You are on page 1of 11

Liquid–liquid phase separation in supersaturated lysozyme solutions and

associated precipitate formation/crystallization


Martin Muschol and Franz Rosenberger

Citation: J. Chem. Phys. 107, 1953 (1997); doi: 10.1063/1.474547


View online: http://dx.doi.org/10.1063/1.474547
View Table of Contents: http://jcp.aip.org/resource/1/JCPSA6/v107/i6
Published by the American Institute of Physics.

Additional information on J. Chem. Phys.


Journal Homepage: http://jcp.aip.org/
Journal Information: http://jcp.aip.org/about/about_the_journal
Top downloads: http://jcp.aip.org/features/most_downloaded
Information for Authors: http://jcp.aip.org/authors

Downloaded 30 Mar 2013 to 128.233.210.97. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jcp.aip.org/about/rights_and_permissions
Liquid–liquid phase separation in supersaturated lysozyme solutions
and associated precipitate formation/crystallization
Martin Muschola) and Franz Rosenberger
Center for Microgravity and Materials Research, University of Alabama in Huntsville, Huntsville,
Alabama 35899
~Received 16 October 1996; accepted 2 May 1997!
Using cloud point determinations, the phase boundaries ~binodals! for metastable liquid–liquid
~L–L! separation in supersaturated hen egg white lysozyme solutions with 3%, 5%, and 7%
(w/ v ) NaCl at pH54.5 and protein concentrations c between 40 and 400 mg/ml were determined.
The critical temperature for the binodal increased approximately linearly with salt concentration.
The coexisting liquid phases both remained supersaturated but differed widely in protein
concentration. No salt repartitioning was observed between the initial and the two separated liquid
phases. After the L–L separation, due to the presence of the high protein concentration phase,
crystallization occurred much more rapidly than in the initial solution. At high initial protein
concentrations, a metastable gel phase formed at temperatures above the liquid binodal. Both crystal
nucleation and gel formation were accelerated in samples that had been cycled through the binodal.
Solutions in the gel and L–L regions yielded various types of precipitates. Based on theoretical
considerations, previous observations with other proteins, and our experimental results with
lysozyme, a generic phase diagram for globular proteins is put forth. A limited region in the
(T,c) plane favorable for the growth of protein single crystals is delineated. © 1997 American
Institute of Physics. @S0021-9606~97!51130-5#

I. INTRODUCTION existing crystal solubilities. These data, together with obser-


vations of gel formation, delineate the phase regions useful
There is growing evidence that supersaturated protein for crystal growth.
solutions can undergo a thermally induced separation into Recent phase diagram calculations for hard spheres with
two metastable liquid phases of widely different concentra- attractive interactions indicate that liquid phase separation
tion. This was first indicated by light scattering from unbuf- can occur underneath the solidus line as the range of the
fered lysozyme–salt solutions,1 and later extended to solu- potential is sufficiently reduced.12,13 In crystallizing protein
tions buffered to pH'7.2,3 The most detailed investigations solutions, the range of attractive interactions appears to fall
were concerned with g-crystallin and its natural variants.4–9 in the relevant range.14 The high salt ~precipitant! concentra-
We are interested in the impact of the demixing transi- tions typically used, reduce the range of the electrostatic re-
tion on protein crystallization. In the high protein- pulsion between the macroions such that the attractive inter-
concentration phase resulting from the demixing, crystalliza- actions dominate.3,15,16 These considerations suggest that
tion is typically too rapid for the growth of single crystals metastable L–L phase separation in crystallizing protein so-
with low defect densities. In addition, the question arises to lutions may be widespread.
what extent various forms of precipitate10 and the fractal
aggregates11 that have been observed in protein solutions are
II. MATERIALS AND METHODS
related to liquid–liquid phase separations? Answers to this
question are complicated by the noticeable lack of character- Our earlier work on lysozyme’s crystallization kinetics
ization of precipitates and their morphologies, which are and its relation to impurity concentrations has shown signifi-
typically considered a nuisance rather than a source of valu- cant differences with respect to various source materials.17–19
able information. Six times recrystallized and lyophilized lysozyme, obtained
Hence, we investigated the existence and salt- from Seikagaku America, showed kinetics comparable to
concentration dependence of the demixing transition in that obtained with highly purified material. Hence, we used
lysozyme solutions under crystallization conditions, and the this material ~lots E94203 and E94Z05! without further pu-
resulting precipitate morphologies. Using light scattering rification. Given the high salt concentrations of our solutions
measurements of the cloud point temperature, we have de- ~see below!, contamination of the solutions by the small
termined coexistence curves for binary liquid lysozyme so- amounts of salt introduced by the added protein was ne-
lutions at pH54.5 and three different NaCl concentrations glected. All other chemicals used were reagent grade or
that have been used in crystal growth studies of lysozyme. purer.
This allowed for comparison of the cloud point curves with The solutions were buffered with 0.1 M sodium acetate/
acetic acid ~NaAc! at pH54.5 as determined with an Orion
a!
Current address: Dept. of Neuroscience, University of Pennsylvania School Sa 520 pH meter. After dissolution of either 3%, 5%, or 7%
of Medicine, Philadelphia, PA 19104. NaCl (w/ v ) in the buffered solution, the pH was readjusted.

J. Chem. Phys. 107 (6), 8 August 1997 0021-9606/97/107(6)/1953/10/$10.00 © 1997 American Institute of Physics 1953

Downloaded 30 Mar 2013 to 128.233.210.97. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jcp.aip.org/about/rights_and_permissions
1954 M. Muschol and F. Rosenberger: Phase separation in lysozyme solutions

Up to protein concentrations of c<150 mg/ml, the lyo-


philized lysozyme powder was directly dissolved in the
buffer/salt mixtures. To remove undissolved protein, air
bubbles, and dust, samples were centrifuged in a Savant
HSC10K centrifuge at 9000 g ~7500 rpm! for 5–10 min and
filtered through a 0.22 mm Millipore Millex-GV filter into a
clean sample vial. For the 5% and 7% NaCl samples, buffer/
salt solutions were preheated to about 45–50 °C to expedite
dissolution of the protein and to keep sample temperatures
well above the cloud point. To obtain higher protein concen-
trations ~up to 400 mg/ml!, about 2 ml of a lysozyme solu-
tion with c'150 mg/ml was prepared as described above.
Then, after insertion of the vial into a 15 ml plastic centri-
fuge tube containing an ice/water mixture for cooling, it was
centrifuged at 3500 g for 5–10 min in a Hermle Labnet Z232
centrifuge. The concurrent cooling and centrifugation re-
sulted in two visibly distinct layers of different protein con-
centration. Lack of adequate temperature control and partial
remixing after centrifugation prevented the use of these sepa-
rated phases for direct determination of the binodal. After
FIG. 1. Time traces of sample temperature and light scattering signal for
decanting of the top layer ~low c!, this process was repeated lysozyme at c5360 mg/ml in 3% NaCl/0.1 M NaAc buffer at pH54.5;
until the desired c ~as measured by uv absorption of dilute T sat553 °C estimated from Eq. ~2!. Note the two reversible increases in
samples; a 28052.64 ml/mg cm!20 was obtained. Prior to the scattering intensity associated with brief decreases in sample temperature by
cloud point measurements, these samples were homogenized 0.5 °C. T cloud57.0 °C.
by centrifugation at 9000 g for 3 min at some elevated tem-
perature.
To check for salt repartitioning possibly associated with originates from the appearance of liquid droplets ~see below!
the L–L separation, we phase separated a solution containing or the onset of spinodal decomposition. Hence, we designate
110 mg/ml lysozyme and 3% NaCl, and determined the this temperature as T cloud for this sample composition. On
Na1 and Cl2 concentrations in the initial and resulting two raising the temperature shortly after the L–L separation by as
phases. Na1 determinations were carried out by atomic ab- little as 0.5 °C above T cloud, the cloudiness responsible for
sorption spectroscopy. The Cl2 concentrations were obtained the strong signal peak readily disappears. The second peak in
from a colorimetric procedure employing the Sigma Chemi- Fig. 1 shows that the cloud point determination is reproduc-
cals diagnostics kit #461-3. ible within a few tenths of a degree.
The coexistence curves for the binary liquid phases were The L–L phase separation enhances the crystal nucle-
determined through light scattering intensity measurements. ation rate in the solution. This is illustrated by Fig. 2, that
A light scattering cell of approx. 200 ml volume and 1 mm was obtained from a sample of initial c5140 mg/ml with 5%
path length ~Starna Cells, Inc.; type 37! was filled with the NaCl. Signal peak A, resulting from the demixing transition
solution, glass stoppered, and placed into a miniaturized and associated with the small dip in temperature from 22 to
computer-controlled scintillation cell setup similar to one de- 21.5 °C, suggests complete re-establishment of the initial,
scribed elsewhere.21 This setup provides for programmed homogeneous solution state. However, about 10 min later
control (60.1 °C) and monitoring of the sample tempera- ~point B! the scattering intensity steeply increased, even after
ture while simultaneously recording the 90° light scattering the temperature had been increased to 25 °C. This strong rise
signal. Determination of the cloud temperature T cloud as a in signal is due to scattering from small crystals that nucleate
function of protein concentration c at a given NaCl concen- upon phase separation in the high-c phase and continue to
tration maps out the coexistence curve in the (T,c) plane. grow at any T below the saturation temperature for the
solid–liquid transition T sat. For this combination of protein
III. RESULTS AND DISCUSSION and salt concentrations, T sat was estimated from a van’t Hoff
extrapolation of the experimental solubility data ~see below!
A. Cloud points and phase diagram
to be 67 °C. Thus it is not surprising that, as Fig. 2 shows,
Figure 1 shows time traces of the sample temperature raising the temperature to 38 °C could not remove the tur-
and corresponding scattering signal obtained from a solution bidity. ~The final plateau in the scattering intensity is due to
with c5360 mg/ml in 3% NaCl. One sees that on lowering the competition between enhanced scattering and enhanced
of the temperature, at first the scattering intensity increases absorption of light in these turbid samples, rather than a
slowly. This reflects an increase in concentration fluctuations steady state in crystal population.!
~due to attractive protein interactions! with decreasing ther- At higher salt and protein concentrations, enhanced
mal energy. Then, at about 7.0 °C, a jump in scattering in- nucleation was often noticeable right after cycling of the
tensity occurs. This sudden increase in sample turbidity sample through the L–L region: The scattering intensity

J. Chem. Phys., Vol. 107, No. 6, 8 August 1997

Downloaded 30 Mar 2013 to 128.233.210.97. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jcp.aip.org/about/rights_and_permissions
M. Muschol and F. Rosenberger: Phase separation in lysozyme solutions 1955

FIG. 2. Time traces for a sample with c5140 mg/ml; T sat567 °C, as esti-
mated from Eq. ~2!. Point A: small dip in temperature reversibly induces
liquid–liquid phase separation. Point B: spontaneous crystal nucleation.

FIG. 3. Cloud point data for lysozyme at pH54.5 in 0.1 M NaAc buffer and
above T cloud
remained elevated above the level prior to the three different NaCl concentration. Dashed lines: fit to Eq. ~1!. Open down-
ward pointing triangles: Cloud point data for lysozyme at pH56.0 in 0.6 M
phase separation and continued to grow. Such measurements sodium phosphate from Ref. 2.
were excluded from our cloud point data. It should be em-
phasized that without cycling through the binodal, nucleation
in the above solution at 25 °C would have required a much
longer time. In general, we have observed, that after cycling
through the binodal, nucleation induction times ~as judged by 610% and 62.5%, respectively, no changes from the initial
the appearance of crystallites! were reduced from hours to ion concentrations were found. Hence, this metastable phase
minutes. Two factors contribute to this pronounced lowering separation in lysozyme is essentially binary and the T cloud
of the kinetic barrier to crystallization: ~a! as we will see data can be justifiably presented in a pseudobinary ~protein-
below, the supersaturation in the high concentration phase is solution! phase diagram. This is fundamentally different
about an order of magnitude higher than in the starting solu- from coacervation, where the phase separation of a solution
tion; and ~b! the difference in surface energy between the component other than the protein leads to a repartitioning of
high-c liquid and the solid phases is reduced due to the the protein across that phase boundary. Coacervation has
closer match in protein densities. been reported for phosphoglucomutase/polyethylene-glycol
Direct evidence for the L–L separation was obtained mixtures23 and is frequently observed with membrane pro-
from microscopic observations as well as centrifugation of teins when using non-ionic detergents.24
the samples below T cloud. Microscopy revealed that the high Our cloud point data obtained at the three different NaCl
concentration protein phase emerged either in the form of
concentrations are presented as full symbols in T – c plots in
spherical droplets or as bulging tubes intertwined with the
Fig. 3. At the highest c’s and thus high supersaturations,
low-c phase. Both morphologies are characteristic of sepa-
T cloud was exceedingly difficult to determine due to the onset
rating liquid phases.22 Centrifugation of samples at T
of crystallization shortly after sample preparation. Hence, the
,T cloud, followed by reheating above T cloud, yielded two
transparent liquid layers ~see also Ref. 2!. Without centrifu- scarcity of data above c'300 mg/ml. With the broad flat
gation the high concentration liquid droplets tended to trans- maxima of the binodal curves, the descending branch is
form into microcrystals before settling out. barely present. Yet, the above phase separation on centrifu-
In order to detect whether the protein demixing is ac- gation unambiguously established the existence of a de-
companied by salt repartitioning, we determined the Na1 and scending high-c branch. For comparison, we have included
Cl2 concentrations in the two liquid phases forming from a in Fig. 3 the L–L transition data previously obtained at
solution with c5110 mg/ml and 3% NaCl. At this lowest of pH56 and 0.6 M sodium phosphate.2
the salt concentrations used, salt repartitioning associated The dashed lines in Fig. 3 are fits of our three cloud
with protein–salt interactions should result in the largest point data sets to the scaling relation for binary demixing
relative changes in salt levels. However, within the experi- from renormalization-group theory with a critical exponent
mental errors of our measurement techniques ~see Sec. II! of of b 50.32525,26

J. Chem. Phys., Vol. 107, No. 6, 8 August 1997

Downloaded 30 Mar 2013 to 128.233.210.97. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jcp.aip.org/about/rights_and_permissions
1956 M. Muschol and F. Rosenberger: Phase separation in lysozyme solutions

FIG. 5. Phase diagram for lysozyme with 5% NaCl at pH54.5 in 0.1 M


FIG. 4. Phase diagram for lysozyme with 3% NaCl at pH54.5 in 0.1 M
NaAc buffer. Full circles: cloud point data from this work. Dashed line: fit
NaAc buffer. Full triangles: cloud point and solubility data from this work.
to Eq. ~1!. Open circles and diamonds: solubility of tetragonal and ortho-
Dashed line: fit to Eq. ~1!. Open, upward pointing triangles solubility of
rhombic lysozyme, respectively, from Refs. 29 and 30. Solid lines: extrapo-
tetragonal lysozyme from Ref. 21. Open, downward pointing triangles: solu-
lated fits to van’t Hoff equation @Eq. ~2! with c 0 53.4 mg/ml, DH
bility of tetragonal lysozyme from Refs. 29 and 30. Solid line: fit to van’t
526.7 k B T 0 at T 0 5298.15 K#. Points A, B, C mark composition and tem-
Hoff equation @Eq. ~2! with c 0 510.3 mg/ml, DH530.5 k B T 0 at T 0
perature of samples used in crystal morphology study; for details see the
5298.15 K# and extrapolation.
text.

H U
T cloud5T crit 12A
c crit2c p
c crit
U J
1/b
~1!
resent fits to the solubility data and extrapolations to higher
and T crit, c crit, and A as adjustable parameters. Within the
protein concentrations based on an expansion of van’t Hoff’s
large fitting uncertainties, all salt concentrations yielded a
relation around the reference temperature T 0 5298 K
critical concentration c crit'~255630! mg/ml. T crit increased
approximately linearly with salt concentration from well be-
low to well above ambient temperature. Interestingly, the
amplitude A for the 3% NaCl solutions is about twice that for
the higher salt concentrations. Despite the experimental un-
c sat5c 0 exp S DH T sat2T 0
kT 0 T0 D ~2!

certainties associated with the measurements of metastable


states, we feel that this difference lies outside the experimen- with c 0 5c sat(298 K). Values of the fitting parameters c 0 and
tal errors. Note that lysozyme in growth from solution un- DH are given in the respective figure captions. To check the
dergoes a transition from the tetragonal to the orthorhombic validity of this extrapolation, we determined the c sat of two
crystal structure around T'25 °C.27 Hence, it is reasonable high concentration lysozyme samples in 3% NaCl solutions;
to expect some change in the solution structure as well. Re- see the upper full triangles in Fig. 4. Predicted and measured
cent Monte Carlo simulations of protein interactions with values agree reasonably well. The resulting T sat’s lie safely
attractive square well potentials reproduced the experimental below the unfolding temperature for lysozyme of 83 °C ~Eq.
values for c crit in g-crystallin but did not yield the large 4 in Ref. 31!. Again, having the transition from the tetrago-
width of the two-phase liquid coexistence region.28 The au- nal to the orthorhombic crystal structure at higher tempera-
thors suggested spatial anisotropy of the protein interactions tures in mind we have plotted solubility curves for both
as possible cause for the broadness of the experimental data. forms in Figs. 5 and 6. For our ensuing discussion, however,
It will be interesting to see if modifications in such interac- only the generic shape of the solubility curves and their lo-
tion anisotropies are, in turn, responsible for the change in cation with respect to the demixing boundary is of signifi-
amplitude of the phase separation curve and the tetragonal– cance. Most recently, a similar phase diagram was reported
orthorhombic structure transition in lysozyme crystals. for lysozyme at pH57.8 and 0.5 M NaCl.3 The authors com-
To complete the phase diagram for lysozyme, we have mented on the strong correlations of shifts in T cloud and T sat
added in Figs. 4–6 the crystal solubility data for the same pH with salt concentration. Such correlations are equally appar-
and the three salt concentrations.21,29,30 The solid lines rep- ent in our experimental data.

J. Chem. Phys., Vol. 107, No. 6, 8 August 1997

Downloaded 30 Mar 2013 to 128.233.210.97. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jcp.aip.org/about/rights_and_permissions
M. Muschol and F. Rosenberger: Phase separation in lysozyme solutions 1957

FIG. 6. Phase diagram for lysozyme with 7% NaCl at pH54.5 in 0.1 M


NaAc buffer. Full squares: cloud point data from this work. Dashed line: fit
to Eq. ~1!. Open squares and diamonds: solubility of tetragonal and ortho-
rhombic lysozyme, respectively, from Refs. 29 and 30. Solid lines: extrapo-
lated fits to van’t Hoff equation @Eq. ~2! with c 0 51.9 mg/ml, DH
524.3 k B T 0 at T 0 5298.15 K#.
FIG. 7. ‘‘Sea urchin’’ crystal morphology obtained in lysozyme solution
with c520 mg/ml and 7% NaCl at room temperature after prior cooling
B. Relevance for protein crystal growth below T cloud.
As described above, the solution demixing severely im-
pacts crystallization. When a typical crystal growth solution
(c!c crit) is cycled through the two-phase region, crystallites evolve from high-c solutions after short cycling through the
develop within minutes compared to hours required for binodal at temperatures slightly above T crit. Three 5% NaCl
nucleation slightly above T cloud. Most solutions that were left containing solutions, designated A, B, and C, with respec-
in the two-phase liquid region formed white, microcrystal- tive c’s5147, 250, and 330 mg/ml, were, after the cloud
line precipitates. Even brief cycling through the binodal lead point measurements, held at 26 °C. As can be seen from Fig.
to the undesirable formation of a large number of irregular 5, these concentrations were well below, close to and well
crystallites. Both morphologies are unsuitable for x-ray above c crit, respectively. Figure 8 shows low-magnification
structure determinations. images of these samples taken within a few hours after
At low protein concentrations with 5% and 7% salt, cy- preparation and cycling through T cloud. Sample A, depicted
cling resulted in spherulitic crystal morphologies that con- in Fig. 8~a!, yielded predominately tetragonal lysozyme crys-
sisted of thin needles growing radially outward from a nucle- tals, with some granularly structured background ~precipi-
ation center; see Fig. 7. This ‘‘sea urchin’’ morphology, tate! which disappeared within a day. In sample B, that is
which has also been obtained in high-pressure crystal growth close to the critical concentration, collections of beads of
studies of lysozyme,32 can readily be understood in terms of about 200 mm diameter formed instead, presumably from the
the L–L demixing. In low-c solutions, only few submicron- high-c phase; see Fig. 8~b!. Probing with a needle revealed a
sized droplets of the high-c phase appear upon phase sepa- mechanically highly resistant gel nature of these beads. This
ration. Through multiple nucleation within a highly super- morphology persisted for many hours before crystals ap-
saturated droplet the protein is rapidly transformed into a peared. As control, a fraction of the same sample was left
polycrystalline solid. In the surrounding low-c phase, nucle- above T cloud. It showed no gel formation. In sample C, the
ation is much less likely, but since this phase is also super- whole solution gelled with small crystallites becoming ap-
saturated, further growth of the polycrystalline spheres that parent under a microscope with crossed polarizers; Fig. 8~c!.
emerge from the high-c droplets is still supported. Competi- To test the impact on crystallization of passing through
tion for nutrient supply among the crystals with various ori- the two-phase boundary, a control sample with same c as in
entations will select only those favorably oriented to grow A was held above T cloud at all times. This sample, labeled
radially outwards from the polycrystalline center. A2, yielded predominately crystals. While much less pro-
We have also investigated the crystal morphologies that nounced than in sample A, A2 also contained ramified clus-

J. Chem. Phys., Vol. 107, No. 6, 8 August 1997

Downloaded 30 Mar 2013 to 128.233.210.97. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jcp.aip.org/about/rights_and_permissions
1958 M. Muschol and F. Rosenberger: Phase separation in lysozyme solutions

ters; see Fig. 9~a!. Figure 9~b! shows one of the crystals that
grew in A2. Scanning of the focal plane through the crystal
revealed that it had incorporated some ramified clusters. The
cluster formation at c’s below those causing the whole solu-
tion to gel @Figs. 9~a! and 9~b!# is strongly suggestive of a
sol–gel transition in lysozyme.33
In solutions with c>c crit and 7% NaCl, extensive gela-
tion occurred even without cycling through T cloud. Thus gel
formation is not subject to a prior L–L phase separation,
although it appears to be enhanced by passage through the
binodal. These gels can be clearly distinguished from both
the high-c liquid and solid phases: They withstand high shear
stress without showing any macroscopic order. Also, gelled
samples turned translucent or slightly milky compared to the
transparent fluid state and were much more difficult to redis-
solve upon dilution. Furthermore, gels were considerably
more stable towards transformation into the crystalline phase
than the high-concentration liquid droplets that form during a
L–L separation. Highly twinned, irregular protein crystals
emerged from gels within a day, but significant portions of
the sample remained gelled for several days.
The 3% NaCl solutions showed no gelation, neither be-
low nor above room temperature. This suggests that the en-
hanced short-range protein–protein attraction at higher salt
concentration triggers or substantially expands the gelation
region for lysozyme solutions. Alternatively, one might in-
terpret the gelation as a consequence of thermal or salt-
induced denaturation. However, this is not supported by the
finding that lysozyme, obtained from dissolved orthorhombic
crystals grown at pH54.7 and at temperatures as high as
60 °C, was found to posses full biological activity.34
Besides the microscopic inspection, the L–L separation
and gel formation can also be detected from changes in the
speckle patterns obtained with a HeNe laser beam passing
through a sample. For instance, a solution with c
550 mg/ml and 5% NaCl showed no significant forward
scattering before the sample temperature reached T cloud.
Once cooled below this binodal, however, a fluctuating
speckle pattern in the forward direction appeared, and per-
sisted long after rewarming the sample above T cloud. The
time scale t for changes in the speckle intensity is given by
t 51/(Dq 2 ). 35 Here, D is the diffusivity, q
5(4 p n/l)sin(u/2) the magnitude of the scattering vector,
n51.33 the solution’s refractive index, l5633 nm the laser
wave length, and u the scattering angle which in our arrange-
ment was 1°–3°. With an estimated t '1 s we obtain an
upper bound for D<231028 cm2/s. The corresponding
lower bound on the cluster radius can be estimated from the
Stokes–Einstein relation D5kT/6p h r h as r h >100 nm.
From this crude observation, however, we can not distin-
guish between crystal nuclei or gel clusters as scatterers.
For practical purposes, it is interesting to note, that the
FIG. 8. Photomicrographs of morphologies obtained in samples A, B, and C forward scattering intensity in the above observations de-
in Fig. 5 ~c5147, 250 and 330 mg/ml, respectively, 5% NaCl! incubated at pended on the route of mixing the protein and buffer/salt
26 °C after the cloud point measurements. ~a! Sample A: tetragonal solutions. As mentioned in Sec. II, for our measurements
lysozyme crystals coexist with a granular background. ~b! Sample B: spheri-
appropriate amounts of protein were directly dissolved into
cal, isotropic gel beads, several hundred mm in diameter. ~c! Sample C
between crossed polarizers: crystals ~bright spots! that continued to grow the salt/buffer solution. However, when samples of the same
from within this gelled sample. final composition were prepared by mixing buffered stock

J. Chem. Phys., Vol. 107, No. 6, 8 August 1997

Downloaded 30 Mar 2013 to 128.233.210.97. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jcp.aip.org/about/rights_and_permissions
M. Muschol and F. Rosenberger: Phase separation in lysozyme solutions 1959

FIG. 10. Typical ~c, NaCl!-phase diagram for lysozyme. Heavy solid line:
L–L coexistence curve ~based on two values, see solid triangles, for c cloud at
15 °C in 5% and 7% NaCl!. Dashed line: Solubility curve based on data
~open, downward pointing triangles! for tetragonal lysozyme at 15 °C from
Ref. 29. Arrows indicate two different pathways for obtaining a lysozyme
solution with c540 mg/ml and 4% NaCl. Vertical upward arrow: direct
dissolution of protein into solution. Inclined arrows: mixing of protein and
salt stock solutions of twice the final concentrations. Large lens-shaped re-
gion: possible local sample compositions during second mixing process,
with parts of the sample passing through the cross-hatched demixing region
even though the final solution lies outside.

solutions of protein and salt at twice their final concentration,


we observed much stronger forward scattering and much
shorter delay times for crystal nucleation. We tentatively as-
sign this enhanced forward scattering to the transient forma-
tion of gel clusters. As schematically depicted in Fig. 10, the
transient concentration heterogeneities unavoidably associ-
ated with the second mixing path are bound to pass part of
the sample through the L–L phase separation region, thus
enhancing gel cluster formation.
Based on these considerations and the observations pre-
sented in Fig. 8, we suggest that gelation underlies some of
the ‘‘amorphous precipitate’’ formation, which plagues
many protein crystal growth experiments. It should be em-
phasized that lysozyme is atypical for globular proteins in
this respect, since, as our experiments show, amorphous pre-
cipitation hardly occurs at the c’s <70 mg/ml that are com-
monly used for crystal growth. However, dynamic light scat-
tering data on lysozyme aggregation in supersaturated
solutions at 20 °C with 5% NaCl revealed two types of ag-
gregates: micron-size fractal clusters and compact clusters
intermediate in size between the lysozyme monomer and the
fractal clusters.36 Since in these experiments the samples
FIG. 9. Photomicrographs obtained from sample A2 (c5147 mg/ml in 5%
NaCl, incubation at 35 °C overnight, T. above T cloud at all times!. Gel were mixed from concentrated stock solutions, based on Fig.
clusters ~a! coexist with crystals ~b! which appear to have incorporated such 10, we tentatively identify these two populations as gel clus-
clusters. ters and crystal nuclei, respectively.

J. Chem. Phys., Vol. 107, No. 6, 8 August 1997

Downloaded 30 Mar 2013 to 128.233.210.97. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jcp.aip.org/about/rights_and_permissions
1960 M. Muschol and F. Rosenberger: Phase separation in lysozyme solutions

IV. PHASE BEHAVIOR OF OTHER PROTEINS AND


GENERIC PHASE DIAGRAMS

The protein most thoroughly studied for its phase-


separation behavior is g-crystallin. Cloud point data7 and
solubility curves6 were determined for several of its native
variants. Similar to lysozyme, the authors report a two-phase
liquid region which is metastable with respect to crystalliza-
tion. Pronounced shifts in solubility and cloud point curves
were observed among the four variants of g-crystallin stud-
ied. The overall shape and relative positions of both phase
boundaries, however, remained essentially unchanged. In ad-
dition, the c crit for all variants were similar, and quite close to
the value found for lysozyme. No gel formation was re-
ported. FIG. 11. Schematic temperature-density phase diagrams for hard spheres
Recently we have performed screening experiments for ~diameter s! with Yukawa attraction ~range d! after Ref. 12. ~a! Standard
phase diagram with a stable gas–liquid coexistence region below the critical
crystallization conditions in apoferritin/CdSO4 solutions. point ~CP!. The sublimation curve intersects the gas–liquid curve at the
Again, a gel phase appeared which was preceded by sample triple point ~TP!. ~b! Phase diagram for d < s /7. The sublimation curve
turbidity. This phase transition occurred within the meta- moves above the then metastable gas–liquid coexistence region.
stable zone for crystallization.37 Qualitatively similar behav-
ior was observed in our laboratory with concanavalin A,
where crystals grew out of murky, turbid solutions. Further- suspensions.41 The gas phase can then be identified with the
more, ‘‘metastable phase diagrams’’ in the ~pH, c!-plane fluid state of uniform protein dispersion, the gas–liquid co-
have been reported for bovine serum albumin, with the shape existence curve as the liquid–liquid binodal and the sublima-
of the demixing and gelation boundary ~see Fig. 60 in Ref. tion line as the crystal solubility curve. Similarly, the pres-
38! closely resembling our (c,T) diagram for lysozyme. The ence of a sol–gel transition has been anticipated in models
authors noted that these states were metastable, without ex- employing adhesive hard spheres.42,43 The percolation
panding on the origin of the metastability. Another example threshold obtained in these calculations originated along the
for this complex phase behavior is the gelation of sickle cell L–L phase boundary, as suggested by our experimental ob-
hemoglobin HbS. Under the same conditions that promote servations.
the gelation of deoxy HbS, eventually crystals are formed.39 Monte Carlo simulations for g-crystallin and lysozyme
In addition, both a spinodal L–L phase separation line and a provide considerable evidence that the interaction between
gelation line have been reported for the HbS system.40 the macroions in protein solutions is limited to the short
Note that the above examples span proteins of widely range underlying the diagram of Fig. 11~b!.28 Most recently,
different molecular weight, isoelectric point, quaternary attractive Yukawa potentials were used44 to model experi-
structure, and biological function. These observations sug- mental x-ray structure factors of lysozyme and g-crystallin
gest that the features in the metastable phase diagram of and to reproduce the spinodal lines from previous phase
lysozyme are quite common to globular proteins. This can be separation experiments in g-crystallin.9 The range of the at-
understood by considering the prevailing attractive protein tractive interactions thus extracted from the experiments was
interactions under high salt conditions, which underlie the only of the order of one-tenth of the protein diameter. Hence,
observed phase behavior. Figure 11 shows schematic metastability of the L–L coexistence region can be expected
temperature-density phase diagrams based on Monte Carlo in general for systems of ~colloidal! particles with dominat-
simulations for hard spheres with attractive Yukawa ing short-ranged centrosymmetric attraction. Colloidal mod-
potentials.12 In Fig. 11~a!, gas is the stable low-density phase els of protein interactions have also proven successful for
at high temperatures. On lowering the temperature, a phase descriptions of the solution behavior of proteins, including
separation into coexisting gas and liquid phases occurs. At lysozyme’s solubility,14 phase separation behavior,3 and in-
high densities the gas–liquid and liquid–solid coexistence teraction effects on light scattering.16
curves intersect at the triple point. Further increase in density Based on the above experimental and theoretical consid-
leads to complete solidification. This standard phase diagram erations we put forth a generic phase diagram for globular
undergoes pronounced changes @Fig. 11~b!# as the range d of proteins. As depicted in Fig. 12~a!, three zones are identified
the interactions is shortened to about one-seventh of the within the solid–liquid coexistence region. Zone II represent
hard-core diameter s. The triple-point vanishes and the gas– the metastable L–L region and Zone III the gelation region.
liquid coexistence region moves underneath the gas–solid Note that the latter is either very narrow or absent for 3%
coexistence curve, forming a metastable state. An analytical NaCl–lysozyme solutions. Zone I is the region most suitable
theory based on hard spheres with attractive square-well po- for protein crystal growth. Previous static and dynamic light
tentials yielded the same result for d ,0.25s . 13 To relate the scattering16,45 and low angle x-ray scattering46 studies in this
results of Refs. 12 and 13 to protein solutions, we refer to zone showed that aggregation during the induction period for
Mayer and McMillan’s treatment of solutions as gaseous nucleation was insignificant. This is not surprising, since su-

J. Chem. Phys., Vol. 107, No. 6, 8 August 1997

Downloaded 30 Mar 2013 to 128.233.210.97. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jcp.aip.org/about/rights_and_permissions
M. Muschol and F. Rosenberger: Phase separation in lysozyme solutions 1961

with different NaCl concentration. Combining these results


with crystal solubility data for the same solution conditions,
phase diagrams for lysozyme were obtained. Relations be-
tween crystal growth and metastable demixing were ex-
plored. Within the two-phase liquid region, nucleation was
enhanced and microcrystalline precipitates formed. Outside
of this region, a sol–gel transition was found at higher salt
concentrations. Both, the L–L phase separation as well as
gelation appear closely related to two common forms of un-
desirable morphologies frequently encountered in protein
crystallization: cloudy and amorphous precipitate, respec-
FIG. 12. Postulated generic phase diagram for globular proteins. ~a! Normal tively. We associate cloudy precipitates with L–L phase
solubility, ~b! retrograde solubility. Solution conditions below and above the separations. Amorphous precipitate formation, on the other
solubility curve in ~a! and ~b!, respectively, are metastable with respect to
crystallization. Zone II: separation into two metastable immiscible liquid hand, is often ascribed to the rapid aggregation prevailing at
phases ~formation of cloudy precipitate!. Gelation in Zone III. Cross- excessive supersaturations.10 We rather suggest as a possible
hatched area: region of incomplete gelation ~amorphous precipitation!. Op- origin of amorphous precipitate formation the presence of an
timal region for protein crystallization: Zone I outside cross-hatched area. additional sol–gel transition within the solid–liquid coexist-
ence region. The intimate coupling between demixing, gela-
tion and crystallization results in complex kinetics of the
persaturations are relatively modest and interfering second-
crystallization process. As we illustrated with an example on
ary phase transitions are absent. The shaded area in Zone I,
mixing procedures, details of the thermal and preparative
however, indicates the possibility that pregelation clustering
history of a sample can result in differences in the outcome
can affect crystal growth even in this zone. Sample A2 in our
of crystal growth experiments under seemingly equivalent
experiments ~see Fig. 9! displayed such behavior with the
experimental conditions.
simultaneous presence of crystals and ramified gel clusters.
Most significantly, the presence of a metastable two-
The extent of the shaded area is likely to depend on the
phase liquid region and a sol–gel transition within supersatu-
specific protein and sample preparation procedure.
rated lysozyme solutions appear to be intrinsic features for a
The phase diagram in Fig. 12~a! applies to proteins with
large variety of globular proteins. This finding led us to a
normal temperature dependence of the solubility, i.e., with
generic shape for phase diagrams of globular proteins, with a
an increase in c sat with temperature. However, both HbS and
limited region suitable for high-quality crystal growth.
apoferritin have retrograde solubility, i.e., c sat decreases with
Clearly, further experimental and theoretical work is re-
increasing temperature. Interestingly, the reported tempera-
quired to test our hypothesis.
ture dependence of the metastable states in HbS follows re-
versed temperature dependence as well.40 Therefore, as sche-
matically depicted in Fig. 12~b!, the phase diagram of ACKNOWLEDGMENTS
proteins with retrograde solubility appears to be a mirror
image of Fig. 12~a!. It is a pleasure to acknowledge valuable experimental
As our work and the above references suggest, the loca- support and suggestions by B. R. Thomas, and M. Banish
tion of the various phase boundaries can vary considerably who also carried out the atomic absorption spectroscopy
with changes in solution conditions ~lysozyme, bovine serum measurements. We are equally grateful to C. Carter, Jr., A.
albumin! or differences in the surface groups of the protein Chernov, F. Ferrone, K. Murphy, R. Nagel, and J. Wienzek
~hemoglobin and g-crystallin!. Therefore, some proteins for pointing out some of the references cited in this work. L.
could develop metastable states, which are undesirable for Carter has expertly prepared the figures. This research has
controlled crystal growth, at rather low c’s. The overall been supported by NASA ~Grants No. NAG8-1161 and No.
shape of the phase diagram, however, should be retained. NAG8-1168! and the State of Alabama through the Center
Hence, any search for advantageous protein crystallization for Microgravity and Materials Research at the University of
conditions can take advantage of information on precipitate Alabama in Huntsville.
formation and associated solution conditions. To date, this
information has frequently been considered useless. From 1
C. Ishimoto and T. Tanaka, Phys. Rev. Lett. 39, 474 ~1977!.
2
our investigation it appears, however, that such data can pro- V. G. Taratuta, A. Holschbach, G. M. Thurston, D. Blankschtein, and G.
vide valuable guidance for adjustments of pH, temperature or B. Benedek, J. Phys. Chem. 94, 2140 ~1990!.
3
M. L. Broide, T. M. Tomine, and M. D. Sasowsky, Phys. Rev. E 53, 6325
precipitating salt concentration to move the solution condi- ~1996!.
tions towards the ‘‘crystallization slot’’ of zone I in the sche- 4
B. M. Fine, J. Pande, A. Lomakin, O. Ogun, and G. B. Benedek, Phys.
matic phase diagram. Rev. Lett. 74, 198 ~1995!.
5
C. Liu, A. Lomakin, G. M. Thurston, D. Hayden, A. Pande, J. Pande, O.
Ogun, N. Asherie, and G. B. Benedek, J. Phys. Chem. 99, 454 ~1995!.
VI. CONCLUSIONS 6
C. R. Berland, G. M. Thurston, M. Kondo, M. L. Broide, J. Pande, O.
Ogun, and G. B. Benedek, Proc. Natl. Acad. Sci. USA 89, 1214 ~1992!.
We have determined the cloud point curves for the meta- 7
M. L. Broide, C. R. Berland, J. Pande, O. Ogun, and G. B. Benedek, Proc.
stable liquid–liquid phase transition in lysozyme solutions Natl. Acad. Sci. USA 88, 5660 ~1991!.

J. Chem. Phys., Vol. 107, No. 6, 8 August 1997

Downloaded 30 Mar 2013 to 128.233.210.97. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jcp.aip.org/about/rights_and_permissions
1962 M. Muschol and F. Rosenberger: Phase separation in lysozyme solutions

8 25
P. Schurtenberger, R. A. Chamberlin, G. M. Thurston, J. A. Thomson, and J. V. Senger, in Phase Transitions, edited by M. Lévy, J. C. Le Guillou,
G. B. Benedek, Phys. Rev. Lett. 63, 2064 ~1989!. and J. Zinn-Justin ~Plenum, New York, 1980!, p. 95.
9 26
J. A. Thomson, P. Schurtenberger, G. M. Thurston, and G. B. Benedek, E. Stanley, Introduction to Phase Transitions and Critical Phenomena
Proc. Natl. Acad. Sci. USA 84, 7079 ~1987!. ~Oxford, New York, 1971!.
10
A. Ducruix and R. Giegé, in Crystallization of Nucleic Acids and Proteins. 27
J. Jollès and J. Berthou, FEBS Lett. 23, 21 ~1972!.
A Practical Approach, edited by A. Ducruix and R. Giegé ~Oxford, New 28
A. Lomakin, N. Asherie, and G. B. Benedek, J. Chem. Phys. 104, 1646
York, 1992!, Chap. 4. ~1996!.
11
Y. Georgalis, A. Zouni, W. Eberstein, and W. Saenger, J. Cryst. Growth 29
E. Cacioppo and M. L. Pusey, J. Cryst. Growth 114, 286 ~1991!.
126, 245 ~1993!. 30
F. Ewing, E. Forsythe, and M. L. Pusey, Acta Cryst. D 50, 424 ~1994!.
12
M. J. Hagen and D. Frenkel, J. Chem. Phys. 101, 4093 ~1994!. 31
F. P. Schwarz, Thermochim. Acta 147, 71 ~1989!.
13
N. Asherie, A. Lomakin, and G. B. Benedek, Phys. Rev. Lett. 77, 4832 32
B. Lorber, G. Jenner, and R. Giegé, J. Cryst. Growth 158, 103 ~1996!.
~1996!.
14
33
D. Stauffer, A. Coniglio, and M. Adam, Adv. Polym. Sci. 44, 103 ~1982!.
D. Rosenbaum, P. C. Zamora, and C. F. Zukoski, Phys. Rev. Lett. 76, 150 34
J. Berthou and P. Jollès, Biochim. Biophys. Acta 336, 222 ~1974!.
~1996!. 35
B. J. Berne and R. Pecora, Dynamic Light Scattering: With Applications to
15
A. George and W. W. Wilson, Acta Cryst. D 50, 361 ~1994!.
Chemistry, Biology, and Physics ~Wiley, New York, 1976!.
16
M. Muschol and F. Rosenberger, J. Chem. Phys. 103, 10424 ~1995!. 36
17 Y. Georgalis, J. Schüler, J. Frank, M. D. Soumpasis, and W. Saenger,
B. R. Thomas, P. G. Vekilov, and F. Rosenberger, Acta Cryst. D 52, 776
~1996!. Adv. Colloid Interf. Sci. 58, 57 ~1995!.
18
P. G. Vekilov and F. Rosenberger, J. Cryst. Growth 158, 540 ~1996!.
37
M. Muschol ~unpublished!.
19
P. G. Vekilov, L. A. Monaco, B. R. Thomas, V. Stojanoff, and F. Rosen-
38
A. H. Clark and S. B. Ross-Murphy, Adv. Polym Sci. 83, 57 ~1987!.
berger, Acta Cryst. D 52, 785 ~1996!.
39
W. A. Eaton and J. Hofrichter, Adv. Prot. Chem. 40, 63 ~1990!.
20
A. J. Sophianopoulos, C. K. Rhodes, D. N. Holcomb, and K. E. Van
40
P. L. San Biagio and M. U. Palma, Biophys. J. 60, 508 ~1991!.
Holde, J. Biol. Chem. 237, 1107 ~1962!.
41
D. A. McQuarrie, Statistical Mechanics ~Harper and Row, New York,
21
F. Rosenberger, S. B. Howard, J. W. Sowers, and T. A. Nyce, J. Cryst. 1976!, Chap. 15.
Growth 129, 1 ~1993!.
42
Y. C. Chiew and E. D. Glandt, J. Phys. A. 16, 2599 ~1983!.
22
P. Guenoun, P. R. Gastaud, F. Perrot, and D. Beysens, Phys. Rev. A 36,
43
W. G. Kranendonk and D. Frenkel, Mol. Phys. 64, 403 ~1988!.
44
4876 ~1987!. M. Malfois, F. Bonneté, L. Belloni, and A. Tardieu, J. Chem. Phys. 105,
23
W. J. Ray, Jr. and C. E. Bracker, J. Cryst. Growth 76, 562 ~1986!. 3290 ~1996!.
24
F. Reiss-Husson, in Crystallization of Nucleic Acids and Proteins. A Prac- 45
M. Muschol and F. Rosenberger J. Cryst. Growth 167, 738 ~1996!.
tical Approach, edited by A. Ducruix and R. Giége ~Oxford, New York, 46
A. Ducruix, J. P. Guilloteau, M. Ries-Kautt, and A. Tardieu, J. Cryst.
1992!, Chap. 8. Growth 168, 28 ~1996!.

J. Chem. Phys., Vol. 107, No. 6, 8 August 1997

Downloaded 30 Mar 2013 to 128.233.210.97. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://jcp.aip.org/about/rights_and_permissions

You might also like