International
Tables for
Crystallography
Volume C
Mathematical, physical and chemical tables
Edited by E. Prince

International Tables for Crystallography (2006). Vol. C. ch. 3.1, pp. 148-151

Section 3.1.1. Crystallization

P. F. Lindleya

a ESRF, Avenue des Martyrs, BP 220, F-38043 Grenoble CEDEX, France

3.1.1. Crystallization

| top | pdf |

3.1.1.1. Introduction

| top | pdf |

The preparation of single crystals probably constitutes the most important step in a crystal structure analysis, since without high-quality diffraction data many analyses will prove problematical, if not completely intractable; time and effort invested in crystallization procedures are rarely wasted. There is a wealth of literature available on the subject of growing crystals and this includes the Journal of Crystal Growth (Amsterdam: Elsevier). This section does not intend to be a comprehensive review of the subject, but rather to provide some key lines of approach with appropriate references. The field of crystallizing biological macromolecules is itself a growth area and, in consequence, has been given a special emphasis.

Useful general references for growing crystals for structure analysis include Bunn (1961[link]), Stout & Jensen (1968[link]), Blundell & Johnson (1976[link]), McPherson (1976[link], 1982[link], 1990[link]), Ducruix & Giegé (1992[link]) and Helliwell (1992[link]). Volume D50 (Part 4) of Acta Crystallographica (1994) reports the Proceedings of the Fifth International Conference on Crystallization of Biological Macromolecules (San Diego, California, 1993) and is essential reading for crystallization experiments in this area. A biological macromolecular database for crystallization conditions has also been initiated (Gilliland, Tung, Blakeslee & Ladner, 1994[link]).

3.1.1.2. Crystal growth

| top | pdf |

Crystallization has long been used as a method of purification by chemists and biochemists, although lack of purity can severely hamper the growth of single crystals, particularly if the impurities have some structural resemblance to the molecule being crystallized (Giegé, Theobald-Dietrich & Lorber, 1993[link]; Thatcher, 1993[link]). The process of crystallization involves the ordering of ions, atoms, and molecules in the gas, liquid, or solution phases to take up regular positions in the solid state. The initial stage is nucleation, followed by deposition on the crystallite faces. The latter can be considered as a dynamic equilibrium between the fluid and the crystal, with growth occurring when the forward rate predominates. Factors that affect the equilibrium include the chemical nature of the crystal surface, the concentration of the material being crystallized, and the nature of the medium in and around the crystal. Relatively little research has been done concerning the process of nucleation, but crystal formation appears to be conditional on the appearance of nuclei of a critical size. Too small aggregates will have either a positive or an unfavourable free energy of formation, so that there is a tendency to dissolution, whilst above the critical size the intermolecular interactions will, on average, lead to an overall negative free energy of formation. The rate of nucleation will increase considerably with the degree of supersaturation, and, in order to limit the number of nuclei (and therefore number of crystals growing), the degree of supersaturation must be as low as possible. Supersaturation must be approached slowly, and, when a low degree has been achieved, it must be carefully controlled. Many factors can influence crystallization, but a conceptually simple explanation of crystal growth has been described in detail by Tipson (1956[link]) and elaborated, for example, by Ries-Kautt & Ducruix (1992[link]). These latter authors provide a useful schematic description of the two-dimensional solubility diagram relating the concentration of the molecule being crystallized to the concentration of the crystallizing agent. The presence of foreign bodies, such as dust particles, makes the nucleation process thermodynamically more favourable, and these should be removed by centrifugation and/or filtration. The addition of seed crystals can often be used to control the nucleation process (Thaller, Eichelle, Weaver, Wilson, Karlsson & Jansonius, 1985[link]). In the case of the formation of crystals of macromolecules in solution, Ferré-D'Amaré & Burley (1994[link]) have described the use of dynamic light scattering to screen crystallization conditions for monodispersity. Empirical observations suggest that macromolecules that have the same size under normal solvent conditions tend to form crystals, whereas those systems that are polydisperse, or where random aggregation occurs, rarely give rise to ordered crystals.

3.1.1.3. Methods of growing crystals

| top | pdf |

General strategies for crystallizing low-molecular-weight organic compounds have been reported by van der Sluis, Hezemans & Kroon (1989[link]) and are listed in Table 3.1.1.1[link]. Many of these strategies are also applicable to inorganic compounds. In the case of biological macromolecules, the main methods utilize one or more of the factors described in Subsection 3.1.1.5[link] and include batch crystallization, the hot-box technique, equilibrium dialysis, and vapour diffusion (see, for example, Blundell & Johnson, 1976[link]; Helliwell, 1992[link]). The growth of macromolecular crystals in silica hydrogels minimizes convection currents, turbidity, and any strain effects due to the presence of the crystallization vessel. Heterogeneous and secondary nucleation are also reduced (Robert, Provost & Lefaucheux, 1992[link]; Cudney, Patel & McPherson, 1994[link]; García-Ruiz & Moreno, 1994[link]; Thiessen, 1994[link]; Robert, Bernard & Lefaucheux, 1994[link]; Bernard, Degoy, Lefaucheux & Robert, 1994[link]; Sica et al., 1994[link]). Various apparata have been described for use with the vapour diffusion technique (see also Subsection 3.1.1.6[link]) and include a simple capillary vapour diffusion device for preliminary screening of crystallization conditions (Luft & Cody, 1989[link]), a double-cell device that decouples the crystal nucleation from the crystal growth, facilitating the control of nucleation and growth (Przybylska, 1989[link]), microbridges for use with sitting drops in the 35–45 µl range (Harlos, 1992[link]), and diffusion cells with varying depths, in order to control the time course of the equilibration between the macromolecule and the reservoir solution (Luft et al., 1994[link]).

Table 3.1.1.1| top | pdf |
Survey of crystallization techniques suitable for the crystallization of low-molecular-weight organic compounds for X-ray crystallography (adapted from van der Sluis, Hezemans & Kroon, 1989[link])

TechniqueAdvantage(s)Limitation(s)
Evaporation from a single solventSimple
Inexpensive
Limitation to solvents with adequate vapour pressure
Crust formation on tube walls
Crystals that are dried are less suitable as seeds, may lose included solvent and become tightly adhered to the crystallization vessel
Difficult to reproduce
Limited number of solvents give concentration 5–200 mg ml−1 for a particular compound
Evaporation from a binary mixture of solvents (volatile solvent and non-volatile precipitant)No crust formation on the tube walls
Crystals are not dried
Stringent demands on solubility, miscibility and volatility of the two solvents
Difficult to reproduce
Batch crystallizationNo demands on the volatility of the solvent or precipitant
Repeated seeding by thermal treatment is easy
Metastable zone with regard to supersaturation must be large
High and almost uncontrollable crystallization rate
Solvents must be miscible
Liquid–liquid diffusionFavourable change in supersaturation at the interface during crystallization
Repeated seeding by thermal treatment is easy
Density differences required for the two liquids (less stringent if capillaries are used)
Viscosity of the liquids greater than water
Solvents must be miscible
High and almost uncontrollable crystallization rate
Sitting-drop vapour-phase diffusionCrystallization rate can easily be controlled by changing the diffusion path, solvent, precipitant, or pH
Repeated seeding easily implemented
Highest number of independent variables to obtain wide variety of conditions
Solvents must be miscible
Solvent preferably less volatile than precipitant
Hanging-drop vapour-phase diffusionCrystallization rate can easily be controlled by changing solvent, precipitant, or pH
Easy examination of crystallization outcome in array-like set-up
Only applicable in case of water-based solvents
Diffusion rate is fast and difficult to control
See previous method
Temperature changeEasily controllable parameter
Repeated seeding extremely easily and accurately carried out
With Dewar flask inexpensive and simple
Limited to thermally stable compounds and (pseudo)polymorphs
Gel crystallizationSuited for sparingly soluble or easily nucleating compoundsLimited variety of solvents possible
Sampling of crystals difficult
Laborious
SublimationNo inclusion of solvent of crystallizationLimited to small hydrophobic molecules
Laborious
SolidificationFor liquids and gases the only applicable methodLimited to thermostable compounds
High change of amorphicity
Laborious

3.1.1.4. Factors affecting the solubility of biological macromolecules

| top | pdf |

There are many factors that influence the crystallization of macromolecules (McPherson, 1985a[link]; Giegé & Ducruix, 1992[link]; Schick & Jurnak, 1994[link]; Tissen, Fraaije, Drenth & Berendsen, 1994[link]; Carter & Yin, 1994[link]; Spangfort, Surin, Dixon & Svensson, 1994[link]; Axelrod et al., 1994[link]; Konnert, D'Antonio & Ward, 1994[link]; Forsythe, Ewing & Pusey, 1994[link]; Diller, Shaw, Stura, Vacquier & Stout, 1994[link]; Hennig & Schlesier, 1994[link]), but the following are particularly important with respect to solubility (Blundell & Johnson, 1976[link]).

Ionic strength. The solubility of macromolecules in aqueous solution depends on the ionic strength, since the presence of ions modifies the interactions of the macromolecule with the solvent. At low ion concentrations, the solubility of the macromolecule is increased, a phenomenon termed `salting-in'. As the ionic strength is increased, the ions added compete with one another and the macromolecules for the surrounding water. The resulting removal of water molecules from the solute leads to a decrease in the solubility, a phenomenon termed `salting-out'. Different ions will affect the solubility of the protein in different ways. Small highly charged ions will be more effective in the salting-out process than large low-charged ions. Commonly used ionic precipitants are listed in Table 3.1.1.2[link], column (a) (McPherson, 1985a[link]).

Table 3.1.1.2| top | pdf |
Commonly used ionic and organic precipitants, adapted from McPherson (1985a[link])

(a) Ionic compounds(b) Organic solvents
Ammonium or sodium sulfateEthanol
Sodium or ammonium citrateIsopropanol
Sodium, potassium or ammonium chloride2-Methyl-2,4-pentanediol (MPD)
Sodium or ammonium acetateDioxane
Magnesium sulfateAcetone
Cetyltrimethylammonium saltsButanol
Calcium chlorideDimethyl sulfoxide
Ammonium nitrate2,5-Hexanediol
Sodium formateMethanol
Lithium chloride1,3-Propanediol
 1,3-Butyrolactone
 Poly(ethylene glycol) 600–20000 (PEG)
The volatility of solvents such as ethanol and acetone may cause handling problems.
Ammonium sulfate can cause problems when used as a precipitant, since pH changes occur owing to ammonium transfer following ammonium/ammonia equilibrium; this effect has been studied in detail by Mikol, Rodeau & Giegé (1989[link]). Monaco (1994[link]) has suggested that ammonium succinate is a useful substitute for ammonium sulfate.

pH and counterions. The net charge on a macromolecule in solution can be modified either by changing the pH (adding or removing protons) or by binding ions (counterions). In general terms, the protein solubility will increase with the overall net charge and will be least soluble when the net charge is zero (isoelectric point). In the latter case, the molecules can pack in the crystalline form without an overall, destabilizing accumulation of charge.

Temperature . Temperature has a marked affect on many of the factors that govern the solubility of a macromolecule. The dielectric constant decreases with increase in temperature, and the entropy terms in the free energy tend to dominate the enthalpy terms (Blundell & Johnson, 1976[link]). The temperature coefficient of solubility varies with ionic strength and the presence of organic solvents. McPherson (1985b[link]) gives a useful account of protein crystallization by variation of pH and temperature.

Organic solvents. Addition of organic solvents can produce a marked change in the solubility of a macromolecule in aqueous solution (care should be taken to avoid denaturation). This is partly due to a lowering of the dielectric constant, but may also involve specific solvation and displacement of water at the surface of the macromolecule. Generally, the solubility decreases with decrease of temperature when substantial amounts of organic solvent are present. Commonly used organic precipitants are listed in Table 3.1.1.2[link], column (b) (McPherson, 1985a[link]).

3.1.1.5. Screening procedures for the crystallization of biological macromolecules

| top | pdf |

Optimal conditions for crystal growth are often very difficult to predict a priori, although many proteins crystallize close to their pI. In order to surmount the problem of testing a very large range of conditions, Carter & Carter (1979[link]) devised the incomplete factorial method, in which a very coarse matrix of crystallization conditions is explored initially. Finer grids are then investigated around the most promising sets of coarse conditions. This technique has been further refined to yield the sparse-matrix sampling technique described by Jancarik & Kim (1991[link]). Table 3.1.1.3[link] lists the crystallization parameters used by these authors. The 50 conditions constituting the sparse matrix are given in Table 3.1.1.4[link]. A recent update of this matrix and a set of stock solutions in the form of a crystal screen kit can be obtained commercially from Hampton Research (1994[link]). Further developments in screening methods are described in Volume D50 (Part 4) of Acta Crystallographica (1994).

Table 3.1.1.3| top | pdf |
Crystallization matrix parameters for sparse-matrix sampling, adapted from Jancarik & Kim (1991[link])

Precipitating agents
Non-volatileSaltsVolatileMixture
2-Methyl-2,4-pentanediol (MPD)Na, K tartrate2-PropanolNH4 sulfate + PEG
Poly(ethylene glycol) (PEG) 400NH4 phosphate 2-Propanol + PEG
PEG 4000NH4sulfate  
PEG 8000Na acetate  
 Li sulfate  
 Na formate  
 Na, K phosphate  
 Na citrate  
 Mg formate  
Range of pH: 4.6, 5.6, 7.5, 8.5
Salts, additives:Ca chloride, Na citrate, Mg chloride, NH4 acetate, NH4 sulfate, Mg acetate, Zn acetate, Ca acetate

Table 3.1.1.4| top | pdf |
Reservoir solutions for sparse-matrix sampling (Jancarik & Kim 1991[link])

No.SaltBufferPrecipitant
(% by mass)
10.02 M Ca chloride0.1 M Acetate30% MPD
2  0.4 M Na, K tartrate
3  0.4 M NH4 phosphate
4 0.1 M Tris2.0 M NH4 sulfate
50.2 M Na citrate0.1 M Hepes40% MPD
60.2 M Mg chloride0.1 M Tris30% PEG 4000
7 0.1 M Cacodylate1.4 M Na acetate
80.2 M Na citrate0.1 M Cacodylate30% 2-Propanol
90.2 M NH4 acetate0.1 M Citrate30% PEG 4000
100.2 M NH4 acetate0.1 M Acetate30% PEG 4000
11 0.1 M Citrate1.0 M NH4 phosphate
120.2 M Mg chloride0.1 M Hepes30% 2-Propanol
130.2 M Na citrate0.1 M Tris30% PEG 400
140.2 M Ca chloride0.1 M Hepes28% PEG 400
150.2 M NH4 sulfate0.1 M Cacodylate30% PEG 8000
16 0.1 M Hepes1.5 M Li sulfate
170.2 M Li sulfate0.1 M Tris30% PEG 4000
180.2 M Mg acetate0.1 M Cacodylate20% PEG 8000
190.2 M NH4 acetate0.1 M Tris30% 2-Propanol
200.2 M NH4 sulfate0.1 M Acetate25% PEG 4000
210.2 M Mg acetate0.1 M Cacodylate30% MPD
220.2 M Na acetate0.1 M Tris30% PEG 4000
230.2 M Mg chloride0.1 M Hepes30% PEG 400
240.2 M Ca chloride0.1 M Acetate20% 2-Propanol
25 0.1 M Imidazole1.0 M Na acetate
260.2 M NH4 acetate0.1 M Citrate30% MPD
270.2 M Na citrate0.1 M Hepes20% 2-Propanol
280.2 M Na acetate0.1 M Cacodylate30% PEG 8000
29 0.1 M Hepes0.8 M Na, K tartrate
300.2 M NH4 sulfate 30% PEG 8000
310.2 M NH4 sulfate 30% PEG 4000
32  2.0 M NH4 sulfate
33  4.0 M Na formate
34 0.1 M Acetate2.0 M Na formate
35 0.1 M Hepes1.6 M Na, K phosphate
36 0.1 M Tris8% PEG 8000
37 0.1 M Acetate8% PEG 4000
38 0.1 M Hepes1.4 M Na citrate
39 0.1 M Hepes2% PEG 400, 2.0 M Na sulfate
40 0.1 M Citrate20% 2-Propanol + 20% PEG 4000
41 0.1 M Hepes10% 2-Propanol + 20% PEG 4000
420.05 M K phosphate 20% PEG 8000
43  30% PEG 1500
44  0.2 M Mg formate
450.2 M Zn acetate0.1 M Cacodylate18% PEG 8000
460.2 M Ca acetate0.1 M Cacodylate18% PEG 8000
47 0.1 M Acetate2.0 M NH4 sulfate
48 0.1 M Tris2.0 M NH4 sulfate
491.0 M Li sulfate 2% PEG 8000
501.0 M Li sulfate 15% PEG 8000

Abbreviations: tris: 2-amino-2-(hydroxymethyl)-1,3-propanediol; hepes: 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid. Buffers: Na acetate buffer, pH = 4.6; Na citrate buffer, pH = 5.6; Na cacodylate buffer, pH = 6.5; Na hepes buffer, pH = 7.5; tris/HCl buffer, pH = 8.5.

3.1.1.6. Automated protein crystallization

| top | pdf |

Several liquid-handling systems have been described that can automatically set up, reproducibly, a range of crystallization conditions (different protein concentrations, ionic strengths, amounts of organic precipitant, etc.) for the hanging-drop, sitting-drop, and microbatch methods. A useful introduction describing a system for mixing both buffered protein solutions and the corresponding reservoirs is given by Cox & Weber (1987[link]). Chayen, Shaw Stewart, Maeder & Blow (1990[link]) describe an automatic dispenser involving a bank of Hamilton syringes driven by stepper motors under computer control that can be used to set up small samples (2 µl or less) for microbatch crystallization (or hanging drops). Further systems have been described by Oldfield, Ceska & Brady (1991[link]), Eiselé (1993[link]), Soriano & Fontecilla-Camps (1993[link]), Sadaoui, Janin & Lewit-Bentley (1994[link]), and Chayen, Shaw Stewart & Baldock (1994[link]).

3.1.1.7. Membrane proteins

| top | pdf |

Integral membrane proteins can be considered as those whose polypeptide chains span the lipid bilayer at least once. The external membrane segments exposed to an aqueous environment are hydrophilic, but it is the tight interaction of the hydrophobic segments of the chain with the quasisolid lipid bilayer that constitutes the major problem in their crystallization. Crystallization trials require disruption of the membrane, isolation of the protein, and solubilization of the resultant hydrophobic region (McDermott, 1993[link]). Organic solvents, chaotropic agents, and amphipathic detergents can be used to disrupt the membrane, but detergents such as β-octyl glucoside are most commonly used, since they minimize the loss of protein integrity. The several classes of detergent employed tend to be non-ionic or zwitterionic at the pH used, have a maximum hydrocarbon chain length of 12 carbon atoms, and possess a critical micelle concentration. The key to crystallizing membrane protein–detergent complexes appears to be the attainment of conditions in which the protein surfaces are moderately supersaturated and, in addition, the detergent micellar collar is at, or near, its solubility limit (Scarborough, 1994[link]). Most successful integral membrane protein crystallizations are near the micellar aggregation point of the detergent (Garavito & Picot, 1990[link]).

References

First citation Axelrod, H. J., Feher, G., Allen, J. P., Chirino, A. J., Day, M. W., Hsu, B. T. & Rees, D. C. (1994). Crystallization and X-ray structure determination of cytochrome c2 from rhodobacter sphaeroides in three crystal forms. Acta Cryst. D50, 596–602.Google Scholar
First citation Bernard, Y., Degoy, S., Lefaucheux, F. & Robert, M. C. (1994). A gel-mediated feeding technique for protein crystal growth from hanging drops. Acta Cryst. D50, 504–507.Google Scholar
First citation Blundell, T. L. & Johnson, L. N. (1976). Protein crystallography, Chap. 3, pp. 59–82. New York: Academic Press.Google Scholar
First citation Bunn, C. W. (1961). Chemical crystallography: an introduction to optical and X-ray methods, 2nd ed., Chaps. 2, 3, 4, pp. 11–106. Oxford University Press.Google Scholar
First citation Carter, C. W. Jr & Carter, C. W. (1979). Protein crystallisation using incomplete factorial experiments. J. Biol. Chem. 254, 12219–12223.Google Scholar
First citation Carter, C. W. Jr & Yin, Y. (1994). Quantitative analysis in the characterization and optimization of protein crystal growth. Acta Cryst. D50, 572–590.Google Scholar
First citation Chayen, N. E., Shaw Stewart, P. D. & Baldock, P. (1994). New developments of the IMPAX small-volume automated crystallization system. Acta Cryst. D50, 456–458.Google Scholar
First citation Chayen, N. E., Shaw Stewart, P. D., Maeder, D. L. & Blow, D. M. (1990). An automated system for micro-batch protein crystallization and screening. J. Appl. Cryst. 23, 297–302.Google Scholar
First citation Cox, M. J. & Weber, P. C. (1987). Experiments with automated protein crystallization. J. Appl. Cryst. 20, 366–373.Google Scholar
First citation Cudney, B., Patel, S. & McPherson, A. (1994). Crystallization of macromolecules and silica gels. Acta Cryst. D50, 479–483.Google Scholar
First citation Diller, T. C., Shaw, A., Stura, E. A., Vacquier, V. D. & Stout, C. D. (1994). Acid pH crystallization of the basic protein lysin from the spermatozoa of red abalone (Haliotis refescens). Acta Cryst. D50, 627–631.Google Scholar
First citation Ducruix, A. & Giegé, R. (1992). Editors. Crystallisation of nucleic acids and proteins: a practical approach. Oxford University Press.Google Scholar
First citation Eiselé, J.-L. (1993). Preparation of protein crystallization buffers with a computer-controlled motorized pipette: PIPEX. J. Appl. Cryst. 26, 92–96.Google Scholar
First citation Ferré-D'Amaré, A. R. & Burley, S. K. (1994). Use of dynamic light scattering to assess crystallisability of macromolecules and macromolecular assemblies. Structure, 2(5), 357–359.Google Scholar
First citation Forsythe, E., Ewing, F. & Pusey, M. (1994). Studies on tetragonal lysozyme crystal growth rates. Acta Cryst. D50, 614–619.Google Scholar
First citation Garavito, R. M. & Picot, D. (1990). Methods: a companion to methods in enzymology, Vol. 1, pp. 57–69. New York: Academic Press.Google Scholar
First citation García-Ruiz, J. M. & Moreno, A. (1994). Investigations on protein crystal growth by the gel acupuncture method. Acta Cryst. D50, 484–490.Google Scholar
First citation Giegé, R. & Ducruix, A. (1992). Crystallisation of nucleic acids and proteins: a practical approach, edited by A. Ducruix & R. Giegé, pp. 1–15. Oxford University Press.Google Scholar
First citation Giegé, R., Theobald-Dietrich, A. & Lorber, B. (1993). Novel trends in protein and nucleic acid crystallisation: biochemical and physico-chemical aspects. Data collection and processing. Proceedings of the CCP4 Study Weekend, edited by L. Sawyer, N. Isaacs & S. Bailey, pp. 12–19. SERC Daresbury Laboratory, Warrington WA4 4AD, England.Google Scholar
First citation Gilliland, G. L., Tung, M., Blakeslee, D. M. & Ladner, J. E. (1994). Biological macromolecule crystallization database, Version 3.0: new features, data and the NASA archive for protein crystal growth data. Acta Cryst. D50, 408–413.Google Scholar
First citation Hampton Research (1994). Crystallisation research tools, Vol. 4, No. 2. Hampton Research, 5225 Canyon Crest Drive, Suite 71–336, Riverside, CA 92597, USA.Google Scholar
First citation Harlos, K. (1992). Micro-bridges for sitting-drop crystallizations. J. Appl. Cryst. 25, 536–538.Google Scholar
First citation Helliwell, J. R. (1992). Macromolecular crystallography with synchrotron radiation, Chap. 2, pp. 11–23. Cambridge University Press.Google Scholar
First citation Hennig, M. & Schlesier, B. (1994). Crystallization of seed globulins from legumes. Acta Cryst. D50, 627–631.Google Scholar
First citation Jancarik, J. & Kim, S.-H. (1991). Sparse matrix sampling: a screening method for crystallization of proteins. J. Appl. Cryst. 24, 409–411.Google Scholar
First citation Konnert, J. H., D'Antonio, P. & Ward, K. B. (1994). Observation of growth steps, spiral dislocations and molecular packing on the surface of lysozyme crystals with the atomic force microscope. Acta Cryst. D50, 603–613.Google Scholar
First citation Luft, J. & Cody, V. (1989). A simple capillary vapor diffusion apparatus for surveying macromolecular crystallization conditions. J. Appl. Cryst. 22, 396.Google Scholar
First citation Luft, J. R., Arakali, S. V., Kirisits, M. J., Kalenik, J., Waarzak, I., Cody, V., Pangborn, W. A. & DeTitta, G. T. (1994). A macromolecular crystallization procedure employing diffusion cells of varying depths as reservoirs to tailor the time course of equilibration in hanging- and sitting-drop vapor-diffusion and microdialysis experiments. J. Appl. Cryst. 27, 443–452.Google Scholar
First citation McDermott, G. (1993). Crystallisation of membrane proteins. Data collection and processing. Proceedings of the CCP4 Study Weekend, edited by L. Sawyer, N. Isaacs & S. Bailey, pp. 20–27. SERC Daresbury Laboratory, Warrington WA4 4AD, England. Google Scholar
First citation McPherson, A. (1976). The growth and preliminary investigation of protein and nucleic acid crystals for X-ray diffraction. Methods Biochem. Anal. 25, 249–345.Google Scholar
First citation McPherson, A. (1982). The preparation and analysis of protein crystals. New York: John Wiley. Google Scholar
First citation McPherson, A. (1985a). Crystallisation of macromolecules: general principles. Methods in enzymology, Vol. 114, pp. 112–120. New York: Academic Press.Google Scholar
First citation McPherson, A. (1985b). Crystallisation of proteins by variation of pH and temperature. Methods in enzymology, Vol. 114, pp. 125–127. New York: Academic Press.Google Scholar
First citation McPherson, A. (1990). Current approaches to macromolecular crystallisation. Eur. J. Biochem. 189, 1–23.Google Scholar
First citation Mikol, V., Rodeau, J.-L. & Giegé, R. (1989). Changes of pH during biomacromolecule crystallization by vapour diffusion using ammonium sulfate as the precipitant. J. Appl. Cryst. 22, 155–161.Google Scholar
First citation Monaco, H. L. (1994). The use of ammonium succinate in protein crystallography. J. Appl. Cryst. 27, 1068.Google Scholar
First citation Oldfield, T. J., Ceska, T. A. & Brady, R. L. (1991). A flexible approach to automated protein crystallization. J. Appl. Cryst. 24, 255–260.Google Scholar
First citation Przybylska, M. (1989). A double cell for controlling nucleation and growth of protein crystals. J. Appl. Cryst. 22, 115–118.Google Scholar
First citation Ries-Kautt, M. & Ducruix, A. (1992). Crystallisation of nucleic acids and proteins: a practical approach, edited by A. Ducruix & R. Giegé, pp. 195–218. Oxford University Press.Google Scholar
First citation Robert, M. C., Bernard, Y. & Lefaucheux, F. (1994). Study of nucleation-related phenomena in lysozyme solutions. Application to gel growth. Acta Cryst. D50, 496–503.Google Scholar
First citation Robert, M. C., Provost, K. & Lefaucheux, F. (1992). Crystallisation of nucleic acids and proteins: a practical approach, edited by A. Ducruix & R. Giegé, pp. 127–143. Oxford University Press.Google Scholar
First citation Sadaoui, N., Janin, J. & Lewit-Bentley, A. (1994). TAOS: an automated system for protein crystallization. J. Appl. Cryst. 27, 622–626.Google Scholar
First citation Scarborough, G. A. (1994). Large single crystals of the nerospora crassa plasma membrane H+-ATPase: an approach to the crystallization of integral membrane proteins. Acta Cryst. D50, 643–649. Google Scholar
First citation Schick, B. & Jurnak, F. (1994). Extension of the diffraction resolution of crystals. Acta Cryst. D50, 563–568.Google Scholar
First citation Sica, F., Demasi, D., Mazzarella, L., Zagari, A., Capasso, S., Pearl, L. H., D'Auria, S., Raia, C. A. & Rossi, M. (1994). Elimination of twinning in crystals of sulfolobus solfataricus alcohol dehydrogenase holo-enzyme by growth in agarose gels. Acta Cryst. D50, 508–511.Google Scholar
First citation Sluis, P. van der, Hezemans, A. M. F. & Kroon, J. (1989). Crystallization of low-molecular-weight compounds for X-ray crystallography. J. Appl. Cryst. 22, 340–344. Google Scholar
First citation Soriano, T. M. B. & Fontecilla-Camps, J. C. (1993). ASTEC: an automated system for sitting-drop protein crystallization. J. Appl. Cryst. 26, 558–562.Google Scholar
First citation Spangfort, M. D., Surin, B. P., Dixon, N. E. & Svensson, L. A. (1994). Internal symmetry of the molecular chaperone cpn60 (GroEL) determined by X-ray crystallography. Acta Cryst. D50, 591–595.Google Scholar
First citation Stout, G. H. & Jensen, L. H. (1968). X-ray structure determination: a practical guide, Chap. 4, pp. 62–82. London: Macmillan. Google Scholar
First citation Thaller, C., Eichelle, G., Weaver, L. H., Wilson, E., Karlsson, R. & Jansonius, J. N. (1985). Seed enlargement and repeated seeding. Methods in Enzymology, Vol. 114, pp. 132–135. New York: Academic Press.Google Scholar
First citation Thatcher, D. R. (1993). Protein purification and analysis for crystallographic studies. Data collection and processing. Proceedings of the CCP4 Study Weekend, edited by L. Sawyer, N. Isaacs & S. Bailey, pp. 2–11. SERC Daresbury Laboratory, Warrington WA4 4AD, England.Google Scholar
First citation Thiessen, M. J. (1994). The use of two novel methods to grow protein crystals by microdialysis and vapor diffusion in an agarose gel. Acta Cryst. D50, 491–495.Google Scholar
First citation Tipson, R. S. (1956). Techniques of organic chemistry, Vol. III, Part I, Chap. 3. New York: Interscience.Google Scholar
First citation Tissen, J. T. W. M., Fraaije, J. G. E. M., Drenth, J. & Berendsen, H. J. C. (1994). Mesoscopic theories for protein crystal growth. Acta Cryst. D50, 569–571.Google Scholar








































to end of page
to top of page