled droplet with buffer, salt, and protein at concentrations CB,
Cs, and Cp respectively.
6. Use the droplets as in usual hanging drop or sitting drop techniques.
Seal the capillaries and proceed as for a batch technique, i.e. by
keeping them at constant temperature. One can also apply a regulated
temperature variation to increase supersaturation.
gel: 10 CG
protein: 2 Cp and CB
salt: 4 Cs
buffer: 5 CB
162
6: Crystalization in gels and related methods
Figure 8. One-step procedure for crystal growth in a CG agarose gel. The volume
proportions are given in order to have final concentrations of protein CP, salt CE, and
buffer CB, For calculating Xg, take Cg in %. Agarose sol uptake needs reverse pipetting.
163
M.-C. Robert et al.
Method
1. Prepare a crystallizing agent/agarose solution by thoroughly mixing
salt, buffer, water, and gel in the proportions indicated in Figure 8d.
This solution is kept at T > TG with the protein solution (Figure 8e).
2. Sample equal volumes of protein and crystallizing agent/agarose solu-
tions in an Eppendorf tube and gently mix with a Pipetman (Figure 8ft.
3. Use this preparation as you do for classical solution growth tech-
niques (see Chapter 4).
4. Decrease the temperature under TG to allow the gel to set.
165
M.-C. Robert et al.
process, being filled either with a chemical buffer, with a porous network or
with any other fluid such as water. Batch crystallization into capillary volumes
was used by Feher et al. (22) to illustrate diffusional transport in protein
crystallization. The use of capillaries for counter-diffusion methods dates
back to the work of Zeppezauer et al. (23), and others reviewed by Phillips
(24) (see also Chapter 5). It seems that these methods using capillaries were
designed with the aim of reaching the critical supersaturation for nucleation
very slowly, looking for a single nucleation event. Attempts to make use of
Otswald ripening processes were also considered (25). Unlike these, recent
studies have tried, starting from conditions far enough from equilibrium, to
search for multiple nucleation events under conditions progressively approach-
ing equilibrium. To illustrate the method we can use a simple technical im-
plementation, the gel acupuncture technique, stressing that our discussion
also applies for counter-diffusion arrangements using gelled protein solutions
and microgravity experiments.
3.2.3 Crystallization outside the gel: gel acupuncture method
The gel acupuncture method is based upon the properties of gels, which are
used to act as the mass transport medium for the precipitating agent and also
to hold capillaries containing the ungelled protein solution. The experimental
set-up is as simple as shown in Protocol 5 (see ref. 26 for specific recipes to
crystallize several proteins).
Method
1. Mix the sodium silicate as commercially supplied with four parts of
water. Mix slowly under continuous stirring 12.5 ml of this solution
with 10 ml of 1 M acetic acid. For this step, you can use a 50 ml vessel.
In a few hours the silica gel is set with a pH about 5.8-6.
2. Prepare the following solutions while the gel sets down:
(a) 20% (w/v) sodium chloride: pour 4.5 g of salt into 22.5 ml of water
and stir until complete dissolution.
(b) 100 mg/ml protein solution: weigh 100 mg of lysozyme into a
small tube (e.g. an Eppendorf tube) and pour 1 ml of
water. Stir gently until complete dissolution.
3. Fill the capillaries with the protein solution. Introduce one of the ends
of the capillary into the protein solution. You will see that the solution
166
6: Crystallization in gels and related methods
flows up by capillarity. Once it reaches a level 1 cm from the other end
of the capillary, remove it from the solution. The solution holds inside
the capillary. Then seal the upper end of the capillary with a small
piece of Plasticine. The next step is to punch the capillary in the gel
layer (but be sure that the gel is set!). Insert the capillary into the gel
about 0.5-1 cm, enough to maintain it straight.
4. Pour the solution of salt (22.5 ml) onto the gel layer and cover the
experiment with a large vessel turned upside down.
Figure 10. Protein crystallization by the gel acupuncture method. The diffusive path of
the salt towards the capillary filled with protein solution is illustrated as well as the
crystal size distribution obtained along the capillary. Picture at right shows an actual
ferritin crystallization experiment using the gel acupuncture method.
Figure 11. Rod-shaped crystal of tetragonal HEW lysozyme obtained by gel acupuncture
method. The isolated crystal was grown until it completely filled the capillary and then
growth continued at both ends of the cylinder. Capillary diameter 0.3 mm. Protein
concentration 100 mg/ml. Salt concentration 10% (w/v) pH 4.5.
In the search for single crystals by this method, large protein concentrations
are recommended to start with. From the point of view of classical crystal-
lization methods, this can be surprising. However, it should be realized that in
the get acupuncture method the precipitation system itself searches for the
best crystallization conditions. Thus the idea, when using large protein con-
centrations, is to trigger the nucteation of an initial amorphous precipitate in
the lower part of the capillary and then leave the system to search for optimal
growth conditions which typically occurs in a time seale of days. The crystal-
lizing agent concentration is another important variable. A very high initial
concentration will exhaust the protein in the capillary with amorphous pre-
cipitation, while a very low initial concentration will produce a batch-type
precipitation behaviour due to the small salt gradients inside the capillary.
Thus, an intermediate concentration, in the range known to precipitate the
protein, is recommended. Finally, the punctuation depth affects the waiting
time and the overall salt gradient inside the capillary. The suggested value of
8 mm has been experimentally found to be a good compromise between
waiting time (longer for higher punctuation depth values) and mechanical
stability of the capillary.
The gel acupuncture technique has been demonstrated for proteins of
different types and diverse molecular weights. To work properly, the tech-
nique must avoid sedimentation of the crystals, as well as the buoyancy driven
convection created as soon as the protein concentration falls in the lower part
of the capillary. It is recommended to use, as the reservoir to be filled with gel.
169
M.-C. Robert et al.
rectangular boxes made of two glass plates separated by a rubber frame hold
with clips (30, 31). With this simple arrangement, the capillaries can be
orientated perpendicular to the gravity field. Nevertheless, very large (up to
10 mm) crystals have already been obtained with this technique, showing a
high diffraction resolution limit and very low mosaicity (1.2 A and 9 arc
second, respectively, for tetragonal HEW lysozyme crystals) (32, 33). These
results are expected to be enhanced in future with in situ measurements
during the growth process, ensuring mechanical stability (34). Finally, use of
the technique for preparation of heavy-atom derivatives and search of crystal-
lization conditions by pH variation (a method advocated by McPherson) (35)
also seems promising.
5. Related methods
5.1 Microgravity
Space experiments share with growth in gels the ability to reduce buoyancy-
driven convection, to reduce impurity concentration on the crystal surface,
and to avoid sedimentation of crystals as well as the secondary nucleation of
3D protein clusters. In addition, the microgravity scenario removes the
plausible chemical interaction of the gel with the reactants used in the chem-
170
6: Crystallization in gels and related methods
ical protocol, including the protein itself. In short, microgravity conducted
experiments may be considered as 'clean' gel experiments.
As with gels, all techniques used in protein crystallization can be im-
plemented under microgravity. A diverse range of facilities, covering vapour
diffusion and liquid counter-diffusion techniques, are currently offered by
several Space Agencies to grow single protein crystals under microgravity
conditions (see refs 38-40 for a full description of facilities).
Because of the youth of microgravity science and the limited number of
opportunities to fly experiments, space crystallization of biological macro-
molecules still faces a number of technical and conceptual problems, which
need to be solved.
After ten years of microgravity-conducted experiments, some improve-
ments of crystal size and crystal quality have been reported (41-43, see ref. 38
for a review), and in a few cases reduced mosaicity was reported. Often,
however, the quality of the space-grown crystals, as evaluated by Wilson-type
plots or mosaicity measurements, does not show a dramatic increment of
quality, especially in terms of limit of resolution.
Current evaluation of space crystallization is basically performed by com-
paring space-grown crystals and crystals grown on earth using the same type
of reactor and the same crystallization conditions. Considering that the typical
dimensions of the reactors are large enough to allow density-driven and
thermal convection on earth, it is evident that space-grown crystals would be
expected to be of higher quality. In any case, what needs to be explained is
why in some cases the reverse result was found. In the future, space crystal-
lization has to face comparison with on-ground techniques emulating micro-
gravity conditions, that is crystal growth within gels and/or capillary volumes.
The similarity between the geometry and mass transport properties of the
gel acupuncture technique and the microgravity facilities is strong enough
(Figure 9) to permit an extrapolation of the above discussion on super-
saturation spatiotemporal evolution (see Section 3.2.3). The immediate
advantage of microgravity conducted experiments on capillary methods is
that, for the same path length, larger volumes of protein solutions can be
used, which could yield larger crystals. Unfortunately, the typical linear
dimensions in the direction of the diffusion path in the facilities currently
available are too short (15 mm for PCF or 8 mm for APCF) to exploit the
advantage of convection free counter-diffusion in large volumes. Therefore
the typical and useful spatial heterogeneity of the counter-diffusion tech-
niques is lost and only a limited set of the wide range of crystallization
conditions that could be reached is tested in practice. This enforces the use of
lower concentrations in all the experiments performed to date, converting the
free interface diffusion and dialysis experiments into batch experiments
because the characteristic time for the nucleation and transport processes are
similar (44, 45). It is clear therefore that the use of longer protein chambers
will permit the exploration of a larger set of local growth conditions.
171
M.-C. Robert et al.
An early criticism of space crystallization (46) was that the limited number
of opportunities to fly experiments makes it impossible to employ the 'trial
and error' methodology used so far on earth in the search for crystallization
conditions. However, this restriction has, in fact, been a major driving force
for the development of the current trend to rationalize protein crystal growth,
as the advance in the understanding of fundamental aspects of protein crystal-
lization (including the knowledge of the growth environment and crystal
quality evaluation) directly derived from space-induced research has been
substantial. Today, it seems evident that such a rationale will be obtained only
by coupling on-ground and space crystallization research. In particular, the
numerical simulation of mass transport and precipitation phenomena (27,
44-50), the development of growth techniques emulating microgravity con-
ditions on-ground, the characterization of nucleation phenomena (51, 52),
growth mechanisms, and surface kinetics (53) will be required in the future. In
addition, the search for a relationship between growth history and crystal
quality requires the improvement of the existing X-ray characterization tools
(54, 55).
Finally, to evaluate properly the future of protein crystallization in space, it
should be considered that beside producing crystals of improved perfection
(with reduced mosaicity), another current interest is basically motivated by
the need for high quality crystals larger than the size (tenths of millimetre)
required for structural studies. The crystallization of biological macro-
molecules will face in the future the optimization of the final crystal size for
purposes not only related to structural biology (e.g. neutron diffraction
measurements) but also the characterization of their physical properties
because their use for technological application is still an unexplored and
exciting field. To grow large macromolecular single crystals, microgravity
offers the best scenario without the limitations of capillary volumes, but first
we need to learn how to properly control impurity distribution and its effect
on the cessation of growth.
5.2 Hypergravity
Contrary to microgravity experimentation, kilogravity experimentation in
centrifuges is a rather unexplored field, although successful results were
reported as early as 1936 for tobacco mosaic virus (56).
In a review on this subject (57), Schlichta emphasized that gravity might be
regarded as a variable in crystal growth and material processing. Besides
convection and sedimentation which are drastically altered by increasing
gravity forces, one has also to consider the increased compressive and
shear stresses, as well as hydrostatic pressure (which can induce solubility
variations) (58).
All these factors influence directly or indirectly the growth parameters. For
example, due to forced sedimentation effect, supersaturated zones appear in
172
6: Crystallization in gels and related methods
an initially undersaturated solution, leading to crystal nucleation and growth
(59). However, the role played by the gravity field may promote additional
effects:
(a) Centrifugation can separate the molecules having different molecular
weights (foreign macromolecule, impurities) or different conformer.
(b) Centrifugation will progressively modify the spatial distribution of the
different populations of oligomeric species or aggregates or interacting
monomers contained in a supersaturated macromolecular solution. Thus,
one could set some local solution compositions suitable for nucleation.
The devices needed for hypergravity experiments are not necessarily
complicated: the crystal growth presented in ref. 59 required a centrifuge
currently available in any biochemistry laboratory.
More fundamental studies need ultracentrifuges equipped with observation
set-ups such as Schlieren optics. As a matter of fact, the development of
centrifugal crystal growth is related to the development of basic studies. A
current feeling is that crystal grown under kilogravity would suffer from
plastic deformation and would not diffract at very high resolution. This does
not seem valid, at least considering the few examples found in the literature
(57, 60-61).
References
1. Robert, M.-C. and Lefaucheux, F. (1988). /. Cryst. Growth, 90, 358.
2. Vekilov, P. G., Alexander, J. I. D., and Rosenberger, F. (1996). Phys. Rev. E, 54,
6650.
3. Miller, T. Y. and Carter, D. C. (1992). J. Cryst. Growth, 122, 306.
4. Sica, F., Demasi, D., Mazzarella, L., Zagari, A., Capasso, S., Pearl, L. H., et al.
(1994). Acta Cryst., D50, 508.
5. Vidal, O., Robert, M.-C., and Boue, F. (1998). J. Cryst. Growth, 192, 257.
6. Vidal, O., Robert, M.-C., and Boue, F. (1998). J. Cryst. Growth, 192, 271.
7. Her, R. K. (1979). The chemistry of silica. Wiley. Interscience, New York.
8. Ramzi, M. (1996). PhD Thesis, University de Strasbourg.
9. Vidal, O. (1997). PhD Thesis, Universite de Paris.
10. Henisch, H. K. (1988). Crystal growth in gels and Liesegang rings. Cambridge
University Press.
11. Garcia-Ruiz, J.-M., Gavira, J. A., Otalora, F., Guasch, A., and Coll, M. (1998).
Mat. Res. Bull., 33, 1593.
12. Vidal, O., Robert, M.-C., Arnoux, B., and Capelle, B. (1999). J. Cryst. Growth, 196,
559.
13. Bernard, Y., Degoy, S., Lefaucheux, F., and Robert, M.-C. (1994). Acta Cryst.,
D50, 504.
14. Cudney, B., Patel, S., and McPherson, A. (1994). Acta Cryst., D50, 479.
15. Provost, K. and Robert, M.-C. (1991). J. Cryst. Growth, 110, 258.
16. Thiessen, K. J. (1994). Acta Cryst., D50, 491.
173
M.-C. Robert et al.
17. McPherson, A., Malkin, A. J., and Kuznetsov, Y. G. (1995). Structure, 3, 759.
18. Georgalis, Y., Umbach, P., Raptis, J., Saenger, W. (1997). Acta Cryst., D53, 691
and 703.
19. Vidal, O., Bernard, Y., Robert, M.-C., and Lefaucheux, F. (1996). J. Cryst.
Growth, 168, 40.
20. Provost, K. and Robert, M.-C. (1995). / Cryst. Growth, 156, 112.
21. Hirschler, J. and Fontecilla-Camps, J. C. (1996). Acta Cryst., D52, 806.
22. Feher, G. and Kam, Z. (1985). In Methods in enzymology (eds H. W. Wyckoff,
C. H. W. Hirs, and S. N. Timasheff), Academic Press, Inc., London, Vol. 114,
p. 77.
23. Zeppezauer, M., Eklund, H., and Zeppezauer, E. S. (1968). Arch. Biochem. Anal.,
126, 564.
24. Phillips, G. N. (1985). In Methods in enzymology (eds. H. W. Wyckoff, C. H. W.
Hirs, and S. N. Timasheff) Academic Press, Inc., London. Vol. 114, p. 128.
25. Weber, P. (1991). Adv. Protein Chem., 41, 16.
26. Garci'a-Ruiz, J.-M., Moreno, A., Otalora, F., Vieoma, C., Rondon, D., and
Zautscher, F. (1998). J. Chem. Educ., 75, 442.
27. Otalora, F. and Garcfa-Ruiz, J.-M. (1996). J. Cryst. Growth, 169, 361.
28. Moreno, A., Rondon, D., and Garcia-Ruiz, J.-M. (1996). J. Cryst. Growth, 166,
919.
29. Garci'a-Ruiz, J.-M. and Moreno, A. (1994). Acta Cryst., D50, 484.
30. Garci'a-Ruiz, J.-M., Lopez Martinez, C., and Martin-Vivaldi Caballero, J. L.
(1985). Cryst. Res. Techn., 20, 1615.
31. Garcia-Ruiz, J.-M. and Moreno, A. (1997). J. Cryst. Growth, 178, 393.
32. Otalora, F., Garcia-Ruiz, J.-M., and Moreno, A. (1996). J. Cryst. Growth, 168, 93.
33. Otalora, F., Capelle, B., Ducruix, A., and Garcia-Ruiz, J.-M. (1999). Acta Cryst.,
D55, 644.
34. Otalora, F., Gavira, J. A., Capelle, B., and Garcia-Ruiz, J. M. (1999). Acta Cryst.,
D55, 650.
35. McPherson, A. (1985). In Methods in enzymology (eds. H. W. Wyckoff, C. H. W.
Hirs, and S. N. Timasheff), Academic Press, Inc., London. Vol. 114, 125.
36. DeLucas, L. J., Long, M. M., Moor, K. M., Rosenblum, W. M., Bray, T. L., Smith,
C., et al (1994). J. Cryst. Growth, 135, 183.
37. Lorber, B., Sauter, C., Ng, J. D., Zhu, D. W., Giege, R., Vidal, O., Robert, M.-C,
Capelle, B, et al. (1999). J. Cryst. Growth, in press.
38. Giege, R, Drenth, J., Ducruix, A, McPherson, A, and Saenger, W. (1995). Prog.
Cryst. Growth Charact., 30, 237.
39. McPherson, A. (1996). Cryst. Rev., 6, 157.
40. Pletser, V, Stapelmann, J, Potthast, L., and Bosch, R. (1999). J. Cryst. Growth,
196, 638.
41. Koszelak, S., Day, J, Leja, C, Cudney, R., and McPherson, A. (1995). Biophys. J.,
69, 13.
42. Ng, J. D, Lorber, B., Giege, R., Koszelak, S, Day, J, Greenwood, A, et al.
(1997). Acta Cryst., D53, 724.
43. Snell, E. H, Weisgerber, S., and Helliwell, J. R. (1995). Acta Cryst., D51, 1099.
44. Otalora, F. and Garcia-Ruiz, J.-M. (1997). J. Cryst. Growth, 182, 141.
45. Garcia-Ruiz, J.-M. and Otalora, F. (1997). J. Cryst. Growth, 182, 155.
46. Leberman, R. (1985). Science, 230, 370.
174
6: Crystallization in gels and related methods
47. Wagner, G. and Linhardt, R. (1991). J. Cryst. Growth, 110, 114.
48. Fowlis, W. W., DeLucas, L. J., Twigg, P. J., Howard, S. B., Meehan, E. J., and
Baird, J. K. (1988). J. Cryst. Growth, 90, 117.
49. Huo, C, Ge, P., Xu, Z., and Zhu, Z. (1991). J. Cryst. Growth, 114, 486.
50. Savino, R. and Monti, R. (1996). J. Cryst. Growth, 165,308.
51. Malkin, A. J., Cheung, J., and McPherson, A. (1993). J. Cryst. Growth, 126, 544.
52. Georgalis, J., Schiller, P., Frank, M., Soumpasis, D., and Saenger, W. (1995). Adv.
Colloid Interface Sci., 58, 57.
53. Land, T. A., Malkin, A. J., Kuznetsov, Yu. G., McPherson, A., and Yoreo, J. J.
(1995). Phys. Rev. Lett., 75, 2774.
54. Helliwell, J. R. (1988). J. Cryst. Growth, 90, 259.
55. Chayen, N. E., Boggon, T. J., Casseta, A., Deacon, A. Gleichmann, T., Habash, J.
et al. (1996). Q. Rev. Biophys., 29, 227.
56. Wyckoff, R. W. G. and Corey, R. B. (1936). Science, 84, 513.
57. Schlichta, P. J. (1992). J. Cryst. Growth, 119, 1.
58. Lorber, B., Jenner, G., and Giege, R. (1996). J Cryst. Growth, 158, 103.
59. Pitts, J. E. (1992). Nature, 355, 117.
60. Behlke, J. and Knespel, A. (1996). J. Cryst. Growth, 158, 388.
61. Barynin, V. V. and Melik-Adamyan, V. R. (1982). Sov. Phys. Crystallogr., 27, 588.
175
This page intentionally left blank
7
Seeding techniques
E. A. STURA
1. Introduction
A seed provides a template for the assembly of molecules to form a crystal
with the same characteristics as the crystal from which it originated. Seeding
has often been used as a method of last resort, rather than a standard practice.
Recently, these techniques have gained popularity, in particular, macro-
seeding, used to enlarge the size of crystals. Seeding has many more applica-
tions, and the use of seeding in crystallization can simplify the task of the
crystallographer even when crystals can be obtained without it. We will
explore the various seeding techniques, and their applications, in the growth
of large single crystals and the methods by which we may attempt to obtain
crystals that diffract to higher resolution.
Crystallogenesis can be divided into two separate phases. The first being
the screening of crystallization conditions to obtain the first crystals, the
second consisting of the optimization of these conditions to improve crystal
size and quality. Seeding can be used advantageously in both these situations.
The first stage in crystallogenesis consists of the discovery of initial crystals,
crystalline aggregates, or microcrystalline precipitate. This may result from a
standardized screening method (1, 2), a systematic method (3), an incomplete
factorial search (see Chapter 4 and refs 4 and 5), or by extensive screening of
many conditions. This may be bypassed by starting with seeds from crystals of
a related molecule that has been previously crystallized. Molecules that have
been obtained by genetic or molecular engineering of a previously crystallized
macromolecule fall in this category. This method is termed cross-seeding. It
has been used to obtain crystals of pig aspartate aminotransferase starting
with crystal from the chicken enzyme (6) and between native and complexed
Fab molecules (7).
Whatever the method used to obtain the initial crystals, seeding may pro-
vide a fast and effective way to facilitate the optimization of growth conditions
without the uncertainty which is intrinsic in the process of spontaneous
nucleation. The streak seeding technique can be used to carry out a search
quickly and efficiently over a wide range of growth conditions. Later the use
E. A. Stum
of macroseeding and microseeding methods can be used to grow large crystals
with a high degree of reproducibility.
2. Seeding
2.1 Supersaturation and nucleation
Details on the use of precipitants, together with general and theoretical
considerations, and practical methods for macromolecule crystallization, are
to be found in other chapters and in other publications (8-12). Here we will
consider some of the aspects of crystallization that directly affect the
application of seeding techniques. It is useful to separate the events leading to
the spontaneous formation of a crystal nucleus and those conditions that
allow a crystal or nucleus to grow. While both events depend on the degree of
supersaturation of the protein, the physical processes involved are very dif-
ferent, and the degree of supersaturation required for nucleation is generally
higher than that required for growth onto an existing crystal plane. In the case
of spontaneous nucleation a new seed must be generated while other events
are taking place and is driven by the requirement to lower the free energy of
the supersaturated state. These other events involve the aggregation of
molecules into various phases. During aggregation, reversible and irreversible
processes are at work simultaneously. The formation of ordered nuclei may
be competing for protein with irreversible processes that produce amorphous
aggregates (such as precipitates and protein skins) and hence constantly lower
the degree of supersaturation of the macromolecule. As supersaturation is
decreased the chance of forming a stable nucleus is reduced. Since the
occurrence of spontaneous nucleation depends on the relative rates at which
these various competing events take place, crystals might never form even
under conditions which might otherwise support crystal growth. The principle
that there is a lower energy requirement in adding to an existing crystal
surface than in creating a new nucleus has important consequences. If the
inverse were true, the protein would partition into a very large number of
small nuclei and large crystals would never grow. Instead, in many cases it is
possible to grow large crystals in the absence of seeding. Since spontaneous
nucleation is a statistical phenomenon, whose probability increases with
increasing degree of supersaturation, the nucleation and growth of crystals is
a process with negative feedback. As a nucleus is formed its growth reduces
the degree of supersaturation of the solution, and hence decreases the prob-
ability that other nuclei will form. To take advantage of this, supersaturation
must be achieved slowly. The degree of supersaturation should be just
sufficient to obtain a small number of nucleation centres. With the proper
choice of crystallization conditions and good control over the environmental
factors, it is often possible to fulfil all of the above conditions. Determining
the appropriate conditions can require many crystallization trials and is con-
sequently time-consuming involving the use of many milligrams of macro-
178
7: Seeding techniques
molecule and other materials. Seeding can be used efficiently and effectively
during the crystallization trials to minimize the quantity of protein required
for this analysis.
3. Crystallization procedures
Crystallization by vapour diffusion is a relatively simple technique (see
Chapter 5). This section deals with the establishment of crystallization
procedures which are suited to the application of seeding.
i. Temperature
To achieve success in crystallization and seeding it is important to control
the overall environment of the set-up. This must include temperature.
Temperature regulation can be achieved by the combined use of a sitting drop
vapour diffusion set-up using a glass pedestal and a constant temperature
incubator. The sitting drop environment provides better temperature control
than its hanging drop counterpart because of heat conduction between the
reservoir solution and the protein solution in the inverted glass pot (see
Figure 1). The problem of condensation on the glass coverslip, caused by
temperature gradients and convection currents in the sealed set-up, are less
likely to affect a sitting drop experiment where the protein drop is situated
close to the surface of the reservoir. In contrast, in the hanging drop environ-
ment, the drop is effectively in thermal contact with the outside air. The thin
coverglass absorbs the heat of condensation and dissipates it to the outside.
Short-lived changes in temperature, such as opening the door of the constant
temperature incubator containing the experiment, will rapidly vary the
temperature of the hanging drop but not that of the reservoir because of its
higher heat capacity. During a rise in temperature, vapour will distil away
from the drop, increasing the degree of supersaturation, which may result in a
shower of microcrystals. This will be more common for crystals that are grown
at high salt concentrations. When the temperature decreases more vapour
condenses onto the drop diluting the protein solution. It is not uncommon in
low salt, or in crystallization trials using hanging drop vapour diffusion under
low PEG concentrations, to observe an increase rather than a decrease in the
volume of the protein-precipitant drop. The use of the sitting drop method
reduces these problems.
Figure 1. Schematic illustration of the sitting drop vapour diffusion plates. (A) A glass cup
is held between long sized forceps over a Bunsen burner with the closed end down
towards the flame. (B) The cup is heated until the glass becomes soft. (C) When the bottom
is malleable the pot is inverted and a depression is made in it with a glass plunger. (D) A
ring of silicone vacuum grease is placed in the bottom of each of the wells of the microtitre
plate. Pots are placed onto top of the ring and pressed down. (E) Each depression is
182
7; Seeding techniques
siliconized, washed repeatedly with distilled water, and the coat baked in an oven. The
siliconized pots can now be placed in each of the wells in the multiwell cluster on top of
the silicone grease ring which holds them in position. (G) The rim of the individual well
is smeared with petroleum jelly to seal the well once the coverglass is placed on top. (H)
The precipitant solution is placed around the inverted pot, the protein in the depression
and the desired amount of precipitant mixed with it before sealing the experiment.
183
E. A. Stura
seeding environment. Such features can greatly enhance growth of quality
single crystals. The tray consists of a 24-well Costar tissue culture plate from
Hampton Research combined with 0.6 ml glass cups from Fisher Scientific.
The wells are significantly smaller than those in the more commonly used
Linbro plates, which use 22 mm circular coverglass slips instead of the 18 mm
required for the Costar plates. The plates are made as described in Protocol 1
and illustrated in Figure 1.
Method
1. Make a depression in the bottom of the 0.6 ml cups by heating over a
Bunsen flame and pressing down on the cylindrical base of each of the
inverted cups with a rounded-end glass plunger.
2. Hold the resulting cup in place at the centre of each well of the tissue
culture plate with Corning silicone vacuum grease. The open end of
the cup sits on the bottom and the depression in the cup is at the
volumetric centre of each individual well (sitting drop rods with
depressions, made by Perpetual System Corporation, and micro-
bridges are available from Hampton Research).
3. Siliconize the cavity created by the depression. Up to 100 ul of protein-
precipitant mixture can be used with this set-up. The reservoir
solution, typically 1 ml, occupies an annulus around the inverted cup.
4. Smear the edges of each well in the Costar tissue culture plate (avail-
able from Hampton Research) with petroleum jelly. Silicone grease
can be used instead of petroleum jelly to give a better seal, although
this will make covers harder to remove for seeding.
5. Place a 18 mm round microscope coverglass on top of each well
ensuring that an airtight seal is achieved.
Since this method is similar in many ways to the hanging drop method,
crystallization conditions determined for hanging drop experiments require
little modification for implementation with sitting drop vapour diffusion.
Another added advantage is that larger volumes can be used. Other sitting
drop methods provide similar advantages although glass pedestals are
recommended for temperature stability.
184
7: Seeding techniques
Figure 2. Schematic drawing of the stages in the application of the streak seeding
technique for analytical seeding. (A) A probe made from a cut pipette tip mounted on a
wooden shaft. On the end of the tip a short segment of an animal whisker is attached
with molten wax. The probe so constructed is used to pick up seeds from an existing
crystal, precipitate, or other ordered aggregate by simply touching it and displacing
seeds from it. (B) The seeds remain attached to the whisker and can be transferred to a
pre-equilibrated drop by running the end of the probe across the drop. Some seeds are
deposited along the path, where they will either grow into large crystals or dissolve into
the solution. (C) Growth of crystals along the streak line indicates that the conditions may
be suitable for the application of other techniques such as micro- or macroseeding. Self-
nucleated crystals will appear away from the streak line.
186
7: Seeding techniques
of supersaturation which allows for sufficient crystal growth without self-
nucleation is determined experimentally by observing the growth (or lack of
growth) of seeded crystals along the streak line (Figure 2C).
Protocol 2. Microseeding
189
E. A. Stura
Protocol 2. Continued
crystals obtained, inconsistent with the a tenfold dilution, extra
precipitant should be added to the stabilizing solution. If nucleation is
independent of the dilutions, buffer should be added to the seed
solution to reduce the precipitant concentration. Also test the drops by
streak seeding from a crystal (Protocol 3) to ensure that the reservoir
solution is appropriate for seeding.
Method
1. By changing the angle at which the whisker is drawn out of the we can
affect the size and number of seeds loaded onto the fibre. This should
be kept constant for reproducibility. The whisker is lifted vertically
upwards, maintaining it perpendicular to the drop's surface to mini-
mize seed retention, seeds are scooped up to maximize size and the
number of seeds picked up.
2. Pre-equilibrate the drops before seeding under conditions previously
determined analytically by streak seeding.
3. Streak seed subsequent drops without loading the probe with new
seeds to achieve seed dilution. To obtain greater dilutions the probe
190
7: Seeding techniques
can be dipped in and out of the reservoir in between streaks, to allow
some seeds to drop into the precipitant solution.
4. Reduce the time the probe spends in the air by opening all the
chambers to be seeded just before picking up the seeds. Streak
the drops sequentially, as speed is important to prevent drying of the
solution on the fibre.
5. Cover all the chambers without delay to reduce evaporation from the
seeded drops.
The seed stock and seed dilutions prepared as previously described, can be
used in streak seeding reliably by dipping the whisker into each of the diluted
solutions including the seed stock and applying the seeds to new drops
(Figure 3). It is common to start by dipping the probe in the most dilute
solution first, streaking one drop, then progressing up the dilution series to the
seed stock. Typically the results are analysed two days to one week after
streaking.
4.2 Macroseeding
In macroseeding a single crystal is introduced into a suitably pre-equilibrated
solution. Single prismatic crystals, which are free from twinning or any other
crystallites, are most suitable for this technique. As in other seeding protocols
it is important to take steps to maintain constant conditions, as even a slight
dehydration of the drop being seeded could temporarily change the state of
supersaturation and induce unwanted nucleation. Performing the experiment
in a very humid environment and by using large drops can reduce dehydra-
tion. A beaker with a filter paper cylinder soaked in distilled water is such an
environment. However it is more practical to use sitting drop multiwell trays,
which have been used with high success in our laboratory. Macroseeding is
done under a dissecting microscope where the small amount of heat gener-
ated from the microscope stage light bulb may actually be slightly beneficial in
increasing the humidity level around the drop. The heat raises the tempera-
ture of the reservoir faster than that of the drop increasing the rate of evapor-
ation from the reservoir, and counteracting evaporation into the room. Seeds
are washed in a slightly dissolving solution to remove the top layer of protein,
which contains possible defects, from the surface of the seed without causing
excessive etching or cracking. They are then transferred to a stabilizing
solution to re-equilibrate the crystals (Figure 4). Older seeds benefit the most
from this treatment, whereas freshly grown crystals may be put directly
through a series of washes in stabilizing solution. From the final wash solution
each seed is then transferred to the protein-precipitant drop to be seeded
(Figure 4).
191
Figure 3, Diagrammatic illustration of the steps involved in microseeding. (A) Crystals of
good morphology are crushed in a glass tissue homogenizer. The resulting seeds are
washed into the bottom of the tube and stored in a test-tube. (B) The seed stock is diluted
to produce a dilution series. (C) Seeds can be picked up from the diluted solutions by
using a probe, or precipitant can be added to these, so that they may be mixed with
protein solution (D). The wells are seated by replacing the coverglass and the seeds are
allowed to grow for several days.
192
7: Seeding techniques
Figure 4. Illustration of the steps involved in macroseeding. (A) A single crystal is picked
up from a drop. Crystals should be of good morphology and free from defects. (B) A
series of washes is performed by repeatedly transferring the crystal from one depression
to another, taking care not to damage the seed. (C) The seed is finally transferred to a pre-
equilibrated drop for further enlargement.
193
E. A. Stura
Method
1. Connect a glass or quartz capillary to a 1 ml glass syringe with a short
piece of rubber tubing such as c-flex (Fisher, 14-169-5c) which gives an
excellent seal.
2. Snap open the end of the capillary with tweezers or scissors, and
siliconize if the experimental situation can benefit from diminished
adhesion of the solution to the glass capillary. After siliconizing it
should be extensively washed.
3. Pick up the crystals under a dissecting microscope, using a magnifica-
tion of x 10 to x 100.
Crystals from hanging drops should first be washed into a secondary
vessel; in a sitting drop vapour diffusion set-up it is possible to do
this directly from the depression in which the crystals have been
growing.
After the coverglass sealing the vapour diffusion chamber is lifted off,
the tip of the capillary is inserted into the drop and a crystal is drawn
into the capillary.
If the crystals are adhering to the well, by withdrawing liquid from the
drop and gently ejecting it onto a chosen crystal, it is often possible to
dislodge the crystal. Unfortunately some crystals have severe
adhesion problems and cannot be dislodged without breaking them.
To reduce this problem, the depressions in glass pots should be
coated with a thin film of Corning vacuum silicone grease before the
protein-precipitant drop is added. The seeds will remain suspended on
top of the grease, and the final crystals are mounted for X-ray
diffraction use without the recurrence of this problem.
4. Once in the capillary the crystal is brought to the middle, and then
allowed to sink and adhere sufficiently to the inside wall of the
capillary so that the liquid can be moved over the crystal.
194
7; Seeding techniques
(a) For crystals that fail to sink to the bottom and adhere to the
capillary wall, a long hair or whisker may be wedged against the
crystal to stop movement while liquid is drawn out.
(b) For soft crystals there is the danger that during this procedure
microseeds may be dislodged from them by the hair with obvious
consequences. A series of washes will minimize the number of
microseeds that will be transferred but not eliminate the risk that
one or more will still be present in the solution which is
transferred together with the crystal.
5. Once the crystal has been separated from the bulk of the mother
liquor, the hair withdrawn and the mother liquor ejected from the
capillary and returned to the original drop, the crystal should remain in
a small pool of liquid inside the capillary. Removing more of the
remaining solution from around the crystal may help diminish the
number of microseeds and aggregated material, and since the
solution may now be at a higher precipitant concentration due to
evaporation during the handling, transferring less of this solution may
avoid creating conditions which are unsuited to the seeding.
Method
1. Fill four depressions of a multiwell sitting drop plate with about 100 ul
of solution from the reservoir where the seeds originated (or prepare a
solution identical to this).
2. Fill the reservoirs around these drops with 1.5 ml of distilled water to
maintain a high degree of moisture around the solution (Figure 4).
3. The crystal is repeatedly transferred and picked up from each of these
stabilizing solution drops until finally, it is picked up into the capillary.
Because the addition of stabilizing solution to the new drop would
unnecessarily modify the equilibrium, or dilute the equilibrated
protein-precipitant solution, it is best to minimize the amount of liquid
that remains around the crystal.
195
E. A. Stura
Protocol 5. Continued
4. Remove the excess liquid with a small thin strip of filter paper or a
very thin capillary.
5. Resuspend the crystal in new mother liquor drawn in from the drop to
be seeded and return it to the well for equilibration and further growth.
Alternatively, after the series of washes, the crystal is allowed to sink
towards the open end of the capillary, so that when the capillary
touches the solution of the drop being seeded, the crystal falls directly
into this solution with little transfer of wash solution.
5. Heterogeneous seeding
The principle that there is a lower energy requirement in adding to an existing
surface than in creating a new nucleus (Section 1.2) holds for many surfaces.
Such aggregation onto surfaces may be considered more of a problem than an
advantage. However, regular surfaces may offer a charge distribution pattern
which is complementary to a possible protein layer and could provide a
suitable starting point for the nucleation of new crystals. The work of
McPherson (21) with various inorganic minerals provides strong support for
the idea that regular planes are able to catalyse the nucleation of crystals of
macromolecules, even if the lattice dimensions of the crystalline minerals
differ from those of the resulting protein crystals. In these studies, nucleation
occurred preferentially on the mineral substrate at a lower degree of super-
saturation than was required for the same crystals to nucleate in the absence
of the minerals. Crystals of related macromolecules can also be used to induce
nucleation of proteins; the resulting crystals may maintain some, but not all,
of the lattice dimensions or symmetry axes of the initial seeds. In such cases,
196
7: Seeding techniques
where the protein in the crystals from which the seeds are obtained is related
to the protein in the solution being seeded, the operation is termed cross-
seeding. When crystals of the same macromolecule are used to induce a
related crystal form under different crystallization conditions, an epitaxial
jump (by analogy with quantum jumps) has been achieved.
5.1 Cross-seeding
5.1.1 Cross-seeding between Fab-peptide complexes
In X-type light chain dimers, the dimers pack so as to form an infinite B-sheet
maintaining one cell dimension in common, 72.4 (± 0.2) A, along one of the
21 axes (22). Such packing in preferred planes for certain classes of protein
molecules, may indeed provide suitable surfaces for nucleation for other
members of that class. A similar observation was made in the course of our
work with different anti-haemagglutinin monoclonal Fab-peptide complex
crystals, where it was noticed that these have a common crystal lattice plane
with cell dimensions 73.0 (± 1.0) A along the 21 axis and 66.4 (± 2.5) A along
one of the other axes (19, 23, 24).
Within the description of the three-dimensional structure of the complex of
Fab 26/9 that recognizes the same six residue epitope of an immunogenic
peptide from influenza virus haemagglutinin (HA1 75-110) as Fab 17/9, it was
possible to understand the hierarchy in the crystal contacts responsible for the
differences and similarities between these crystal forms (25). In brief, 26/9 and
17/9 antibodies are very similar, but their interaction with the peptide are
slightly different. Structural and sequence analysis suggests that amino acid
differences near the peptide binding site are responsible for altering slightly
the specificity of 26/9 for three peptide residues. Since the peptide is essential
for one of the crystal interactions, we can understand the influence of peptide
length on the crystallization and the similarity in crystallization between these
antibodies. Cross-seeding, using the streak seeding method can bridge the gap
between the various peptide complexes of 26/9. Initial crystals were obtained
by spontaneous nucleation with a nine residue peptide (HA1 100-108). The
quality of these crystals was improved by using the streak seeding technique
as a microseeding method. Seeds from these crystals were used to search for
growth conditions of complex crystals of Fab 26/9 with longer peptides, for
which no conditions for spontaneous nucleation had been found. Both the 13-
mer (HA1 98-110) and the 23-mer (HA1 88-110) peptide-Fab mixtures
responded positively to the seeding. Seeds obtained from this cross-seeding
were used to seed repeat experiments to dilute out the effect of the hetero-
geneous seeds and optimize the crystallization conditions for the new
complexes. The crystals obtained from this second seeding diffract to 2.5 A
resolution.
A third anti-peptide antibody 21/8 also belonging to this same panel but
their heavy chains belong to different classes; Fab 21/8 is derived from an
197
E. A. Stum
IgG
2b, while 26/9 is cleaved from an IgG2a. Nevertheless, seeds from Fab
26/9-13-mer complex crystals have been used to induce crystallization in
solutions of Fab 21/8-13-mer mixtures, under identical crystallization con-
ditions. The Fab 21/8-13-mer crystals first obtained by the streak seeding
experiment (Figure 5D) were very thin needles and although optimization of
the conditions resulted in better crystals, the real breakthrough was achieved
from the refined conditions which yielded large crystals of an unrelated form
by spontaneous nucleation.
Cross-seeding does not need to be carried out by streak seeding as, for
example, large crystals from the chicken mitochondrial aspartate amino-
transferase were used as seeds in the cross-seeding of the pig enzyme (6)
(initial cross-seeded crystals of the chicken enzyme were badly twinned but
were improved to X-ray quality in a second cycle of macroseeding). However
streak seeding can substantially increase the speed with which crystals can be
obtained.
5.1.2 Cross-seeding from native Fab to Fab-peptide complex
Anti-peptide Fab 50.1, that recognizes an epitope of the gp120 surface glyco-
protein of HIV-1, is an example where native crystals were used to seed the
Fab-peptide complexes. The native crystals can be grown in three different
morphologies. Spontaneous nucleation has not been observed for any of the
peptide complexes tested. The peptide lengths vary from 13-40 residues.
Crystals of the Fab-13-mer complex have now been obtained by streak
seeding with the native Fab crystals in 12-24% PEG 10000 pH 5-8 (26). In
the Fab-peptide solution most of the protein is found partitioned in a gel
phase covering the bottom of the drop. On addition of native seeds, crystals
grow by acquiring protein from the gel phase surrounding the seeds. The
morphology of the Fab-peptide complex crystals differs substantially from
that of native crystals (Figure 5E). It is also worth noting that while the
crystals of the native Fab are very mosaic, data have been collected on well
198
7: Seeding techniques
ordered crystals of the seeded complex that diffract to better than 2.8 A
(Figure 5F) sufficient for X-ray structure determination (27).
199
E. A. Stura
which serine 221, the active serine, of the bacterial protease (28-30), is con-
verted into a selenolcysteine (31) is one such example. Even after extensive
efforts to better purify the engineered enzyme (32) crystals could only be
obtained by cross-seeding from crystals of the commercially available native
subtilisin. Although the quality of these seed crystals, obtained from com-
mercial grade enzyme was rather poor, the crystals of the selenolsubtilisin
obtained from the cross-seeding experiment were of good morphology and
size. The crystallization conditions under which the cross-seeding was done
were similar to those for the native subtilisin. After further optimization it
was possible to obtain good quality crystals, which were used for the structure
determination of the modified enzyme (32). Under the optimized conditions
some preparations nucleated spontaneously.
201
E. A. Stura
Protocol 6. Continued
3. Check for the development of a line. If a line develops use the crystals
obtained to seed other drops. If growth is slow, increase the
precipitant concentration.
6. Crystallization of complexes
6.1 Considerations in the crystallization of complexes
When crystallizing complexes, such as a receptor-ligand, an enzyme-
inhibitor, or an Fab-antigen complexes, it is important to consider the
resultant heterogeneity of the system. Both members of the complex will be
somewhat heterogeneous, and the resulting mixture will be composed of
202
7: Seeding techniques
complexed and uncomplexed molecules in different ratios depending on the
molar ratio of the two molecules in the solution and the dissociation constant.
To reduce the heterogeneity it is important to optimize the number of
complexed versus the uncomplexed molecules.
For protein complexes with small ligands it is possible to increase the
number of complexed protein molecules by adding an excess of ligand.
Theoretically, the excess ligand that is necessary to achieve the desired ratio
of bound to unbound may be calculated from the dissociation constant if
known. In practice it is best to set up experiments at different protein:ligand
ratios, typically 1:1 to 1:20. Larger excesses are generally unnecessary and
may even inhibit crystal growth.
Co-crystallization of enzyme substrate complexes presents a different level
of complexity. Catalysis of the substrate into product will result in a mixture
of free enzyme, enzyme-product complex, and enzyme-substrate complex.
Such experiments are best attempted by using non-productive substrate
analogues, inhibitors that mimic the transition state, and undissociable end-
products. Triphosphate nucleotides, such as ATP, are easily hydrolysed,
hence, non-hydrolysable analogues such AMP-PNP, adenosine diphosphate
y-S (ADPyS), and their analogous guanosine derivatives are now commonly
used for crystallization instead of ATP or GTP. Vanadate, molybdate, and
tungstate are commonly used as phosphate mimics (36).
When the complex consists of two or more macromolecules of comparable
size the addition of an excess of one increases rather than decreases the
heterogeneity of the system. In such cases, if the affinity between the
macromolecules is 108 or better, the complex can be purified, otherwise it is
best to mix the macromolecules in the appropriate stoichiometric ratio. When
the stoichiometry of the system is not known it is best to set up the
crystallization of the complex at different receptor:ligand ratios. Uncom-
plexed molecules are likely to adhere to the lattice of the complex crystal, and
interfere with the growth of such crystals. In order for the crystal to continue
its growth without defects the unbound molecule must either become com-
plexed while still maintaining its lattice contacts, else it must break all lattice
bonds and diffuse away from the crystal surface to be replaced by a
complexed molecule. The energy involved in each of the lattice interactions
that must be broken will determine the inhibitory effect of the uncomplexed
molecule with respect to the growth of the complex crystal. Assuming that the
number of lattice bonds is proportional to the surface area of the molecule,
we expect that the inhibiting effect of an uncomplexed molecule will be
roughly proportional to molecular weight. Hence a small excess of the smaller
molecule is expected to be less damaging to the crystallization of the complex
than an excess of the larger molecule. When the crystallization proceeds
slowly, the crystal is less likely to incorporate unbound molecules as defects. If
we consider the deleterious effect of incorporating a defect into the lattice the
203
E. A. Stum
absence of the larger molecule will carry a greater energy penalty, and hence
it will be less likely to occur. Therefore, it is better to have a slight excess of
the smaller macromolecule when growing complex crystals, except during the
initial search, when suitable growth conditions have not been established, and
the most ordered nucleus is likely to occur with a slight excess of the larger
macromolecule.
When the affinity between the molecules in the complex is not high, and the
off rate is substantial, we must screen for possible crystallization conditions
where the relative solubility of the complex is lower than that of the
uncomplexed molecules. It is also important to use high concentrations of
both macromolecules to push the equilibrium in favour of the complex,
although the use of lower concentrations will allow for a larger number of
trials. The number of trials can be minimized by using a screening approach as
described in ref. 3. When crystals are obtained, streak seeding can be used to
determine whether the complex or either macromolecule has crystallized.
Method
1. Separate several crystals from the mother liquor from which they have
been growing, and wash them to remove residual mother liquor. The
procedure described for the handling of crystals for macroseeding
may be used here. Alternatively, the crystals can be lifted from the
drop using the same probe used for steak seeding or a loop as for
cryo-crystallography (Chapter 13).
2. If a skin has formed first remove the skin by running the probe around
the drop. Skins are often correlated with oligomerization through the
formation of disulfides (it is advisable to analyse the sample under
both reducing and non-reducing conditions). Place the probe under
the crystal and lift it out from the solution. Repeated attempts are
generally needed.
3. Dissolve the crystals in distilled water typically with a final volume of a
few microlitres. If a probe is being used, just touch the end of the
probe with the crystal on top and the crystal will drop into the water
and dissolve. More crystals can then be picked up if the crystals are
small.
4. Check under a microscope that the crystals have dissolved. Crystals
that do not dissolve under these conditions may have to be dissolved
under more acidic, more basic, or higher salt conditions, or by adding
urea or SDS directly to the crystals. When using urea or high salt to
dissolve the crystals, the urea or the salt will have to be dialysed out
before running the polyacrylamide gel. Silver staining of the gel may
be needed if the crystals are small.
5. Run the solutions used for crystallization in one of the lanes. Some
oligomerization will occur as a result of boiling in SDS under non-
reducing conditions. This should be more accentuated for the control,
which will be at higher protein concentration than the dissolved
crystals.
The comparison between the control and the dissolved crystals should be
able to determine the macromolecular content of the crystals.
205
E. A. Stum
7. Concluding remarks
The application of seeding methods in macromolecular crystallization has
proven invaluable for obtaining high resolution X-ray quality crystals when
conventional methods have failed. It provides a means of analysing many
conditions without requiring large amounts of protein solution. Streak seed-
ing is particularly valuable as it provides a fast method of analytical seeding
with easy visualization of the results. Cross-seeding is a powerful tool to
crystallize a given protein with seeds from a related protein. The application
of micro- and macroseeding methods can result in the production of large
single crystals for X-ray structure determination. Without such methods many
projects would not have been viable.
206
7: Seeding techniques
Acknowledgements
The contribution of Dr Ian Wilson to the editing of the previous version as
well as the support provided through his grants by National Institutes of
Health Grants AI-23498, GM-38794, and GM-38419 is here acknowledged.
The French Atomic Energy Commission (CEA) has provided support for the
revisions.
References
1. Stura, E. A., Nemerow, G. R., and Wilson, I. A. (1992). /. Cryst. Growth, 122, 273.
2. Jancarik, J. and Kim, S.-H. (1991). J. Appl. Cryst., 24,409.
3. Stura, E. A., Satterthwait, A. C, Calvo, J. C, Kaslow, D. C., and Wilson, I. A.
(1994). Acta Cryst, D50, 408.
4. Carter, C. W., Jr. and Carter, C. W. (1979). J. Biol Chem., 254, 12219.
5. Betts, L., Frick, L., Wolfenden, R., and Carter, C. W., Jr. (1989). J Biol. Chem.,
264, 6737.
6. Eichele, G., Ford, G. C., and Jansonius, J. N. (1979). J. Mol. Biol, 135, 513.
7. Wilson, I. A., Rini, J. M., Fremont, D. H., and Stura, E. A. (1990). In Methods in
enzymology (ed. Langone, J. J.), Academic Press, London. Vol. 203, p. 270.
8. McPherson, A. (1982). Preparation and analysis of protein crystals. Wiley, New
York.
9. Blundell, T. L. and Johnson, L. N. (1976). Protein crystallography. Academic
Press, New York.
10. Arakawa, T. and Timasheff, S. N. (1985). Methods in enzymology, (eds Wychoff,
H. W., Hirs, C. H. W., and Timasheff, S. N.), Academic Press, London, Vol. 114,
p. 49.
11. Feher, G. and Kam, Z. (1985). In Methods in enzymology, ibid ref. 10. Vol. 114,
p. 77.
12. McPherson, A. (1985). In Methods in enzymology, ibid ref. 10. Vol. 114, p. 112.
13. Vekilov, P. G., Monaco, L. A., Thomas, B. R., Stojanoff, V., and Rosenberger, F.
(1996). Acta Cryst., D52, 785.
14. Rosenberger, F., Vekilov, P. G., Muschol, M., and Thomas, B. R. (1996). J. Cryst.
Growth, 168,1.
15. Thaller, C., Weaver, L. H., Eichele, G., Karlsson, R., and Jansonius, J. (1981).
J. Mol. Biol., 147,465.
16. Thaller, C., Eichele, G., Weaver, L. H., Wilson, E., Karlsson, R., and Jansonius,
J. N. (1985). In Methods in enzymology, ibid, ref 10, Vol. 114, p. 132.
17. Stura, E. A. and Wilson, I. A. (1991). J. Cryst. Growth, 110, 270.
18. Stura, E. A. and Wilson I. A. (1990). Methods, 1, 38.
19. Stura, E. A., Stanfield, R. L., Fieser, T. M., Balderas, R. S., Smith, L. R., Lerner,
R. A., et al. (1989). J. Biol. Chem., 264, 15721.
20. Stura, E. A., Johnson, D. L., Inglese, J., Smith, J. M., Benkovic, S. J., and Wilson,
I. A. (1989). /. Biol. Chem., 264, 9703.
21. McPherson, A. (1989). Sci. Am., 260, 3, 62.
22. Schiffer, M., Chang, C.-H., and Stevens, F. J. (1985). J. Mol. Biol., 186,475.
207
E. A. Stum
23. Schulze-Gahmen, U., Rini, J. M., Arevalo, J. H., Stura, E. A., Kenten, J. H., and
Wilson, I. A. (1988). J. Biol. Chem., 263,17100.
24. Wilson, I. A., Bergmann, K. F., and Stura, E. A. (1986). In Vaccines '86 (ed. R. M.
Channock, R. A. Lerner, and F. Brown), pp. 33-7. Cold Spring Harbor
Laboratory, Cold Spring Harbor, NY.
25. Churchill, M. E., Stura, E. A., Pinilla, C., Appel, J. R., Houghten, R. A., Kono, D.
H., et al. (1994). J Mol. Biol., 241, 534.
26. Stura, E. A., Stanfield, R. L., Rini, J. M., Fieser, G. G., Silver, S., Roguska, M., et
al (1992). Proteins, 14,499.
27. Rini, J. M., Stanfield, R. L., Stura, E. A., Salinas, P. A., Profy, A. T., and Wilson, I.
A. (1993). Proc. Natl. Acad. Sci. USA, 90, 6325.
28. Markland, F. S., Jr. and Smith, E. (1971). In The enzymes, 3rd edn (ed. P. D.
Boyer), Vol. III, pp. 561-608. Academic Press, New York.
29. Kraut, J. (1971). In The enzymes, 3rd edn (ed. P. D. Boyer), Vol. III, pp. 547-60.
Academic Press, New York.
30. Neidhart, D. J. and Petsko, G. A. (1988). Protein Eng., 2, 271.
31. Wu, Z.-P. and Hilvert, D. (1989). J. Am. Chem. Soc., III, 4513.
32. Syed, R., Wu, Z. P., Hogle, J. M., and Hilvert, D. (1993). Biochemistry, 32, 6157.
33. Stura, E. A., Ruf, W., and Wilson, I. A. (1996). J. Cryst. Growth, 168, 260.
34. Stura, E. A., Ghosh, S., Garcia-Junceda, E., Chen, L., Wong, C.-H., and Wilson, I.
A. (1995). Proteins, 22, 67.
35. Stura, E. A., Charbonnier, J. B., and Taussig, M. J. (1999). J. Cryst. Growth, 196,
250.
36. Arvai, A. S., Bourne, Y., Williams, D., Reed, S. I., and Tainer, J. A. (1995).
Proteins, 21,70.
37. Laemmli, U. K. (1970). Nature, 227, 680.
38. Andrews, A. T. (1981). In Electrophoresis. Theory, techniques and biochemical
and clinical applications (ed. A. R. Peacocke and W. F. Harrington), p. 63.
Clarendon Press, Oxford.
39. Hames, B. D. (1981). In Gel electrophoresis: a practical approach (ed. B. D.
Hames and D. Rickwood). IRL Press Limited, London, Washington DC 4.
40. Stura, E. A., Arevalo, J. H., Feinstein, A., Heap, R. B., Taussig, M. J., and Wilson,
I. A. (1987). Immunology, 62, 511.
208
8
1. Introduction
At first glance crystallizing nucleic acids poses the same problems as crystal-
lizing proteins since most of the variables to investigate are alike. It is thus
astonishing that crystallization data banks (1) that describe so many successful
protein crystallizations are so poor in information on nucleic acids. This relies
on the physico-chemical and biochemical characteristics of nucleic acids dis-
tinguishing them from proteins. The aim of this chapter is to underline features
explaining the difficulties often encountered in nucleic acid crystallization and
to discuss strategies that could help to crystallize them more readily, either as
free molecules or as complexes with proteins. Other general principles, in
particular for RNA crystallization, are discussed in ref. 2.
Among natural nucleic acids only the smaller ones provide good candidates
for successful crystallizations. Large DNAs or RNAs can a priori be excluded
because of their flexibility that generates conformational heterogeneity not
compatible with crystallization. Thus the smaller RNAs with more compact
structures (with 75-120 nt), especially transfer RNAs (tRNAs), but also 5S
RNA, were the first natural nucleic acids to be crystallized (3, 4). At present
attempts are being made with other RNA systems, such as ribozymes and
introns, fragments of mRNA, viroids, viral and other tRNA-like RNAs,
SELEX-evolved RNAs, and crystallization successes leading to X-ray struc-
ture determinations were reported for RNA domains of up to 160 nt long,
with the resolution of the P4-P6 domain of the self-splicing Tetrahymena
intron (5).
The recent excitement in nucleic acid crystallography, and particularly in
RNA crystallography, have partly been due to technological improvements in
the preparation methods of the molecules. Advances in oligonucleotide
chemical synthesis provide opportunity for making large amounts of pure
desoxyribo- and more recently of ribo-oligomers of any desired sequence.
This led to the crystallization of a number of DNA and RNA fragments and
was followed by the co-crystallization of complexes between proteins and
such synthetic fragments. Transcription methods of RNAs from synthetic
DNA templates were also essential for rejuvenating the structural biology of
A.-C. Dock-Bregeon et al.
RNAs. In the case of complexes of proteins with RNAs, the main difficulty
was to purify large quantities of homogeneous biological material with well
defined physico-chemical properties. The problem has now been overcome in
many cases and problems of larger complexity are now addressed such as
improved crystallizations of ribosome. Virus crystallization will also be briefly
discussed.
Method
1. Construct an insert ending with restriction sites, and containing the T7
promoter and the RNA sequence. This is made by ligation of synthetic
DNA oligomers chosen to hybridize unambiguously in tandem so as to
give the correct, double-stranded, sequence.
2. Insert this synthetic gene into a plasmid, digested with the appropriate
restriction enzymes.
3. Amplify the plasmid by cell culture.
4. Extract the DNA.
5. Linearize the DNA template at the restriction site.
6. Transcribe the DNA template in an appropriate medium containing the
polymerase and the nucleotide monomers.
7. Remove the RNA from the transcription medium, and if needed purify
by gel electrophoresis or HPLC.
Method
1. Prepare phenol by adding 50% (w/w) water. When melted add a few
drops of 1 M KOH to bring the pH of the supernatant around 7.0.
2. Suspend the cells (5 ml/g) in extraction buffer.
214
8: Nucleic acids and their complexes
3. Add to the cell suspension the same volume of phenol. Shake
vigorously for 30 min at room temperature and centrifuge 5 min at
3000 g (room temperature). Recover the upper phase.
4. Add 10-20% of the initial volume of extraction to the phenol phase,
shake vigorously, and recover again the upper phase. Mix the aqueous
phases.
5. Repeat steps 3 and 4 (with some cells, like T. thermophilus, it is
advisable to add 0.1% SDS in the aqueous phase for the second
extraction). Phenol may be removed from aqueous phase by ether
extraction. Caution: ether is volatile and easily flammable.
6. Precipitate RNAs with ethanol (see Protocol 3).
where e = 25 ml/mg (this e value applies for most RNAs, but for exact
measurements it may be necessary to determine it experimentally) (50), / is
the optical path in cm, and c the concentration in mg/ml.
216
8: Nucleic acids and their complexes
Two ways described in Protocol 3 and Protocol 4 can be used to concentrate
RNAs. The method described in Protocol 4 is convenient to change the
solvent. When the RNA is concentrated to a small volume, dilute it in the new
solvent and concentrate again. Repeat several times.
Method
1. Prepare the solution. It should contain Mg2+ ions (> 2 mM) and a Na
salt such as Na acetate (> 10 mM, generally at pH 6.0). For good
recovery, the RNA (e.g. tRNA) solution should be > 0.1 mg/ml. If not,
raise the Na acetate concentration to > 100 mM. Take care that the
solution does not contain too much salt (i.e. after a chromatography,
dialyse first in water).
2. Add two or three volumes of ethanol (best quality). The precipitate
forms.
3. Leave to precipitate completely at -20°C (2 h or more) or at -80°C (20
min or more). For the shortest oligonucleotides or low concentrations,
use the lowest temperature.
4. Centrifuge at the lowest possible temperature, 10 min at > 5000 g
should be sufficient.
5. Dry the pellet under vacuum in the presence of solid KOH.
6. Dissolve the pellet in the desired amount of buffer.
217
A.-C. Dock-Bregeon et al.
Protocol 4. Continued
Method
1. Prepare a set-up of the type Amicon (for large volumes) or Centricon
(for volumes of a few millilitres). Use membranes of correct cut-off
(usually 10000).
2. Concentrate by pushing the solvent through the membrane under
nitrogen pressure (for the Amicon set-up) or by centrifugation (for
Centricon). The RNA (or DNA oligonucleotide) concentrates on the
membrane.
3. Recover the solution when the desired volume is obtained.
DNA B-form
CGCATATATGCG
MPD 10vs40% - 0.5 mM - 0.4 mM 22 mM P212121 57
Mg(Ac)2 2.2 A
CCAAGATTGG, with G:A mismatch
MPD 45% 4 3.0 mM - None 0.7 M C2 58
1.3 A
5'-ACCGGCGCCACA
TGGCCGCGGTGT-5'
MPD 40% 4 1.0 mM 50 mM 1.2 mM 18 mM R3 8
pH 6.0 Mg(Ac)2 2.8 A
Me
CCAGGC CTGG
MPD 40% 4 2.0 mM 20 mM 0.0 mM 50 mM P6 59
microdialysis pH 7.5 2.25 A
218
8: Nucleic acids and their complexes
Table 1. Continued
CGCGCG
Isopropanol 5% 2 mM 30 mM 10 mM 15 mM P21212, 60
pH 7.0 0.9 A
5 5
m CGTAm CG
MPD 8 vs 50% - 4 mM 30 mM 7 mM 10 mM P212121 61
pH 7.0 1.2 A
(5BrCG)3
MPD 10 vs 60% 18 or 37 0.5 mM 20 mM - 200 mM P212121 62
pH 6.5 NaCI 1.4 A
m5CGUAm5CG
MPD8.5 vs 30% Room 4mM 28 mM - 15mM P212121 63
pH 7.0 1.3 A
cccc
MPD 20% - 2.7 mM 100 mM - - I23 64
pH 5.5 2.3 A
RNA:DNA hybrid
r(GCG)d(TATACGC)
MPD 40% - 1.5 mM 30 mM 8 mM 15 mM P212121 65
pH 6.0 1.9 A
RNA
U(UA)6A
MPD 35% 35 4 mM 40 mM None 0.4 M P212121 32
pH 6.5 2.25 A
5'-GGCC(GAAA)GGCC-3', with internal loop
PEG 400 30% Room 2 mM 50 mM - 5 mM P6522 18
Tris MnCI2 2.3 A
pH 7.5 + 20mM
NaCI
5'-GGGGCUAc 25 1 mM 12.5 mM 1 mM 50 mM C2 19
CCUCGAU-5' pH 6.5 MgSO4 1.7 A
MPD 6 vs 35-45%
DNA:drug complexes
CGCG + ditercalinium Room 0.7 mM 16.8 mM 0.3 mM 0.8 mM P41212 66
MPD6 vs 30% + 0.2mM pH 6.0 +14mM 1.7A
drug NH4Ac
CGCGAATTCGCG 5 3 mM 10 mM None 30 mM P2,2,2, 67
+ berenil + 2 mM pH 7.0 2.5 A
MPD 20 vs 50% drug
a
When the crystallization medium is buffered, the buffer is always sodium cacodylate.
b
Most often, MgCI2; in other cases, the salt is specified.
c
Amino acid accepting stem of tRNAAla with G:U mismatch.
219
A.-C. Dock-Bregeon et al.
51 and 52. For tRNAs, a compilation of crystallization conditions is given in
ref. 3. Other general ideas on RNA crystallization can be found in refs 2 and
53.
Putrescine H2N-(CH2)4-NH2
Cadaverine H2N-(CH2)5-NH2
Spermidine H2N-(CH2)3-NH-(CH2)4-NH2
Thermine H2N-(CH2)3-NH-(CH2)3-NH-(CH2)3-NH2
Spermine H2N-(CH2)3-NH-(CH2)4-NH-(CH2)3-NH2
221
A.-C. Dock-Bregeon et al.
4.1.5 Additives
For this part we enter in more specific problems linked to the nature of the
systems investigated. MgCl2 and CaCl2 are the most common additives used.
Phosphate salts have to be avoided for two reasons: they often lead to
insoluble compounds and act as competitors for nucleic acid binding sites.
When existing, cofactors or small substrates (like ATP, GTP, L-tryptophan)
should be used as an important variable in crystallization screenings.
a The concentration in the reservoir is given, or initial concentrations in the form: C(drop) versus C(reservoir).
bThe molar ratio of protein monomers versus DNA duplex is given in brackets.
c
When several DNA duplexes were tried, the conditions indicated are those producing the best crystals.
8: Nucleic acids and their complexes
The main problem with binary complexes is the choice of the best DNA
sequence and of its optimal length. Two major constraints have to be taken
into account: the biological relevance of the sequence and the stability of the
duplex. An effect of the number of base pairs (which should have been a
multiple of seven) was thought to be important after the co-crystallization of
the DNA binding domain of phage 434 represser and its operator (99). Later
examples were no longer in this line. Some co-crystallizations were made with
blunt-ended oligomers: e.g. phage A Cro represser with a 17-mer operator
(98) or phage 434 repressor with a 14-mer operator (99). Others underline the
importance of overhanging nucleotides. These could reinforce the end-to-end
stacking of DNA duplexes which seem to be a common mode of packing.
Clearly, there is no generally applicable rationale that specifies the optimal
length and terminal structure of the oligonucleotides to be used in
crystallizing protein:DNA complexes. The principal limitation of the choice
seems to be the production of the oligonucleotides, especially if the required
sequence is large. This problem has been nicely overcome in the crystalliza-
tion of the CAP protein complexed with DNA (102). Ten oligonucleotides, up
to 20 nt in length, were synthesized. These are able to self-hybridize and were
mixed to generate 19 different double-stranded segments (of 28-36 bp) with
symmetric overhangs of zero, one, or two bases. Crystallization conditions
were examined with 26 different DNA segments, 28 or more bp in length, that
explored a variety of sequences (symmetric or not), length, and extended 5'-
or 3'-termini. Crystals of variable quality were produced, one of them
diffracting to 3.0 A resolution.
Plant viruses
CCMV 20-50 mg/ml Room Succinate PEG 8000 NaN31 mM, 129
0.3 M pH 3.3 3.7-4.0% EDTA 1 mM
CpMV 35 mg/ml 20 K phosphate PEG 8000 (NH 4 )S0 4 0.4M 130
50 mM pH 7.0 2%
STMV 20 mg/ml 23 Cacodylate, (NH4)SO4 NaCI or 131
Na phosphate, 10-18% NaC2H4O4
or Tris 40 mM
pH 6,6.5, or 7
STNV 10-12 or 7-8 Na phosphate PEG6000 Mg 2+ 1 mM 132
mg/ml 50 mM pH 6.5 0.4%
TBSV 30 mg/ml 4 None (NH4)SO4 133
0.5 M
TYMV 25 MES (NH4)H2P04 134
100mMpH3.7 1.11-1.15M
Insect and bacterial viruses
BBV 8 mg/ml 20 Na phosphate (NH4)S04 130
50 mM 13.5%
pH 6.9-7.2
FHV 18 mg/ml Room bisTris PEG 8000 CaCI2 20 mM 135
10 mM pH 6.0 2.8%
MS2 1% 37 Na phosphate PEG 6000 NaN30.02% 136
0.4 M pH 7.4 1.5%
DNA viruses
CPV 10 mg/ml Room Tris PEG 8000 CaCI2 6mM 137
10 mM pH 7.5 0.75%
$X174 8 mg/ml 20,4 bisTris methane PEG 8000 138
90-93 mM 1.5-2.0%
pH 6.8
a
BBV, black beetle virus; BEV, bovine enterovirus; CCMV, cucumber chlorotic mottle virus; CpMV, cowpea mosaic virus; CPV,
canine parvovirus; FHV, flock house virus; HRV, human rhinovirus; STMV, satellite tobacco mosaic virus; STNV, satellite tobacco
necrosis virus; TBSV, tomato bushy stunt virus; TYMV, turnip yellow mosaic virus; MS2 and 0 X 174, two bacteriophages.
A.-C. Dock-Bregeon et al.
first crystallized biological materials (Chapter 1). Now many viruses have
been crystallized and more than 20 structures of spherical viruses are deter-
mined (124). A list of typical representatives that yielded highly ordered
crystals is given in Table 5. For the best, diffraction limit often exceeds 3.0 A
resolution, probably as a consequence of their symmetric and isometric
structure.
The importance of the external capsid and hence the non-effect of RNA or
DNA on crystal formation is nicely demonstrated with cowpea mosaic virus
(CpMV). The genome of this virus consists of two RNA molecules, RNA1
(5.9 kb) and RNA2 (3.5 kb), which are encapsidated in separate particles.
Empty capsids are also formed in vivo. All three components are of the same
size and appear to have identical surfaces. Isomorphous crystals were
obtained with each of the isolated components or with a mixture of the three
components, and the same ratio of components was found in the crystals and
in the crystallizing solution (130).
As for other macromolecular systems, a wide diversity of conditions led to
crystal formation. Details on crystallization conditions can be found in Table 5
and in ref. 124. PEGs (2-3%), alone or mixed with (NH4)2SO4 in the 0.5 M
range, are the most currently used crystallizing agents. Interestingly, their
concentration range is low when compared to other systems. Crystallizations
are usually done at room temperature (20 °C). The pH range is larger than for
nucleic acids and reaches the acidic domain (i.e. 3.3-7.5). It is only limited by
the stability of the capsid. Thus, turnip yellow mosaic virus, an RNA spherical
virus, was crystallized at pH 3.7 (134). Finally, attempts leading to pre-
cipitation should not be discarded since crystals of viruses can also grow from
heavy precipitates by Ostwald ripening mechanisms as exemplified for tomato
bushy stunt virus (TBSV) (139). However, under such circumstances duration
of crystal growth can be long (several weeks and more).
References
1. Gilliland, G. L. and Ladner, J. E. (1996). Curr. Opin. Struct. Biol., 6,595.
2. Lietzke, S. E., Barnes, C. L., and Kundrot, C. E. (1995). Curr. Opin. Struct. Biol.,
5,645.
3. Dock, A.-C., Lorber, B., Moras, D., Pixa, G., Thierry, J.-C., and Giege, R. (1984).
Biochimie, 66,179.
4. Abdel-Meguid, S. S., Moore, P. B., and Steitz, T. A. (1983). J. Mol. Biol., 171, 207.
5. Cate, J. H., Gooding, A. R., Podell, E., Zhou, K., Golden, B. L., Kundrot, C. E., et
al. (1996). Science, 273, 1678.
6. Timsit, Y. and Moras, D. (1992). In Methods in enzymology (ed. D. M. J. Lilley
and J. E. Dahlberg), Vol. 211, pp. 409-29. Academic Press, London.
7. Wing, R., Drew, H., Takano, T., Broka, C., Tanaka, S., Itakura, K., et al. (1980).
Nature, 287, 755.
8. Timsit, Y., Westhof, E., Fuchs, R .P. P., and Moras, D. (1989). Nature, 341, 459.
9. Westhof, E., Auffinger, P., and Gaspin, C. (1996). In DNA and protein sequence
238
8: Nucleic acids and their complexes
analysis: a practical approach (ed. M. Bishop and C. Rawlings), pp. 255-78. IRL
Press, Oxford.
10. Sigman, D. S., Mazumder, A., and Perrin, D. M. (1993). Chem. Rev., 93, 2295.
11. Ehresmann, C., Baudin, F., Mougel, M., Romby, P., Ebel, J.-P., and Ehresmann,
B. (1987). Nucleic Acids Res., 15, 9109.
12. Kolchanov, N. A., Titov, I. I., Vlassova, I. E., and Vlassov, V. V. (1996). Prog.
Nucleic Acid Res. Mol. Biol., 53, 131.
13. Westhof, E. and Michel, F. (1995). In RNA-protein interactions (ed. K. Nagai and
I. W. Mattaj), pp. 25-51. IRL Press, Oxford.
14. Gold, L., Polisky, B., Uhlenbeck, O., and Yarus, M. (1995). Annu. Rev. Biochem.,
64, 763.
15. Leonard, G. A., McAuley, H. K., Ebel, S., Lough, D. M., Brown, T., and Hunter,
W. N. (1994). Structure, 2, 483.
16. Pley, H. W., Flaherty, K. M., and McKay, D. M. (1994). Nature, 372, 68.
17. Schindelin, H., Zhang, M., Bald, R., Furste, J.-P., Erdmann, V. A., and
Heinemann, U. (1995). J. Mol. Biol., 249, 595.
18. Baeyens, K. J., De Bondt, H. L., Pardi, A., and Holbrook, S. R. (1996). Proc. Natl.
Acad. Sci. USA, 93,12851.
19. Ott, G., Dorfler, S., Sprinzl, M., Muller, U., and Heinemann, U. (1996). Acta
Cryst.,D52,871.
20. Ferre-D'Amare, A. R., Zhou, K., and Doudna, J. A. (1998). J. Mol. Biol., 279,621.
21. Oubridge, C., Ito, T., Teo, C. H., Fearnley, L, and Nagai, K. (1995). J. Mol. Biol.,
249, 409.
22. Pingoud, A., Fliess, A., and Pingoud, A. (1989). In HPLC of macromolecules: a
practical approach (ed. R. W. A. Oliver), pp. 183-208. IRL Press, Oxford.
23. Davanloo, P., Rosenberg, A. H., Dunn, J. J., and Studier, W. (1984). Biochemistry,
81, 2035.
24. Milligan, J. F., Groebe, D. R., Witherell, G., and Uhlenbeck, O. C. (1987). Nucleic
Acids Res., 15, 8783.
25. Frugier, M., Florentz, C., Hosseini, M. W., Lehn, J.-M., and Giege, R. (1994).
Nucleic Acids Res., 22, 2784.
26. Yin, Y. and Carter, C. W., Jr (1996). Nucleic Acids Res., 24, 1279.
27. Lapham, J. and Crothers, D. M. (1996). RNA, 2, 289.
28. Giege, R., Sissler, M., and Florentz, C. (1998). Nucleic Acids Res., 26, 5017.
29. Price, S. R., Ito, N., Oubridge, C., Avis, J. M., and Nagai K. (1995). J. Mol. Biol.,
249, 398.
30. Perreault, J. P., Wu, T., Cousineau, B., Ogilvie, K. K., and Cedergren, R. (1990).
Nature, 344, 565.
31. Anderson, A. C., Scaringe, S. A., Earp, B. E., and Frederick, C. R. (1996). RNA, 2,
110.
32. Dock-Bregeon, A.-C., Chevrier, B., Podjarny, A., Johnson, J., de Bear, J. S.,
Gough, G. R., et al. (1989). J. Mol. Biol., 209, 459.
33. Cruse, W. B., Saludjian, P., Biala, E., Strazewski, P., Prange, T., and Kennard, O.
(1994). Proc. Natl. Acad. Sci. USA, 91, 4160.
34. Monier, R. (1971). In Procedure in nucleic acids research (ed. G. L. Cantoni and
D. R. Davies), Vol. 2, pp. 618-28. Harper and Row, New York.
35. Morikawa, K., Kawakami, M., and Takemura, S. (1982). FEBS Lett., 145, 194.
36. Giege, R. and Ebel, J.-P. (1968). Biochim. Biophys. Acta, 161, 125.
239
A.-C. Dock-Bregeon et al.
37. Kowalak, J. A., Dalluge, J. J., McCloskey, J. A., and Stetter, K. O. (1994).
Biochemistry, 33, 7869.
38. Yonath, A., Frolow, F., Shoham, M., Mussig, J., Makowski, I., Glotz, C., et al.
(1988). J. Cryst. Growth, 90,231.
39. Perona, J. J., Swanson, R., Steitz, T. A., and Soll, D. (1988). J. Mol. Biol., 202,121.
40. Dirheimer, G. and Ebel, J.-P. (1967). Bull. Soc. Chim. Biol., 49,1679.
41. Nishimura, S. (1971). In Procedure in nucleic acids research (ed. G. L. Cantoni and
D. R. Davies), Vol. 2, pp. 542-64. Harper and Row, New York.
42. Spencer, M., Neave, E. J., and Webb, N. L. (1978). J. Chromatogr., 166,447.
43. Gillam, I., Millward, S., Blew, D., von Tigerstrom, M., Wimmer, E., and Tener, G.
M. (1967). Biochemistry, 6,3043.
44. Singhal, R. P., Griffin, G. D., and Novelli, G. D. (1976). Biochemistry, 15, 5083.
45. Nishimura, S., Shindo-Okada, N., and Crain, P. F. (1987). In Methods in
enzymology (ed. R. Wu), Vol. 155, pp. 373-8. Academic Press, London.
46. Holmes, W. M., Hurd, R. E., Reid, B. R., Rimerman, R. A., and Hatfield, G. W.
(1975). Proc. Natl. Acad. Sci. USA, 72, 1068.
47. Dudock, B. S. (1987). In Molecular biology of RNA. New perspectives (ed. M.
Inouye and B. S. Dudock), pp. 321-9. Academic Press Inc., New York and
London.
48. Brown, D. M. (1974). In Basic principles in nucleic acids chemistry (ed. P. O. P.
Ts'O), Vol. 2, pp. 1-90. Academic Press Inc., New York and London.
49. Werner, W., Krebs, W., Keith, G., and Dirheimer, G. (1976). Biochim. Biophys.
Acta, 432, 161.
50. Gueron, M. and Leroy, J.-L. (1978). Anal. Biochem., 91, 691.
51. Kennard, O. and Hunter, W. N. (1989). Q. Rev. Biophys., 22, 327.
52. Neidle, S. (1996). In Methods in molecular biology (ed. C. Jones, B. Mulloy, and
M. Sanderson), Vol. 56, pp. 267-92. Humana Press Inc., Totowa, NJ.
53. Wahl, M. C., Ramakrishnan, B., Ban, C., Chen, X., and Sundaralingam, M. (1996).
Acta Cryst., D52, 668.
54. Wang, A. H. J., Fujii, S., van Boom, J. H., and Rich, A. (1982). Proc. Natl. Acad.
Sci. USA, 79, 3968.
55. Hunter, W. N., Langlois D'Estaintot, B., and Kennard, O. (1989). Biochemistry,
28, 2444.
56. Langlois d'Estaintot, B., Dautant, A., Courseille, C., and Precigoux, G. (1993).
Eur. J. Biochem., 213, 673.
57. Yoon, C., Prive, G. G., Goodsell, D. S., and Dickerson, R. (1988). Proc. Natl.
Acad. Sci. USA, 85, 6332.
58. Prive, G. G., Heinemann, U., Chandrasegaran, S., Kan, L. S., Kopka, M., and
Dickerson, R. (1987). Science, 238, 498.
59. Heinemann, U. and Alings, C. (1991). EMBO J., 10, 35.
60. Wang, A. H. J., Quigley, G. J., Kolpak, F. J., Crawford, J. L., van Boom, J. H., van
der Marel, G., et al. (1979). Nature, 282, 680.
61. Wang, A. H. J., Hakoshima, T., van der Marel, G., van Boom, J. H., and Rich, A.
(1984). Cell, 37, 321.
62. Chevrier, B., Dock, A.-C., Hartmann, B., Leng, M., Moras, D., Thuong, M. Y., et
al. (1986). J. Mol. Biol., 188, 707.
63. Zhou, G. and Ho, P. S. (1990). Biochemistry, 29, 7229.
64. Chen, L., Cai, L., Zhang, X., and Rich, A. (1994). Biochemistry, 33, 13540.
240
8: Nucleic acids and their complexes
65. Wang, A. H. J., Fujii, S., van Boom, J. H., van der Marel, G., van Boeckel, S. A.
A., and Rich, A. (1982). Nature, 299,601.
66. Gao, Q., Williams, L. D., Egli, L., Rabinowitch, D., Chen, S.-L., Quigley, G. J., et
al. (1991). Proc. Natl. Acad. Sci. USA, 88, 2422.
67. Brown, D. G., Sanderson, M. R., Skelly, J. V., Jenkins, T. C, Brown, T., Garman,
E., et al. (1990). EMBO J., 9,1329.
68. Kim, S. H., Quigley, G., Suddath, F. L., McPherson, A., Sneden, D., Kim, J. J., et
al. (1973).J. Mol Biol, 75, 421.
69. Johnson, C. D., Adolph, K., Rosa, J. J., Hall, M. D., and Sigler, P. B. (1970).
Nature, 226,1246.
70. Giege, R., Moras, D., and Thierry, J.-C. (1977). J. Mol. Biol., 115, 91.
71. Baeyens, K. J., Jancarik, J., and Holbrook, S. R. (1994). Acta Cryst., D50,764.
72. Mikol, V., Rodeau, J.-L., and Giege, R. (1989). J. Appl. Cryst., 22, 155.
73. Sakai, T. T. and Cohen, S. S. (1976). Prog. Nucleic Acid Res. Mol. Biol., 17,15.
74. Young, J. D., Bock, R. M., Nishimura, S., Ishikura, H., Yamada, Y., Rajbandhary,
U. L., et al. (1969). Science, 166,1527.
75. Sauter, C., Ng, J. D., Lorber, B., Keith, G., Brion, P., Hosseini, M.-W., et al.
(1999). /. Cryst. Growth, 196, 365.
76. Ladner, J. E., Finch, J. T., Klug, A., and Clark, B. F. C. (1972). J. Mol. Biol, 72, 99.
77. Bancroft, D., Williams, L. D., Rich, A., and Egli, M. (1994). Biochemistry, 33,
1073.
78. Holbrook, S. R., Sussman, J. L., Warrant, W. R., and Kim, S. H. (1978). J. Mol.
Biol., 123,631.
79. Gessner, R. V., Quigley, G. J., Wang, A. H. J., van der Marel, G., van Boom, J. H.,
and Rich, A. (1985). Biochemistry, 24, 237.
80. Brennan, R. G., Westhof, E., and Sundaralingam, M. (1986). J. Biomol. Struct.
Dyn.,3,649.
81. Einspahr, H., Cook, W. J., and Bugg, C. E. (1981). Biochemistry, 20, 5788.
82. Hingerty, B. E., Brown, R. S., and Klug, A. (1982). Biochim. Biophys. Acta, 697,
78.
83. Doudna, J. A., Grosshans, C., Gooding, A., and Kundrot, C. E. (1993). Proc. Natl.
Acad. Sci. USA, 90,7829.
84. Scott, W. G., Finch, J. T., Grenfell, R., Fogg, J., Smith, T., Gait, M. J., et al. (1995).
J. Mol. Biol., 250, 327.
85. Dickerson, R. E., Goodsell, D. S., Kopka, M. L., and Pjura, P. E. (1987). J.
Biomol. Struct. Dyn., 5, 557.
86. Moras, D. and Bergdoll, M. (1988). J. Cryst. Growth, 90,283.
87. Anderson, A. C., Earp, B. E., and Frederick, C. A. (1996). J. Mol. Biol., 259,696.
88. Dock-Bregeon, A.-C. and Moras, D. (1987). Cold Spring Harbor Symp. Quant.
Biol., 52,113.
89. Golden, B, Gooding, A. R., Podell, E. R., and Cech, T. R. (1996). RNA, 2,1295.
90. Freemont, P. S., Friedman, J. M., Beese, L., Sanderson, M. R., and Steitz, T. A.
(1988). Proc. Natt. Acad. Sci. USA, 85, 8924.
91. Brown, D. G. and Freemont, P. S. (1996). In Methods in molecular biology (ed. C.
Jones, B. Mulloy, and M. Sanderson), Vol. 56, pp. 293-318. Humana Press Inc.,
Totowa, NJ.
92. Grable, J., Frederick, C. A., Samudzi, C., Jen-Jacobson, L., Lesser, D., Greene, P.,
et al. (1984). /. Biomol. Struct. Dyn., 1,1149.
241
A.-C. Dock-Bregeon et al.
93. Suk, D., Lahm, A., and Oefner, C. (1988). Nature, 332,464.
94. Luger, K., Mader, A. W., Richmond, R. K., Sargent, D. F., and Richmond, T. J.
(1997). Nature, 389, 251.
95. Lorber, B., Giege, R., Ebel, J.-P., Berthet, C., Thierry, J.-C, and Moras, D.
(1983). J. Biol. Chem., 258, 8429.
96. Jordan, S. R., Whitcombe, T. V., Berg, J. M., and Pabo, C. O. (1985). Science,
230, 1383.
97. Ruff, M., Cavarelli, J., Mikol, V., Lorber, B., Mitschler, A., Giege, R., et al.
(1988).J. Mol. Biol, 201,235.
98. Brennan, R. G., Takeda, Y., Kim, J., Anderson, W. F., and Matthews, B. W.
(1986). J. Mol. Biol, 188, 115.
99. Anderson, J., Ptashne, M., and Harrison, S. C. (1984). Proc. Natl. Acad. Sci.
USA, 81, 1307.
100. Aggarwal, A. K., Rodgers, D. W., Drottar, M., Ptashne, M., and Harrison, S.
(1988). Science, 242, 899.
101. Joachimiak, A., Marmorstein, R. Q., Schevitz, R. W., Mandecki, W., Fox, J. L.,
and Sigler, P. B. (1987). J. Biol. Chem., 262, 4917.
102. Schultz, S. C., Shields, G. C., and Steitz, T. A. (1990). J. Mol. Biol., 213,159.
103. Ellenberger, T. E., Brandel, C. J., Struhl, K, and Harrison, S. C. (1992). Cell, 71,
1223.
104. Pavletich, N. P. and Pabo, C. O. (1993). Science, 261, 1701.
105. Clark, K. L., Halay, E. D., Lai, E., and Burley, S. K. (1993). Nature, 364, 412.
106. Yang, W. and Steitz, T. A. (1995). Cell, 82,193.
107. Schwabe, J. W. R., Chapman, L., Finch, T., and Rhodes, D. (1993). Cell, 75,
567.
108. Kim, Y., Geiger, J. H., Hahn, S., and Sigler, P. B. (1993). Nature, 365,512.
109. Glover, J. N. M. and Harrison, S. C. (1995). Nature, 373, 257.
110. Nikolov, D. B., Chen, H., Halay, E. D., Usheva, A. A., Hisatake, K., Lee, D. K.,
et al. (1995). Nature, 377,119.
111. Sousa, R. (1995). Acta Cryst., D51,271.
112. Boeglin, M., Dock-Bregeon, A.-C., Eriani, G., Gangloff, J., Ruff, M., Poterzman,
A., et al. (1996). Acta Cryst., D52,211.
113. Arnez, J. and Steitz, T. A. (1994). Biochemistry, 33, 7567.
114. Cusack, S., Yaremchuk, A., and Tukalo, M. (1996). EMBO J., 15, 6321.
115. Reshetnikova, L., Khodyreva, S., Lavrik, O., Ankilova, V., Frolow, F., and Safro,
M. (1993). /. Mol Biol, 231, 927.
116. Yaremchuk, A. D., Tukalo, L, Krikliviy, N., Malchenko, N., Biou, V., Berthet-
Colominas, C., et al. (1992). FEBS Lett., 310,157.
117. Nissen, P., Reshetnikova, L., Siboska, G., Polekhina, G., Thirup, S., Kjeldgaard,
M., et al. (1994). FEBS Lett, 356,165.
118. Oubridge, C., Ito, N., Evans, P. R., Teo, C. H., and Nagai, K. (1994). Nature, 372,
432.
119. Record, M. J. J., Anderson, C. F., and Lohman, T. M. (1978). Rev. Biophys., 11,
103.
120. Giege, R., Lorber, B., Ebel, J.-P., Moras, D., Thierry, J.-C., Jacrot, B., et al.
(1982). Biochimie, 64, 357.
121. Trakhanov, S., Yusupov, M., Shirokov, V., Garber, M., Mitschler, A., Ruff, M., et
al (1989). J. Mol Biol, 209,327.
242
8: Nucleic acids and their complexes
122. Thygesen, J., Krumholz, S., Levin, I., Zaytzev-Bashan, A., Harms, J., Bartels, H.,
et al. (1996). /. Cryst. Growth, 168, 308.
123. Ban, N., Freeborn, B., Nissen, P., Penczek, P., Grassucci, R. A., Sweet, R., et al.
(1998). Cell, 93,1105.
124. Fry, E., Logan, D., and Stuart, D. (1996). In Methods in molecular biology (ed. C.
Jones, B. Mulloy, and M. Sanderson), Vol. 56, pp. 319-63. Humana Press Inc.,
Totowa, NJ.
125. Smyth, M., Fry, E., Stuart, D., Lyons, C, Hoey, E., and Martin, S. J. (1993). J.
Mol. Biol., 231, 930.
126. Li, T., Zhang, A., luzuka, N., Nomoto, A., and Arnold, E. (1992). J. Mol. Biol.,
223,1171.
127. Arnold, E., Erickson, J. W., Fout, S. G., Frankenberger, E. A., Hecht, H. J., Luo,
M., et al. (1984). J. Mol. Biol., 177, 417.
128. Luo, M., Vriend, G., Kamer, G., Minor, I., Arnold, E., Rossman, M. G., et al.
(1987). Science, 235,182.
129. Speir, J. A., Munshi, S., Baker, T. S., and Johnson, J. E. (1993). Virology, 193,
234.
130. Sehnke, P. C., Harrington, M., Hosur, M. V., Li, Y. R., Usha, R., Tucker, R. C., et
al (1988). /. Cryst. Growth, 90, 222.
131. Koszelak, S., Dodds, J. A., and McPherson, A. (1989). J. Mol. Biol., 209,323.
132. Jones, T. A. and Liljas, L. (1984). J. Mol. Biol., 177, 735.
133. Harrison, S. C. and Jack, A. (1975). J. Mol. Biol., 97,173.
134. Canady, M. A., Larson, S. B., Day, J., and McPherson, A. (1996). Nature Struct.
Biol., 3,771.
135. Fisher, A. J., McKinney, B. R., Wery, J. P., and Johnson, J. E. (1992). Acta Cryst.,
48,515.
136. Valegard, K., Unge, T., Montelius, I., Strandberg, B., and Fiers, W. (1986). J.
Mol. Biol., 190, 587.
137. Luo, M., Tsao, J., Rossmann, M. G., Basak, S., and Compans, R. W. (1988). J.
Mol. Biol., 200, 209.
138. Willingmann, P., Krishnaswamy, S., McKenna, R., Smith, T. J., Olson, N. H.,
Rossmann, M. G., et al. (1990). J. Mol. Biol., 212,345.
139. Ng, J. D., Lorber, B., Witz, J., Theobald-Dietrich, A., Kern, D., and Giege, R.
(1996). J. Cryst. Growth, 168, 50.
243
This page intentionally left blank
9
Crystallization of membrane
proteins
F. REISS-HUSSON and D. PICOT
1. Introduction
Crystallization of membrane proteins is one of the most recent developments
in protein crystal growth; in 1980, for the first time, two membrane proteins
were successfully crystallized, bacteriorhodopsin (1) and porin (2). Since then,
a number of membrane proteins (about 30) yielded three-dimensional
crystals. In several cases, the quality of the crystals was sufficient for X-ray
diffraction studies. The first atomic structure of a membrane protein, a
photosynthetic bacterial reaction centre, was described in 1985 (3), followed
by the structure of about ten other membrane protein families. Crystallization
of membrane proteins is now an actively growing field, and has been discussed
in several recent reviews (4-8).
The major difficulty in the study of membrane proteins, which for years
hampered their crystallization, comes from their peculiar solubility properties.
These originate from their tight association with other membrane compo-
nents, particularly lipids. Indeed integral membrane proteins contain hydro-
phobic surface regions buried in the lipid bilayer core, as well as hydrophilic
regions with charged or polar residues more or less exposed at the external
faces of the membrane. Disruption of the bilayer for isolating a membrane
protein can be done in various ways: extraction with organic solvents, use of
chaotropic agents, or solubilization by a detergent. The last method is the
most frequently used, since it maintains the biological activity of the protein if
a suitable detergent is found. This chapter will be restricted to specific aspects
of three-dimensional crystallizations done in micellar solutions of detergent.
In some cases, it is possible to separate soluble domains from the membrane
protein either by limited proteolysis or by genetic engineering. Such protein
fragments can then be treated as soluble proteins and so will not be discussed
further in this chapter. We refer to Chapter 12 and the review by Kuhlbrandt
(9) for the methodology of two-dimensional crystallization used for electron
diffraction.
F. Reiss-Husson and D. Picot
2. Crystallization principles
The general principles discussed in this book for the crystallization of soluble
biological macromolecules apply for membrane proteins; the protein solution
must be brought to supersaturation by modifying its physical parameters
(concentrations of constituents, ionic strength, and so on), so that nucleation
may occur. The main differences from the behaviour of soluble proteins stem
from the following two points:
(a) The entity which is going to crystallize is the protein-detergent complex,
not the protein alone. Yet, most of the detergent found in the crystal is
disordered. This has been demonstrated for three detergents (C10DAO,
C12DAO, and C8G) associated with two bacterial reaction centres (10,11)
and OmpF porin (12). Usually, only a few ordered detergent molecules
are seen in the electron density maps. But the amount of disordered
detergent in the crystals is fairly high; about 200 molecules of detergent
are associated with one reaction centre protein, and form a ring around
the hydrophobic transmembrane a helices. The detergent ring is inter-
connected with its neighbours by bridges. Thus ribbon-like detergent
structures run throughout the crystal. These findings explain why the
characteristics of the detergent molecules (such as their length) are so
crucial in the crystallization process. Indeed they should fit around the
hydrophobic regions of the protein without hindering the interprotein
contacts.
(b) The solubility of the protein-detergent complex is governed not only by
the protein properties, but also (and mainly) by those of the detergent
micellar solution. Generally, as will be discussed below, this detergent is
non-ionic; its micellar solution therefore only exists in a limited range of
concentration and temperature, defined in a phase diagram (Figure 1).
Outside of this range, the micellar solution may spontaneously break
apart into two immiscible aqueous phases; one being enriched in deter-
gent, the other one remaining essentially depleted in detergent. The
temperatures and concentrations at which phase separation is observed
define a curve, called the consolution boundary. Depending on the
detergent and the crystallizing agent used, this boundary may be reached
starting from the micellar solution either by increasing or by decreasing
the temperature as shown in Figure 1. When phase separation takes place,
the solubilized membrane protein generally partitions into the detergent-
rich phase (but exceptions are known for glycoproteins) (13). Phase
separation is a function of all constituents of the solution such as deter-
gent, protein, nature and concentration of salt, concentration of a crystal-
lizing agent like PEG. Phase separation seems to play a major role in the
crystallization because it is quite often observed that crystallization takes
place right before phase separation occurs. Choosing crystallization con-
246
9: Crystallization of membrane proteins
ditions close to the consolution boundary thus appears equivalent to
bringing the protein-detergent complexes into a supersaturated state.
In conclusion, the chosen detergent plays a crucial role in membrane pro-
tein crystallization and it is therefore important to gain knowledge of its
properties before embarking on crystallization experiments.
247
F. Reiss-Husson and D. Picot
synthesized for this purpose. Most of them, however, were already used by
biochemists for solubilizing and purifying membrane proteins. As such they
have been reviewed in a recent volume of this series (14). With a few except-
ions, they are either non-ionic, with uncharged polar groups, or zwitterionic at
the pH used. Furthermore they all are short aliphatic compounds, the length
of their hydrophobic part not exceeding that of a normal C12 hydrocarbon
chain. This last feature results in moderate or high critical micellar
concentration (CMC) values (the CMC being the concentration limit between
molecular and micellar solutions, compare Figure 1). In a micellar solution,
micelles are in dynamic equilibrium with the monomeric detergent still pre-
sent in concentration equal to the CMC. Thus, the higher the CMC, the larger
this exchange. Such a mobile character might play a role in the crystallization
of the detergent-protein complexes.
One further requirement of these detergents is purity and chemical
homogeneity, which are important for:
(a) Reproducibility of experiments. When a detergent is heterogeneous (e.g.
it contains a mixture of various hydrocarbon chains) or impure (traces of
248
9: Crystallization of membrane proteins
fatty alcohol, and so on), its composition is badly defined and may vary
from batch to batch. This may lead to the intolerable situation that, with a
new detergent batch, crystals may no longer be obtained or still grow but
with lower quality.
(b) Quality of the crystals. For example trace impurities in C8G have been
reported to interfere with crystal growth of bacteriorhodopsin.
Thus it is advisable to pay attention to purity of the selected detergent and
be able to check it. Thin-layer chromatography (TLC) is one of the useful
tools, simple and fast to perform (Protocol 1). Trace impurities may however
escape detection by TLC.
Method
1. Equilibrate a silica gel plate in a TLC tank with enough ethyl acetate:
methanol (4:1, v/v) to wet the bottom of the plate over 5 mm.
2. Dissolve 1 mg detergent in 100 ul ethanol in a microcentrifuge tube.
3. Using a microsyringe, spot 10 ul of this solution at the bottom of the
plate, 1 cm from the edge, and let it dry.
4. Put the plate back in the tank and wait until the solvent front reaches
within 1 cm of the top.
5. Take the plate out and let it dry.
6. Put the plate in the desiccator containing iodine vapours and let it
stain.b Only one spot should be present.
Method
1. Prepare a working solution of 10 uM ANS in water by dilution of a
stock 400 (AM ANS solution.
2. Prepare 500 ul of the stock detergent solution in 10 (uM ANS at a
detergent concentration about 100 times the expected CMC. Mix
thoroughly. Fluorescence of this solution is taken as the 100% control.
3. Read the fluorescence emission at 490 nm with excitation at 370 nm
while titrating a 2 ml sample of 10 (uMANS with small aliquots of stock
detergent solution (up to 50 ul). Use the fluorescence of 10 uM ANS as
the blank.
4. Plot the relative fluorescence versus detergent concentration. A steep
increase in fluorescence indicates the onset of micellization. Draw a
straight line through the points in the steep increase region; its
intersection with the x axis is the CMC of the detergent.
Method
1. Pour the column with the ion exchanger resin.
2. Wash the column with 120 ml ethanol then with 600 ml water. Stop the
flow when water is draining the gel surface.
3. Dissolve 5 g C8G in 50 ml water and put the solution on the gel. Elute
at a flow rate of 0.2 ml/min. Then wash with water at the same rate.
4. Collect the first 100 ml of eluate. Lyophilize and store at -20°C.
3.2 n-Alkyl-thioglucosides
Although these detergents have not been used very often in crystallizations
(23, 24), they could be assayed instead of the glucosides. The C6, C7, C8, and
C10compounds are commercially available.
251
Table 3. Crystallization conditions for some membrane proteins
Method
1. Stir for 2 h a 10% (w/v) solution of the detergent in distilled water with
SnCI2 (0.5% (w/v) final concentration).
2. Add NaCI to 10% (w/v) final concentration, then add an equal volume
of dichloromethane and mix thoroughly. Let stand until the two layers
separate.
3. Discard the upper water layer and recover the lower organic layer
which contains the detergent.
4. Extract the organic phase with an equal volume of 1% NaOH, 10%
NaCI, then three times with 10% NaCI (the pH of the final NaCI layer
must be 7). Each time discard the water layer.
5. Dry the organic phase for 24 h over anhydrous Na2SO4, then in a flash
evaporator at 40 °C.
6. Store the purified detergent at -20°C.
Method
1. Mix the protein solution (~ 1 mg/ml) with 20 vol. of hexane:
isopropanol (3:2, v/v).
2. Shake well and then centrifuge at low speed (5000 g) for 20 min.
3. Recover the supernatant. Repeat steps 1 and 2 on the pellet.
4. Combine the supernatants and dry this under a stream of N2 or with a
flash evaporator.
5. Dissolve the residue in the minimal volume of chloroform.
6. Carry out TLC analysis of the lipid extract as described in Protocol 1,
except use chloroform:methanol:water (65:25:4, by vol.) as the solvent.
The neutral lipids run near the solvent front and the other lipids are
fractionated into various classes. They are identified with reference to
known standards and published Rf values, in addition to the use of
specific stains (28) instead of iodine staining or H2SO4 charring.
256
9: Crystallization of membrane proteins
Homogeneity of the preparation requires also its monodispersity, i.e. all the
protein-detergent complexes should have the same composition. This is
verified by gel filtration experiments, e.g. on FPLC columns (Superose gels,
Pharmacia) or HPLC ones (such as the TSK-SW or TSK-PW gels, Toso
Haas). From these experiments one can estimate the size of the whole com-
plex, including detergent. On the other hand, the amount of bound detergent
may be determined by several techniques (see ref. 32 for a review). Com-
bining these results allows the aggregation state of the protein-detergent
complex to be determined and controlled.
4.2.1 Dialysis
This is the simplest procedure but not applicable in all cases. The meaningful
parameter of a dialysis membrane is its cut-off value. It must be low enough
for retention of the protein-detergent complex. Thus for large complexes,
highly permeable membranes can be used; e.g. for a complex of 100 kDa, a
Spectrapor 7 membrane with a cut-off value of 50 kDa may be used. With
such a pore size, exchange of detergents with CMC values higher than 1 mM
is relatively rapid (a few days). On the other hand, for a small complex (e.g.
15 kDa) a Spectrapor 1 membrane (cut-off value 6 kDa) should be chosen and
only detergents of high CMC (10 mM or so) will exchange at an acceptable
rate. Since the diffusion rate between two detergents may differ by several
orders of magnitude, care should be taken in order to avoid detergent depletion
leading to irreversible aggregation or increase of detergent concentration
causing inactivation of the protein.
In favourable cases dialysis may be performed in two steps; e.g. it is possible
with the bacterial reaction centre, to exchange 0.1% C12DAO for 0.8% C8G
as follows:
(a) First dialyse for 48 h against a detergent-free buffer, with several changes
of reservoir (Spectrapor semi-microtubing, cut-off 12 kDa). Removal of
C12DAO results in increased turbidity of the sample.
(b) Transfer the bag in a C8G containing buffer for another 24 h. Loss of
turbidity indicates redissolution of the protein.
From a practical point of view, dialysis may be performed either in the
familiar closed bags, or for small volumes (< 500 ul) in microdialysis
257
F. Reiss-Husson and D. Picot
cells, either home-built (see Chapter 5) or commercially available (Pierce,
Amicon). Some dialysis membranes, particularly those stored wet (e.g.
Spectrapor type 7, Amicon) are specially prone to fungi contamination.
Before use, a good precaution is to boil membranes for 1 min in 1% (w/v)
NaHCO3, then to soak them three times in highly pure water (Milli Q grade),
and to use them immediately afterwards.
4.2.2 Chromatography
The sample which contains detergent 1 is chromatographed on a column
equilibrated with detergent 2, and eluted with detergent 2. This method is
feasible with any type of detergent, with various chromatographic supports.
(a) Gel filtration is gentle and can be used with all detergents. However, it
usually dilutes the sample appreciably.
(b) Ion exchange chromatography (with DEAE, CM exchangers, or some-
times hydroxyapatite) is restricted to non-ionic detergents; it has the
advantage of concentrating the sample when elution is done by a steep
salt increase, but one must check that the stability of the protein is not
affected by the shift of the CMC, which is induced by the higher salt
concentration.
If the presence of salt in the final sample is not wanted, a mixed column
consisting of ion exchanger superposed on gel filtration matrix (Sephadex
G25) will exchange the detergent and desalt the sample altogether (33).
Microcolumns built from Pasteur pipettes are useful for such ion exchange
procedure.
4.2.3 Precipitation
The protein in the presence of detergent 1 is first precipitated with cold ethanol,
which is a solvent of detergent 1; the precipitate is washed to eliminate deter-
gent 1, and redissolved in detergent 2. This method has been used only for
porins as it requires a very sturdy protein. Salt or PEG precipitation may also
be used in some cases but care should be taken to avoid phase separation.
Furthermore, since precipitated protein still bound a large amount of
detergent, several precipitation cycles are needed.
The efficiency of these procedures can be judged from the absence of
detergent 1 in the final sample. Unfortunately, very few detergents exist in a
labelled form and those are expensive. Colorimetric determination is possible
for glucosides and maltosides with reagents specific for reducing sugars (34).
In other cases, TLC of extracts is the only method.
5. Crystallization protocols
For conducting crystallization trials with a membrane protein, similar
strategies, like the incomplete factorial design (see Chapters 4 and 5), to those
developed for the soluble proteins can be used (7). Several parameters have
to be chosen (see Chapter 1, Table 7) to which should be added the nature
and concentration of the detergent. Some of these parameters have been
discussed earlier with reference to soluble proteins (see Chapter 5). One of
these parameters, the purification of the protein, will influence the crystal-
lization condition to such an extent that it may be as important to modify the
purification as the crystallization protocol. This has been the case for porin
(35) and prostaglandin H synthase (31), for which the change of the detergent
used for solubilization was critical to obtain good crystals, even if an other
detergent was then used for the crystallization.
Protein-protein, protein-detergent, and detergent-detergent interactions
can be observed in membrane protein crystals. Depending on the nature of
the protein and the detergent, the crystallization process will be more in-
fluenced by one or another type of interaction. Protein-protein interactions
are more specific than detergent-detergent interactions and should thus yield
better crystals. This may explain why crystals of the reaction centre from Rps.
viridis diffract better than those from Rb. sphaeroides since the former is
crystallized with an additional soluble subunit. Therefore, larger proteins, but
with a larger polar domain, may be easier to crystallize than their more hydro-
phobic counterparts. An increase of the hydrophilic surface of cytochrome c
oxidase of Paracoccus denitrificans has been obtained by forming a complex
with a conformation specific engineered Fv fragment (36). This has allowed
well diffracting crystals to be grown using the detergent C12M, that is able to
maintain the activity. Thus, the Fv fragment counterbalances the disadvantage
of C12M, i.e. to form large micelles (37). We will stress now a few specific
points based on published crystallization protocols (Table 3).
259
F. Reiss-Husson and D. Picot
5.1 Detergent
The choice of detergent is still empirical. The first criterion is to maintain the
functional and structural integrity of the protein. Not only the type but also
the detergent concentration are important. Furthermore, the behaviour and
stability of the protein will be very different below and above the CMC as well
as above and below the consolution boundaries. The optimal stability of a
membrane protein is often observed around the CMC, which may therefore
be a good starting detergent concentration for a crystallization experiment. A
number of membrane proteins crystallize with a wide variety of detergents. In
one case, a systematic search has been done over 23 detergents with Omp F,
an E. coli porin (23). Among them, 16 non-ionic detergents (from classes
described in Table 7) could be used successfully. Interestingly, OmpF
crystallized also in micellar solutions of short chains lecithins (diC6- or
diC7-glycerophosphatidylcholines) or lysolecithins (monoC14- or monoC16-
glycerophosphatidylcholines) used as detergents. However, crystals could not
be obtained with ionic detergents, nor with detergents derived from bile salts
(cholate, CHAPS, or CHAPSO). A non-ionic, non-steroid polar group and a
short alkyl chain seemed thus to be the only requirement, without narrow
specificity.
The reverse situation may prevail, and strict requirements may exist for
chain length or detergent type. For example, a light-harvesting chloroplast
protein, LHCII, crystallizes reliably with C9G and poorly with C8G (4).
Therefore, for an unknown protein, screening should be done using at least
two homologues of each detergent class. However, before beginning an ex-
tensive screening, one could try the most popular C8G and C12DAO deter-
gents. Indeed they allowed a number of successful crystallizations such as
bacterial reaction centres and light-harvesting complexes; E. coli porins gave
crystals with one of them or both. From a certain point of view, these two
detergents may be considered as equivalent: in two bacterial reaction centre
crystals, the regions they occupied respectively around the hydrophobic a
helices could be nearly superimposed (10, 11).
C12DAO should be tried alone and also in the presence of an additive (see
Section 7) which was required in some cases. C8G has been used either alone,
or mixed with low amounts of other short chain detergents; these were
however not essential for crystallization but improved crystal growth. Thus
first trials may be done with pure C8G only.
The initial concentration of detergent in the sample should be chosen only
slightly higher than the CMC (see Table 2). Under this condition, the deter-
gent is present either as monomers or as part of protein-detergent complexes,
with very few pure detergent micelles. For porin crystallization, the optimal
range of C8G concentration is narrow: 8-9 mg/ml. Below the CMC (less than
7 mg/ml) or well above it (more than 10 mg/ml) growth rate and nucleation
are excessive (38). The same optimal range was found for C8G with bacterial
260
9: Crystallization of membrane proteins
reaction centres, and it does not seem to be very sensitive to the protein
concentration.
5.2 Additives
Small molecules with amphiphilic character have been sometimes added to
the crystallization media (see Table 3). Most often used is heptane-l,2,3-triol
(high melting point isomer); hexane-l,6-diol, benzamidine, glycerol, and
triethylamine phosphate were also used. Their effects are various.
(a) They may be absolutely required: best example is heptane-l,2,3-triol,
essential for crystallizing Rps. viridis reaction centre with C12DAO.
(b) They improve crystal growth and quality, but crystallization still takes
place in their absence; this is the case of heptane-l,2,3-triol for Rb.
sphaeroides reaction centre with C12DAO.
(c) In other cases their presence has no effect whatsoever; it is the case of
E. coli porin with C8G.
These 'additives' have been usually used at quite high molarities, when
compared to those of detergent; in the cases cited above heptane-l,2,3-triol
was present at about 0.1 M. Their mode of action is still poorly understood.
One hypothesis is that their small size and amphiphilic character allow them
to localize between neighbouring protein molecules, in regions inaccessible to
detergent, filling thus voids in the lattice (39). Another explanation, which has
some experimental support, is they modify the micellar structure of the deter-
gent by partitioning into the micelles (40); thus they change the consolution
boundaries and could bring them in a favourable temperature range.
Detergents with short aliphatic chains and a CMC too high to be used alone
may also be used as additive (41). The addition of another detergent at low
concentration (C8E4) may also have a similar effect to the other type of
additive (31).
Whatever the case, such additives may be tried when all previous trials
done in their absence have failed. It is better to test them beforehand on the
protein in solution; indeed they may have a denaturing effect.
5.4 Optimization
Once crystals (most often microcrystals) have been observed in trial experi-
ments, crystallization conditions have to be improved for crystal size and
quality. The strategy is based on the same principles as for soluble proteins
(see Chapter 4). Excessive nucleation, leading to a 'shower' of microcrystals,
should be avoided; at the same time, growth rate should be kept low enough,
as crystal defects are frequently observed when the rate is too high (hollow
crystals may even be obtained). Practically, this implies repeating the trials
with the different parameters (pH, concentrations, and so on) slightly modi-
fied around their initially positive values, over a fine grid. At this stage, use of
an additive or of a small amount of a second detergent (compare Section 5.2)
may be included as a further variation.
The crystal form of several membrane proteins has been shown to depend
on several parameters: type of detergent, pH, nature of the buffer, and the
ionic strength when PEG together with salt are present (38). By varying these
parameters, it has been possible to select a form which grows better, or is
more suitable for structure determination because of its symmetry or unit cell
dimensions.
6. Experimental techniques
Crystallization of membrane proteins may be performed with all the experi-
mental set-ups described in Chapter 5. Vapour diffusion and microdialysis
have been more frequently used than batch crystallization and free liquid-
liquid interface (see Table 3). Crystallization of cyt bc1 in gel of agarose has
also been described (26).
Because of the wetting properties of detergents, their drops tend to spread
263
F. Reiss-Husson and D. Picot
when formed on a planar glass slide, and to fall out when the slide is inverted.
Therefore, vapour diffusion with hanging drops is restricted to drop volumes
less than 10 ul. With sitting drops formed on depression slides, there is no
restriction in volume.
Microdialysis is performed in capillaries or microtubes closed by dialysis
membrane with sample volumes less than 150 ul, equilibrated against a
reservoir. Choice of the cut-off value of the membrane should take into
account the molecular weights of the components of the sample (see Section
4.2.1). Depending on this cut-off value, and also on the thickness of the
membrane, dialysis rate, and diffusable species may be controlled; e.g. PEG
4000 and 6000 diffuse (but slowly) through a membrane of cut-off value 25 000
daltons, together with water and salts, but not if a cut-off value of 2000 daltons
is used. We have observed that with these two membrane types, all other
conditions being the same, crystallization of a bacterial reaction centre with
PEG 4000 does not occur similarly (unpublished experiments).
Choosing between microdialysis and vapour diffusion is often a matter of
personal preference. Cost of microdialysis is higher when the detergent is
expensive, as detergent must be present in the microdialysis reservoir but may
be omitted from the vapour diffusion reservoir. One of the advantages of the
microdialysis method over vapour diffusion for screening experiments is the
possibility of changing individual constituents of the mixture; furthermore
the detergent concentration may be kept constant throughout the crystalliza-
tion, by putting it at the same concentration in the sample and in the reservoir.
Changing the dialysis reservoir is also very easy.
The main disadvantage of microdialysis is the rapid equilibration between
sample and reservoir (much faster than through vapour diffusion), which may
be troublesome if growth rate has to be slowed down. In that case, double
dialysis (see Chapter 5) is recommended.
Finally, when crystals are obtained, it important to realize that they are
extremely fragile and that their stabilization and manipulation are often diffi-
cult. For example, crystals of prostaglandin H synthase are stable in artificial
mother liquor only with the detergent at its CMC; this value is critical enough
that changes of the CMC due to the addition of salt or sucrose have to be
taken into account (6). Recently, it has been shown that the cryo-crystallo-
graphic techniques used for soluble protein (see Chapter 13) may successfully
be applied to membrane protein, provided that suitable stabilization conditions
are found (46, 49, 50, 76).
7. Conclusion
The structures of several membrane proteins have been solved during the past
few years, some of them to high resolution (51, 52). This has shown that the
methodology originally developed for porin and bacteriorhodopsin has a
more general validity. Furthermore, crystallization conditions worked out for
264
9: Crystallization of membrane proteins
one protein have been successfully used with other proteins, opening the way
to the design of more systematic strategies. Various detergents are suitable.
Their role is important: their properties and their phase diagrams influence
the crystallization conditions; they are still associated with the protein in the
crystal lattice. The methodology requires (as for soluble proteins) a systematic
search over the different parameters, including the nature of the detergent.
This adds one more factor to this empirical analysis. However, good quality
crystals are still not easy to obtain; the difficulties encountered with bacterio-
rhodopsin provide the more vivid example. This has stimulated the search for
alternative approaches; one of them takes advantage of the bicontinuous
cubic phases of lipids where the lipid molecules are arranged in curved three-
dimensional bilayers. Such a phase could incorporate bacteriorhodopsin and
was used as a matrix for its crystallization (53). This allowed well-ordered
crystals to be grown with an improved quality as compared to those pre-
viously grown in detergent solutions (54). It is to be hoped that this method
could be of general use for other membrane proteins. On the other hand, the
problem of maintaining a pure and active protein in solution has been
recently tackled with the use of polymeric amphiphiles (55). Nevertheless,
finding proper expression system to overexpress these types of protein is still a
difficult task (56) that will have to be overcome before membrane protein
structures could flood the Protein Data Bank.
References
1. Michel, H. and Oesterhelt, D. (1980). Proc. Natl. Acad. Sci. USA, 77, 1283.
2. Garavito, R. M. and Rosenbusch, J. P. (1980). J. Cell Biol., 86, 327.
3. Deisenhofer, J., Epp, O., Miki, K., Huber, R., and Michel, H. (1985). Nature, 318,
618.
4. Kuhlbrandt, W. (1988). Q. Rev. Biophys., 21, 429.
5. Michel, H. (ed.) (1991). Crystallization of membrane proteins. CRC Press, Boca
Raton.
6. Garavito, R. M., Picot, D., and Loll, P. J. (1996). J. Bioeng. Biomemb., 28, 13.
7. Song, L. and Gouaux, J. E. (1997). In Methods in enzymology (ed. C. W. Carter
and R. M. Sweet), Academic Press, London. Vol. 276, p. 60.
8. Ostermeier, C. and Michel, H. (1997). Curr. Opin. Struct. Biol., 7, 697.
9. Kuhlbrandt, W. (1992). Q. Rev. Biophys., 25, 1.
10. Roth, M., Lewit-Bentley, A., Michel, H., Deisenhofer, J., Huber, R., and
Oesterhelt, D. (1989). Nature, 340, 659.
11. Roth, M., Arnoux, B., Ducruix, A., and Reiss-Husson, F. (1991). Biochemistry, 30,
9403.
12. Pebay-Peyroula, E., Garavito, R. M., Rosenbusch, J. P., Zulauf, M., and Timmins,
P. A. (1995). Structure, 3, 1051.
13. Bordier, C. (1981). J. Biol. Chem., 256, 1604.
14. Findlay, J. B. C. (1990). In Protein purification applications: a practical approach
(ed. E. L. V. Harris and S. Angal), pp. 59-82. IRL Press, Oxford.
15. Zulauf, M., Furstenberger, U., Grabo, M., Jaggi, P. M. R., and Rosenbusch, J. P.
265
F. Reiss-Husson and D. Picot
(1989). In Methods in enzymology, (ed. S. Fleischer and B. Fleischer), Academic
Press, London. Vol. 172, p. 528.
16. Rosenthal, K. S. and Koussale, F. (1983). Anal. Chem., 55, 1115.
17. De Vendittis, E., Palumbo, G., Parlato, G., and Bocchini, V. (1981). Anal.
Biochem.,115, 278.
18. Van Aken, T., Foscall-Van Aken, S., Castleman, S., and Ferguson-Miller, J.
(1986). In Methods in enzymology, ibid ref. 15. Vol. 125, p. 27.
19. Herrmann, K. W. (1966). J. Coll. Int. Sci., 22, 352.
20. De Grip, W. J. and Bovee-Geurts, R. (1979). Chem. Phys. Lipids, 23, 321.
21. Lund, S., Orlowski, S., de Foresta, B., Champeil, P., le Maire, M., and Moller, J. V.
(1989).J. Biol. Chem., 264, 4907.
22. Lorber, B., Bishop, J. B., and DeLucas, L. J. (1990). Biochim. Biophys. Acta, 1023,
254.
23. Eisele, J.-L. and Rosenbusch, J. P. (1989). J. Mol. Biol., 206, 209.
24. Soulimane, T., Gohlke, U., Huber, H., and Buse, G. (1995). FEBS Lett., 368, 132.
25. Ashani, Y. and Catravas, G. N. (1980). Anal. Biochem., 109, 55.
26. Yu, C.-A., Xia, D., Deisenhofer, J., and Yu, L. (1994). J. Mol. Biol., 243, 802.
27. Garavito, R. M., Markovic-Housley, Z., and Jenkins, J. (1986). J. Cryst. Growth,
76, 701.
28. Kates, M. (1972). In Techniques in lipidology (ed. T. S. Work and E. Work), Vol.
3, Part II. North Holland, Amsterdam.
29. Radin, N. S. (1981). In Methods in enzymology, (ed. J. M. Lowenstein), Academic
Press, London. Vol. 72, p. 5.
30. Bartlett, G. R. (1959). J. Biol. Chem., 234, 466.
31. Garavito, R. M. and Picot, D. (1990). Methods: a companion to methods in
enzymology, 1, 57.
32. M011er, J., Le Maire, M., and Andersen, J. P. (1986). In Progress in protein lipid
interactions (ed. A. Watts and J. J. De Pont), Vol. 2, pp. 147-96. Elsevier,
Amsterdam.
33. Rivas, E., Pasdeloup, N., and Le Maire, M. (1982). Anal. Biochem., 123, 194.
34. Rao, P. and Pattabiraman, T. N. (1989). Anal. Biochem., 181, 18.
35. Kreusch, A., Weiss, M. S., Welte, W., Weckesser, J., and Schulz, G. E. (1991).
J. Mol. Biol., 217, 9.
36. Kleyman, G., Ostermeier, C., Ludwig, B., Skerra, A., and Michel, H. (1995).
Biotechnology, 13, 155.
37. Ostermeier, C., Iwata, S., Ludwig, B., and Michel, H. (1995). Nature Struct. Biol.,
2, 842.
38. Garavito, R. M. and Rosenbusch, J. P. (1986). In Methods in enzymology, ibid ref.
15. Vol. 125, p. 309.
39. Michel, H. (1983). Trends Biochem. Sci., 8, 56.
40. Timmins, P. A., Hauk, J., Wacker, T., and Welte, W. (1991). FEBS Lett., 280, 115.
41. Forst, D., Schulein, K., Wacker, T., Diederichs, K., Kreutz, W., Benz, R., et al.
(1993). J. Mol. Biol., 229, 258.
42. Krauss, N., Schubert, W.-D., Klukas, O., Fromme, P., Witt, H. T., and Saenger, W.
(1996). Nature Struct. Biol., 3, 965.
43. Blaauw, M., Dekker, N., Verheij, H. M., Kalk, K. H., and Dijkstra, B. W. (1995).
FEBS Lett., 373, 10.
44. Ducruix, A., Arnoux, B., and Reiss-Husson, F. (1988). In The photosynthetic
266
9: Crystallization of membrane proteins
bacterial reaction center: structure and dynamics (ed. J. Breton and A. Vermeglio),
Vol. 149, pp. 21-5. Plenum, New York.
45. Ray, W. J. and Puvathingal, J. M. (1985). Anal. Biochem., 146, 307.
46. Kurumbail, R., Stevens, A. M., Gierse, J. K., McDonald, J. J., Stegeman, R. A.,
Pak, J. Y., et al. (1996). Nature, 384, 644.
47. Berry, E. A., Shulmeister, V. M., Huang, L.-S., and Kim, S.-H. (1995). Acta Cryst.,
D51, 235.
48. Weckstrom, K. and Zulauf, M. (1985). J. Chem. Soc. Faraday Trans. 1, 81,
2947.
49. Lee, J. W., Chan, M., Law, T. V., Kwon, H. J., and Jap, B. K. (1995). J. Mol. Biol.,
252, 15.
50. Yu, C.-A., Xia, J.-Z., Kachurin, A. M., Yu, L., Xia, D., Kim, H., et al. (1996).
Biochim. Biophys. Acta, 1275, 47.
51. Weiss, M. S. and Schulz, G. E. (1992). J. Mol. Biol., 221, 493.
52. Song, L., Hobaugh, M. R., Shustak, C., Cheley, S., Bayley, H., and Gouaux, J. E.
(1996). Science, 274, 1859.
53. Landau, E. M. and Rosenbusch, J. P. (1996). Proc. Natl. Acad. Sci. USA, 93,
14532.
54. Pebey-Peyroula, E., Rummel, G., Rosenbusch, J. P., and Landau, E. M. (1997).
Science, 277, 1676.
55. Tribet, C., Audebert, R., and Popot, J.-L. (1996). Proc. Natl. Acad. Sci.USA, 93,
15047.
56. Grisshammer, R. and Tate, C. G. (1995). Q. Rev. Biophys., 28, 315.
57. Frindi, M., Michels, B., and Zana, R. (1992). J. Phys. Chem., 96, 8137.
58. Degiorgio, V. (1985). In Physics of amphiphiles: micelles, vesicles and microemulsions
(ed. V. Degiorgio and M. Cord), pp. 303-35. North Holland, Amsterdam.
59. Michel, H. (1982). J. Mol. Biol., 158, 567.
60. Chang, C.-H., Schiffer, M., Tiede, D., Smith, U., and Norris, J. (1985). J. Mol.
Biol., 186, 201.
61. Allen, J. P. and Feher, G. (1984). Proc. Natl. Acad. Sci. USA, 81, 4795.
62. Yeates, T. O., Komiya, H., Rees, D. C., Allen, J. P., and Feher, G. (1987). Proc.
Natl. Acad. Sci. USA, 84, 6438.
63. Arnoux, B., Ducruix, A., Reiss-Husson, F., Lutz, M., Norris, J., Schiffer, M., et al.
(1989). FEBS Lett., 258, 47.
64. Allen, J. P. (1994). Proteins, 20, 283.
65. Buchanan, S. K., Fritzsch, G., Ermler, U., and Michel, H. (1993). J. Mol. Biol., 230,
1311.
66. Katayama, N., Kobayashi, M., Motojima, F., Inaka, K., Nozawa, T., and Miki, K.
(1994). FEBS Lett., 348, 158.
67. Feiek, R., Ertlmaier, A., and Ermler, U. (1996). FEBS Lett., 396, 161.
68. Papiz, M. Z., Hawthornthwaite, A. M., Cogdell, R. J., Woolley, K. J., Wightman,
P. A., Ferguson, L. A., et al. (1989). J. Mol. Biol., 209, 833.
69. Koepke, J., Hu, X., Muenke, C., Schulten, K., and Michel, H. (1996). Structure, 4,
581.
70. Pauptit, R. A., Zhang, H., Rummel, G., Schirmer, T., Jansonius, J. N., and
Rosenbusch, J. P. (1991). J. Mol. Biol., 218, 505.
71. Kreusch, A., Neubuser, A., Schiltz, E., Weckesser, J., and Schultz, G. E. (1994).
Protein Sci., 3, 58.
267
F. Reiss-Husson and D. Picot
72. Tucker, A. D., Jackman, S., Parker, M. W., and Tsernoglou, D. (1991). J. Mol.
Biol., 222, 881.
73. Stauffer, K. A., Page, M. G. P., Hardmeyer, A., Keller, T. A., and Pauptit, R. A.
(1990). J. Mol. Biol., 211, 297.
74. Hirsch, A., Wacker, T., Weckesser, J., Diederichs, K., and Welte, W. (1995).
Proteins, 23, 282.
75. Tsukihara, T., Aoyama, H., Yamashita, E., Tomizaki, T., Yamaguchi, H.,
Shinzawa-Itoh, K., et al. (1995). Science, 269, 1069.
76. Luong, C., Miller, A., Barnett, J., Chow, J., Ramesha, C., and Browner, M. F.
(1996). Nature Struct. Biol., 3, 927.
77. Gouaux, J. E., Braha, O., Hobaugh, M. R., Song, L., Cheley, S., Shustak, C., et al.
(1994). Proc. Natl. Acad. Sci. USA, 91, 12828.
268
10
From solution to crystals with a
physico-chemical aspect
M. RIES-KAUTT and A. DUCRUIX
1. Introduction
Biological macromolecules follow the same thermodynamic rules as inorganic
or organic small molecules concerning supersaturation, nucleation, and crystal
growth (1). Nevertheless macromolecules present particularities, because the
intramolecular interactions responsible of their tertiary structure, the inter-
molecular interactions involved in the crystal contacts, and the interactions
necessary to solubilize them in a solvent are similar. Therefore these different
interactions may become competitive with each other. In addition, the bio-
logical properties of biological macromolecules may be conserved although
the physico-chemical properties, such as the net charge, may change depend-
ing on the crystallization conditions (pH, ionic strength, etc.). A charged
biological macromolecule requires counterions to maintain the electro-
neutrality of the solution; therefore it should be considered as a protein (or
nucleic acid) salt with its own physico-chemical properties, depending on the
nature of the counterions.
To crystallize a biological macromolecule, its solution must have reached
supersaturation which is the driving force for crystal growth. The understand-
ing of the influence of the crystallization parameters on protein solubility of
model proteins is necessary to guide the preparation of crystals of new
proteins and their manipulation. Only the practical issues are developed in
this chapter, and the reader should refer to recent reviews (2-4) for a de-
scription of the fundamental physical chemistry underlying crystallogenesis.
2.1 Solubility
Solubility is defined as the amount of solute dissolved in a solution in equilib-
rium with its crystal form at a given temperature. For example, crystalline
ammonium sulfate dissolves at 25°C until its concentration reaches 4.1 moles
per litre of water, the excess remaining non-dissolved. More salt can be
dissolved when raising the temperature, but if the temperature is brought
back to 25°C, the solution becomes supersaturated, and the excess of salt
crystallizes until its concentration reaches again its solubility value at 25°C
(4.1 moles per litre of water).
In the case of biological macromolecules, the solubility is additionally
defined by the characteristics of the solvent. Proteins are mostly solubilized in
water which acts through hydrogen bonds. In some cases another protic
solvent (an alcohol) or an aprotic solvent (e.g. acetone, DMSO, dioxane,...)
is added at low concentration. In addition, the solvent solutions contain at
least the ubiquitous buffer used to fix the pH of the solution and therefore the
net charge of the protein. Salts are added not only to ensure an ionic strength
but most often to reach supersaturation.
Throughout this chapter, protein solubility is defined as the concentration
of soluble protein in equilibrium with the crystalline form at given tempera-
ture and pH values, and in the presence of a given concentration of solvent
compounds others than the protein (i.e. water, buffer, crystallizing agents,
stabilizers, additives). The solubility values depend on the physico-chemical
characteristics of the protein itself (hydrophilicity, net charge, type of solvent
exposed residues) and of the solvent (pH, dielectric constant, ionic strength,
concentration, and nature of the additives).
Figure 1 illustrates the variability of protein solubilities, depending on the
protein itself or on the protein salt (e.g. different lysozyme salts). The
solubility of the three proteins: bovine pancreatic trypsin inhibitor (BPTI) (5)
in ammonium sulfate, collagenase from Hypoderma lineatum (Hl) in
ammonium sulfate (6), and hen egg white (HEW) lysozyme in NaCl (7), cover
a very large range of both ionic strength and solubility values, in their
standard crystallization conditions. Furthermore the solubility of a same
protein, HEW lysozyme, can be changed drastically when changing the nature
of the crystallizing salt, as shown by the solubility curves of lysozyme/KSCN,
lysozyme/NaCl, and lysozyme/NH4OAc.
In the literature other conventions of defining solubility are encountered; it
may be the protein concentration measured before the actual equilibrium is
reached, or it can be evaluated in the presence of precipitate instead of crystals
(8). Their applications are discussed at the end of this section.
270
10: From solution to crystals with a physico-chemical aspect
Figure 1. Solubility curves at 18°C of BPTI (5), collagenase from Hypoderma lineatum (6),
and HEW lysozyme (7, 10). The crystallizing agents are AS (ammonium sulfate), NH4OAc
(ammonium acetate), NaCI (sodium chloride), and KSCN (potassium thiocyanate).
The zone of the solubility diagram where crystals appear (nucleation zone)
depends on the supersaturation, which is the ratio, Cp/Cs, of the protein
concentration over the solubility value, but also on the kinetics to reach these
conditions.
The protein purity (see Chapter 2) must be checked before doing any
screening experiments. In a mixture of proteins, the first crystals contain the
most supersaturated protein and this may not be the most concentrated one.
Furthermore the solubility of the major protein may be affected differently
from the contaminant when changing a parameter.
Method
1. Define one parameter to vary (e.g. salt concentration), keeping all
others strictly constant (e.g. pH, temperature, nature of salt and
buffer).
2. Choose at least four values of the variable (different salt concentra-
tions over a large range), because solubility curves usually do not fit
with linear curves.
3. Set up the batch experiments (> 10 ul) in duplicate at two or three
different initial protein concentrations for a given parameter value.
4. Follow the decrease of the protein concentration of the supernatant,
by withdrawing a crystal-free aliquot of the duplicate set-up, each
week for optical density (OD) measurements. If microcrystals are
present, filter or centrifuge the aliquot before the dilution for the OD
measurement.
Once crystallization has started, the protein concentration in the
supernatant will converge to a constant value, solubility, with time.
This value is identical for the different initial protein concentration at a
same ionic strength.
5. When the protein concentrations remain constant for at least two
weeks and are identical for the different initial protein concentration at
the same ionic strength, confirm the measurement by testing the
original undisturbed set of experiments.
i. Column method
A solution of either supersaturated or undersaturated protein solution is
poured in a microcolumn filled with crystals. An aliquot of the solution is
periodically withdrawn from the bottom of the column to follow the change of
the protein concentration by optical density measurements. This method (11)
is based on the maximization of the exchange between the available crystal-
line surface area and minimal free solution volume to reach equilibrium. It
overcomes the problem of prolonged equilibration time, as equilibrium, i.e.
the solubility value, is reached within one to five days.
Two microcolumns are run in parallel: one for crystallization (super-
saturated state), one for dissolution (undersaturated state). This method is
mostly appropriate for the study of the influence of temperature. To deter-
mine protein solubility at different concentrations of a given salt, crystals can
be prepared from a same batch, but must then be equilibrated carefully at
respective salt concentrations. When solubility is checked in different salts, a
batch of crystals is prepared in each appropriate salt.
iii. Scintillation
A thermoregulated cell (50-100 ul) is filled with the solution to crystallize.
The bath temperature is changed until crystallites occur inducing scintillation
which is detected by a photodiode signal. The temperature is changed back-
wards and forwards to define the solubility limit, defined by the appearance
and disappearance of the crystallite detected by the scintillation signal. One
solubility value is obtained within approximately 12-24 hours. In addition this
technique (13) allows crystallization induction times to be measured which
were shown to follow supersaturation.
v. Michelson interferometry
A Michelson interferometer is used for the observation of concentration
gradients around a crystal to determine whether the crystal is growing or
dissolving when changing the temperature (15). The volume of the cell is
about 70 ul. The equilibrium temperature is obtained within two hours.
Method
1. Open the coverslip with the drop in which crystals have grown for at
least two weeks.
2. Withdraw 1 (preferably 2) ul of clear supernatant under the binocular.
If too many microcrystals are present, centrifuge the drop and take the
aliquot from the clear supernatant to avoid diluting crystals.
3. Dilute to the minimal volume required for an OD measurement at
280 nm. Eventually measure the protein concentration by the Bradford
method using a coloured dye to increase the absorbance of the
protein-dye complex.
aWhen using the dialysis technique, an aliquot can be withdrawn from the dialysis cell with a
Hamilton syringe. The dialysis cell can of course no longer be immersed in the reservoir, but it can
be rescued in a vessel with some reservoir solution around in order to avoid the solution drying.
Figure 3. Residual protein concentration in a drop after a vapour diffusion process, and
selection of the seeding conditions.
Once the residual protein concentration is measured, news drops are pre-
pared at protein concentrations ranging from P1 to P5, as indicated on
Figure 3, directly at the salt concentration of the reservoir (batch). The
concentrations P3 to P5 aim at covering the searched metastable zone.
The protein concentration P1 (< solubility) is added to have a better
estimate of the lower limit of the metastable zone. Crystals dissolve in
undersaturated drops (< P1) and remain unchanged in saturated drops
(between P1 and P2). They grow in slightly supersaturated drops (P2 to
P4, where you should seed) whereas new nucleation occurs in more
supersaturated conditions (> P4-P5).
(f) Mounting the crystals. When a crystal is recovered from a drop, it is in a
solution of a given protein and crystallizing agent concentrations as
shown in Figure 2. Very often reservoir solution is used to transfer the
crystal. In fact, this can be done safely only if the remaining protein
concentration in the drop is lower than = 0.5 mg/ml (e.g. B in Figure 2). If
the solubility value is higher (e.g. A in Figure 2), then the crystal would
start to dissolve as the reservoir contains no protein. Knowing the
residual protein concentration in the mother liquor gives the amount of
protein to introduce in additional mounting solutions.
Similarly protein should be added when soaking crystals in cryo-
protectant solutions for cryo-crystallography when crystals dissolve. Cryo-
277
M. Ries-Kautt and A. Ducruix
protectants often change the protein solubility. Although the knowledge of
the residual protein concentration of the crystallization drop is not repre-
sentative of the solubility of the protein in the cryo-protectant solution,
crystals should be soaked in a series of cryo-protectant solutions containing
a protein concentration higher than the residual protein concentration.
278
10: From solution to crystals with a physico-chemical aspect
reservoir solution without the macromolecule leads to the dissolution of
the macromolecule crystals.
(b) Below the solubility curve the solution is undersaturated, the system is
thermodynamically stable, and the biological macromolecule will never
crystallize.
(c) Above the solubility curve, the concentration of the biological macro-
molecule is higher than the concentration at equilibrium. This corresponds
to the supersaturation zone. A supersaturated macromolecule solution
contains an excess of macromolecule which will appear as a solid phase
until the macromolecule concentration reaches the solubility value in the
solution (supernatant). In some cases the excess of macromolecule may
concentrate in oily drops in a liquid-liquid separation. The rate of
supersaturation is defined as the ratio of the biological macromolecule
concentration over the solubility value. The higher the supersaturation
rate, the faster this solid phase appears.
It is often difficult to understand how supersaturated macromolecule
solutions are achievable. In terms of molarity, it must be remembered that
macromolecule solubilities are very low (uM to mM) compared to small
molecules or inorganic molecules (mM). This also corresponds to very low
volume fractions of solute which allow macromolecule solutions to be
prepared at supersaturations as high as 10 to 20 times the solubility. For small
molecules supersaturation of only 1.1 to about 1.5 are achievable. Recently
we observed crystallization of HEW lysozyme at supersaturations around 1.5
when working with 400 mg/ml (28 mM) protein which correspond to 30%
volume fraction (17).
However the higher the supersaturation, the faster the solid phase appears
in the solution, as described below. Contrary to macromolecule purification
which implies precipitation, crystallization requires an accurate control of the
level of supersaturation. This allows nucleation of crystals, i.e. a solid phase
with a three-dimensional periodicity, by controlling the nucleation rate to
yield few single crystals.
2.3.1 Precipitation zone
Precipitation occurs at very high supersaturation (= 30 to 100 times the
solubility value for HEW lysozyme). Insoluble macromolecules rapidly sep-
arate from the solution in an amorphous state. If the solution is centrifuged,
the supernatant is in fact still supersaturated and crystallization may occur. To
differentiate amorphous precipitate from microcrystals, fresh drops can be
seeded (see Chapter 7) with this material; amorphous precipitate dissolves
whereas microcrystals grow.
2.3.2 Nucleation zone
At a sufficient supersaturation, nucleation spontaneously occurs, once critical
activation free energy is overcome. This is called homogeneous nucleation.
279
M. Ries-Kautt and A. Ducruix
Crystallization often occurs at lower saturation when it is induced by vibra-
tions or the presence of particles (dust, precipitate, irregularities of crystal-
lization cell); it is then called heterogeneous nucleation. The latter is usually
characterized by non-reproducibility, therefore it is recommended to filter all
solutions and blow the coverslips with an air stream before setting up the
hanging drops.
Nucleation requires a lower supersaturation than precipitation. To give an
order of magnitude, the nucleation range for HEW lysozyme is = 5 times the
solubility for dialysis and batch crystallizations, and about 10 times the
solubility for vapour diffusion. Crystals appear faster and in larger numbers
with increasing supersaturation. High supersaturation may be useful to find
the nucleation zone, but growing crystals for X-ray diffraction may benefit
from a search of the optimal supersaturation where few but large crystals are
grown.
The nucleation rate, defined as the number of nuclei formed per unit
volume and unit time, is linked (1, 18) to:
(a) Supersaturation, as illustrated by the curve A of Figure 5. For super-
saturations higher than B*, the critical supersaturation, nucleation occurs.
When increasing the supersaturation, the number of crystals increases.
Figure 5. Nucleation rate versus supersaturation for A, high solubility, and B, low
solubility. The curves A and B delimit the metastable zone from the nucleation one.
280
10: From solution to crystals with a physico-chemical aspect
(b) The number of molecules per unit volume. When the solute is sparingly
soluble, the solution remains in a metastable state over long periods.
Nucleation requires much higher supersaturation to occur (Figure 4).
Once B* is reached, the nucleation rates becomes drastic, as illustrated by
the curve B in Figure 5. The different curves A and B of Figure 5
correspond respectively to the situations A and B of Figure 2.
Crystallization conditions for which the solubility is very low should be
avoided. The precipitation curve is then very close to the solubility curve, the
domain of crystallization becomes very narrow, which brings difficulties in
defining the right conditions for growing large crystals. This is the case for the
crystallization of HEW lysozyme/KSCN, where the crystallization zone is
limited to a range of 100 mM KSCN. On the counterpart, HEW lysozyme/
Nad crystallizes over a broad range of 1400 mM. To enlarge the nucleation
zone, solubility must be increased. This can be done by:
• decreasing the ionic strength while increasing the macromolecule concen-
tration, if the slope of the solubility curve is smooth enough
• using another salt in which the solubility is higher (e.g. from HEW
lysozyme/KSCN to HEW lysozyme/NaCl shown in Figure 1)
• changing the pH to increase the protein net charge
• changing the temperature.
Figure 6. Variations of the protein and salt concentration in four drops A to D during
equilibration with their reservoirs for a classical 1:2 ratio. The bold arrows start at the
initial conditions in the drops, and end at the expected final conditions, if no
crystallization occurs during the drop/reservoir equilibration.
282
10: From solution to crystals with a physico-chemical aspect
with a typical 1:2 ratio for the salt concentration initially in the drop and in the
reservoir. In condition A the salt gradient between the initial drop conditions
and the reservoir is 0.5 M whereas it is 0.8 M in D. This implies a faster
equilibration for A than for D. As a consequence the drop/reservoir equilib-
ration may become much faster than the time lag for nucleation. Therefore
nucleation occurs at a higher supersaturation, implying a higher nucleation
rate. To help the system to crystallize before the drop/reservoir equilibration
is achieved, the experiment may be run in two steps of lower gradient, or with
a larger drop/reservoir distance, or in ACA instead of Linbro plates. Similarly
for dialysis, capillaries may be preferred to Cambridge buttons, because the
protein solution equilibrates more slowly with the reservoir solution, allowing
nucleation to occur before equilibration is achieved.
2.4.3 Equilibration to reach the solubility value
Once crystallization has started, the protein concentration in the solution
decreases until it reaches the solubility value. These kinetics depend on the
growth kinetics of a given crystal form of a given biological macromolecule,
the number of crystals growing, the amount of protein to crystallize, and
stirring (or not) the solution.
For the photochemical reaction centre of Rhodobacter sphaeroides (20), the
solubility value was reached within 12 days even though the unstirred batch
method was used. In this case the crystals grew very quickly.
For tetragonal HEW lysozyme crystals, equilibration requires up to nine
months if using unstirred batch methods. The delay can be reduced to two
months by stirring (17) the crystallization vials. However this was not
successful for very high protein concentrations which are very viscous. These
experiments were kept for one month at low temperature to accelerate the
crystallization before letting them equilibrate at 18°C.
3. Proteins as polyions
Throughout the process of crystallization and of structure determination, a
protein must be considered as a polyion of a given net charge, surrounded by
counterions. The number of counterions is at least equal to the net charge to
ensure the electrostatic compensation. Even though the biochemical activity
may not be altered, the physico-chemical properties of a protein, and hence its
behaviour in crystallization, may be changed significantly by variations of its
net charge or by adsorption of small molecules or ions onto its surface. It is
useful to begin a new crystallization project by first calculating the probable
net charge of the protein as a function of pH, although it is an approximation.
Additives are well defined solvent constituents (chemical nature and con-
centration); otherwise they are impurities. To improve the reproducibility of
crystallization experiments and the reliability during a structural investigation
involving different protein batches, impurities should be eliminated whenever
283
M. Ries-Kautt and A. Ducruix
possible (see Chapter 2). This is usually done for biological contaminants, but
seldom for small molecules or salts. After a routine purification step, aqueous
protein solutions contain various additives, including at least a buffer, various
salts from elution gradient, and additives to prevent oxidation of free SH
groups, EDTA, NaN3, stabilizers, etc. Some compounds or ions may be bound
by the protein. Proteins from commercial sources often contain organic or in-
organic compounds whose nature depends on the source, and the amount on
the batch. Up to 14% (w/w) salt were observed for commercial HEW lysozyme
(21), although the purity was otherwise excellent in terms of biological activity.
All the uncontrolled organic and inorganic solutes should be eliminated or
replaced by co-ions and counterions of known concentration and nature.
a The charge contribution z of the charged residue is given in each column for the indicated pH value. For a given protein. note: ni, the number of the amino acid
in each corresponding column. Multiply zi by ni of each column. Sum all charges for a given pH (row) to obtain the net charge at this pH.
M. Ries-Kautt and A. Ducruix
In Table 1 the contribution zi of a given type of charged residue at a given
pH value appears in each column. For a given protein, each column has to be
multiplied by n, the number of a given type of amino acid. Summing all
charges for a given pH (row) gives the net charge at this pH.
The estimation of a protein net charge is also accessible on the web at:
• http://www-biol.univ-mrs.fr/d_abim/compo-p.html
• http://www.expasy.ch/sprot/protparam.html
• http://www.infobiogen.fr/service/deambulum
The calculated pI (pH for a net charge of zero) should be supplemented by
the electrophoretic measurement of the experimental pI. If a difference is
observed between the estimated pI and the experimental one, it means that
one of the conditions detailed for the estimation is not met or that some
additives of the experimental conditions interact with the protein.
Water soluble proteins can be classified broadly, according to their pI
(experimental or estimated from the content of charged residues), as:
(a) Acidic proteins, having a higher content of Asp and Glu, than His, Lys,
and Arg. Their pI is lower than 6. This arbitrary value is linked to the pKa
value of the basic group His.
(b) Basic proteins, with a higher content of His, Lys, and Arg, than Asp and
Glu. Their pI is above 7.5-8.
(c) Neutral proteins containing roughly equal numbers of acidic and basic
residues, and therefore presenting a pI near the neutrality.
In the pH range between pH 6-8, acidic proteins bear a negative net charge,
basic proteins a positive one, and neutral proteins a net charge of about zero.
This classification does not include membrane proteins, which naturally
occur in hydrophobic, lipid environments and which are known to be poorly
soluble in water (see Chapter 9). Their solubilization requires detergents, and
the physical chemistry of protein-detergent systems is very different from the
discussion of this chapter. Nevertheless, the preceding discussion does apply
to their water soluble surfaces, and should therefore also be considered in
those studies.
Method
1. Fill two 5 ml syringes with 1.5-3 ml of respectively Bio-Rad AG 50W-X8
20-50 mesh, H+ form for the cation exchange, and Bio-Rad AG 1-X8
20-50 mesh, OH" form for the anion exchange. Rinse five times with
287
M. Ries-Kautt and A. Ducruix
Protocol 3. Continued
1 ml of pure water. Minimize the dead volume of water to avoid
dilution of the protein sample.
2. Aspirate the dialysed protein solution (< 1 ml) in the syringe
containing the cation exchange resin. Shake the syringe for = 5 min,
then remove the solution from the syringe through a 0.22 um filter.
The pH of the solution becomes more acidic (pH = 3-4), depending on
how extensive the dialysis was.
3. Aspirate the acidic protein solution in the syringe containing the anion
exchange resin. Shake the syringe for = 5 min, then remove the
solution from the syringe through a 0.22 um filter.
4. Aspirate 1 ml of pure water into the first syringe and shake for a few
minutes to recover protein remaining in the dead volume. Remove this
solution and rinse the second syringe. This step is repeated twice for a
better recovery of the protein.
5. Isoionic protein solutions can be rapidly deep-frozen in liquid nitrogen
and freeze-dried for storage. The freeze-dried protein is stored at
-80°C.
6. To prepare protein solutions, solubilize the isoionic protein powder in
pure water, centrifuge the solution, and filter it to remove insoluble
protein. Adjust to the required pH and add desired additives.
Figure 7. Example of the change of the protein net charge versus pH for three basic
proteins. The pH for which the net charge is zero is the pI. At pH 4.5, the net charge of
BPTI is higher than for the two other proteins.
Figure 8. Crystallization conditions (bold segments) for three basic proteins, BPTI,
erabutoxin, and HEW lysozyme, at pH 4.5 and 18°C.
290
10: From solution to crystals with a physico-chemical aspect
4.2 pH
A change of pH implies a change of the protein net charge:
(a) Near the pKa of most numerous charged residues, solubility varies very
rapidly.
(b) Outside the range of the pKa values of charged residues, the solubility
changes smoothly.
(c) Solubility is minimal at the pI of the protein as shown in the case of insulin
(28) (Figure 9), egg albumin (29), haemoglobin (30), and B-lactoglobulin
(31). Conversely solubility is higher when the net charge increases (17).
293
M. Ries-Kautt and A. Ducruix
Figure 9. Variation of the solubility of insulin versus pH (redrawn from ref. 28). Note that
decreasing the pH by 1 unit (i.e. from pH 5 to pH 4), the net charge changes from +1 to
+5 and the solubility from = 0.4 to = 1.2 mg/ml. When increasing the pH by the same
increment (+1 unit, but from pH 5 to pH 6), the net charge changes from +1 to only-0.5
and the solubility from ~ 0.4 to only = 0.6 mg/ml.
4.3.1 Salting-in
Solubility data of carboxyhaemoglobin (Figure 10a) showed that protein
solubility first increased (salting-in) and then decreased (salting-out) with
increasing ionic strength (34). This phenomenon is explained by the decrease
of the chemical activity of the protein when the ionic strength of its environ-
ment increases (36). It is worth emphasizing that the solubility variation of
carboxyhaemoglobin at 25°C and pH 6.6 corresponds to its minimal
solubility, i.e. near the pI.
As for HEW lysozyme (Figure 10b) bearing a net charge different from
zero, no salting-in could be evidenced (17). Here, the screening of the salt on
the electrostatic protein-protein interactions seems to dominate the effect of
the protein chemical activity. Furthermore, salting-in may be reduced or
emphasized depending whether co-ions or counterions bind to the proteins, as
will be discussed in Section 4.4.
4.3.2 Salting-out
Salting-out corresponds to a decrease of protein solubility at high ionic
strength, where the protein behaves as a neutral dipole and solubility is
mainly governed by hydrophobic effects. Theoretically a crystallizing agent
added to the protein-water (solute-solvent) system can either bind to the
protein (preferential binding) or be excluded (preferential exclusion) depend-
ing on preferential protein-additive or protein-water interactions (37). The
net interaction of salting-out is preferential exclusion, even though molecules
or additives may bind to the protein.
Protein solubility has been expressed (34) according to:
where S is the protein solubility (in mg/ml), m the molal salt concentration
(g salt/1000 g water), and 3 the intercept at m = 0. 3 is a constant at high salt
concentration and function of the net charge of the protein, thus strongly pH-
dependent. Therefore it is minimal at the isoelectric point. The magnitude of
3, as well as charge distribution, varies with temperature. Ks is the salting-out
constant. It is independent of pH and temperature, but depends on the nature
of the salt.
However, experimentally defined solubility curves rarely fit with a linear
296
10: From solution to crystals with a physico-chemical aspect
Figure 10. Variation of the solubility versus ionic strength. (a) Salting-in for Carboxy-
haemoglobin near its pl and in the presence of various salts (redrawn from ref. 34). (b) No
salting-in for positively charged lysozyme in the presence of NaCl and at different pH
values (17).
297
M. Ries-Kautt and A. Ducruix
function (Figure 1). This non-linearity may be essentially due to the following
reasons:
(a) The solubility tends to the solubility value of the protein in the buffer at
low concentrations of the crystallizing agent, the efficiency of the crystal-
lizing agent, and of the buffer becoming comparable. The curves of HEW
lysozyme solubility in the presence of various salts (10) converge at low
salt concentrations toward its solubility curve in the sodium acetate
buffer.
(b) Higher amounts of protein are required for crystallization at low salt
concentrations, so the solubility value can be affected by the presence of
higher amounts of protein related salts; they can either be counterions, or
salts which were not eliminated by a previous dialysis step.
(c) Protein binding of counterions can no more be neglected compared to
preferential exclusion. The solubility is then affected by the change of
both the net charge of the protein and the ionic strength.
Method
1. Prepare five dialysis buttons with protein solution at a concentration
as high as possible.
2. Prepare 2 ml reservoir solutions at 0.05, 0.1, 0.2, and 0.5 M of KSCN,
and at 0.1, 0.5, 1.0, and 2.0 M ionic strength of NaNO3, NaCI, NaOAc,
and NH4S04 in the same buffer as the protein.
300
10: From solution to crystals with a physico-chemical aspect
3. Place a dialysis button in the reservoir at the lowest concentration of
each salt. Close the well with a coverslip.
4. Transfer it after at least 2 h to the next concentration. Respect the
same delay for each change.
5. As soon as precipitation is observed, place the dialysis button in the
previous reservoir and prepare an intermediate concentration to refine
the value of the precipitation limit.
The relative position of phosphate and citrate versus sulfate in the series
may change, depending on the pH, since their ionic strength varies rapidly
around their pKa values with Zi2 according to Equation 1.
4.5 Temperature
The variation of protein solubility with temperature may be either direct,
i.e. increasing with temperature, or retrograde. The behaviour of protein
solubility with temperature cannot been foreseen, and is not characteristic,
neither of the protein, nor of the crystallizing agent, but of the protein salt as
illustrated with BPTI. This protein has a retrograde solubility change with
temperature in the presence of ammonium sulfate (5) and sodium chloride
(52), but direct with potassium thiocyanate (5) (Figure 11),
Figure 11. Solubility behaviour of different BPTI salts versus temperature. It is retrograde
with ammonium sulfate (AS) and NaCl, but direct with KSCN. Redrawn from refs 5 and 52.
Method
1. Take a series of screening experiments (vapour diffusion, batch, or
dialysis) covering the nucleation zone, i.e. from clear drops to ones
containing slight precipitation at 18-20°C.
302
10: From solution to crystals with a physico-chemical aspect
2. Introduce the Linbro plate in a Styrofoam box and place these
experiments in an incubator at 4°C.
3. On the next day, check whether the nucleation or the precipitation has
increased or is reduced compared to the initial observations.
4. Bring the experiments again to 18-20°C for a day or two to check the
reversibility.
5. Place these experiments in an incubator at 37°C for another day or
two.
6. Check again whether the nucleation or the precipitation has increased
or is reduced.
7. Analyse the results:
(a) No change whatever the temperature. This may indicate that:
• The sampling of ionic strength or pH is too large (and the
nucleation zone very small). Repeat the experiments with drops
differing by smaller steps of ionic strength or pH.
• The ionic strength is too high. Repeat the experiments with
drops at lower ionic strength and higher protein concentration.
• The pH of crystallization is too close to the pl. Repeat the experi-
ments with drops at lower or higher protein net charge.
(b) The lower the temperature, the more intense the precipitation.
Solubility is directly related to temperature. Verify nevertheless
the reversibility of the precipitation by bringing the experiments to
a higher temperature, to avoid confusion between precipitation
which is reversible, with denaturation which may not be
reversible.
(c) The lower the temperature, the less intense the precipitation.
Solubility is retrograde with temperature, at least with this salt
combination. Repeat the experiments when changing the
crystallizing agent.
Likely to the effect of pH, the variation of protein solubility with tempera-
ture is amplified at low ionic strength (33, 53). To benefit the effect of
temperature this means working at rather low ionic strength, but conversely
carrying out experiments at higher ionic strength for the transport of crystals
with their crystallization solution to stabilize them (i.e. to avoid further
nucleation or dissolution of the crystals).
Method
1. Choose a range of crystallizing agent concentration to cover the
nucleation zone, i.e. from clear drops to those containing slight pre-
cipitation, at a constant protein concentration. As an example, these
reference conditions may be 20 mg/ml protein and reservoirs from
0.8-1.3 M ammonium sulfate (with 0.1 M steps for the reservoirs).
2. Prepare a series of six drops in the reference conditions in the first row
of a Linbro plate (A1 to A6).
3. Set up the second row at the same protein concentration (B1 to B6),
but with reservoirs where 0.2 M ammonium sulfate is replaced by
0.3 M NaCI. It is recommended to take into account the ionic strength
rather than molarity (see Section 4.3) when replacing crystallizing
agents.
4. Observe whether crystallization occurs at higher or at lower ionic
strength in the second row compared to the first one:
(a) The nucleation zone is shifted by one column to lower ionic
strength. NaCI is slightly more efficient than ammonium sulfate.
Continue by setting up the third row while replacing 0.4 M
ammonium sulfate by 1.2 M NaCI.
(b) The nucleation zone starts already in B1. NaCI is much more
efficient than ammonium sulfate. Set up the third row to centre
again the nucleation zone by replacing 0.4 M ammonium sulfate
by only 0.4 M NaCI and starting at lower ionic strength (from
1.2-2.7 M total ionic strength instead of 2.4-3.9 M total ionic
strength).
5. Set up the last row at the same protein concentration (D1 to D6), but
with reservoirs containing only NaCI at concentrations chosen
depending on the previous results to centre the nucleation zone.
5. Crystallization
The aim of crystallization experiments is first to locate the nucleation zone,
then to optimize the physico-chemical parameters and the kinetics to grow
large single crystals. Testing a large number and combination of variables may
305
M. Ries-Kautt and A. Ducruix
yield different crystal forms. It is worth optimizing different polymorphs
because the number of molecules in the asymmetric unit and the diffraction
quality may be very different.
Method
1. According to the estimation of the variation of net charge with pH (see
Section 3.1), select one pH close to the pI, and two on both sides of the
pI so that the net charge is about the same value, but of opposite sign.
When this is in conflict with the stability of the protein, select three
other pH values for which the variation of the protein net charge is as
large as possible. Taking the example of BPTI (Figure 7), a pH higher
than the pI would not be well suited. Therefore one would select pH
10.5, 9, and 4.5 for which the net charge of this protein is 0,
=>7 + , and = 14+ respectively.
2. Prepare three Linbro plates one for each pH. Fill the six reservoirs of
the first row (A) of each Linbro plate with 2 ml of buffer corresponding
to the pH of each plate.
3. Prepare three dialysis buttons filled with the stock protein solution.
Introduce one of them in A1 of each Linbro plate. Observe if the
protein solution remains clear for a few days. If precipitation occurs:
(a) It happens for the lower net charge. Prepare a dialysis button at
the lower protein concentration until the drop remains clear for a
week.
(b) It seems not to be linked to the net charge. Transfer the sample in
the original buffer to search for reversibility of the precipitation.
Replace the buffer solution twice or more to ensure a good
exchange of the buffer solution. If the precipitate remains, check
for possible denaturation and choose another pH for the following
steps.
4. Prepare 15 dialysis buttons filled with the stock protein solution.
307
M. Ries-Kautt and A. Ducruix
Protocol 7. Continued
Introduce them in the remaining A2 to A6 reservoirs. Let them stand
for one day. Eventually change the reservoir when using large protein
volumes to ensure the buffer exchange.
5. Fill the six reservoirs of the row B with six different crystallizing agents
(2 ml), each in the appropriate buffer of a given Linbro plate. The
crystallizing agents should be chosen according to the protein net
charge (see Section 4.4.1). They should preferably be of different
chemical types; thiocyanate, halide (Cl-, Br-, l-, or F-), carboxylate
(acetate, citrate, tartrate), sulfate (or phosphate), PEG, divalent cation
(Mg2+ or Ca2+). As a rule of thumb, the concentration of the first
reservoir may be 0.1-0.5 M, or 5-10% PEG.
6. Transfer the dialysis buttons from row A to row B. After two to five
days, observe the protein solutions:
(a) Case 1: the solution is clear. Prepare the next reservoir C at twice
the concentration of the one in B.
(b) Case 2: the solution precipitates. Prepare the next reservoir C at
half the concentration of B. Transfer the dialysis button, first back
to row A to dissolve the precipitate, then to C.
(c) Case 3: the solution B is neither clear, nor precipitated. Wait for
another period of two to five days to decide whether the next
reservoir concentration should be increased or decreased by only
10%.
7. Continue until the limits between clear solutions and precipitation (i.e.
lower and upper limits of the nucleation zone) are defined.
8. Set up a new set of experiments to refine the concentration of each crys-
tallizing agent at each pH with a small step in between the nucleation
zone limits. At this stage, the vapour diffusion may be more suitable.
5.1.4 Optimization
Once the nucleation zone is defined, the optimal conditions to grow large
single crystals must be sought. At this step the protein/reservoir equilibration
kinetics should be included among the variables to be adjusted. The tools to
perform optimization can also be found in Chapter 4.
5.2 Polymorphism
Table 4 illustrates the variety of crystal forms observed for lysozyme when
changing the crystallization conditions (temperature, pH, nature of the
crystallizing agent) or a combination of them. Apart from the crystal form,
also the number of molecules in the asymmetric unit can change. Both can
present an advantage for the crystallographer.
308
Table 4. Polymorphism of HEW lysozyme
References
1. Boistelle, R. and Astier, J.-P. (1988). J. Cryst. Growth, 90, 14.
2. Rosenberger, F., Vekilov, P. G., Muschol, M., and Thomas, B. R. (1996). J. Cryst.
Growth, 168,1.
3. Chernov, A. A. (1997). Phys. Rep., 288,61.
4. Ries-Kautt, M. and Ducuix, A. (1997). In Methods in enzymology (ed. C. Carter
and R. Sweet), Academic Press, London. Vol. 276, Part A, Chap 3, p. 23.
5. Lafont, S., Veesler, S., Astier, J.-P., and Boistelle, R. (1997). J. Cryst. Growth, 173,
132.
6. Carbonnaux, C., Ries-Kautt, M., and Ducruix, A. (1995). Protein Sd., 4, No. 10,
2123.
7. Broutin, I, Ries-Kautt, M., and Ducruix, A. (1995). J. Appl. Crystallogr., 28,
614.
8. Schein, C. H. (1990). Biotechnology, 8, 308.
9. Ataka, M. and Tanaka, S. (1986). Biopolymers, 25,337.
10. Ries-Kautt, M. and Ducruix, A. F. (1989). J Biol. Chem., 264, 745.
11. Cacioppo, E., Munson, S., and Pusey, M. L. (1991). J. Cryst. Growth, 110, 66.
12. Boistelle, R., Astier, J.-P., Marcis-Mouren, G., Desseaux, V., and Haser, R.
(1992). J. Cryst. Growth, 123,109.
13. Rosenberger, F., Howard, S. B., Sower, J. W., and Nyce, T. A. (1993). J. Cryst.
Growth, 129,1.
14. Gripon, C., Legrand, L., Rosenman, L, Vidal, O., Robert, M.-C., and. Boue, F.
(1997). J. Cryst. Growth, 177,238.
15. Sazaki, G., Kurihara, K., Nakada, T., Miyashita, S., and Komatsu, H. (1996).
J. Cryst. Growth, 169, 355.
16. Darcy, P. A. and Wiencek, J. M. (1998). Acta Cryst., D54, 1387.
17. Retailleau, P., Ries-Kautt, M., and Ducruix, A. (1997). Biophys. J., 73, 2156.
18. Feher, G. and Kam, Z. (1985). Methods in enzymology (ed. S. P. Colowick and N.
O. Kaplan), Vol. 114, pp. 77-112.
19. Luft, J. R. and DeTitta, G. T. (1997). Methods in enzymology, ibid ref. 4. Vol. 276,
p. 110.
20. Gaucher, J. F., Ries-Kautt, M., Reiss-Husson, F., and Ducruix, A. (1997). FEBS
Lett, 401,113.
21. Jolivalt, C., Ries-Kautt, M., Chevallier, P., and Ducruix, A. (1997). J. Synchrotron
Rad., 4,28.
310
10: From solution to crystals with a physico-chemical aspect
22. Gilliland, G. L., Tung, M., Blakeslee, D. M., and Ladner, J. E. (1994). Acta Cyst.,
D50,408.
23. von Hippel, P. H. and Schleich, T. (1969). Ace. Chem. Res., 2,257.
24. von Hippel, P. H. and Schleich, T. (1969). In Structure and stability of biological
macromolecules (ed. S. N. Timasheff and G. D. Fashman), pp. 417. Marcel
Dekker, Inc., New York.
25. Israelachvili, J. (1992). In Intermolecular and surface forces. Academic Press,
London, pp. 246.
26. Israelachvili, J. (1992). Surface Sci. Rep., 14,113.
27. Salemme, F. R., Genieser, L., Finzel, B. C., Hilmer, R. M., and Wendoloski, J. J.
(1988). J. Cryst. Growth, 90, 273.
28. Fredericq, E. and Neurath, J. (1950). Am. Chem. Soc., 72,2684.
29. S0rensen, S. P. L. and H0yrup, M. (1915-17). Compt. Rend. Trav. Lab. Carlsberg,
12, 312.
30. Green, A. A. (1931). J. Biol. Chem., 93, 517.
31. Gronwall, A. (1942). Compt. Rend. Trav. Lab. Carlsberg, 24, 8.
32. Tanford, C. and Wagner, M. (1954). J. Am. Chem. Soc., 76, 3331.
33. Mikol, V. and Giege, R. (1989). J. Cryst. Growth, 97, 324.
34. Green, A. A. (1932). J. Biol. Chem., 95, 47.
35. Melander, W. and Horvath, C. (1977). Arch. Biochem. Biophys., 183, 200.
36. Arakawa, T. and Timasheff, S. N. (1985). In Methods in enzymology (ed. S. P.
Colowick and N. O. Kaplan), Academic Press, London. Vol. 114, p. 49.
37. Timasheff, S. N. and Arakawa, T. (1988). J. Cryst. Growth, 90, 39.
38. Hofmeister, F. (1888). Arch. Exp. Pathol. Pharmakol, (Leipzig) 24, 247.
39. Collins, K. D. and Washabaugh, M. W. (1985). Q. Rev. Biophys., 18, 323.
40. Cacace, M. G., Landau, E. M., and Ramsden, J. J. (1997). Q. Rev. Biophys., 30,
241.
41. Guilloteau, J.-P., Ries-Kautt, M., and Ducruix, A. (1992). J. Cryst. Growth, 122,
223.
42. Veesler, S., Lafont, S., Marcq, S., Astier, J. P., and Boistelle, R. (1996). J. Cryst.
Growth, 168, 124.
43. Lafont, S. (1996). These de I'Universite d'Aix-Marseille III.
44. Guilloteau, J. P., Fromage, N., Ries-Kautt, M., Reboul, S., Bocquet, D., Dubois,
H., etal. (1996). Proteins: Structure, Function, Genetics, 25,112.
45. Menez, R. and Ducruix, A. (1993). J Mol. Biol., 232,997.
46. Diller, T. C., Shaw, A., Stura, E. A., Vacquier, V. D., and Stout, C. D. (1994). Acta
Oysr.,D50,620.
47. Leavis, P. C. and Rothstein, F. (1974). Arch. Biochem. Biophys., 161, 671.
48. Ries-Kautt, M. and Ducruix, A. (1991). J. Cryst. Growth, 110, 20.
49. Saludjian, P., Prange, T., Navaza, J., Guilloteau, J.-P., Ries-Kautt, M., Menez, R.,
et al. (1992). Acta Cryst., B48,520.
50. Howell, P. L. (1995). Acta Cryst., D51, 654.
51. Ries-Kautt, M., Ducruix, A., and Van Dorsselaer, A. (1994). Acta Cryst., D50,
366.
52. Lafont, S., Veesler, S., Astier, J.-P., and Boistelle, R. (1994). J. Cryst. Growth, 143,
249.
53. Howard, S. B., Twigg, P. J., Baird, J. K., and Meehan, E. J. (1988). J. Cryst.
Growth, 90,94.
311
M. Ries-Kautt and A. Ducruix
54. Alderton, G. and Fevold, H. L. (1946). J. Biol. Chem., 164,1.
55. Palmer, K. J. (1947). Struct. Rep., 11,729.
56. Steinrauf, L. K. (1959). Acta Cryst., 12, 77.
57. Jolles, P. and Berthou, J. (1972). FEBS Lett, 23,21.
58. Haas, D. J. (1967). Acta Cryst., 23,666.
59. Vuillard, L., Rabilloud, T., Leberman, R., Berther-Colominas, C., and Cusack, S.
(1994). FEBS Lett., 353,294.
60. Elgersma, A. V., Ataka, M., and Katsura, T. (1992). J. Cryst. Growth, 122,31.
312
11
Diagnostic of pre-nucleation and
nucleation by spectroscopic
methods and background on the
physics of crystal growth
S. VEESLER and R. BOISTELLE
1. Introduction
Unlike the crystallization of small inorganic molecules, the problem of protein
crystallization was first approached by trial and error methods without any
theoretical background. A physico-chemical approach was chosen because
crystallographers and biochemists needed criteria to rationally select crystal-
lization conditions. In fact, the problem of the production of homogeneous
and structurally perfect protein crystals is set the same as the production of
high-quality crystals for opto-electronic applications, because, in both cases,
the crystal growth mechanisms are the same. Biological macromolecules and
small organic molecules follow the same rules concerning crystallization even
if each material exhibits specific characteristics.
This chapter introduces the fundamentals of crystallization: super-
saturation, nucleation, and crystal growth mechanisms. Phase diagrams are
presented in Chapter 10. Special attention will be paid to the behaviour of the
macromolecules in solution and to the techniques used for their analysis: light
scattering (LS), small angle X-ray scattering (SAXS), small angle neutron
scattering (SANS), and osmotic pressure (OP).
where C and Cs are the actual concentration and the saturation concentration,
i.e. the solubility, respectively. This ratio is dimensionless but its value
depends somewhat on the concentration units (g/litre, mol/litre, mol fraction,
activities, and so on). For protein crystallization, the concentrations are
mostly expressed as mg/ml, i.e. g/litre, which is the easiest way but probably
not the best for explaining crystallization kinetics. Since activities of proteins
cannot yet be calculated, molar fractions are the more appropriate units.
Unfortunately, due to the complexity of the protein solutions, they are seldom
used.
For the sake of simplicity, supersaturation is mostly defined as B, the ratio
defined in Equation 2, or as another dimensionless ratio a = B - 1:
As we will see in the sequel, several growth rate equations contain the term
InfJ included in Equation 1. Traditionally, for the growth of crystals made
314
11: Diagnostic of prenucleation
from small molecules, ln(3 is approximated by a. This is permitted because
supersaturations are often low or very low. For a supersaturation of 5%, i.e. p
= 1.05, we haveCT= 0.05 and lnB ~ 0.049. However, proteins often nucleate
and grow from highly supersaturated solutions, e.g. at supersaturation (3 = 3,
we have ln(3 ~ 1.099, whereas a = 2. It is self-evident that in that case
replacing lnB by a should be avoided. In general, lnB is significantly different
from a as soon as B exceeds about 1.5 which is a very low value for protein
crystallization.
It is also worth noting that supersaturation is sometimes defined as the
difference C - Cs. In this case, its value drastically depends on the con-
centration units. The difference C - Cs = 100 mg/ml, for example, reduces to
about 1 X 10-2 if the concentrations are expressed as mol fractions, for a
molar weight of the solute Mw = 10000. In general, it is more suitable to use
B or cr to solve the nucleation or growth rate equations. However this may
conceal the specific influence of the concentration on crystallization. As an
example, let us consider the case for which the solubility of the protein
decreases when increasing the concentration of the crystallization agent, salt,
or poor solvent. Thereby, in the case of BPTI in NaCl solutions (1), a
supersaturation of twice the solubility, P = 2, can be achieved in the area of
the solubility diagram where solubility is large (44 mg/ml in 1.4 M NaCl
solutions at 25 °C) or low (3 mg/ml in 2.3 M NaCl solutions at 25 °C). In the
former case the mass of solute which will be deposited is 44 mg/ml whereas in
the latter case it is only 3 mg/ml. Despite the same B value, nucleation and
growth will be favoured in the former case.
3. Nucleation
When a solution is supersaturated, the solid phase forms more or less rapidly
depending on the conditions: concentration of solute, crystallization agent,
pH, supersaturation, temperature, nature and concentration of impurities,
stirring, presence of solid particles. Primary nucleation occurs in a solution
that is clear, without crystals. It is called homogeneous nucleation if the nuclei
form in the bulk of the solution. On the other hand, it is called heterogeneous
if the nuclei preferentially form on substrates such as the wall of the
crystallizer, the stirrer, or solid particles (dust particles, and so on). Secondary
nucleation which is induced by the presence of already existing crystals is less
frequent during protein crystallization because the crystallizers are rarely
equipped with stirrers which generate attrition, shear at the crystal surface.
This is the steady state rate of nucleation, J being a number of clusters per
unit time and unit volume of solution. It is further assumed that, at
equilibrium between the size classes, the rates at which a monomer leaves or
sticks on a cluster are equal, so that:
As it will be seen hereafter the small clusters turn into stable nuclei only if
they contain a critical number of monomers. Accordingly, the nucleation rate
J is mainly dependent on the class sizes around i*. It is the product of the
nuclei concentration times the frequency at which they exceed the critical
number i* by addition of a monomer. It expresses as:
where AG* is the activation free energy for forming a nucleus of critical size
and Z the so-called Zeldovich factor:
where i is the number of molecules in the nucleus, Aj the area of the nucleus,
and 7J its interfacial free energy with respect to the solution. The first term
represents the energy to create the volume whereas the second term is the
excess energy to create the surface. To simplify the demonstration we can also
suppose that the nucleus is a sphere so that:
At equilibrium, when dAG/jr = 0, the nucleus has the critical radius r*, as
shown in Figure 1,
Figure 1. Variation of the activation free energy for three-dimensional nucleation versus
nucleus size.
317
S. Veesler and R. Boistelle
The critical activation free energy for creating the nucleus with critical radius
r* is one-third of the energy required for creating its surface. As shown in
Figure 7, at the critical size r*, the nucleus is in a very labile equilibrium. If it
gains one molecule so that r > r* it grows. But if it loses one molecules so that
r < r*, then it spontaneously dissolves. In both cases there is a gain in energy.
Inserting Equation 13 or Equation 14 into Equation 9 allows for the
calculation of the nucleation rate if nucleation is homogeneous.
If we name St the area of the nucleus and Sa the area of the interface between
the nucleus and the substrate, the activation free energy for heterogeneous
nucleation is:
318
11: Diagnostic of prenucleation
At equilibrium, when 8AGhet /8r = 0, the radius of the critical nucleus is:
which is the product of AG* for homogeneous nucleation times a term depend-
ing on the contact angle. It is worth noting that for a = 180°, AG het = AG*.
The substrate does not have any effect on nucleation. For a = 90°, AGhhet =
AG*/2. If a tends toward zero, then AG*het tends also toward 0. That means
that the substrate induces nucleation even at very low supersaturation since
less and less energy is required to form the nucleus. The nucleation rate
Equation 9 drastically increases when the contact angle a decreases and
subsequently the activation free energy for nucleation.
3.4 Examples
Let us first imagine a system for which nucleation of small molecules is
homogeneous. To solve Equation 9 we suppose that, in Equation 13, each
molecule occupies a volume V = (5 X 5 x 5) X 10-24 cm3. If the solubility is
rather high (typically 10-50 g/litre) then the interfacial free energy, -ft, is
rather low, e.g. 10 erg cm"2 (10 mJ m-2). Typically in Equation 9 one has uN1 =
1020 cm-3 s-1. Inserting T = 293 K and kB = 1.38 X 10-16 erg/K yields:
In that case a very large supersaturation is needed to obtain 1 nucleus cm-3 s-1.
In fact B equals 7718 which is completely unrealistic! However it is a good
illustration of the difficulty often encountered for nucleating proteins. Further-
more, it is difficult to consider that the kinetic coefficient, taken as 1020 cm-3 s-1
in the previous example, would be greater for macromolecules than for small
molecules. Hence, the only way to obtain reasonable values of J and of (3 is to
assume that the surface free energy of a protein crystal is significantly lower
than the energies usually encountered for crystals of small molecules. If to
calculate Equation 21, yi = 1 erg cm-2 is used instead of 10 erg cm-2 then the
J values are given by Equation 20 and subsequently the same low p values
are obtained. Consequently lower surface free energies can compensate
higher molecular volumes in the case of protein crystallization. Values of
•yj = 0.5-0.7 erg cm"2 for thaumatin (Mw = 22000 Da) were observed by
Malkin et al. (5).
As a concluding remark, it should be emphasized that there is no special
reason that, a priori, all proteins should have a low or very low surface free
energy. If the general rule holds for proteins, it might be, then for sparingly
soluble proteins the surface free energies are relatively high. But, these
proteins also nucleate, sometimes even after rather short induction periods.
The only explanation would be that nucleation is heterogeneous. As a matter
of fact, it cannot be homogeneous because the required supersaturation
would be much too high. The existence of heterogeneous nucleation can be
checked with a very well known model protein, i.e. with hen egg white
lysozyme (HEWL). With p values of ~ 3, the solution deposits several tens of
crystallites per cubic centimetre of solution within a few hours if the solution
is not carefully filtered. On the other hand, it deposits sometimes only one or
320
11: Diagnostic of prenucleation
two crystals after one or two days, if most of the solid particles are removed
by filtration.
Finally, in order to estimate the nucleation rate and the interfacial free
energy -yj, it is possible to measure the time lag for nucleation, or induction
time tj, as a function of different supersaturations. Assuming that after the
time tj J = 1 cm'3 s"1 on has:
so that:
Plotting Intj versus l/ln2p should give a straight line, the slope of which is
proportional to yl which is the only unknown in the term on the right side of
Equation 23. This method was often used for determining the •/! values of
crystals of small molecules. Due to the uncertainties of the measurements of t;
it only gives a good order of magnitude.
4.1 Methods
Depending on the techniques used, different information can be obtained on
the solutions. Since we only give a brief survey of these techniques, the reader
can refer to the different monographs where they are widely described. The
references are given hereafter. In this section, solution scattering and osmotic
pressure techniques will be presented. These techniques aim at obtaining
information on molecules in solution: molecular weight, size, aggregation
states, polydispersity, and interactions. As will be seen in Section 4.3 probing
the protein interactions is very important in the field of protein crystallization.
In dilute solution interactions include excluded volume term, repulsive
321
S. Veesler and R. Boistelle
electrostatic term, and attractive Van der Waals term, the last two terms are
described in the so-called DLVO theory of colloidal stability (6).
4.1.1 Solution scattering techniques
For all scattering experiments in solution the principle is the same: a mono-
chromatic beam of visible light, X-rays, or neutrons impinges on the protein
molecules and induces an oscillating polarization of their electrons. The
molecules then serve as a secondary source which is radiated and scattered.
For neutrons the interaction with the matter is different, because neutrons are
scattered by the atomic nuclei, the scattered intensity depending on the
scattering length density.
i. Light scattering (7, 8)
Depending on the way the data are analysed, two types of experiments are
possible: elastic or static light scattering (SLS) and quasi-elastic scattering or
dynamic light scattering (DLS). The experimental set-up is shown in Figure 3.
Static light scattering (SLS)
The experiment consists in measuring the photons intensity scattered by the
solution at different angles and for different concentrations. Let us consider
two cases depending on the size of the protein with respect to the wavelength
of the light.
(a) The light is scattered by spherical particles which are small compared to
the wavelength of the light, d < X/10, where d is the characteristic size of
the protein and the wavelength of the light (400 < X < 650 nm). In that
case the particle is assumed to be a punctual source of light, and the
intensity scattered is independent of the angle. The experiments are
usually carried out at 90°, then:
322
11: Diagnostic of prenucleation
AI the difference between the intensity scattered by the solvent and the
solution at 90°.
Information obtained: Mw (the molecular weight of the particle) and A2
(the second virial coefficient), also noted B22 in the literature. The sign of
the second virial coefficient is indicative of the type of interactions: it is
negative when the interactions between molecules are attractive and
positive if the interactions are repulsive.
(b) The light is scattered by larger particles, A/10 < d < X; in that case the
particle is assumed to be a multiple source of the light so that there is a
phase difference between the light scattered by different portions of the
particle at any time, the intensity scattered is angular dependent, and then:
where C is the protein concentration, R the molar gas constant, 8.31 107 erg
mor1 K-1, and T the absolute temperature.
Information obtained: Mw and A2-
324
11: Diagnostic of prenucleation
concentration dependence of the measured parameter. Furthermore, the
experimenter has to check whether the particles are all the same in the whole
concentration range under consideration.
4.2.2 LS
The volume of solution required for a measurement is about 100-300 ul. A
typical experimental protocol is described in Protocol 1. Prior to the experi-
ment it is necessary to control several points:
(a) Absence of fluctuation in the laser intensity.
(b) Absence of parasitic light due to reflections or refractions. These con-
ditions are difficult to obtain in experiments carried out at angles below
30° with a commercial set-up.
(c) Absence of dust, air bubbles, glass particles, and other foreign tiny
materials in solution.
(d) Measurement achieved at the proper angle.
(e) Good transparency of the solution is required in order to avoid the
multiple diffusion, if needed a dilution must be done.
Therefore, before any experiment it is necessary to check the laser quality,
the optical trajectory, and to filter and/or centrifuge the solution. To treat the
signal it is also essential to know the refractive index and viscosity of the
crystallization medium (buffer + crystallization agent). In addition, for SLS
experiments, it is essential to know the increment of the refractive index of
the solution as a function of the protein cpncentration.
4.2.3 SAXS-SANS
These experiments can only be carried out using a synchrotron radiation or a
nuclear reactor. Runs are always allocated by a program committee to which
the application must be submitted. Since the number of runs is limited, it is
recommended to test the sample quality before using one of these techniques.
For instance DLS is a good tool for checking whether the molecules are
aggregated or not. If the polydispersity is high the SAXS and SANS will fail.
(a) SAXS: the volume of solution required for one measurement is ~ 100 ul.
The electron density of the crystallization agent should be as low as
possible, otherwise the signal due to the particles which is under investi-
gation disappears in the background. Thus electron-rich buffer at high
concentration (e.g. ammonium sulfate) should be avoided.
(b) SANS: the volume of solution required for one measurement is about
150 ul. In these experiments it is often necessary to dissolve the protein
in D2O solutions. Moreover, Broutin et al. (15) and Gripon et al. (16)
recently showed a shift of the solubility of lysozyme when H2O is replaced
by D2O. Furthermore, D2O affects the interactions between the particles
325
S. Veesler and R. Boistelle
in solution, and care must be taken when working with materials which
have a tendency to aggregate. Accordingly, the experimenter has to check
the aggregation behaviour of the protein solution before any experiment.
4.2.4 OP
The volume of solution required for a measurement is about 120 ul. Actually,
three measurements performed with 40 ul are necessary.
(a) The main point of this experiment concerns the equilibration on both
sides of the membrane. The higher the salt concentration is, the more
difficult the equilibration is. Practically the upper concentration limit is
about 500 mM.
(b) High viscosities solutions can generate very long equilibrium time.
(c) It is important to check the temperature and the pH stability.
4.3 Examples
DLS is the most widely used method for the characterization of protein
solutions and was first proposed as a diagnostic tool for protein crystallization
by Kam et al. (17). Zulauf et al. (18) studied 15 proteins, in dilute solutions and
in the absence of crystallization agent, and suggested that the detection of
aggregates indicates that crystallization will not be successful. Ferre-D'Amare
et al. (19) have determined the crystallizability of three different RNAs by
DLS with this criterion. In addition, more recent studies showed the absence
of large molecular aggregates in supersaturated solutions for different proteins
(1, 20, 21), contrary to Georgalis et al. (22, 23) who observed the formation of
326
11: Diagnostic of prenucleation
praggs (precipitating aggregates) or craggs (crystallizing aggregates) depend-
ing on the conditions.
Measurements of A2 (or B22) by SLS for different proteins in different sol-
vents have shown that conditions which promote crystallization are grouped
within a narrow range of A2 values (24, 25). Moreover different studies have
shown a systematic relationship between solubility and second virial co-
efficient (25). Recently Bonnete et al. (26) showed the complementary nature
of DLS, SAXS, and OP in order to probe interaction in protein solution.
Rather than reviewing all the outputs for crystallogenesis of the above
methods, we discuss here an example of the application of DLS with the
associated experimental protocol, and another example dealing with an
application of SAXS.
4.3.1 DLS study applied to porcine pancreatic a-amylase (20)
In this example special attention was paid to the polydispersity of under- and
supersaturated solutions of a-amylase. The results are presented in Figure 4,
and can be summarized as follows: polydispersity is very high (v > 10%) when
the protein concentration is much lower than solubility whereas it is very low
(v < 10%) when the protein concentration is nearly equal or even slightly
higher than solubility. Even more important, monodispersity is a prerequisite
for obtaining good crystals. Some polydispersity seems to be acceptable if
there are no large aggregates in solution.
Method
1. Prepare a solution of 300 ul of the protein at the desired concentration.
2. Filter three times the solution with the same LCR13 (Millipore) filter
and pour the solution in a glass cell. Dust can also be removed by
centrifugation (20000-30000 g, for 1-2 h).
3. Put the glass cell in the sample holder.
4. Switch on the laser and check the optical trajectory. Avoid continual
illumination of the solution for several hours because of potential
protein denaturation.
5. Run the analysis for 1-3 min; the correlator receives the signal from
the detector (photomultiplier).
6. Analyse the data: the cumulant method (27) directly gives the diffusion
coefficient and the polydispersity.
327
S. Veesler and R. Boistelle
Figure 5. X-ray scattering curves recorded at 100 mg/ml lysozyme concentration in acetate
buffer plus added salt. Anion series as indicated in the figure with Na+ as counterion,
ionic strength = 0.150 (17).
5. Crystal growth
When a nucleus grows and transforms into a crystal, the different faces of the
growing crystal exhibit growth mechanisms and rates that depend on external
factors (supersaturation, impurities, temperature, and so on) and internal
factors (structure, bonds, defects, and so on). According to the periodic bond
chain (PBC) theory (29-32) there are three types of crystal faces (Figure 6).
• F (flat) faces: they contain at least two PBCs in the slice of thickness dhkl,
where dhkl is the interplanar distance of the face (hkl).
• S (stepped) faces: they contain only one PBC in the slice dhkl.
• K (kinked) faces: they do not contain any PBC in the slice dhkl.
Let us just recall that a PBC is an interrupted chain of strong bonds running
along a crystallographic direction in the crystal. Since all sites on the K faces
are growth sites, more commonly called kinks, the K faces grow by direct
incorporation of the growth units which hit them. The growth rate is high and,
normally, these faces do not occur on the crystal morphology, because the
growth form of the crystal is made up only of the faces which have the slowest
growth rate.
Conversely, the F faces are poor in kinks. They grow by lateral spreading of
Figure 6. Schematic representation of a crystal exhibiting flat (F), stepped (S), and kinked
(K) faces. The front face exhibits a polygonized growth spiral, whereas the top face
exhibits a two-dimensional nucleus.
329
S. Veesler and R. Boistelle
the growth layers. For being integrated into the crystal, the solute molecules
must first adsorb on the surface, and later on diffuse toward the step of a
growth layer along which they migrate toward a kink. Such faces grow either
by a two-dimensional mechanism or a spiral growth mechanism (Figure 6).
Since the number of kinks is low, the growth rates are low too.
At last, the S faces are in an intermediate situation. Their growth rate is
lower than that of the K faces but higher than that of the F faces. Normally,
the S faces do not appear on the crystal morphology, except when their
growth rate is slowed down by adsorption of an impurity for example.
The growth mechanisms have been discussed in detail elsewhere (33).
Hereafter we summarize the general trends.
where X is the so-called edge free energy expressed here per molecule in the
edge. In the mononuclear model, there is only one 2D nucleus which spreads
across the surface so that the growth rate R of this face is:
where d is the height of the growth layer, S the area of the face, and B2 the 2D
nucleation rate, i.e. the number of nuclei forming per unit time and unit area
(cm-2 s-1). B2 can be written:
where n1 is the number of growth units adsorbed per unit area of the face and
6 the frequency at which the 2D nucleus of critical size become supercritical
and grows. AG2* is the activation free energy for 2D nucleation:
Inserting Equation 31 into Equations 30 and 29 shows that the growth rate is
an exponential function of supersaturation. As in the case of three-
dimensional nucleation, there is a critical supersaturation below which the
growth rate is zero or nearly zero. A dead zone is observed at low super-
330
11: Diagnostic of prenucleation
saturation when measuring R as a function of p. Once this critical super-
saturation is exceeded, the growth rate drastically increases with increasing
supersaturation. Growth is difficult to control.
If several nuclei spread at the same time across the crystal face, growth is
determined by a 2D polynuclear mechanism. In that case the expression for
the growth rate is somewhat more complicated (33-35).
5.1.2 Growth by a spiral mechanism
When a screw dislocation emerges on a crystal face, it generates a growth
spiral (Figures 6 and 7). Since the growth spiral is made of a parallel sequence
of steps, growth can take place even at low supersaturation since the growth
units which adsorb onto the crystal face easily find growth sites where they are
incorporated into the crystal. It can be seen (Figure 7} that the growth rate of
the face can schematically be written as:
where v is the lateral velocity of the steps, d their height, and y their
equidistance. If the spiral is circular one has:
where A. is also the edge free energy of the steps, a being the distance between
two molecules in the step. Considering Figure 7 and Equation 33, we see that
an F face which exhibits a growth spiral is really flat only if the super-
saturation is low (y large). Conversely, it takes a conical outline when the
supersaturation is high (y small) due to the high step density.
The theories of the spiral growth mechanism were extensively discussed
elsewhere (36-39). Here, we give only a few possibilities which can be derived
from the general growth rate equation that is not commented here. Depend-
ing on the influence of the different parameters, growth depends on surface
diffusion, kink integration kinetics, and so on. As an example, let us suppose
331
S. Veesler and R. Boistelle
that growth is controlled by surface diffusion of the growth units toward the
steps. Then we have two possibilities.
At low supersaturation:
In both equations, Ds is the surface diffusion coefficient and xs the mean free
path of diffusion; n1 is again the number of adsorbed growth units per unit
area, and V the volume of a molecule in the crystal.
Inserting Equation 33 into Equation 34 shows that R is proportional to a X
InB. With the approximation lnB = a the so-called primary quadratic growth
rate law is obtained at low supersaturation:
whereas at high supersaturation the primary linear growth rate law is obtained:
where a is the length of the elementary jump of the growth units which diffuse
toward the kinks. Once more, if we insert Equation 33 into Equation 39, and
332
11: Diagnostic of prenucleation
with the assumption that lnB = a a secondary quadratic growth rate law is
obtained:
where n0 is the number of growth units per unit solution volume, Dv the
volume diffusion coefficient, and 8 the thickness of the boundary layer. This
equation is similar to that derived from Fick's law (33, 40):
333
S. Veesler and R. Boistelle
Plotting R/U1/2 versus R1/n for all curves obtained at different supersaturations,
provides the highest possible growth rates (for U = oo) by extrapolating the
straight lines thus obtained to R/U1/2 = 0.
As concerns proteins crystallization, it is practically always carried out in
stagnant systems. The reason for this is the missing of instrumentation, but
perhaps also the fragility of proteins. Moreover, Dv of proteins are two order
of magnitude smaller than the ones of small molecules. Accordingly, the
growth rates are mainly controlled by volume diffusion (Equations 39, 42) or
by the kink integration kinetics (Equation 39).
a few seed crystals were introduced in the crystallization cell and the displace-
ment of the faces was recorded as a function of time. A linear dependence
with supersaturation was obtained and growth was interpreted as a process
controlled by volume diffusion. On a more microscopic scale, direct measure-
ments of the step velocity are possible using laser Michelson interferometry
(42) or in situ atomic force microscopy (5). Such measurements are especially
interesting if the step equidistance can be related to supersaturation in order
to deduce the edge free energy of the step of the growth layer (Equation 33).
Knowing the step velocity, also allows a better estimation of the parameters
involved in the growth rate equations (mean free path for diffusion, surface
diffusion coefficient, relaxation time for entering into the kinks, and so on).
References
1. Lafont, S., Veesler, S., Astier , J.-P., and Boistelle, R. (1994). J. Cryst. Growth,
143,249.
2. Abraham, F. F. (1974). Homogeneous nucleation theory. Academic Press, New
York.
3. Zettlemoyer, A. C. (1969). Nucleation. Marcel Dekker, New York.
4. Katz, J. L. (1982). In Interfacial aspects of phase transformations (ed. B.
Mutaftschiev), p. 261. D. Reidel, Dordrecht, The Netherlands.
5. Malkin, A. J., Kuznetsov Yu, G., Glantz, W., and McPherson, A. (1996). J. Phys.
Chem., 100, 11736.
338
11: Diagnostic of prenucleation
6. Israelachvili, J. N. (1991). Intermolecular and surface forces. Academic Press,
London.
7. Tanford, C. (1961). In Light scattering in physical chemistry of macromolecules,
p. 275. John Wiley & Sons, Inc., New York, London.
8. Pecora, R. (1985). Dynamic light sacttering: application of photon correlation
spectroscopy. Plenum Press, New York and London.
9. Guinier, A. and Fournet, G. (1955). Small-angle scattering of X-rays. John Wiley,
New York.
10. Tardieu, A. (1994). In Neutron and synchrotron radiation for condenses matter
studies (ed. J. Baruchel, J. L. Hodeau, M. S. Lehmann, J. R. Regnard, and C.
Schlenker), Vol. III, p. 145. Les Editions de Physique, Springer-Verlag.
11. Lindner, P. and Zemb, T. (1991). Neutron X-ray and light scattering: introduction
to an investigative tool for colloidal and polymeric systems. North-Holland,
Amsterdam.
12. Jacrot, B. and Zaccai, G. (1981). Biopolymers, 20, 2413.
13. Einsenberg, H. (1976). Biological macromolecules and polyelectrolytes in solution.
Clarendon Press, Oxford.
14. Parsegian, V. A., Rand, R. P., Fuller, N. L., and Rau, D. C. (1986). In Methods in
enzymology, (ed. L. Packer), Academic Press, London. Vol. 127, p. 400.
15. Broutin, L, Ries-Kautt, M., and Ducruix, A. (1995). J. Appl. Cryst., 28, 614.
16. Gripon, C., Legrand, L., Rosenman, I., Vidal, O., Robert , M.-C., and Boue, F.
(1996). C. R. Acad. Sci. Paris, 323,215.
17. Kam, Z., Shore, H. B., and Feher, G. (1978). J. Mol. Biol., 123, 539.
18. Zulauf, M. and D'Arcy, A. (1992). J. Cryst. Growth, 122, 102.
19. Ferre-d'Amare, A. R., Zhou, K. Z., and Doudna, J. A. (1998). J. Mol. Biol., 279,
621.
20. Veesler, S., Marcq, S., Lafont, S., Astier, J.-P., and Boistelle, R. (1994). Acta
Cryst., D50, 355.
21. Muschol, M. and Rosenberger, F. (1996). J. Cryst. Growth, 167, 738.
22. Georgalis, Y., Zouni, A., and Saenger, W. (1992). J. Cryst. Growth, 118, 360.
23. Georgalis, Y., Zouni, A., Eberstein, W., and Saenger, W. (1993). J. Cryst. Growth,
126, 245.
24. George, A. and Wilson, W. W. (1994). Acta Cryst., D50, 361.
25. George, A., Chiang, Y., Guo, B., Arabshahi, A., Cai, Z., and Wilson, W. W.
(1997). In Methods in enzymology, (ed. C. W. Carter and R. M. Sweet), Academic
Press, London. Vol. 276, p. 100.
26. Bonnete, F., Malfois, M., Finet, J., Tardieu, A., Lafont, S., and Veesler, S. (1997).
Acta Cry'St., D53, 438.
27. Koppel, D. E. (1972). J. Chem. Phys., 57, 4814.
28. Ducruix, A., Guilloteau, J.-P., Ries-kautt, M., and Tardieu, A. (1996). J. Cryst.
Growth, 168, 28.
29. Hartman, P. and Perdok, W. G. (1955). Acta Cryst., 8, 49.
30. Hartman, P. (1973). In Crystal growth: an introduction (ed. P. Hartman), p. 367.
North-Holland, Amsterdam.
31. Hartman, P. (1982). Geol. Mijnbouw, 61, 313.
32. Hartman, P. and Bennema, P. (1980). J. Cryst. Growth, 49,145.
33. Ohara, M. and Reid, R. C. (1973). Modeling crystal growth rates from solution.
Prentice Hall, Englewood Cliffs, NJ.
339
S. Veesler and R. Boistelle
34. Hillig,W. B. (1966). ActaMetalL, 14,1868.
35. Madsen, L. and Boistelle, R. (1979). J. Cryst. Growth, 46, 681.
36. Burton, W. K., Cabrera, N., and Frank, F. C. (1951). Phil. Trans. Roy. Soc., 243,
299.
37. Chernov, A. A. (1961). Sov. Phys. Usp., 4, 116.
38. Gilmer, G. H., Ghez, R., and Cabrera, N. (1971). J. Cryst. Growth, 8, 79.
39. Bennema, P. and Gilmer, G. H. (1973). In Kinetics of crystal growth, an
introduction (ed. P. Hartman), p. 263. North-Holland, Amsterdam.
40. Nielsen, A. E. (1964). Kinetics of precipitation. Pergamon, Oxford.
41. Boistelle, R., Astier, J.-P., Marchis-Mouren, G., Desseaux, V., and Haser, R.
(1992). J. Cryst. Growth, 123, 109.
42. Vekilov, P. G., Ataka, M., and Katsura, T. (1993). J. Cryst. Growth, 130,317.
43. Boistelle, R. (1982). In Interfacial aspects of phase transformation, p. 621. Erice,
Sicily.
44. Parker, R. L. (1970). In Solid state physics, Vol. 25, p. 151. Academic Press, New
York.
45. Kern, R. (1968). Bull. Soc. Fr. Mineral. Cristallogr., 91, 247.
46. Giege, R., Dock, A.-C, Kern, D., Lorber, B., Thierry, J.-C, and Moras, D. (1986).
J. Cryst. Growth, 76, 554.
47. Vekilov, P. G. (1993). Prog. Cryst. Growth, 26, 25.
48. Lorber, B., Skouri, M., Munch, J.-P., and Giege, R. (1993). J. Cryst. Growth, 128,
1203.
49. Abergel, C., Nesa, P. M., and Fontecilla-Camps, J. C. (1991). J. Cryst. Growth,
110, 11.
340
12
Two-dimensional crystallization of
soluble proteins on planar lipid
films
A. BRISSON, O. LAMBERT, and W. BERGSMA-SCHUTTER
1. Introduction
Electron crystallography of protein two-dimensional (2D) crystals constitutes
a fast-expanding method for determining the structure of macromolecules at
near-atomic resolution (1, 2). The main limitation in the application and
generalization of this approach remains in obtaining highly ordered 2D
crystals, as is the case of 3D crystals in X-ray crystallography.
Several methods of 2D crystallization are available which can be classified
into two families, depending on the type of proteins under investigation,
either membrane proteins (3, 4) or soluble proteins (5, 6). In both cases, 2D
crystallization is a self-organization process which spontaneously occurs
between macromolecules which are restricted to diffusing by translation and
rotation in a 2D space, with a fixed orientation along the normal to this plane.
The scope of this chapter is restricted to the 2D crystallization of soluble
proteins on planar lipid films, by the so-called 'lipid monlayer crystallization
method' (5). Our aim is to present a step-by-step description of the experi-
mental procedures involved in the application of this method.
342
12; Two dimensional crystallization of soluble proteins
Method
1. Handle Teflon discs with gloves.
2. Place a disc in a bath of 1% Hellmanex II for 2 h.
3. Rinse the disc extensively with warm tap-water for 1 h.
4. Rinse with several baths of deionized water.
346
12: Two dimensional crystallization of soluble proteins
5. Smash the disc against a piece of tissue paper lying on a table, to get
rid of the water.3 The disc is ready for use.d
6. For prolonged storage, keep the discs either in water or dry, in a box.
Perform the treatment (steps 1-5) before use.
aA good indication of the state of hydrophobicity of a Teflon support is obtained by smashing
it on a piece of tissue paper; the wells must look completely dry after two smashes. If some
water remains at the bottom of the wells or at their edges, washing of the disc must be
repeated. When discs have been used repeatedly, a second washing step might not be
sufficient to obtain a hydrophobic surface. It is then recommended to brush the wells in order
to eliminate lipids stuck onto their surface.
'Teflon surfaces tend to become electrostatic, which can lead to several surprising effects; a
microscope grid may 'jump' upon deposition on the lipid-coated wells, or a lipid droplet
approached from the Teflon surface may become deformed or even 'explode'. Although it is
difficult to judge the influence of such a behaviour on crystallization, the best remedy is to
discharge the surface with an antistatic device.
Method
1. Place a freshly conditioned Teflon disc in a Petri dish.
2. Poor water around the disc up to about mid-height.
3. Deposit in each well 17 uJ of the protein solution.
4. Rinse the 10 ul syringe three times with chloroform.
5. Deposit 0.6 ul of the lipid solutiona on top of each protein droplet.
6. Install the lid on the Petri dish and close the hole with a piece of tape,
in order to limit evaporation.
7. Incubate.b-c
a Lipids must not flow over the edges. If this occurs, irreproducible results can be expected and
it is wise to use another well or another Teflon disc.
"The incubation time required for crystal growth is variable and depends on the protein and
lipid system. For example, one hour or less is sufficient to get 20 crystals of streptavidin or
cholera toxin; in the case of annexin V, one hour is also sufficient to get one type of crystal,
with p6 symmetry, while several days are required to get highly ordered p3 crystals (18).
c
Most studies reported until now have been performed at ambient temperature, around 20°C.
It is of course important to use conditions in which the protein is stable.
347
A. Brisfion et al.
348
12: Two dimensional crystallization of soluble proteins
350
12: Two dimensional crystallization of soluble proteins
(b) Evaporation of carbon on the plastic film (Protocol 3, step 9).
(c) Dissolution of the plastic film (Protocol 3, steps 10 and 11).
Method
1. Deposit four droplets of the nitrocellulose solution on a clean glass
slide, with a Pasteur pipette.
2. Form a continuous and homogeneous liquid film by tilting the slide.
3. Eliminate most of the liquid by holding the slide vertically against a
piece of filter paper.
4. Let dry for 10 min.
5. Float off the plastic film on a water surface, by slowly inserting the
slide at glancing angle into the water-bath.
6. Deposit EM grids on top of the plastic film.
7. Deposit a piece of absorbent paper or a piece of Parafilm on top of the
plastic film covered with the grids, while maintaining the paper/
Parafilm by one edge. Wait until a good contact is formed with the
underlying plastic film.
8. Lift up the paper/Parafilm and deposit it on a clean surface. Allow for
a complete drying, under a lamp.
9. Evaporate a thin film of carbon on the nitrocellulose side, using
standard EM procedures (22).a
10. Place the grids, carbon side up, on top of several pieces of filter paper
soaked with amyl acetate, in a glass Petri dish. Close the dish and
leave for several hours to overnight.
11. Transfer the grids on a dry filter paper.b
a Grids prepared up to here can be used for negative staining without removal of the plastic
film. On the other hand, for cryo-microscopy experiments, it is mandatory to remove the
plastic film in order to avoid artefacts.
b Carbon films prepared by this method are in general flat (Figure 3a) (23). The flatness of
carbon films is certainly an important parameter for allowing a good contact with the lipid
chains and thus achieving an efficient transfer.
351
A. Brisson et al.
Reagents
• 0.3% (w/v) formvar solution in dichloroethane
352
12: Two dimensional crystallization of soluble proteins
C. Dissolution of the plastic film
1. Dissolve the formvar film by depositing the grids on top of several
pieces of filter paper soaked with dichloroethane, in a glass Petri dish.
2. Let stand overnight.
3. Transfer the grids on a dry paper.f
353
A. Brisson et al.
Protocol 5. Continued
2. Check each grid with an optical microscope for homogeneity and
integrity of the holey film,c,d
a This method has been adapted from Fukami and Adachi (25) by Chretien et al. (26) (adapted
from ref. 26 with permission).
b It is possible to adjust the size of the holes at this step. A long exposition on the metal block
will produce larger holes. Several trials are necessary to obtain holes with the desired size.
c With holey films prepared by this method, about half of the surface is covered with holes and
the other half with carbon (Figure 3c).
d 'The main advantage of this method, as compared with most other methods of fabrication of
holey films, is that floating of the plastic film is easy and reproducible.
354
12: Two dimensional crystallization of soluble proteins
Method
1. Immediately after the fishing step (Protocol 6, step 4), add a 5 ul
droplet of the uranyl acetate solution to the liquid film present on the
grid.a
2. Wait for 30 sec.
3. Remove the excess liquid by touching a grid border with a filter paper.
4. Allow for complete drying.
5. In the case of holey films, evaporate a thin layer of carbon on the side
of the protein-lipid film.b
6. The grid is ready for observation in the microscope.
a
At the beginning of each new study, it is recommended to compare the results obtained with
several negative stains. Sodium phosphotungstate (2% (w/v) aqueous solution, pM 7) is
another commonly used negative stain.
b
The deposition of a carbon layer enhances the mechanical stability and improves the
conductivity of self-supported interfacial films.
Figure 4. Aspect of interfacial films transferred with continuous carbon films, (a) and |b)
correspond to annexin V and streptavidin, respectively, (a) Domains exhibiting a uniform
greyness and vesicles cover the carbon film support. The vesicles surround the domain
areas and most probably form during the transfer step. Extensive carbon film areas are
devoid of domains and vesicles (*). (b) Domains showing multilayered structures are
characteristic of films transferred with continuous carbon supports. On these low
magnification images (scale bar: 1 um), the crystalline nature of the domains is not visible.
356
12: Two dimensional crystallization of soluble proteins
rule that undoubtedly has exceptions, interfacial films picked up with a con-
tinuous carbon film present domains easily distinguishable from the carbon
background, together with vesicular material (Figure 4). These domains are
often folded or overlap in multilayered structures, and present morphologies
characteristic of each protein-lipid system. On the other hand, interfacial films
transferred with a holey carbon film appear as homogeneous layers of uni-
form greyness and thus uniform thickness, mostly devoid of domains or large
vesicles (Figures 3b, 3c, 5, and 6). It is now commonly accepted that the inter-
facial films are transferred without, or with minor, reorganization when holey
films are used (8, 18, 29), while they are submitted to profound reorganization
upon transfer/drying with continuous carbon films (13, 18, 30). The annexin V
system constitutes an extreme case in this context, as p6 crystals are obtained
with holey films, while p3 crystals are observed with continuous carbon films
(18). Most strikingly, these p3 crystals do not pre-exist at the air-water
interface and their formation is induced by the transfer step. Cholera toxin
constitutes another interesting case as highly ordered 2D crystals are obtained
after transfer with continuous carbon films while close-packed 2D domains
are obtained with holey films, and thus pre-exist at the air-water interface
(18). The coherent picture which emerges from these studies is that upon
specific binding to ligands incorporated into lipid monolayers at the air-water
interface, some proteins form 2D crystals, while many others self-organize in
close-packed assemblies. Upon transfer with a continuous carbon film, these
close-packed assemblies are 'stressed' and may reorganize into more compact
and better ordered 2D crystals.
2. At high magnification (X 50 000), the crystalline nature of the transferred
material can be visualized. However, it is in general not possible to get a
quantitative evaluation of the crystalline order by a mere 'eye' observation.
Even when strongly contrasted stain striations are observed, this information
is of low resolution as it represents most often the accumulation of stain
between molecules (see for example Figure 5). The most objective way to
evaluate the crystalline quality of protein-lipid interfacial films is by optical
diffraction or Fourier transform calculation (see Section 6).
The characterization of interfacial films by EM is one of the most time-
consuming steps in the whole procedure. The main reason is the huge number
of areas of potential interest on each grid and the variability of aspect existing
between grids and also within different areas of one given grid. It is important
to consider that:
(a) In the case of holey films, holes covered by either 2D crystals or close-
packed assemblies present the same aspect, and there are of the order of
104 to 106 holes per grid.
(b) With continuous carbon films, the number of domains of potential
interest present on a grid is even larger.
357
A. Brinnon et al.
Figure 5. Interfacial film of streptavidin transferred with a holey carbon film. The
streptavidin film consists of a mosaic of crystalline domains. Two large crystals, one in
each hole, are indicated by two arrows aligned along the main directions of stain
striations. Their frontiers with adjacent 2D crystals are delineated with dashed lines. Next
to the carbon threads, the streptavidin film is often disordered (*), suggesting that the
crystalline structure is disorganized when the carbon film touches the interfaciai film or
during the 'fishing step'. Scale bar: 1000 A.
358
12: Two dimensional crystallization of soluble proteins
Figure 6. Interfacial film of annexin V transferred with a holey carbon film. Single crystal-
line domains of annexin V cover the holes. The main orientation of the lattice (arrows) is
almost conserved between adjacent holes, suggesting that the interracial film was a
single monocrystal before transfer. The crystal is built up of trimers of annexin V (circles)
assembled with p6 symmetry. Scale bar: 1000 A.
359
A. Brisson et al.
(c) Crystalline domains of 1 um2 are large enough to provide high resolution
information and the surface of a grid is equivalent to about 106 such
domains. This explains why screening EM grids for the presence of crystals
and optimizing crystallization conditions are extremely time-consuming.
7. Conclusion
The lipid-layer crystallization method is a rational and general method for
growing highly ordered 2D crystals of macromolecules. Until now, it has been
Figure 7. Optical diffraction, (a) Scheme of an optical diffraction set-up. The principal
components of an optical diffraction set-up are: a laser source, a convergent lens, and a
screen placed in the diffraction plane of the lens. In the set-up presented here, the lens,
placed at a distance p from the laser source, is illuminated by a non-parallel beam. The
rays emerging from the lens converge at the diffraction plane, located at a distance p'
from the lens, such as: 1/p + 1/p' = 1/f (f: focal length of the lens). The advantage of using
a non-parallel beam illumination is that the size of the diffraction pattern from the EM
negative can be easily adjusted by changing the distances between the lens and the
source and/or between the EM negative and the lens. Sub-areas of the negatives are
evaluated for their crystalline quality and selected for further processing. According to
the diffraction theory, a periodic grating of period d, illuminated by a coherent beam of
wavelength \, gives rise to two diffracted beams forming an angle 9 with the direct beam,
such as: d sin O = y. On the screen placed at a distance L from the negative, two
diffraction peaks will be observed, located at a distance D from the centre, such as: D = L
tg O. As 9 is small, dD = L X = cst. (b) Example of a diffraction pattern of a negatively
stained 2D crystalline domain of annexin V (adapted from ref. 35). The diffraction peaks
are arranged onto a hexagonal lattice. The (0,6) and (6,2) reflections, at 1/13.4 and 1/11.2
A-1 respectively, are circled. Scale: 1 cm = 0.028 A-1.
360
12: Two dimensional crystallization of soluble proteins
References
1. Kimura, Y., Vassylyev, D. G., Miyazawa, A., Kidera, A., Matsushima, M.,
Mitsuoka, K., et al. (1997). Nature, 389, 206.
2. Nogales, E., Wolf, S. G., and Downing, K. H. (1998). Nature, 391, 199.
3. Kiihlbrandt, W. (1992). Q. Rev. Biophys., 25, 1.
4. Jap, B. K., Zulauf, M., Scheybani, T., Hefti, A., Baumeister, W., Aebi, U., et al.
(1992). Ultramicroscopy, 46, 45.
5. Uzgiris, E. E. and Kornberg, R. D. (1983). Nature, 301, 125.
6. Harris, J. R. (1992). Microsc. Anal., 13.
7. Avila-Sakar, A. J., Guan, T. L., Arad, T., Schmid, M. F., Loke, T. W., Yonath, A.,
et al. (1994). J. Mol. Biol., 239, 689.
8. Kubalek, E. W., Kornberg, R. D., and Darst, S. A. (1991). Ultramicroscopy, 35,
295.
362
12: Two dimensional crystallization of soluble proteins
9. Mosser, G., Mallouh, V., and Brisson, A. (1992). J. Mol. Biol., 226, 23.
10. Celia, H., Hoermann, L., Schultz, P., Lebeau, L., Mallouh, V., Wigley, D. B., et al.
(1994). J. Mol. Biol., 236, 618.
11. Avila-Sakar, A. J. and Chiu, W. (1996). Biophys. J., 70, 57.
12. Kornberg, R. D. and Darst, S. A. (1991). Curr. Opin. Struct. Biol, 1, 642.
13. Brisson, A., Olofsson, A., Ringler, P., Schmutz, M., and Stoylova, S. (1994). Biol.
Cell, 80, 221.
14. Gaines, G. L., Jr. (1966). Insoluble monolayers at liquid-gas interphases. Wiley,
New York.
15. Roberts, G. (ed.) (1990). Langmuir-Blodgett films. Plenum Press, New York.
16. Mosser, G. and Brisson, A. (1991). J. Struct. Biol, 106, 191.
17. Venien-Bryan, C., Lenne, P.-F., Zakri, C., Renault, A., Brisson, A., Legrand, J.-F.,
et al. (1998). Biophys. J., 74, 2649.
18. Brisson, A., Bergsma-Schutter, W., Oling, F., Lambert, O., and Reviakine, I.
(1999). J. Cryst. Growth, 196, 456.
19. Blankenburg, R., Meller, P., Ringsdorf, H., and Salesse, C. (1989). Biochemistry,
28, 8214.
20. Frey, W., Schief, W. R., and Vogel, V. (1996). Langmuir, 12, 1312.
21. Reviakine, L, Bergsma-Schutter, W., and Brisson, A. (1998) J. Struct. Biol, 121,
356.
22. Baumeister, W. and Hahn, M. (1978). In Principles and techniques of electron
microscopy: biological applications (ed. M. A. Hayat), Vol. 8, p. 1. Van Nostrand
Reinhold Co., New York.
23. Schmutz, M., Lang, J., Graff, S., and Brisson, A. (1994). J. Struct. Biol, 112, 252.
24. Sjostrand, F. S. (1956). In Stockholm Conf. Electron Microscopy, Proc. 20.
25. Fukami, A. and Adachi, K. (1965). J. Electron Microsc., 14, 112.
26. Chretien, D., Fuller, S. D., and Karsenti, E. (1995). J. Cell Biol, 129, 1311.
27. Asturias, F. J. and Kornberg, R. D. (1995). J. Struct. Biol, 114, 60.
28. Henderson, R., Baldwin, J. M., Ceska, T. A., Zemlin, F., Beckmann, E., and
Downing, K. H. (1990). J. Mol Biol, 213, 899.
29. Voges, D., Berendes, R., Burger, A., Demange, P., Baumeister, W., and Huber, R.
(1994). /. Mol Biol, 238, 199.
30. Darst, S. A., Ahlers, M., Meller, P. H., Kubalek, E. W., Blankenburg, R., Ribi, H.
O., et al. (1991). Biophys. J., 59, 387.
31. Ringler, P., Mttller, W., Ringsdorf, H., and Brisson, A. (1997). Chem. Eur. J., 3,
620.
32. Huetz, P., van Neuren, S., Ringler, P., Kremer, F., van Breemen, J. F. L.,
Wagenaar, A., et al (1997). Chem. Phys. Lipids, 89, 15.
33. Wilson-Kubalek, E. M., Brown, R. E., Celia, H., and Millagan, R. A. (1998). Proc.
Natl Acad. Sci. USA, 95, 8040.
34. Edwards, A. M., Darst, S. A., Hemming, S. A., Li, Y., and Kornberg, R. D. (1994).
Nature Struct. Biol., 1, 195.
35. Olofsson, A., Mallouh, V., and Brisson, A. (1994). J. Struct. Biol., 113, 199.
363
This page intentionally left blank
13
Soaking techniques
E. A. STURA and T. GLEICHMANN
1. Introduction
Once crystals of a macromolecule are obtained there are many circumstances
where it is necessary to change the environment in which the macromolecule
is bathed. Such changes include the addition of inhibitors, activators, sub-
strates, products, cryo-protectants, and heavy atoms to the bathing solution to
achieve their binding to the macromolecule, which may have sufficient free-
dom to undergo some conformational changes in response to these effectors.
In fact, macromolecular crystals have typically a high solvent content which
ranges from 27-95% (1, 2). Although, part of this solvent, 'bound solvent'
(typically 10%) is tightly associated with the protein matrix consisting of both
water molecules and other ions that occupy well defined positions in refined
crystal structure it can be replaced in soaking experiments, at a slower rate
compared to the 'free solvent'.
In this chapter we will consider the relative merits of various methods for
modifying crystals, the restraints that the lattice may impose on the macro-
molecule, and the relative merits of soaking compared to co-crystallization.
A crystal suitable for data collection will contain from 5-100 ug of protein,
a proportionate amount of ligand must be present in the soaking solution
so that a stoichiometric or higher concentration is achieved in an
appropriate solution compatible with crystal stability. The exact molarity
or stoichiometric ratio required will depend on the affinity of the
compound and is likely to vary from case to case.
Method
1. Measure 1-10 mg of ligand. Make a saturated solution of the ligand in
a suitable solvent (see Table 7) by adding solvent to the ligand until
fully dissolved. Calculate the molarity of the solution obtained.
367
E. A. Stura and T. Gleichmann
Protocol 1. Continued
2. Test the solubility by adding 1 ul of the saturated solution to 1 ml of
the same buffer solution used for the protein. Continue adding until
the solution becomes opalescent. The molarity of the resulting
solution can be calculated.
3. If the solubility is in the millimolar range, soaking can be done directly
in the drops or capillaries (Protocol 5). If the solubility is less than
0.2 mM a soaking volume of 300 ul or greater will be needed and spot
plates or vials should be used (Figure 1).
4. Mix the appropriate volume of ligand saturated solution with the
precipitant used in the crystallization experiment so that the desired
molarity or so that ligand-protein stoichiometry will be achieved.
5. Test that the precipitant-solvent mixture to be used for soaking is
compatible with the crystal. For volatile solvents, just replace the
reservoir solution with the precipitant-solvent mixture and allow the
crystal to equilibrate with the reservoir by vapour diffusion. Check for
cracks and if possible test that crystals equilibrated in such a manner
still diffract. Later exchange the mother liquor in the drop with this
solution and repeat checks.
a Typical concentrations and usage for organic solvents and additives. As it is suggested throughout
the table, combinations of these compounds can be more effective to solubilize ligands and less likely
to be incompatible with the crystals. Several organic compounds are suitable both as additives to
crystallization set-ups as for use as cryo-solvents. For example the effect of slow equilibration of MPD
onto crystals of the multisubstrate adduct complex of glycinarnide ribonucleotide transformylase,
which under room temperature conditions diffracted to only 2.0 A was to extend the resolution to 1.96
A at cryogenic temperature collected with a conventional X-ray source (35). The improvement in
resolution may have been due to MPD rather than cryo-cooling.
which have been used to change the mother liquor in which the crystals are
bathed, such as for the introduction of a substrate, are well described
elsewhere (4, 5), pressure cells are used far the incorporation of krypton and
xenon into crystals.
Buffers are exchanged in order to change the pH, to analyse pH-induced
changes, to favour heavy-atom or drug binding, or to avoid conditions of
370
13: Soaking techniques
Figure 1. Schematic drawing of the transfer of crystals from either a sitting drop set-up,
or a hanging drop set-up, to a well for soaking. (A) A crystal is dislodged from the drop
using a probe. If the crystal adheres firmly to the glass (or plastic) use a sacrificial crystal
placed between the probe and the crystal to push against. The crystal is then floated to
the surface to easily be picked up as in Figure 3 or in (B). (B) Crystals can be transferred
directly from a sitting drop to a capillary because the vapour from the reservoir solution
(not shown) protects the drop containing the crystal from dehydration. For a hanging
drop vapour diffusion experiment the coverslip is placed at the centre of a plastic Petri
dish within a ring of filter paper soaked in water. Evaporation of water from the filter
paper will ensure that the drop does not dry out while the crystal is picked up into the
capillary connected to the syringe. (C) The crystal is transferred into a small vial or a well
in a spot plate containing the soak solution. (D) After soaking the crystal is removed from
the soak solution for mounting. The walls of the vial used for soaking the crystal should
not be high, as this will restrict the angle at which the crystal can be picked up from the
soak solution into the capillary, as further restrictions are also imposed by the dissecting
microscope, which also limits the working angle. For soak volumes below 1 ml a spot
plate may be preferable.
Method
1. Connect a glass or quartz capillary tube to a 1 ml glass syringe with a
short piece of rubber tubing such as c-flex (Fisher, 14-169-5c) which
gives an excellent seal.
2. Snap open the end of the capillary with tweezers or scissors. The glass
capillary may be siliconized if the experimental situation can benefit
from a diminished adhesion of the solution to the glass wall such as
when viscous solutions are handled. After siliconizing it should be
extensively washed.
3. To increase the volume available for handling crystals, mother liquor
(20-50 ul) can be added to the drop (some crystals require the mother
liquor to contain protein for stability). If the crystals adhere to the well,
withdraw liquid from the drop and gently eject it onto a chosen crystal.
Check that the desired crystal moves in the flow.
4. If flushing with liquid fails to dislodge the crystals the probes for streak
seeding (Chapter 5) are used for this purpose. Select a thick whisker
with a sharp point or cut a new point if needed. Run the point around
the contour of the crystal, this will detach the crystal from precipitated
or denatured protein in the depression. Now push gently on the crystal
with the wide side of the whisker. Slowly apply pressure and watch for
movement. Should the crystal show signs of breaking up or cracking,
select a smaller crystal that can be sacrificed and utilize it as shown in
372
13: Soaking techniques
Figure 1. Using a loop to dislodge the crystal is another option
(Protocol 4). Unfortunately, some crystals have severe adhesion
problems and cannot be dislodged without breaking them. For such
problematic crystals, glass pots with depressions, or microbridges,
coated with a thin film of Corning vacuum silicone grease should be
used in the original crystallization set-up.
5. Pick up crystals into the capillary by pulling back the plunger of the
syringe.
6. Transfer crystals from the syringe directly into a soak solution from
which they are later picked up and mounted for X-ray studies.
Crystals grown by hanging drop (see Chapter 5) can be flushed with mother
liquor from the coverslip into a larger container and then picked up as
described in Protocol 2 for sitting drops. Since it can be difficult to find small
crystals in a large container, the method described in Protocol 3 may be
preferable.
Method
1. Cut a circular piece of filter paper to fit a Petri dish 4-5 cm inside
diameter.
2. Cut out a small circle from the centre of the filter paper such that the
coverglass from the hanging drop can fit inside this without touching
the paper.
3. Soak the filter paper with distilled water.
4. Place the coverglass with the hanging drop in the centre. Mother
liquor is added to the drop (20-50 ul) and the crystals for soaking can
be picked up and soaked as described in Protocol 2. steps 2-6.
This set-up has been used for the stable transportation of crystals to
synchrotron facilities, by soaking the filter paper with mother liquor instead of
water.
Instead of a capillary and a syringe a loop can be used for the handling of
crystals as described in Protocol 4. This method is widely used for soaking
crystals in cryo-solvents.
373
E. A. Stum and T. Gleichmann
Method
1. Follow Protocol 2, steps 1 and 2 if it is necessary to increase the
volume of the drop. Select a loop with a diameter about 1.5 times the
maximum size of the crystal. Loops can be made with individual fibres
from plain dental floss or can be purchased pre-made from Hampton
Research.
2. Clean the loop in methanol and wash with water. If the crystal is stuck
it can be dislodged as in Protocol 2, steps 3 and 4 or by gently pushing
with the loop. Tease the crystal to the top of the drop. When close to
the surface of the drop, place the loop under the crystal and lift it out
of the drop. Keep the loop only slightly above the drop, to avoid it
drying out and focus the microscope on the loop to ensure that the
crystal is in the loop.
3. Transfer the crystal to the soaking well as rapidly as possible. Drying
out of the solvent around the crystal is the main disadvantage of the
method.
374
13: Soaking techniques
Figure 2, Schematic representation of the various stages involved in capillary soaks. (Al
After the crystal is picked up into the capillary (Figure 1A) the mother liquor is removed
from around the crystal by pushing the liquid out, while holding the crystal in position
with a hair (in many cases the surface tension between the crystal and the capillary is
sufficient to hold the crystal in place). (B) With a thin strip of filter paper, taking care not
to touch the crystal the excess liquid is removed. (C) The soak solution is drawn into the
capillary to bathe the crystal. (D) Paraffin oil or buffer is added to the open end and the
capillary is sealed with molten wax. (E) After soaking the capillary can be snapped open
with thin-nosed forceps and the soak solution is removed using a thin piece of filter
paper. (F) The capillary can now be sealed at both ends and the crystal used for X-ray
diffraction studies. (G) Crystals that have been used for X-ray work can be soaked by
snapping of one end of the sealed capillary with forceps and opening the other end with
a hot needle. A piece of wet filter paper is placed over the crystal to prevent the crystal
from warming up during this procedure. A solution is then introduced at the broken end
of the capillary so that it bathes the crystal. Petroleum jelly is used to seal the experiment
as it is easy to remove prior to resealing the thin walled capillary tube with wax. (H) The
soaked crystal can be used for X-ray diffraction analysis.
375
E. A. Stum and T. Gleichmann
the soak solution are then removed and the crystal used for further X-ray
studies such as for collecting an inhibitor complex data set after the native
protein data have been measured, if the crystal has survived the damage from
the first irradiation.
Method
1. Suck the new solution into the capillary fully immersing the crystal.
2. Add paraffin oil to the open end of the capillary, leaving an air gap
between the oil and the soak solution, for the duration of the soak.
3. After the soak period has elapsed the oil and the solution are removed.
4. Remove the excess solution around the crystal with filter paper.
5. Add either soak solution and or oil to the open end of the capillary to
maintain a moist environment for the crystal.
6. Seal with wax while still attached to the syringe. A wet strip of filter
paper (5 mm wide) can be placed on the outside of the capillary to
keep the crystal cool while the ends are sealed. The crystal is now
mounted for X-ray diffraction work. Other techniques for mounting
crystals can be found in Chapter 14 and elsewhere (10).
376
13: Soaking techniques
Method
1. Follow Protocol 2 using a large depression plate (Figure 1) or large
volume in a soak vial.
2. Soak several crystals for 20 min to several days. Harvest one crystal at
a time, mount in a capillary, and test to determine whether the ligand
or heavy-atom has bound to the crystal.
3. Remove old soak solution leaving the remaining crystals in the vial or
depression plate and add fresh solution. Continue testing the crystals
and replacing or adding more ligand or heavy-atom solution.
3. Soaking application
In this section we will consider some of the more typical applications of
soaking: heavy-atom derivatization and the soaking for cryo-crystallography.
80
Table 2. Hg—mercury compoundsa
a Mercury compounds are targeted to sulfhydryl groups. Short soak times 1-3 h can produce useful
derivatives with concentrations as low as 0.01 mM. Mercury has also a strong tendency to bind to zinc
sites at a histidine nitrogen. The mercury compounds can be grouped in three classes: the ionic group,
the most commonly used are mercury chloride and mercury acetate; the alkyl chain mercury
compounds, ethyl mercury phosphate, ethyl mercury chloride, methyl mercury chloride, methyl
mercury acetate; and the aromatic group, the most popular being EMTS, mersalyl, pCMBS, and
pCMB. K2Hgl4 cannot be grouped with the ionic mercurials as it tends to give different results, but it is
definitely a compound worth trying. Baker's dimercurial consisting of two mercury atoms, and TAMM,
a heavy metal cluster of four mercury atoms which has been used in the phasing of large molecular
assemblies (36-39), have good solubility and are definitely worth trying.
"See Table 13.
380
13: Soaking techniques
381
E. A. Stura and T. Gleichmann
82
Table 5. Pb—lead compoundsa
81
Table 6. TI—thallium compoundsa
382
13: Soaking techniques
76
Table 8. Os—osmium compoundsa
74
Table 10. W—tungsten compoundsa
383
E. A. Stura and T. Gleichmann
" lodination is well described in refs 5 and 50. Certain iodide salts
of heavy metals may derivatize proteins due to I" rather than
the metal itself. Tantalum and niobium compounds are of interest
for their use in phasing large macromolecular assemblies (3).
Brominated and iodinated nucleotides are used in the phasing of
protein nucleotide complexes.
bSee Table 13.
384
13: Soaking techniques
Ac: Acros Chemicals; 711 Forbes Avenue, Pittsburgh, PA 15219, USA. Tel: (800) 227-6701.
http://www.fishersci.com/catalogs
Af: Alfa; Johnson Matthey Catalog Company, Inc., PO Box 8247, Ward Hill, MA 01835-
0747, USA. Tel: (800) 343-0660.
Al: Aldrich Chemical Co.; 1001 West St Paul Avenue, Milwaukee, Wl 53233, USA. Tel:
(800) 558-9160.
http://www.sigma.com/SAWS.nsf/Pages/Aldrich?EditDocument
An: Anatrace Inc.; 434 West Dussel Drive, Maumee, OH 43537-1624, USA. Tel: (800) 252-
1280. http://www.anaTrace.com
CS: Chem Service; PO Box 3108, West Chester, PA 19381-3108, USA. Tel: (610) 692-3026.
Fi: Fisher Scientific; 711 Forbes Avenue, Pittsburgh, PA 15219, USA. Tel: (800) 766-7000.
http://www.fishersci.com/catalogs
FI: Fluka Chemie AG; Industriestrasse 25, CH-9471 Buchs, Switzerland.
http://www.sigma.aldrich.com/SAWS.nsf/Pages/Fluke?EditDocument
I: ICN Pharmaceutical Inc.; 3300 Hyland Avenue, Costa Mesa, CA 92626, USA. Tel: (714)
545-0100.
M: Mallinckodt; 470 Frontage Road, West Haven, CT 06516, USA. Tel: (203) 933-7064.
N: Noah Technologies; 1 Noah Park, San Antonio, TX 28249, USA. Tel: (210) 691-2000.
P: Pfaltz and Bauer; 172 E Aurora Street, Waterbury, CT, USA. Tel: (203) 574-0075.
Si: Sigma Chemical Company; PO Box 14508, St. Louis, MO 63178, USA.
http://www.sigma.aldrich.com/SAWS.nsf/Pages/Sigma7EditDocument
Sp: SPECS and BioSPECS bv; Koninginnegracht, 94-95, 2514 AK The Hague, The
Netherlands. PO Box 85586, 2508 CG The Hague, The Netherlands (mailing address)
Tel: 31-70-355-4473. Fax: 31-70-355-8527.
Brandon/SPECS Inc.; (North American sales company), PO Box 1244, Merrimack,
New Hampshire 03054, USA. Tel: 603-424-2035. Fax: 603-424-2035.
St: Strem Chemical; 7 Mulliken Way, Newburyport, MA 01950, USA. Tel: (508) 462-3191.
aData for Tables 2-12 has been compiled from structures reported in: Macromolecular structures
(1991-4), and Atomic structures of biological macromolecules (1990-3) (ed. W. A. Hendrickson and K.
Wuthrich). Current Biology Ltd., London.
386
13: Soaking techniques
Figure 3. (A) Prepare a loop (cryo-loop) from a rayon fibre. This is best done by placing
both ends of the fibre into a needle and pulling the two ends until a loop of the correct
diameter is obtained. Place a spot of epoxy at the junction of the fibre. (b) Pick up the
crystal from the drop using the loop mounted on a wooden or glass rod. (C) Soak crystal
in spot plate (easier than in vials). For long soaks, place a ring of petroleum jelly or
vacuum grease around the well, and cover with a coverglass. Place the spot plate in a
drawer for platinum and other light-sensitive heavy-atom solutions. Warning: loops used
for heavy-atom work may retain some heavy-atom soaked in the fibre. (D) After the soak,
pick up the crystal from the soak solution. (E) If the drop attached to the loop is very large,
the excess liquid can be removed by touching the outside of the loop with a thin strip of
filter paper, being careful not to get close to the crystal. (F) The crystal is then plunged
into the nitrogen stream or the stream is blocked by a card or ruler and when the crystal
is in position the card is rapidly removed. In the design used at SSRL a thin bent brass
plate clips onto the nozzle and diverts the beam while the crystal is being positioned.
When the crystal is in place the brass plate is made to springs back allowing the stream
to flash-freeze the crystal.
Method
1. Select crystals for freezing of roughly 0.4 mm in each dimension or
smaller. Larger crystals are more problematic as they may develop
cracks, and form crystalline ice due to slow heat transfer. Since the
radiation damage is small or negligible at cryo-temperature, the
strength of the X-ray source and exposure times will be able to
compensate for the smaller crystal size. Fast data collection on a
wiggler line at a synchrotron radiation source is preferable than long
exposure times with conventional sources.
2. Select a cryo-solvent. See Table 1 for suggestions. Garman and
Mitchell (28) give the minimum amount of glycerol to be added to 50
typical crystallization conditions. The cryo-buffer typically is a
combination of the crystallization reservoir and a cryo-solvent. For
crystallizations from small molecular weight PEG (200-600) or MPD no
cryo-solvent is required, although the precipitant concentration may
need to be increased. If the crystallization is from PEG 4000 or higher
molecular weight PEG, replace some of the PEG 4000 by PEG 200 or
just add PEG 200 at varying concentrations. The next most popular
choices are ethylene glycol or MPD. The alcohol sugars are generally
milder, and may be tolerated by those crystals that crack under the
other conditions.
387
E. A. Stura and T. Gleichmann
Protocol 7. Continued
3. A small amount of this buffer is picked up in a loop (Figure 3) and
shock-frozen in the nitrogen stream. Plunge loop into nozzle to flash-
freeze or block stream with a paper card, then remove it quickly. If the
buffer stays transparent it has formed vitreous ice, whereas
opaqueness indicates formation of crystalline ice which will give a
powder diffraction pattern, 'ice-rings'. If there is no loss in resolution
or increased mosaicity (30) or anisotropy minimal ice-rings can be
tolerated as most integration programs can cope with this problem if
not too severe.
4. The next step is to optimize the cryo-solvent. Check whether the cryo-
protectant causes lattice damage, this normally results in cracks, very
fine hair-line cracks manifest themselves as a brown tinge when the
crystal is observed under the microscope. Take a few diffraction
images at various concentrations and with different cryo-protectants
at room temperature with crystals mounted in capillaries to select the
least damaging cryo-solvent and to maximize the resolution limit.
Some cryo-solvents may indeed enhance diffraction. Since this pro-
cess can take a long time, if the crystals are large enough to be
analysed on a conventional source, it is best to perform these tests in
advance of synchrotron data collection. Often the functional plot of
cryo-protectant versus resolution limit has a minimum (30).
5. Pull crystal through cryo-protectant in the loop to transfer the crystal
to the cryo-buffer (Figure 3). In a stepwise transfer using solutions with
increasing amounts of cryo-protectant is important in order to reduce
osmotic shock. It is also important to keep the number of operations
small to reduce damage and subsequent increase of mosaicity. For a
small crystal 20-30 sec between transfers is sufficient and the crystal
should be frozen immediately, else the mosaicity might increase.
Evaporation of buffer also requires speedy transfer to the cryo-stream.
4. Conclusions
Soaking is most commonly used to obtain heavy-atom derivatives, although
crystallization of previously modified proteins either chemically or biologically
have also been used (31, 32). Soaking and co-crystallization are two different
approaches to achieving complexes of macromolecules. The two procedures
are both alternative and complementary to each other. By soaking effectors
into preformed crystals it is possible to analyse the structure of complexes,
only if crystal lattice constraints permit. The problem of cracking of the
crystals which may occur both when binding effectors (15) and heavy-atoms
(33), can be often resolved by the use of cross-linking agents. One must how-
ever understand that complexes obtained by soaking may differ from com-
plexes obtained by co-crystallization. Flow cells in which a constant supply of
388
13: Soaking techniques
substrate is supplied to the enzyme in the crystal and product is washed away
may answer the problem in cases where the rate of product formation is
significantly slower than the rate of diffusion through the crystal (19, 34).
Acknowledgements
We would like to thank Dr Ping Chen for her contribution to the first edition
of this chapter and Dr Ian A. Wilson for reading and support of that work
through his grants by the National Institutes of Health Grants AI-23498, GM-
38794, and GM-38419. T. G. was supported by BMBF grant 05 641BJA 4 (to
Rolf Hilgenfeld) and by the Australian Research Council grant AD 984283
(to B. Kobe). E. S. thanks the French Atomic Energy Commission (CEA) for
support during the revision work.
References
1. Matthews, B. W. (1968). /. Mol. Biol. 33, 419.
2. Cohen, C., Caspar, D. L. D., Parry, D. A. D., and Lucas, R. M. (1971). Cold Spring
Harbor Symp. Quant. Biol., 36, 205.
3. Thygesen, J., Weinstein S., Franceschi, F., and Yonath, A. (19%). Structure, 4, 513.
4. Wyckoff, H. W., Doscher, M. S., Tsernoglou, D., Inagami, T., Johnson, L. N.,
Hardman, K. D., et al. (1967). J. Mol. Biol., 127, 563.
5. Petsko, G. A. (1985). In Methods in enzymology (eds H. W. Wychoff, C. H. W.
Hirs, and S. N. Timasheff), Academic Press, London. Vol. 114, p. 141.
6. Remington, S., Wiegand, G., and Huber, R. (1982). J. Mol. Biol., 158, 111.
7. Kasvinsky, P. J. and Madsen, N. B. (1977). J. Biol. Chem., 251, 6852.
8. Hadju, J., Acharya, K. R., Stuart, D. L, McLaughlin, P. J., Barford, D.,
Oikonomakos, N. G., et al. (1987). EMBO J., 6, 539.
9. McPherson, A. (1982). Preparation and analysis of protein crystals. Wiley, New
York.
10. Blundell, T. L. and Johnson, L. N. (1976). Protein crystallography. Academic
Press, New York.
11. Quicho, F. A. and Richards, F. M. (1964). Proc. Natl. Acad. Sci. USA, 52, 833.
12. Hunter, M. J. and Ludwig, M. L. (1962). /. Am. Chem. Soc., 84, 3491.
13. Browne, D. T. and Kent, S. B. H. (1975). Biochem. Biophys. Res. Commun., 67,
126.
14. Dombradi, U., Hadju, J., Bot, G., and Friedrich, P. (1980). Biochemistry, 19, 2295.
15. Lorek, A., Wilson, K. S., Sansom, M. S. P., Stuart, D. L, Stura, E. A., Jenkins, J.
A., et al. (1984). Biochem. J., 218, 45.
16. Green, D. W., Ingram, V. M., and Perutz, M. F. (1954). Proc. Roy. Soc., A225, 287.
17. Sigler, P. B. and Blow, D. M. (1965). J. Mol. Biol., 14, 640.
18. Mondragon, A., Wolberg, C., and Harrison, S. C. (1989). J. Mol. Biol., 205, 179.
19. Petsko, G. A. (1985). In Methods in enzymology ibid ref. 5. Vol. 114, p. 147.
20. Schiltz, M., Prange, T., and Fourme, R. (1994). J. Appl. Cryst., 27, 950.
21. Stowell, M. H. B., Soltis, S. M., Kisker, C., Peters, J. W., Schindelin, H., Rees, D. C.,
et al. (1996). J. Appl. Cryst., 29, 608.
22. Rodgers, D. W. (1994). Structure, 2,1135.
389
E. A. Stum and T. Gleichmann
23. Tsernoglou, D., Hill, E., and Banaszack, L. J. (1972). J. Mol. Biol, 69, 75.
24. Hope, H., Frolow, R, von Bohlen, K., Makowski, I., Kratky, C., Halfon, Y., et al.
(1989). Acta Cry St., B45,190.
25. Petsko, G. A. (1975). J. Mol. Biol., 96, 381.
26. Douzou, P., Hui Bon Hoa, G., and Petsko, G. A. (1975). J. Mol. Biol., 96, 367.
27. Hadju, J., Machin, P. A., Campbell, J. W., Greenhough, T. J., Clifton, I. J., Zurech,
S., et al. (1987). Nature, 329, 176.
28. Garman, E. F. and Mitchell, E. P. (1996). J. Appl. Cryst., 29, 584.
29. Hope, H. (1988). Acta Cryst., B44,22.
30. Mitchell, E. P. and Garman, E. F. (1994). J. Appl. Cryst., 27,1070.
31. Stura, E. A., Johnson, D. L., Inglese, J., Smith, J. M., Benkovic, S. J., and Wilson,
I. A. (1989). J. Biol. Chem., 264, 9703.
32. Yang, W., Hendrickson, W. A., Crouch, R. J., and Satow, Y. (1990). Science, 249,
1398.
33. Ringe, D., Petsko, G. A., Yakamura, F., Suzuki, K., and Ohmori, D. (1983). Proc.
Natl. Acad. Sci. USA, 80, 3879.
34. Farber, G. K., Glasfeld, A., Tiraby, G., Ringe, D., and Petsko, G. A. (1989).
Biochemistry, 28, 7289.
35. Klein, C., Chen, P., Arevalo, J. H., Stura, E. A., Marolewski, A., Warren, M. S., et
al. (1995). J. Mol. Biol., 249, 53.
36. Deisenhofer, J., Epp, O., Miki, K., Huber, R., and Michel, H. (1984). J. Mol. Biol.,
180, 385.
37. O'Holloran, T. V., Lippard, S. J., Richmond, T. J., and Klug, A. (1987). J. Mol.
Biol., 194, 705.
38. Bentley, G. A., Boulot, G., Riottot, M. M., and Poljak, R. J. (1990). Nature, 348, 254.
39. Reinemer, P., Dirr, H. W., Ladenstein, R., Schaeffer, J., Gallay, O., and Huber, R.
(1991). EMBO J., 10, 1997.
40. Garcia, K. C., Degano, M., Stanfield, R. L., Brunmark, A., Jackson, M. R.,
Peterson, P. A., et al. (1996). Science, 274,209.
41. Kim, S.-H., Shin, W.-C, and Warrant, R. W. (1985). In Methods in enzymology
ibid ref. 5. Vol. 114, p. 156.
42. Stout, C. D., Mizuno, H., Rao, S. T., Swaminathan, P., Rubin, J., Brennan, T., et al.
(1978). Acta Cryst., B34, 1529.
43. Livnah, O., Stura, E. A., Johnson, D. L., Mulcahy, L. S., Wrighton, N. C., Dower,
W. J., et al. (1996). Science, 273, 464.
44. Egloff, M. P., Cohen, P. T., Reinemer, P., and Barford, D. (1995). J. Mol. Biol,
254, 294.
45. Stura, E. A., Satterthwait, A. C., Calvo, J. C., Kaslow, D. C., and Wilson, I. A.
(1994). Acta Cryst., D50, 448.
46. Prasad, G. S., Stura, E. A., McRee, D. E., Laco, G. S., Hasselkus-Light, C., Elder,
J. H., et al. (1996). Protein Sci., 5, 2429.
47. Stura, E. A., Ghosh, S., Garcia-Junceda, E., Chen, L., Wong, C.-H., and Wilson, I.
A. (1995). Proteins, 22, 67.
48. Weaver, T. M., Levitt, D. G., Donnelly, M. L, Stevens, P. P., and Banaszak, L. J.
(1995). Nature Struct. Biol., 2, 654.
49. Ladenstein, R., Bacher, A., and Huber, R. (1987). J. Mol. Biol. 195, 751.
50. Ghosh, D., Enman, M., Sawicki, M., Lala, P., Weeks, D. R., Li, N. et al. (1999).
Acta Cryst. D55, 779.
390
a4
X-ray analysis
L. SAWYER and M. A. TURNER
1. Introduction
This chapter covers the preliminary characterization of the crystals in order to
determine if they are suitable for a full structure determination. Probably
more frustrating than failure to produce crystals at all, is the growth of beauti-
ful crystals which do not diffract, which have very large unit cell dimensions,
or which decay very rapidly in the X-ray beam, though this last problem has
been largely overcome by freezing the sample.
It is impossible in one brief chapter to give more than a flavour of what the
X-ray crystallographic technique entails and it is assumed that the protein
chemist growing the crystals will have contact with a protein crystallographer,
who will carry out the actual structure determination and in whose laboratory
state-of-the-art facilities exist. However, preliminary characterization can
often be carried out with little more than the equipment which is widely avail-
able in Chemistry and Physics Departments and so the crystal grower remote
from a protein crystallography laboratory can monitor the success of their
experiments. The reader should refer to the first edition for protocols useful
for photographic characterization but such techniques are seldom used
nowadays. It must be remembered, in any case, that X-rays are dangerous and
the inexperienced should not try to X-ray protein crystals without help.
2.1 X-rays
2.1.1 Why use X-rays
The scattering or diffraction of X-rays is an interference phenomenon and the
interference between the X-rays scattered from the atoms in the structure
produces significant changes in the observed diffraction in different direc-
tions. This variation in intensity with direction arises because the path
differences taken by the scattered X-ray beams are of the same magnitude as
the separation of the atoms in the molecule. Put another way, to 'see' the
individual atoms in a structure, it is necessary to use radiation of a similar
wavelength to the interatomic distances, typically 0.15 nm or 1.5 A and
radiation of that wavelength lies in the X-ray region of the electromagnetic
spectrum. It is also important to realize that it is the electrons which scatter
the X-rays and so what is in fact observed is the electron density of the sample.
Because the electrons cluster round the atomic nuclei, regions of high
electron density correspond to the atomic positions.
Figure 1. Each of the unit cells shown in this two-dimensional example is a valid choice
for the lattice of points. The cell on the right is a centred cell and has twice the contents
of the others.
393
Table 1. The crystal systems and related data for a chiral molecule
Figure 2. X-rays (X1 X2, X3) reflected from lattice planes A, B, C. To observe a scattered
beam of X-rays in direction R, the thickened path must equal a whole number of
wavelengths. The ray from plane C travels twice as far as that from B, and so on.
Figure 3. The set of planes 123 are shown as they cut a unit cell. The intercepts on b
occur every 1/2 and on c every 1/3.
Figure 4. A diagram illustrating the relationship between sets of planes in a crystal in real
or direct space and points representing a diffracted X-ray beam in reciprocal (or diffrac-
tion) space. Notice that the direction from the origin of reciprocal space (large point) to
any point, e.g. 130, is perpendicular to the planes in the crystal and that the length is
proportional to the reciprocal of the plane spacing.
The points can be seen to make up another lattice (reciprocal lattice) whose
axes and angles are derived from those of the crystal. This idea can be
extended to three dimensions. It is important to realize that each reflection
contains a contribution from every atom in the crystal and, conversely, each
atom in the crystal contributes to every reflection. Thus, as the crystal is
moved about in the X-ray beam, reflections flash out and can be recorded
when the geometrical arrangement of X-ray beam, crystal orientation, and
detector satisfies Bragg's Law.
To help understand diffraction from a crystal, there is a construction intro-
397
L. Sawyer and M. A. Turner
Figure 5. The Ewald construction. For clarity, this is shown as a planar diagram but IXO is
the diameter of a sphere of radius 1/\.
duced by Ewald and shown in Figure 5. As we move the crystal, the reciprocal
lattice also moves about a fixed origin. With the crystal, X, as centre, a sphere
is drawn of radius 1/\. and the origin, O, of the reciprocal lattice is taken as the
point where the X-ray beam leaves the sphere after passing through the
crystal. As the crystal is rotated about the z axis (perpendicular to the page)
the reciprocal lattice rotates until the point P lies on the surface of the sphere.
The point P is the 410 reflection arising from the planes of spacing d410. The
angles at IX and XP, i.e. IXA and BXP are equal to 0 so that OXP = 20 and
OP is perpendicular to the crystal planes AXB. Now OP = 2 X XO X sin 0 =
2 X (1/X) X sin 0. However, OP = l/d410 and so l/d410 = (2/\) X sin 0 which is
Bragg's Law. Thus, the Ewald sphere gives a readily understandable way of
relating the orientation of the crystal to the diffraction pattern observed. In
order to collect a set of X-ray data, it is necessary to move the crystal (and in
some methods, the detector) in such a way that every reciprocal lattice point
passes through the sphere of reflection (Figure 6). There are various ways of
achieving this, some of which are described in Section 4.
The space group in which a molecule crystallizes may impose certain con-
ditions on the reflections which can be observed so that by looking at the
diffraction pattern of the crystal, it is often possible to determine the space
group unambiguously. Furthermore, the higher the symmetry of the crystal,
the less data is actually required to be collected. A diffraction pattern has a
centre of symmetry since reflections in opposite directions from the same
planes must have the same intensity (I(h k l) = l(h k l) is Friedel's Law) (see
Figure 3). Thus the diffraction symmetry shown in Table 2 has a centre
of symmetry even though the space groups do not. The effects of the lattice
type and symmetry elements upon the diffraction pattern are shown in Table 3
398
14: X-ray analysis
Figure 6. The Ewald sphere intersected by several reciprocal lattice layers. As the crystal
is moved, the reciprocal lattice pivots about O but XA and OB remain parallel at all times.
As shown, a film placed perpendicular to XO will record a series of concentric circles. As
the crystal is rotated through a small angle about AX, the circles will become extended
into lunes as the neighbouring spots on each level pass through the sphere of reflection.
and the effect can be explained with reference to Figure 2. If the beam X3
scattered from row C is one wavelength behind X1 scattered from row A, then
X2 scattered from row B is exactly half a wavelength behind and it will cancel
out the reinforcing contributions from rows A and C. Thus, interposing planes
midway between the planes separated by the unit cell repeat as is the case for
a centred lattice, leads to a systematic absence of reflections. Further, if a
twofold screw axis is perpendicular to the planes, there will always be an
identical (but rotated) set of scatterers to row A, on row B. Only when the
index is even along the axial direction will constructive interference occur and
the reflection be observed. Notice that simple rotation axes do not generate
any systematic absences.
aThe reflections listed here are identical . If Friedel's Law holds then l(hkl) = l(hkl) and this generates
an equal number of equivalent reflections. In protein crystallography, anomalous scattering which
leads to a breakdown in Friedel's Law, is used to help with phasing the reflections and so the two sets,
equivalent to I(hkl) and I(hkl) must be kept separate.
bThe underlined reflections are those which are required to specify the Laue symmetry with the
others being generated by repeated application of the symmetry elements.
cThe axes in the trigonal and hexagonal systems referred to here are a = b, c, a = B = 90°, y = 120°
dWhen hexagonal axes are being used, i = hk.
3. Mounting crystals
Mounting a protein crystal is a procedure which requires a reasonable degree
of manual dexterity. It is impossible to be dogmatic about the right and wrong
way, and each person develops their own technique, modifying it as required
from protein to protein depending on the size, strength, temperature behaviour,
need to exclude oxygen, or toxicity. Although early workers did dry their
crystals (14), drying out of mother liquor in the crystal generally disrupts it
such that no useful data can be collected. Mounting methods are therefore
designed to maintain the interstitial mother liquor as it is in the drop from
which the crystal grew. 'Flash-cooling' is a way of greatly reducing radiation
damage (15-17) but it can also help with the problem of fragile crystals by
preventing the loss of the interstitial water necessary to maintain crystal
integrity. Indeed, nowadays many laboratories routinely freeze their crystals.
Figure 8. A diagram of the steps involved in mounting a crystal. The numbers refer to the
steps described in Protocol 7 in the text.
Method
1. Attach a 1.5 cm length of rubber tubing to the end of a disposable
tuberculin syringe (1 ml). With the diamond tool, score a Lindemann
tube near the closed end and break it neatly. Insert the wide end of the
406
14: X-ray analysis
Lindemann tube into the rubber tubing and if necessary, roll back the
ends of the tubing to improve the seal. In this way, you have created a
narrow-bore pipette.
2. Draw the crystal with a small amount of mother liquor or handling
buffer into the Lindemann tube. The tuberculin syringe is small
enough that the apparatus can be held in one hand with the thumb
available for drawing up on the plunger. Remove the end of the tube
from the crystal droplet and continue drawing the crystal further up
the Lindemann tube. With the crystal at the desired height in the tube,
draw a final small plug of mother liquor into the end of the Lindemann
tube.
3. Seal the open end of the Lindemann tube with wax. Soak a small piece
of tissue in water and drape the wet paper over the tube at the height
of the crystal. This is to protect the crystal from heat conduction up the
tube while melted wax is being applied to the end. Seal the end of the
tube with wax. Applying a small piece of Plasticine to the wax makes
handling easier and allows the capillary to be stuck on a microscope
slide or the table top for subsequent manipulation.
4. Score and break with the diamond tool the Lindemann tube a second
time — this time 'above' the crystal. If breaking or cutting a glass tube
without scoring it, add a drop of wax just to the crystal side of where
the break is to be made, but well clear of the crystal, to prevent the
tube collapsing when being broken.
5. Dry the remaining buffer from around the crystal with the aid of a
shred of filter paper inserted through the open end of the tube. If
necessary, larger volumes can be removed with a finely drawn-out
Pasteur pipette or small bore Lindemann/syringe assembly as in step
1, before the drying stage. A dry mount is preferred for two reasons.
The faces are more easily visible when aligning the crystal and the
absence of solvent may reduce the effects of crystal slippage. It should
be kept in mind, however, that in this dry atmosphere, the crystal is
susceptible to solvent loss, thus the following steps should be
performed as quickly as possible.
6. Seal the other end with wax using the wet tissue draped once again
over the tube to protect the crystal.
Many variations are possible at the discretion of the mounter. For example,
it may be preferable to have two plugs of buffer in the Lindemann tube; one
on either side of the crystal. This can be accomplished by adding a small
amount of buffer to the top of the tube before the final wax seal is applied. If
it is necessary to reposition the crystal, opening up the wax plug is most easily
done with a heated needle.
407
L. Sawyer and M. A. Turner
Figure 9. A typical goniometer head with the crystal sealed in a capillary fixed upon it.
The key shown is for adjusting the slides and arcs. It has a fine Alien key at the other end
for locking the arcs after adjustment. The threaded ring at the base will screw onto an
X-ray camera or ctiffractometer.
408
Figure 10. The cryo-loop method of crystal mounting. (a) A goniometer head with the
magnetic base and mounted loop. (b) A close-up of the tip of a typical mounted loop of
about 0.5 mm diameter. (c) A protein crystal flash-frozen in its cryo-solvent, (d) Typical
equipment for handling frozen crystals. Left to right: a goniometer head, a magnetic
base, a CrystalCap with mounted cryo-loop, an 18 mm cryo-vial, a cryo-vial in a plastic
handling tube, a mounted cryo-loop in a plastic pipette tip for mounting a crystal, a
plastic handling tube made from a Pasteur pipette and ideal for filling the cryo-vial with
liquid nitrogen.
409
L. Sawyer and M. A. Turner
nitrogen temperature). As noted already, recent developments in the cryo-
crystallography of biological molecules have meant that in many laboratories
data collection at 100 K is now routine. A general overview of these develop-
ments is given by Rodgers (19). If it becomes apparent that very low tempera-
tures will be required (because conventionally mounted crystals have
unworkably short lifetimes in the X-ray beam) a different mounting pro-
cedure must be applied. Whilst there is some benefit in equilibrating the
protein crystals in cryo-protectant, it is not strictly necessary, though it is
essential to have a suitable cryo-protectant mother liquor. This can often be
obtained by mixing crystal mother liquor with increasing concentrations of
glycerol until a capillary containing the solution remains transparent when
plunged into liquid nitrogen.
In addition to the equipment mentioned in Section 3.2.1, some special
equipment is needed, both for mounting but also for X-ray work. A popular
and convenient device for maintaining the crystal at 100 K whilst in the X-ray
beam is the Cryostream made by Oxford Cryosystems but most X-ray gener-
ator manufacturers provide an equivalent. Much of the equipment for crystal
mounting can conveniently be obtained from Hampton Research but it can
also be hand-made in the laboratory. Protocol 2 and Figure 10 illustrates how
the crystal is mounted in a cryo-loop and also shows a convenient and cheap
way of handling the mounted loop once frozen in liquid nitrogen.
Method
1. Attach a holder (a 1 ml plastic pipette tip is suitable) to the magnetic
base end of a mounted cryo-loop. Insert the vial part of the
CrystalCap into a small (5 x 5 x 1 cm) expanded polystyrene float
with a hole to fit the vial firmly, the open end being up.
2. Select the crystal to be frozen and place on the same slide a drop of
cryo-protectant buffer solution.
3. Submerge the float and vial in the liquid nitrogen until it is cold
(boiling ceases) and the vial is full of liquid nitrogen.
4. Carefully scoop up the crystal with the cryo-loop in which it will be
held by surface tension, and immediately immerse it in the cryo-
410
14: X-ray analysis
protectant. The time required for cryo-protection varies depending on
buffer system and cryo-protectant used. This will need to be deter-
mined by trial and error and it is sensible to begin with some less
good crystals.
5. Carefully scoop up the crystal from the cryo-protectant and plunge it
into the liquid nitrogen in the vial. Allow the magnetic base to cool
down (boiling ceases) before screwing the cap onto the vial.
6. The crystal is now mounted and frozen—it must now be maintained
at approximately this temperature for as long as it is required.
7. Either transfer the vial and cap to a suitable storage Dewar containing
liquid nitrogen if the X-ray work is not to proceed immediately or take
the crystal in the Dewar to the X-ray laboratory. To transfer the crystal
to the X-ray diffractometer, the method will depend upon the exact
arrangement of the goniometer spindle. If vertical (e.g. R-Axis Image
Plate), a goniometer with an extension to permit the frozen crystal to
be positioned so that it points downwards, is essential. If the spindle
is horizontal (e.g. Mar Image Plate) proceed as follows.
8. Place the goniometer head with magnetic base attached on the
spindle and adjust the z-translation so that the cryo-loop when in
position will be in the centre of the cold gas stream. This adjustment
is conveniently done with the mounted loop when finding a suitable
cryo-protectant. It may be helpful to withdraw the nozzle of the cold
stream slightly to allow some extra room for the next stage. Also
arrange for the arc with the largest angular displacement to be
vertical and at the upper extremity of its travel. The crystal when
mounted will then be pointing downwards by some 20-30°
depending on the goniometer head.
9. Attach the mounted loop to the base, unscrew the cap and withdraw
the vial of nitrogen, holding the vial (re-)filled with liquid nitrogen in a
holder so that the metal part of the cap is free to be located on the
magnetic base. The crystal should now be in the stream of cold dry
nitrogen. Care should be taken not to disturb the X-ray back-stop.
10. Reposition the nozzle, if necessary, to be as close to the crystal as
possible without interfering with the X-ray beam. This is conveniently
done by a second person as soon as the vial is removed. Ensure that
there are no drafts in the laboratory which might deflect the flow from
the cryo-cooler and similarly do not breathe at the crystal whilst
mounting it.
11. Begin the X-ray measurements.
4. X-ray data
The fundamental data about a crystal which must be known before the
structure solution can be attempted are the unit cell dimensions and the space
411
L. Sawyer and M. A. Turner
group. Until relatively recently, these data were always determined first, in
order that the strategy for data collection could be optimized, a necessary
prerequisite for crystals with limited X-ray lifetime. Nowadays however, most
data are recorded automatically by the oscillation/rotation method, often
before the space group and cell dimensions are known, and the main purpose
of examining the first images is to determine that the crystal is single, un-
cracked, and diffracts X-rays. In addition, some clues about the space group
can be obtained from the symmetry of the pattern near the principal zones,
but this is not really necessary. The data images are stored on tape or disc and
the images further processed (fairly) automatically by computer. The unit cell
dimensions are calculated and the space group is determined, once the data
have been processed, by plotting out layers as 'mock' precession photos to
observe the systematic absences more easily and hence determine the space
group. Thus the strategy has now become one of shoot first and ask questions
later.
where dmax is the maximum resolution for which data are required, q is the
spacing of planes perpendicular to the X-ray beam (e.g. a when the a axis is
parallel to the X-ray beam), and A, which is typically 0.1-0.3°, is the reflecting
range of the crystal, or mosaic spread. The strategy adopted in current prac-
tice is to use a relatively large rotation range, or as large a range as possible.
412
14: X-ray analysis
However, the high degree of automation available with area detector and
image plate software and the cheapness of disc storage allow oscillation
ranges less than or comparable to the actual diffraction spot size to be used
which in turn allows integration of the spot as it traverses the reflecting
position. Oscillation ranges larger than a typical spot result in the collection of
intensity not just of the spot itself but of background 'in front of and 'behind'
the spot as well, thus reducing the effective signal-to-noise ratio for that spot.
Larger oscillation ranges, however, are used to minimize the time of overall
data collection with image plates because of the relatively large time require-
ment for scanning each image before the plate can be used for collection of
the next frame.
Equipment
• X-ray generator or synchrotron producing • Imaging plate diffractometer
monochromatic radiation around 0.1 nm • Appropriate graphics workstation and
wavelength software
Method
1. With the detector at the desired distance, centre the crystal in the X-ray
beam as described above. The shorter the crystal-to-detector distance,
the higher is the resolution to which data can be measured but the
greater the likelihood of spot overlap. If nothing is known beforehand,
it saves time to use the setting already in use.
2. Check the diffraction pattern by exposing a frame for an arbitrary
length of time and oscillation range, for example, 120 sec and 0.25°
with an area detector, 10 min and 1° with an image plate. If spots are
visible to the edges of the image, it may be desirable to swing the
detector out to a non-zero 20 angle, or decrease the crystal-to-detector
distance. This will help determine, according to Bragg's equation, the
resolution to which the crystal diffracts.
3. Having decided the length of time to be spent exposing each frame
and the oscillation range desired to achieve spot separation even at
the edges of the detector, begin data collection. Depending on the
system and programs used, it is recommended that data processing
be started as soon as possible. It may become obvious while
attempting to process the data that problems exist with the crystal. If
this is the case, the decision can be made to end the measurement and
try another crystal without wasting detector time.
415
L. Sawyer and M, A. Turner
is that used in the program XDS developed by Kabsch (22). The autoindexing
routine begins by assigning a reciprocal-space vector to each spot. Low
resolution differences between these reciprocal lattice points are accumulated
in clusters and are sorted by decreasing population. The first two which are at
an angular separation > 45° are chosen and indices are assigned to them.
These are used as a basis set from which the remaining difference-vector
clusters can be indexed. Originally, it was expected that space group and cell
dimensions of the crystal were known prior to running the autoindexing
routine however contemporary algorithms allow both orientation and un-
known cell dimensions to be determined. Alternative choices of cell dimen-
sions are given with associated agreement factor allowing statistical
consideration of all the possibilities.
The autoindexing routine employed by DENZO (23) uses a different
algorithm coined 'real space indexing' whereby a complete search of all
possible indices of a reflection is carried out using a Fast Fourier transform.
Once the three best linearly independent vectors with minimal unit cell
volume are found, the cell is 'reduced' to describe a standard basis for the
description of the unit cell. In DENZO, a basis set for each of the 14 Bravais
lattices is found and a distortion index is calculated for the peaks in the peak
search list. The user must then, on the basis of the magnitude of the deviation
from ideal Bravais lattice symmetry, decide upon most likely cell dimensions
and the space group.
Method
1. Observe the symmetry of the diffraction pattern of the zero level zones
('mock' precession photographs) which must be consistent with the
unit cell parameters and lattice type already determined. This gives the
diffraction (Laue) symmetry given in Table 2 for the 11 relevant classes
(and helps to ensure that principal zones have indeed been identified).
416
14: X-ray analysis
Diffraction symmetry always has an inversion centre. For example, a
triclinic cell, P1, has -1 diffraction symmetry. It is the appearance of
extra symmetry which allows labelling of crystal class. At this point
the axes can be assigned as a, b, or c such that a, p, and -y are close or
equal to 90° (unless a trigonal or hexagonal cell is suspected) and the
cell is primitive (see Table 2). A zero layer photograph by definition,
arises from either the hk0, h0I, or 0kl sets of planes. These can be
assigned arbitrarily in the case of certain space groups. The Inter-
national tables for X-ray crystallography (24) will help with the task of
assigning axes according to crystallographic convention. In general,
the unique axis is b for monoclinic cells and c for cells of higher
symmetry. The upper layer images, which contain no reciprocal axes,
must also be assigned as hkn, hnl, or nkl where n > 1.
2. Index the spots, h, k, I on each image. Be aware that systematically
absent reflections also require indexing.
3. Analyse the systematically absent reflections in the diffraction pattern.
This pin-points the space group often, but not always, uniquely. Use
the axial absences to identify any screw axes.
4. Check that assignments of systematic absences are consistent with
upper level images as well. The upper layers also allow, for example,
distinction between a sixfold and a threefold axis. (These look the
same on a zero level photograph.)
5. Identify, with use of the International tables, a list of space groups
compatible with the observed diffraction patterns. In some cases there
is no ambiguity: e.g. P212121, whilst in others no distinction is possible
until the structure solution is under way, e.g. 1222 and 1212121 have
identical systematic absences as have the enantiomorphs P3-|21 and
P3221 where only the hand of the screw axis differs.
6. Try to find as high a symmetry space group which is consistent with
your observations and work to lower symmetries as need be.
7. Determine the approximate number of molecules in the unit cell from
the unit cell dimensions, the molecular weight of the molecule, and
Vm. Knowing the crystal system helps in this, e.g. if the crystals are
orthorhombic, there must be a multiple of four molecules in the unit
cell. (Note that a 'molecule' may also be some identically repeated
portion of the protein or polynucleotide.)
5. Concluding remarks
The object of this brief excursion into X-ray crystallography has been to intro-
duce the ideas and methods required to collect the information necessary for
the first publication on a new crystalline material. Such papers should include
not only the purification and crystallizing conditions, which should be repro-
ducible, but also the techniques employed to obtain the X-ray diffraction data
and the crystal lifetime in the X-ray beam and on the shelf. The unit cell
dimensions and space group together with the resolution obtainable from a
crystal have been the main concern of this chapter. Vm, the number of
molecules in the asymmetric unit, the solvent content, and any comments about
the subunit structure are also generally mentioned. Increasingly, molecular
replacement techniques will reveal similarities to known structures and so the
'crystallization note' is often superseded by the preliminary structure,
obtained rapidly from the complete data set which is collected from the first
crystals. Finally, protein crystallographers always enjoy talking about their
418
14: X-ray analysis
subject and the number of groups around the world has risen considerably
since the first edition of this book. You will have discovered that the
technique requires a modicum of dedication and therefore do seek guidance
in getting your project under way.
References
1. Glusker, J. P., Lewis, M., and Rossi, M. (1994). Crystal structure analysis for
chemists and biologists. VCH, New York.
2. Stout, G. H. and Jensen, L. H. (1989). X-ray structure determination, 2nd edn.
John Wiley, New York.
3. McRee, D. E. (1993). Practical protein crystallography. Academic Press Inc., New
York.
4. Drenth, J. (1994). Principles of protein X-ray crystallography. Springer-Verlag,
New York.
5. Blundell, T. L. and Johnson, L. N. (1976). Protein crystallography. Academic
Press, London.
6. Bragg, W. L. (1968). Sci. Am., 219, 58.
7. Glusker, J. P. (1994). Methods Biochem. Anal, 37, 1.
8. Johnson, L. N. and Hajdu, J. (1990). Eur. J. Biochem., 29, 1669.
9. Rossmann, M. G. (1990). Acta Cryst., A46, 73.
10. Hope, H. (1990). Annu. Rev. Biophys. Biophys.Chem.,26, 107.
11. Acharya, R., Fry, E., Stuart, D., Fox, G., Rowlands, D., and Brown, F. (1989).
Nature, 337, 709.
12. Quiocho, F. A. and Richards, F. M. (1964). Proc. Natl. Acad. Sci. USA, 52, 833.
13. Matthews, B. W. (1968). J. Mol. Biol, 33, 491.
14. Hodgkin, D. C. and Riley, D. P. (1968). In Structural molecular biology (ed. A.
Rich and N. Davidson), pp. 15-28. Freeman, San Francisco.
15. Henderson, R. (1990). Proc. Roy. Soc. Land., B241,6.
16. Teng, T. Y. (1990). J. Appl. Cryst., 23, 387.
17. Garman, E. A. and Mitchell, E. P. (1996). J. Appl. Cryst., 29, 584.
18. Hajdu, J., McLaughlin, P. J., Helliwell, J. R., Sheldon, J., and Thompson, A. W.
(1985). J. Appl. Cryst., 18, 528.
19. Rodgers, D. W. (1994). Structure, 2, 1135.
20. Engel, C., Wierenga, R., and Tucker, P. A. (1996). J. Appl. Cryst., 29, 208.
21. Arndt, U. W. and Wonacott, A. (ed.) (1978). The rotation method in crystallo-
graphy. North-Holland Publishers, Amsterdam.
22. Kabsch, W. (1988). J. Appl. Cryst., 21, 67.
23. Otwinowski, Z. and Minor, W. (1997). Methods in enzymology (eds C. W. Carter
and R. M. Sweet), Academic Press, London. Vol. 276, pp. 307.
24. Hahn, T. (ed.) (1987). International tables for X-ray crystallography. D. Reidel
Publishing Co., Dordrecht, Netherlands.
419
This page intentionally left blank
Al
List of suppliers
Aldrich-Chemical Co., Inc., 1001 W. St Paul Avenue, PO Box 355, Milwaukee,
WI 53201, USA. (chemicals)
Alpha Laboratories Ltd., Eastleigh, Hampshire, UK. (multiple liquid dispenser)
American Can Company, Greenwich, CT 06830, USA. (Parafilm® 'M',
laboratory film)
Amersham
Amersham International pic., Lincoln Place, Green End, Aylesbury,
Buckinghamshire HP20 2TP, UK.
Amersham Corporation, 2636 South Clearbrook Drive, Arlington Heights, IL
60005, USA.
Amicon Division, W. R. Grace and Co., 72 Cherry Hill Drive, Beverly, MA
01915, USA. (filters, membranes)
Anderman
Anderman and Co. Ltd., 145 London Road, Kingston-Upon-Thames, Surrey
KT17 7NH, UK.
Applied Biosystems, Inc., 850 Lincoln Center Dr., Foster City, CA 94404,
USA and Birch wood Science Park North, Warrington, Cheshire WA3 7PB,
England, (biochemical instrumentation, chemicals)
Appligene, route du Rhin, BP 72, 67402 Illkirch Cedex, France, (biochemicals)
Bachem, Hauptstrasse 144, CH-4416 Bubendorf, Switzerland, (detergents)
BDH Limited, Broom Road, Poole, BH12 4NN, UK. (electrophoresis
products)
Beckman, 4550 Noris Canyon Road, PO Box 5101, San Ramon, CA 94583,
USA. (centrifugation, pipetting station)
Beckman Instruments
Beckman Instruments UK Ltd., Progress Road, Sands Industrial Estate,
High Wycombe, Buckinghamshire, HP12 4JL, UK.
Beckman Instruments Inc., PO Box 3100, 2500 Harbor Boulevard, Fullerton,
CA 92634, USA.
Becton-Dickinson and Co., Clay Adams Div., 299 Webro Road, Parsippany,
NJ 07054, USA. (Falcon plasticware)
Becton Dickinson
Becton Dickinson and Co., Between Towns Road, Cowley, Oxford OX4 3LY,
UK.
List of suppliers
Becton Dickinson and Co., 2 Bridgewater Lane, Lincoln Park, NJ 07035, USA.
Bender and Hobein GmbH, D-8000 Munchen 2, Lindwurmstrasse 71,
Germany, (free flow electrophoresis)
Bijhoelt and Heuvelen SV, The Netherlands, (transparent and adhesive
plastic foils)
BioBlock Scientific, BP 111, F-67403 Illkirch Cedex, France, (scientific
equipments)
BioRad, 1414 Harbour Way South, Richmond CA 94804, USA. (HPLC, IEF
equipments, biochemicals)
Boehringer Mannheim, PO Box 310120, D-6800 Mannheim 31, Germany.
(biochemicals)
Bio
Bio 101 Inc., c/o Statech Scientific Ltd, 61-63 Dudley Street, Luton, Bedford-
shire LU2 OHP, UK.
Bio 101 Inc., PO Box 2284, La Jolla, CA 92038-2284, USA.
Bio-Rad Laboratories
Bio-Rad Laboratories Ltd., Bio-Rad House, Maylands Avenue, Hemel
Hempstead HP2 7TD, UK.
Bio-Rad Laboratories, Division Headquarters, 3300 Regatta Boulevard,
Richmond, CA 94804, USA.
BioWhittaker, Inc., 8830 Biggs Ford Road, Walkersville, MD 21793, USA.
Boehringer Mannheim
Boehringer Mannheim UK (Diagnostics and Biochemicals) Ltd, Bell Lane,
Lewes, East Sussex BN17 1LG, UK.
Boehringer Mannheim Corporation, Biochemical Products, 9115 Hague Road,
P.O. Box 504 Indianapolis, IN 46250-0414, USA.
Boehringer Mannheim Biochemica, GmbH, Sandhofer Str. 116, Postfach
310120 D-6800 Ma 31, Germany.
British Drug Houses (BDH) Ltd, Poole, Dorset, UK.
Brookhaven Instrument Corp., 750 Blue Point Road, Holtsville, NY 11743,
USA. (light scattering instrumentation)
Bunton Instrument Co., Inc., 615 South Stonestreet Avenue, Rockville, MD
20850, USA. (microgrippers)
Calbiochem Behring Diagnostics, 10933 N. Torrey Pines Road, La Jolla, CA
923037, USA. (biochemicals, detergents)
Cambridge Repetition Engineers Ltd., Green's Road, Cambridge, CB4 3EQ,
UK. (dialysis buttons for crystallization)
CEA verken AB, S-152 01 Strangnas, Sweden. (X-ray films)
Charles Supper Company Inc., 15 Tech Circle, Natick, MA 07160, USA.
(crystallographic equipment)
CJB Developments Limited, Airport Service Road, Portsmouth, Hampshire
PO35PG, UK. (large-scale preparative electrophoretic apparatus)
Cole and Palmer Instrument Co., 7425 N. Oak Park Avenue, Chicago, IL
60648, USA. (scientific equipments)
422
List of suppliers
Coming, Inc., Science Products, MP-21-5-8, Corning, NY 14831, USA.
(glassware, pipettes)
Costar Nucleopore®, One Alewife Center, Cambridge, MA 02140, USA and
Costar Europe, Ltd., PO Box 94, 1170 AB Badhoevedrop Sloterweg 305a,
1171 VC Vadhoevedrop, The Netherlands, (titration and crystallization
plates, pipettors)
Cruachem Ltd., West of Scotland Science Park, Acre Road, Glasgow G20
0UA.
Difco Laboratories
Difco Laboratories Ltd., P.O. Box 14B, Central Avenue, West Molesey, Surrey
KT8 2SE, UK.
Difco Laboratories, P.O. Box 331058, Detroit, MI 48232-7058, USA.
Douglas Instruments Ltd., 255 Thames House, 140 Battersea Park Road,
London SW11 4NB, UK. (automatic batch crystallization system)
Dow Corning Corp., Dow Corning Center, Box 0994, Midland, MI 48686-
0994, USA. (silicone oil, grease)
Dupont de Nemours and Co., Concord Plaza, Wilmington, DE 19898, USA.
(chemicals, instrumentation)
Du Pont
Dupont (UK) Ltd., Industrial Products Division, Wedgwood Way, Stevenage,
Herts, SGI 4Q, UK.
Du Pont Ltd., NEN Life Science Products, PO Box 66, Hounslow TW5 9RT,
UK.
Du Pont Co. (Biotechnology Systems Division), P.O. Box 80024, Wilmington,
DE 19880-002, USA.
Dynatech Laboratories, Inc., 14340 Sullyfield Circle, Chantilly, VA 22021,
USA. (titration plates for crystallization robots)
Eastman-Kodak Co., 343 State St., Rochester, NY 14650, USA and Kodak
House, Station Road, Hemel Hempstead, Herts HP1 1JU, UK. (chemicals,
films)
Enraf Nonius Delft, PO Box 483, 2600 AL Delft, The Netherlands.
(diffractometry)
Euromedex, Produits de Recherche, 29 rue Herder, F-67000 Strasbourg,
France, (chemicals, protease inhibitors)
European Collection of Animal Cell Culture, Division of Biologies, PHLS
Centre for Applied Microbiology and Research, Porton Down, Salisbury,
Wilts SP4 OJG, UK.
Everett's Co., Parkgate, Nr, Southampton, UK. (vacuum wax, seals, and
lubrifiants)
Falcon (Falcon is a registered trademark of Becton Dickinson and Co.).
Fisher Scientific Company, 711 Forbes Avenue, Pittsburgh, PA 15219^1785,
USA. (biochemicals, scientific equipments)
Flow Laboratories, Woodcock Hill, Harefield Road, Rickmansworth, Herts.
WD3 1PQ, UK.
423
List of suppliers
Flow Laboratories International SA, via Lambro 23/25, I-20090 Opera
(MI), Italy, (biochemical equipments, Linbro plate, CrystalPlate and
coverslips)
Fluka
Fluka-Chemie AG, CH-9470, Buchs, Switzerland.
Fluka Chemicals Ltd., The Old Brickyard, New Road, Gillingham, Dorset
SP8 4JL, UK.
Fluka Chemie AG, Industriestrasse 25, CH-9470 Buchs, Switzerland, (bio-
chemicals, detergents)
Genset SA, 1, rue Robert et Sonia Delaunay, 75011 Paris, France.
Genzyme Corporation, 75 Kneeland Street, Boston, MA 02111, USA.
(protease-free deglycosylation enzymes)
Gibco BRL, Bethesda Research Laboratories, Life Technologies, Inc. PO
Box 6009, Gaithersburg, MD 20877, USA. (biochemicals, growth media)
Gibco BRL (Life Technologies Inc.), 3175 Staler Road, Grand Island, NY
14072-0068, USA.
Gibco BRL (Life Technologies Ltd.), Trident House, Renfrew Road, Paisley,
Scotland, PAS 4EF, UK.
Gilson Medical Electronics, Inc., 72 rue Gambetta, BP 45, F-95400 Villers-le-
Bel, France and 3000 W. Beltine Hwy., PO Box 27, Middleton, WI 53562,
USA. (sample changers)
Gow-Mac Inc., PO Box 32, Bound Brook, NJ 08805-0032, USA. (thermal
conductivity detectors)
Hamilton Co., PO Box 10030, Reno, NV 89520-0012, USA. (syringes)
Hampton Research, 27632 El Lazo Road, Suite 100, Laguna Beach, CA
92677-3913, USA.
Heraeus Feinchemikalien und Forchungsbedarf GmbH, Alter Weinberg,
D-7500 Karlsruhe 41-Ho., Germany, (chemicals, reagents for silaniz-
ation)
Hewlett-Packard Co., Analytical Group, Mailstop 20B AE, Palo Alto, CA
94403, USA. (robotics)
Hilgenberg Glass Company, D-3509 Malsfeld, Germany. (X-ray glass/quartz
capillaries)
Arnold R. Horwell, 73 Maygrove Road, West Hampstead, London NW6 2BP,
UK.
Huber Diffraktionstechnik GmbH, D-8219 Rimsting, Germany, (diffractometry)
Hybaid
Hybaid Ltd., 111-113 Waldegrave Road, Teddington, Middlesex TW11 8LL,
UK.
Hybaid, National Labnet Corporation, P.O. Box 841, Woodbridge, NJ. 07095,
USA.
HyClone Laboratories 1725 South HyClone Road, Logan, UT 84321, USA.
IBF Biotechnics, 35 avenue Jean-Jaures, 92290 Villeneuve-la-Garenne,
France. (chromatographic matrices, biochemicals)
424
List of suppliers
ICI Cambridge Research Chemicals, Gadbrook Park, Northwich, Cheshire
CW9 7RA, UK. (chemicals)
ICN Biomedicals, Inc., Micromedica Systems Diagnostic Division, 102 Witmer
Road, Horsham, PA 19044-2281, USA. (robotic protein crystallization
system II, pipetting stations)
ICN Flow, 330 Hyland Avenue, Costa Mesa, CA 92626, USA. (biochemical
equipments, Linbro plate, CrystalPlate® and coverslips)
Imaging Technology Inc., 600 West Cummings Park, Woburn, MA 01801,
USA. (digitizers)
Intermec Corp., 4405 Russell Road, PO Box 360602, Lynnwood, WA 98046-
9702, USA. (barcode printer)
International Biotechnologies Inc., 25 Science Park, New Haven, Connecticut
06535, USA.
Invitrogen Corporation
Invitrogen Corporation 3985 B Sorrenton Valley Building, San Diego, CA.
92121, USA.
Invitrogen Corporation do British Biotechnology Products Ltd., 4-10 The
Quadrant, Barton Lane, Abingdon, Oxon OX14 SYS, UK.
Jouan SA, rue Bobby Sands, F-44800 Saint Herblain, France, (laboratory
equipments)
Keithley Data Acquisition and Control, 28775 Aurora Road, Cleveland, OH
44139, USA. (instrument interfaces)
Kodak: Eastman Fine Chemicals 343 State Street, Rochester, NY, USA.
Kohyo Trading Company, Kyodo Bldg 4-1,2 Chome, Iwando-cho, Chiyoda-ky,
Tokyo, Japan, (detergents)
Leica SARL, see Wild-Leitz.
Leitz/Leica. Ill Deer Lake Road, Deerfield, IL 60015, USA.
Life Technologies Inc., 8451 Helgerman Court, Gaithersburg, MN 20877,
USA.
Marresearch, Grosse Theaterstrasse 42, Postfach 303670, 2000 Hamburg 36,
Germany, (image plate)
Memmert GmbH and Co., Aeussere Ritterbacherstrasse 38, D-8540
Schwabach, Germany, (laboratory equipments, thermostated cabinets)
Merck
Merck Industries Inc., 5 Skyline Drive, Nawthorne, NY 10532, USA.
Merck, Frankfurter Strasse, 250, Postfach 4119, D-64293, Germany.
Merck, Frankfurter Strasse 250, D-6100 Darmstadt, Germany, (chemicals and
biochemicals)
Microflex Technology, Inc., The Millennium Centre, PO Box 31, Triadelphia,
WV 26059, USA. (microgrippers)
Micromedica System, Inc., (see ICN Biomedicals). (pipetting station)
Millipore
Millipore (UK) Ltd., The Boulevard, Blackmoor Lane, Watford, Herts WD1
8YW, UK.
425
List of suppliers
Millipore Corp./Biosearch, P.O. Box 255, 80 Ashby Road, Bedford, MA
01730, USA.
Millipore Waters, PO Box 255, Bedford MA 01730, USA and Zone
Industrielle, F-67120 Molsheim, France. (filtration, membranes, HPLC
equipments)
NAPS Gottingen GmbH, Nucleic Acids Products Supply, Rudolf-Wissel Str.
28, 37070 Gottingen, Germany.
National Institute of Standards and Technology, (Standard Reference Data)
Bldg. 221/A323, Gaithersburg, MD 20899, USA. (Software with crystalliz-
ation data bank)
National Instruments Corp., 12109 Technology Blvd, Austin, TX 78727-6204,
USA. (instrument interfaces, laboratory software)
Neosystem Laboratories, Technopole du Rhin, 21 rue du la Rochelle, F-67100
Strasbourg, France, (peptides)
New England Biolabs (NBL)
New England Biolabs (NBL), 32 Tozer Road, Beverley, MA 01915-5510, USA.
New England Biolabs (NBL), c/o CP Labs Ltd., P.O. Box 22, Bishops Stortford,
Herts CM23 3DH, UK.
Nikon Corporation Instrument Div., Fuji Bldg 2-3, 3-Chome, Maranouchi,
Chiyoda ku, Tokyo 100, Japan, (stereo microscopes)
Nikon Europe BV, Shipholm weg 321, 1171 AE Badhoevedorp, The
Netherlands. (stereo microscopes)
Nunc Inc., 2000, North Aurora Road, Naperville, IL 60566, USA. (plastic
tubes and plates)
Ominifit Ltd., 51 Norfolk Street, Cambridge CB1 2LE, UK and 2005 Park
Street, Box 56, Atlantic Beach, NY 11509, USA. (valves)
Omnilabo Holland BV, Breda, The Netherlands, (multi-well plates)
Oxyl, Peter Henlein Strasse 11, D-8903 Bobingen, Germany, (detergents)
Panasonic Inc., One Panasonic Way, Secaucus, NJ 07094, USA. (optical disc
recorder)
Pentapharm Ltd., Engelgasse 109, CH-4002 Basel, Switzerland, (protease
inhibitors)
Peptide Institute, 476 Ina Miush-shi, Osaka 562, Japan, (protease inhibitors)
Perkin-Elmer
Perkin-Elmer Ltd., Maxwell Road, Beaconsfield, Bucks. HP9 1QA, UK.
Perkin-Elmer Ltd., Post Office Lane, Beaconsfield, Bucks, HP9 1QA, UK.
Perkin-Elmer-Cetus (The Perkin-Elmer Corporation), 761 Main Avenue,
Norwalk, CT 0689, USA.
Perpetual Systems Corporation, 2283 Lewis Avenue, Rockville, Maryland
20851, USA. (sitting-drop rods for cystallization)
PerSeptive Biosystems, City of Dover, Kent County, DL 19901, USA.
Pfanstiel Laboratory, Inc., 1219 Glen Rock Avenue, Wavkega, IL 60085 0439,
USA. (detergents)
426
List of suppliers
Pharmacia Biosystems
Pharmacia Biotech Europe Procordia EuroCentre, Rue de la Fuse-e 62,
B-1130 Brussels, Belgium.
Pharmacia Biosystems Ltd. (Biotechnology Division), Davy Avenue,
Knowlhill, Milton Keynes MK5 8PH, UK.
Pharmacia LKB Biotechnology AB, Bjorngatan 30, S-75182 Uppsala, Sweden.
Phenomenex, 6100 Palos Verdes Drive S., Rancho Palos Verdes, CA 90274,
USA. (HPLC columns for tRNA)
Pierce, PO Box 1512, 3260 BA Oud-Beijerland, The Netherlands, (laboratory
supplies)
Polycrystal book service, PO Box 3439, Dayton, Ohio 45401, USA.
(crystallography books)
PolyLabo Paul Block et Cie, BP 36, F-67023 Strasbourg Cedex, France.
(scientific equipments)
The Product Integrity Company, Enfield, CT 06082, USA. (programs for
factorial analysis)
Prolabo, 12 rue Pelee, F-7511 Paris, France, (chemicals, equipments)
Promega
Promega Ltd., Delta House, Enterprise Road, Chilworth Research Centre,
Southampton, UK.
Promega Corporation, 2800 Woods Hollow Road, Madison, WI 53711-5399,
USA.
Protein Solutions Incorporated, 2300 Commenwealth Drive, Suite 102,
Charlottesville, VA 22901, USA.
Pye Unicam Ltd, York Street, Cambridge CB1 2PX, UK. (Philips X-ray
generator)
Qiagen
Qiagen Inc., do Hybaid, 111-113 Waldegrave Road, Teddington, Middlesex,
TW11 8LL, UK.
Qiagen Inc., 9259 Eton Avenue, Chatsworth, CA 91311, USA.
Radiometer, A/S 49 Krogshojvej, DK 2880 Dagsvaerd, Denmark. (pH-meter,
conductimeter)
Rainin Instrument Co. Inc., Mack Road, Woburn, MA 01801, USA. (filters)
Resolution Technology, 26000 Avenida Aeropuerto 22, San Juan Capistrano,
CA 92675, USA. (time-lapse VCR)
Rigaku, Monschauer Strasse 7, D-4000 Diisseldorf-Heerdt, Germany & 3
Electronics Avenue, Danvers, MA 01923, USA. (X-ray generators, image
plate)
Roucaire, BP 65, F-78143 Velizy-Villacoublay Cedex, France, (scientific
equipments)
Schleicher and Schuell
Schleicher and Schuell Inc., Keene, NH 03431 A, USA.
Schleicher and Schuell Inc., D-3354 Dassel, Germany.
Schleicher and Schuell Inc., c/o Andermann and Company Ltd.
427
List of suppliers
Seikagaku Kogyo Co. Ltd., 1-5, Nihonbashi-Honcho 2-Chome Chuo-ku,
Tokyo, 103, Japan, (biochemicals, glycosylases)
Serva Feinbiochemica GmbH and Co., PO Box 105260, D-6900 Heidelberg,
Germany, (biochemicals)
Setaram, 7 rue de 1'Oratoire, BP. 34, F-69641 Caluire Cedex, France.
(instrumentation, calorimeters)
Shandon Scientific Ltd., Chadwick Road, Astmoor, Runcorn, Cheshire WA7
1PR, UK.
Siemens AG, Mess., Pruf. und Prozesstechnik, Ostl. Rheinbriickenstrasse 50,
D-7500 Karlsruje 21, Germany, (diffractometry)
Sigma Chemical Company
Sigma Chemical Company (UK), Fancy Road, Poole, Dorset BH17 7NH, UK.
Sigma Chemical Company, 3050 Spruce Street, P.O. Box 14508, St. Louis,
MO 63178-9916, USA.
Societe 3412, 65 avenue de Stalingrad, F-95104 Argenteuil, France.
(crystallization boxes)
Sofranel, 59 rue Parmentier, 78500 Sartouville, France. (X-ray glass/quartz
capillaries)
Sorvall DuPont Company, Biotechnology Division, P.O. Box 80022,
Wilmington, DE 19880-0022, USA.
Speciality Chemicals, PO Box 1466, Gainesville, FL 32602, USA. (Prosil®-28
reagent for silanization)
Spectrum Medical Industries, Inc., 8430 Santa Monica Blvd, Los Angeles, CA
90069, USA. (dialysis membranes-Spectrapore®)
Stratagene
Stratagene Ltd., Unit 140, Cambridge Innovation Centre, Milton Road,
Cambridge CB4 4FG, UK.
Strategene Inc., 11011 North Torrey Pines Road, La Jolla, CA 92037, USA.
Tosohaas, 6th and Market Streets, Philadelphia, PA 19105, USA. (HPLC
columns)
Transformation Research Inc., PO Box 241, Framington, MA 01701, USA.
(protease inhibitors)
United States Biochemical, P.O. Box 22400, Cleveland, OH 44122, USA.
Vegatec S.A.R.L., 7 place des Onze Arpents, F-94800 Villejuif, France.
(detergents)
Velmex, Inc., PO Box 38, E. Bloomfield, NY 14443, USA. (stepper motors,
motorized slides)
Wellcome Reagents, Langley Court, Beckenham, Kent BR3 3BS, UK.
Whatman Laboratory Sales Ltd, Unit 1, Colred Road, Parkwood, Maidstone,
Kent, ME15 9XN, UK. (chromatography supports)
Wild-Leitz (Leica SARL), 86 avenue du 18 juin 1940, F-92563 Rueil-
Malmaison Cedex, France and CH-9435 Heerbrugg, Switzerland, (stereo
microscopes)
Wolfgang Miiller, Reierallee 12, D-1000 Berlin 27, Germany. (X-ray glass/
quartz capillaries)
Index
ACA CrystalPlates® 134-6 stoichiometry 226
activation free energy 316-21 streak seeding 204
acupuncture method 166-70 column method of solubility measurement 273
additives 335-8, 369 complexes, see co-crystallizations
for co-crystallizations 227 computer software
divalent cations 222, 223 experimental design 86-8
for membrane protein crystallization 261 net charge estimation 284
monovalent ions 222 statistics 105
for nucleic acid crystallization 221-3 X-ray data processing 414-16
polyamines 221 concentration 124-5
spermine 221, 223 estimation 25-6
agarosegel 150 measurement 125
preparation 159, 161 consolution boundary 246
see also gel crystallization Costar plates 182, 184
ageing 23 countercurrent fractionation 215
ammonium sulfate 9 counter-diffusion techniques 164-6
amorphous precipitate 279 coverslip preparation 131
animal whisker probes 185 critical micellar concentration 248, 249-50
cross-linking 38, 377
cross-seeding 177, 180, 197-200
bacterial expression systems 46; see also cryo-crystallography 379, 381, 385-8
Escherichia coli expression system cryo-protectants 277-8, 369, 379, 381
baculoviral expression system 46-7, 51 mounting crystals for 410-11
batch crystallization 138-41, 142-3 Cryschem plates 136
in gels 161-4 crystallization
biochemical analysis databases 288-9
of crystals 34-6 historical aspects 4-7
of samples 24-6 kinetics 281-3, 334-5, 337-8
Bragg'slaw 395 methods 121, 126-41, 145; see also specific
buffers 122, 220-1, 294 methods
parameters affecting 7-9
practising 143-5
calorimetry for solubility measurement 274 strategy choice 12-13, 306-8
capillary crystals 137 see also crystals
soaking 374-6 crystallography grade purity 8-9, 27
carbon films 349-54 CrystalPlates® 134-6
centrifugation experiments 172-3 crystals
chromatography analysis 34-6
detergent exchange 258 classes 394, 395
of detergents 249 face types 329-30
hydrophobic interaction 216 growth, see growth of crystals
oflipids 256 properties 2-4
co-crystallizations 202-4, 224-34, 367 368 structure 392-5
additives 227 symmetries 393,395
agents 227 systems 393-5
analysis techniques 204-6
DNA:drug 219, 224
DNA:protein 227-31 databases 288-9
homogeneity problems 202-3, 226 density measurement 36-9
protocols 228-30, 232-3 by cross-linking 38
purification 226-7 Ficoll™ method 37-8
RNA:protein 231-4 using molar absorption coefficient 38-9
stability 226 using organic solvents 37
Index
detergent 245, 246-51, 254-5 orthogonal arrays 84-5
choice 260 sampling 82-8
concentration 260-1 see also mathematical models
critical micellar concentration 248, 249-50 expression systems 18-19, 46-55
exchange 257-8 bacterial 46; see also Escherichia coll
purification 248-9, 251, 255 baculoviral 46-7, 51
thin-layer chromatography 249 Escherichia coll 46, 48-50, 51
dialysis techniques methionine auxotroph strains 65
crystallization 126-30, 141-2, 145 strain DL41: 65
detergent exchange 257-8 mammalian cells 47, 51
double 129-30 yeast 47, 51
Hofmeister series testing 300-1
membrane proteins 263-4
microcap 127-8 Fab-peptide complex cross-seeding 197-9
nucleation zone location 306, 307-8 face types 329-30
salt removal 123-4, 286, 288 factorial experimental design 77-82, 103
for screening 306, 307 fractional 84
tubing preparation 124 incomplete 82-4, 103
diffractometry 418 Ficoll™ method 37-8
divalent cation additives 222, 223 floating drops 140-1
DMA fusion tags 50-1, 52
co-crystallization
with drugs 219, 224
with proteins 227-31 gel crystallization 149-70
crystallization protocols 218-19, 228-30 acupuncture method 166-70
purification 22 agarosegel 150
synthesis preparation 159, 161
chemical 211-12 batch method 161-4
fragment design 210 cavity formation 153
see also nucleic acid crystallization and counter-diffusion 164-6
double dialysis 129-30 crystal characteristics 170
crystal preparation 170
diffusion properties 152
gel incorporation into crystals 153
electron microscopy gel preparation 158-61
grid preparation 349-54 and impurities 157
negative staining 355 inside gel 154-7, 161-4
specimen preparation 355 of membrane proteins 263
transfer of films onto grids 354 nucleation 154-7
of two-dimensional crystals 355-60 outside gel 166-70
electrophoresis 24 silica gel 150-2
nucleic acids 24-5 preparation 158-9
energy, activation free 316-21 gel electrophoresis 24
epitaxial nucleation 180, 200-2 nucleic acids 24-5
Escherichia coll expression system 46, 48-50, glass coverslip preparation 131
51 gold compounds 381
methionine auxotroph strains 65 COSSET 86-8
strain DL41: 65 gravity manipulation 170-3
evaporation kinetics 137-8 growth of crystals
Ewald sphere 398 cessation 10, 338
experimental design control ll,330-4
computer-generated 86-8 and face type, 329-30
experimental matrix preparation 90-1 kinetics 281-3, 334-5, 337-8
factorial 77-82, 103 rate 334-5
fractional 84 impurities 337-8
incomplete 82-4, 103 spiral 331-4
Hardin-Sloane 85-6, 87, 88 and temperature 334-5
minimum-prediction variance 85 by two-dimensional nucleation 330
430
andex
handling samples 26-7 lipid layer crystallization 341-63
hanging drops 130 advantages 344
withACACrystalPlates® 134-6 electron microscopy 349-60
with Cryschem plates 136-7 grid preparation 349-54
in Linbro boxes 132-4 negative staining 355
recrystallization 134 specimen preparation 355
soaking 373 transfer of films onto grids 354
Hardin-Sloane designs 85-6, 87, 88 helical crystallization 362
heavy-atom derivatives 224, 366, 377-9 lipid solution preparation, 344-5
soaking compounds 379, 380-4 optical diffraction 360
suppliers 385 protein solution preparation 345-6
heterogeneity, see homogeneity of samples reproducibility 362
HIVintegrase 56 setting up 344-8
Hofmeister series 94, 298-301 Teflon supports 346-7
testing by dialysis 300-1 lipids
homogeneity of samples 28-31 solution preparation 344-5
improving 33—4 thin-layer chromatography 256
probing 31-2 lysozyme crystallization
hosts, see expression systems by gel acupuncture 166-7
hydrophobic interaction chromatography in Linbro boxes 144
216 polymorphism 308-10
hydrophobic ligands, soaking 367-8
hydrostatic pressure 140
hypergravity 172-3 macromolecular crystals, see crystals
macromolecular samples, see samples
macroseeding 178, 180, 191, 193-6
impurities 27-8, 335-8 of needles 196
and gel crystallization 157 MAD 64, 366
see also purification/purity mammalian cell expression systems 47, 51
inclusion bodies 48-50 mathematical models 75-120
INFAC 86, 87 analysis 100-12
insect cell expression systems 46-7, 51 contrast analysis 101—2
interface diffusion 141 crystal property scoring 96-100
interferometry 274 multiple regression analysis 102-12
internet/web sites for optimization 112-16
crystallization databases 288-9 polymorph resolution 116-18
E. coll Genetic Stock Center 65 for screening 82-5, 89-96
experimental design 86-7 stationary point identification 108-12
heavy-atom compound suppliers 385 Matthew's coefficient 403
protein net charge estimates 286 membrane protein crystallization 245-68
ionic strength 295-8, 304 additives 261
iridium compounds 382 agents 261-3
isomorphous replacement 366, 377-9 concentration 258-9
detergent 245, 246-51, 254-5
choice 260
kinetics concentration 260-1
of crystallization and growth 281-3, 334-5, critical micellar concentration 248,
337-8 249-50
of evaporation 137-8 exchange 257-8
krypton 379 purification 248-9, 251, 255
thin-layer chromatography 249
gel techniques 263
lanthanide compounds 384 homogeneity 257
Laue technique 418 lipid analysis and elimination 256
lead compounds 382 microdialysis 263-4
light microscopy 121, 404-5 optimization 263
light scattering techniques 274, 322-5, 326-7 PEG use 261-2
Linbro boxes 132-4, 137 protocols 252-3, 259-64
431
Index
membrane protein crystallization (continued) DNAiprotein 227-31
purification 255-7 homogeneity problems 226
'salting-out' 262-3 protocols 228-30, 232-3
solubility problems 245 purification 226-7
vapour diffusion 263-4 RNA:protein 231-4
without detergent 265 stability 226
mercury compounds 380 stoichiometry 226
methionine auxotroph strains 65-7 concentration 216-18
cell growth 65-7 engineering 224
fermentation medium 66 experimental design 223
starter medium 66 gel electrophoresis 24-5
methionine pathway inhibition 67-8 pH 220
micellar solutions 248 preparation 210-18
critical micellar concentration 248, 249-50 protocols 218-24
Michelson interferometry 274 purification 22, 215-16
microcap dialysis 127-8 co-crystallizations 226-7
microdialysis 126-S screening kits 223
of membrane proteins 263-4 storage of samples 23
microgravity 170-2 synthetic fragment synthesis 210-14
microheterogeneity of samples 28-31 and temperature 220
improving 33-4 see also RNA
probing 31-2
microscopy 121, 404-5; see also electron
microscopy optimization 308
microseeding 178, 180, 188-91, 192 membrane protein crystallization 263
mixed-bed resins 287-8 modelling 112-16
modelling, see mathematical models orthogonal arrays 84-5
monovalent ion additives 222 osmium compounds 383
mother liquor 123 osmotic pressure techniques 324, 325-6
mounting solutions and residual protein overlap extension method 59-61
concentration 277
multiple anomalous dispersion 64, 366
multiple regression 102-12 packing 10
mutagenesis 59-64 PEG 122,261-2
random 61-2, 63-4 purification 122-3
site-directed 59-61 periodic bond chain theory 329
pH 220, 293-5, 304
phase diagrams 141-3, 278-81
net charge 283-6, 288-90 phenol extraction of RNA 214-15
websites 286 platinum compounds 381
neutron scattering 324, 325 polarized light 404-5
niobium compounds 384 polyamine additives 221
nuclease inhibitors 30, 34 polyethylene glycol (PEG) 122, 261-2
nucleation 10, 279-81 purification 122-3
epitaxial 180,200-2 polymorphism 336-7
heterogeneous 280, 315, 318-19 lysozyme 308-10
homogeneous 279, 315, 316-18 resolution by modelling 116-18
primary versus secondary 315 precipitation 279
rate 280-1,315-16 pre-nucleation 321-8
two-dimensional 330 pre-seeding 179-84
nucleic acid crystallization 209-24 pressure manipulation 140
additives 221-3 probes, making and cleaning 185
agents 220 proteases 29, 58
buffers 220-1 inhibitors 30,33
co-crystallizations 224-34 preparation 33-4
additives 227 proteins
agents 227 aggregation 55, 56
DNArdrug 219, 224 co-crystallizations 203, 224-34
432
Index
additives 227 phenol extraction 214—15
agents 227 purification 22,215-16
analysis techniques, 204-6 storage 23
DNA:protein 227-31 synthesis
homogeneity problems 226 chemical 213-14
protocols 228-30, 232-3 fragment design 211
purification 226-7 in vitro 212-13
RNA:protein 231-4 tRNAs
stability 226 co-crystallization with proteins 231-4
stoichiometry 226 modifications 30
streak seeding 204 purification 22, 215-16
concentration estimation 25 sources 214—15
modifications 29-30 see also nucleic acid crystallization
oxidation 56
protein-protein interactions 293
purification 21-2, 283-4
residual, concentration measurement 274-8 salt 298-301
sequence modification 55-64 crystal recognition 405
shortening 57-9 removal
storage 22-3 by dialysis 123-4, 286,288
streak seeding 204 by mixed-bed resins 287-8
tagged 50-1, 52 and solubility 292, 295-301
see also membrane protein crystallization 'salting-in' 296
proteolysis 29 'salting-out' 9, 296, 298
limited 58-9 membrane proteins 262-3
purification/purity 7-9 samples
co-crystallizations 226-7 ageing 23
crystallography grade 8-9, 27 biochemical analysis 24-6
of detergents 248-9, 251, 255 concentration 124-5
improving 32-3 estimation 25-6
of membrane proteins 255—7 measurement 125
of nucleic acids 22, 215-16 handling 26-7
of polyethylene glycol (PEG) 122-3 homogeneity 28-31
probing 31-2 improving 33-4
of proteins 21-2, 283-4 probing 31-2
of RNA 22, 215-16 preparation 121-5
of selenomethionyl proteins 68-9 purification, see purification/purity
techniques 19-22,28 solid particle removal 125
see also impurities sources 18-19
Pyrex plates 136 storage 22-3
sampling 82-8
sandwich drops 130
random mutagenesis 61-2, 63—4 with ACA CrystalPlates® 134-6
refrigerators 121 Schaeffer method 354
residual protein concentration 274—8 scintillation and solubility measurement 274
resins, mixed-bed 287-8 screening 306, 307
response surface 75,77 kits 84, 223
ribosome crystals 234-5 models 82-5, 89-96
RNA SDS-PAGE analysis 204-5
co-crystallization with proteins 231-4 seeding 177-208
concentration 216-18 analytical 185-8
estimation 26 complexes 204
crystallization 210 cross-seeding 177, 180, 197-200
protocols 219 heterogeneous 180, 196-200
handling 216-18 homogeneous 180
natural macroseeding 178, 180, 191, 193-6
purification 215-16 of needles 196
sources 214-15 microseeding 178, 180, 188-91, 192
433
Index
seeding (continued) microscopy 273-4
pre-seeding 179-84 scintillation 274
by vapour diffusion 180-4 temperature controlled light scattering
probes 185 274
residual protein concentration measurement and pH 293-5, 304
276-7 and salts 292, 295-301
streak seeding 177 and temperature 121, 302-3, 304
analytical 185-8 see also supersaturation
complexes 204 solution preparation 121-3
microseeding 190-1 solvents 369
selenolsubtilisin 199-200 space crystallization 170-2
selenomethionine substitution 64-70 space group determination 416-17
crystallization 69 sparse matrix kits 84
crystal storage 69-70 spermine 221, 223
eukaryotes 68 spiral growth 331-4
health and safety measures 70 stationary point identification 108-12
prokaryotes 65-8 statistics packages 105
purification 68-9 stepwise replacement 304-5
silica gel 150-2 storing samples 22-3
preparation 158-9 streak seeding 177
see also gel crystallization analytical 185-8
site-directed mutagenesis 59-61 complexes 204
sitting drops 130 microseeding 190-1
with ACA CrystalPlates® 134-6 subcloning 53
in capillaries 137 supersaturation 9, 178-9, 270, 271, 279,
with Cryschem plates 136 314-15
in Linbro boxes 137 rate 279
making plates for 184
soaking 372-3
small angle neutron scattering 324, 325
small angle X-ray scattering 324, 325, 327-8 tantalum compounds 384
soaking 365-90 Tag DNA polymerase 61
applications 366, 377-88 Teflon supports 346-7
capillaries 374-6 temperature
for cryo-crystallography 379, 381, 385-8 and crystal growth 334-5
in dilute ligand solutions 376-7 and nucleic acid crystallization 220
hanging drops 373 regulation 121, 181
for heavy-atom derivatives 366, 377-9 and solubility 121, 302-3, 304
soaking compounds 379, 380-5 thallium compounds 382
hydrophobic ligands 367-8 thaumatin crystallization 144-5
for isomorphous replacement 366, 377-9 thin-layer chromatography
sitting drops 372-3 of detergents 249
techniques 368-77 oflipids 256
software packages thymol 137
experimental design 86-8 tRNAs
net charge estimation 284 co-crystallization with proteins 231—4
statistics 105 modifications 30
X-ray data processing 414-16 purification 22, 215-16
solubility 9, 269-83 sources 214-15
denned 270 tungsten compounds 383
diagrams 141-3,278-81 two-dimensional crystallization 341-63
factors affecting 291-305 advantages 344
and ionic strength 295-8, 304 electron microscopy 349-60
measurement 271-4 grid preparation 349-54
calorimetry 274 negative staining 355
column method 273 specimen preparation 355
crystallization 271—3 transfer of films onto grids 354
Michelson interferometry 274 lipid solution preparation, 344-5
434
Index
xenon 379
X-ray crystallography 391-419
reproducibility 362 crystal choice 405
setting up 344-8 data 411-18
Teflon supports 346-7 collection 411-14, 417-18
two-dimensional nucleation 330 processing 414-17
resolution 400, 402
diffraction patterns 395-9
diffractometry 418
uranium compounds 383 Laue technique 418
mounting crystals 404-11
for cryo-crystallography 410-11
at low temperatures 408-11
vapour diffusion 130-8, 142, 145 at room temperature 405-7
of membrane proteins 263-4 optical alignment 413
and pre-seeding 180-4 oscillation data 412-16
vectors, see expression systems preliminary investigations 402-4
virus crystals 235—8 pre-nucleation investigations, 324, 325,
327-8
space group determination 416-17
X-ray sources 392
web/internet sites
crystallization databases 288-9
E. coli Genetic Stock Center 65 yeast expression systems 47, 51
experimental design 86-7
heavy-atom compound suppliers 385
protein net charge estimates 286 Zeppenzauer cells 126
435