Subido por Juan Guillermo

Historia de la criopreservacion

Anuncio
The History and Principles of
Cryopreservation
D.E. Pegg, M.D.1
Downloaded by: Collections and Technical Services Department. Copyrighted material.
ABSTRACT
The ability of glycerol to protect cells from freezing injury was discovered accidentally. The subsequent development of cryopreservation techniques has had a huge impact in many fields, most notably in reproductive medicine. Freezing injury has been
shown to have two components, direct damage from the ice crystals and secondary damage caused by the increase in concentration of solutes as progressively more ice is formed.
Intracellular freezing is generally lethal but can be avoided by sufficiently slow cooling,
and under usual conditions solute damage dominates. However, extracellular ice plays a
major role in tissues. Cryoprotectants act primarily by reducing the amount of ice that is
formed at any given subzero temperature. If sufficient cryoprotectant could be introduced, freezing would be avoided altogether and a glassy or vitreous state could be produced, but osmotic and toxic damage caused by the high concentrations of cryoprotectant
that are required then become critical problems. The transport of cryoprotectants into
and out of cells and tissues is sufficiently well understood to make optimization by calculation a practical possibility but direct experiment remains crucial to the development of
other aspects of the cryopreservation process.
KEYWORDS: Cryopreservation, cryoprotectant, freezing
T
he accidental discovery by C. Polge, A.U.
Smith, and A.S. Parkes in 1948 that glycerol would enable fowl spermatozoa to survive freezing to 70°C1 initiated a phase of dramatic development in the techniques we now know as “cryopreservation.” Stories
abound as to precisely how this discovery was made: at
least it is clear that a mistake in the labeling of a bottle of
solution in a refrigerator led to some fowl semen being
frozen in a mixture of glycerol, albumen, and water
rather than the intended solution of levulose. The levulose solution was ineffective but the glycerol solution was
highly effective. At first the researchers were unaware of
the composition of their apparently miraculous solution
but, with the help of an analytical chemist, its true identity was revealed and a few more experiments then
showed that glycerol was the active ingredient.2 A truly
serendipitous discovery, but recall the observation by
Hans Krebs (I think) that “Chance favours the prepared
mind!” Indeed it did, and the subsequent work at the
Medical Research Council’s Mill Hill Institute, particularly by Audrey Smith and a succession of young colleagues, was crucial to establishing this new field of science. It is interesting to note that the title of the paper
describing that fundamental observation was “Revival of
Spermatozoa after Vitrification and Dehydration at
Low Temperatures”; yet these experiments did not in
fact produce vitrification in the sense that is now meant
by that term. Their method would now be termed the
“classical freezing” approach. It is also interesting to note
that much of the subsequent research and perhaps the
greatest practical impact of all cryopreservation work
has been in the areas of reproductive medicine, animal
The Cryobiology of Assisted Reproduction: Gametes and Gonads; Editor in Chief, Bruce R. Carr, M.D.; Guest Editors, S.L. Tan, M.D., Roger
G. Gosden, Ph.D., D.Sc. Seminars in Reproductive Medicine, Volume 20, Number 1, 2002. Address for correspondence and reprint requests:
Professor D.E. Pegg, Medical Cryobiology Unit, Department of Biology, University of York, PO Box 373, York YO10 5YW, United Kingdom.
1Medical Cryobiology Unit, Department of Biology, University of York, York, United Kingdom. Copyright © 2002 by Thieme Medical
Publishers, Inc., 333 Seventh Avenue, New York, NY 10001, USA. Tel: +1(212) 584-4662. 1526-8004,p;2002,20,01,005,014,ftx,en;sre00149x.
5
SEMINARS IN REPRODUCTIVE MEDICINE/VOLUME 20, NUMBER 1 2002
husbandry, and now the conservation of endangered
species. The introduction of the glycerol technique for
freezing bull spermatozoa revolutionized the cattle
breeding industry, as described by Audrey Smith herself
in 19613:
Farmers could (now) select semen from bulls
which might not have been available when required.
Valuable semen was no longer wasted, and semen from
bulls of uncertain value could be stored for several years
until their progeny had matured and been tested for
milk yield and other qualities. Another potential advantage of the method was that, during epidemics of
infections such as foot and mouth disease, there would
be a reserve of semen previously collected which could
be used until all danger of further contagion had
passed. The banks of semen at low temperatures would
permit economy in the number of bulls used for breeding within one country. In addition, the possibility was
opened up of vigorous international trade in frozen
semen to improve stock throughout the world.
This review will consider the mechanisms by
which the classical freezing method seeks to preserve
the viability of cells and tissues and how the more recent
approach of vitrification differs.
Following that initial discovery, nearly all the subsequent developments of the classical freezing method
have relied upon the addition of a cryoprotective compound, such as glycerol, and the empirical optimization
of a number of variables that have been shown experimentally to affect survival. These include the nature and
concentration of the cryoprotectant and the temperature
at which it is added, the rates of cooling and warming,
the storage temperature, and the temperature and rate at
which the cryoprotectant is removed. Despite its largely
empirical nature, this approach was strikingly successful,
and effective methods were soon developed for a wide
range of cells, including the spermatozoa of several
species, erythrocytes, lymphocytes, and hemopoietic
cells, various endocrine cells, and many strains of tissue
culture cell. These early successes are well documented
in A.U. Smith’s 1961 monograph “Biological Effects of
Freezing and Supercooling.”3 These practical successes
stimulated a considerable volume of more fundamental
work that has uncovered a number of the mechanisms
that are involved: the fundamental importance of the
total quantity of ice that is formed, the location of the ice
crystals in relation to the cells, the toxicity of cryoprotectants and the temperature dependence of that toxicity,
and the magnitude of osmotically induced changes in
volume. These factors will be considered in more detail
and then the ability of predictive modeling to specify
preservation methods without recourse to experimentation will be addressed.
DAMAGE BY ICE
When a dilute aqueous solution is frozen, the ice that
forms is essentially pure crystalline water; ice has negligible ability to dissolve solutes. Solutes are therefore rejected and concentrate in the dwindling volume of unfrozen liquid, depressing the chemical potential of water
and achieving equilibrium with ice at each temperature.
Thus, the familiar freezing point depression curve of an
aqueous solution (that for NaCl/water is shown in Fig.
l) also describes the dependence of solution composition upon temperature. In the presence of ice and at
equilibrium, composition is determined solely by temperature and is therefore independent of initial composition; however, initial composition does control the
amount of ice that forms and the factor by which the
concentration increases at a given temperature. It is important to realize that the increase in concentration of
NaCl that occurs during the freezing of isotonic saline
is enormous, actually 32-fold at 21°C.
The question then arises: Is it the ice, the elevated salt concentration, or both that damage cells during progressive freezing? A rather definite answer was
given in the early 1950s4–7 by Jim Lovelock, who was a
colleague of Polge, Smith, and Parkes at Mill Hill.
Lovelock showed that the extent of hemolysis observed
in human erythrocytes that were frozen to various temperatures in solutions of sodium chloride could be accounted for quantitatively by the effect of the salt concentration produced by freezing to each temperature. A
duplication of that experiment in our laboratory is illus-
Figure 1 The phase diagram of sodium chloride/water. From
left to right the curve shows the freezing points of solutions of
sodium chloride of increasing strength. The lowest possible
freezing point is 21.4°C, at which temperature the concentration is almost 24% in w/w terms. If solutions less concentrated
than this are cooled, when they reach the curve, sufficient ice
will crystallize to hold the solution composition on the line. For
concentrations greater than 24% w/w, the curve describes the
dependence upon temperature of the concentration of saturated solutions of sodium chloride.
Downloaded by: Collections and Technical Services Department. Copyrighted material.
6
Figure 2 Graph showing the extent of hemolysis when
human erythrocytes suspended in isotonic saline were frozen
to and thawed from the indicated temperatures (solid line). For
comparison, the broken line shows the hemolysis produced by
exposing the cells to the corresponding concentrations of sodium chloride followed by resuspension in isotonic saline.
trated in Figure 2, and it does indeed show an impressive correspondence between the effect of the two treatments. However, it should be noted that an experiment
of this sort cannot prove causation: I am told that there
is a very close correlation between the production of pig
iron in the United States in the 1980s and the birth rate
in the United Kingdom, yet I doubt if anyone would
argue that either one caused the other! But in this case
a rational causative mechanism was postulated and
the close correlation convinced most workers. Lovelock
went on to explain the cryoprotective action of glycerol and of other highly soluble, nontoxic, penetrating
solutes by their ability to moderate the increase in salt
concentration that occurs during freezing; this is demonstrated for the system NaCl/glycerol/water in Figure 3 and tended to reinforce acceptance of the “saltdamage” theory of freezing injury. In addition, we have
shown that the addition of glycerol has the same influence on the effect of exposure to high concentrations of
salt as it does on the effect of salt concentration by
freezing,8 and we have argued that this makes it even
more probable that the salt-damage theory is valid.
However, the fundamental logical position is unchanged and Mazur has pointed out9 that there is a fixed
relationship between salt concentration and the amount
of ice formed at any given subzero temperature when solutions of a given cryoprotectant in isotonic saline are
frozen: although cryoprotectants actually moderate the
Figure 3 The effect of glycerol on the increase in solute concentration during progressive freezing. Each solution contained
isotonic saline plus the initial concentration of glycerol that is
shown against each curve in mole fraction terms. The corresponding molar concentrations are 0 and approximately 1, 2, 3,
and 4 mol/L. The concentration factor for each solution at each
temperature applies both to glycerol and sodium chloride.
rise in salt concentration by reducing the amount of ice,
the existence of the phenomenon of cryoprotection does
not support either the salt- or the ice-damage theory.
Further, Mazur has provided extensive experimental evidence that ice may indeed have a direct damaging action.9–12 He did this by freezing erythrocytes in solutions
of glycerol and sodium chloride in which the initial concentration of salt varied between 0.6 X and 4.O X isotonic, which made it possible to separate, to a considerable extent, the increase in salt concentration from the
quantity of ice that formed. Mazur found a stronger correlation between severe damage and the amount of ice
rather than the salt concentration.
One problem with this sort of experiment is that
the means used to separate the variables (different initial tonicities) may also affect the behavior of the cells:
cells destined to be subjected to low “unfrozen fractions” enter the experiment in a swollen state, and we
have suggested that such cells may behave differently
from normal cells.13 In fact, we were able, with erythrocytes suspended in solutions of NaCl/glycerol/water, to
demonstrate remarkably similar responses to freezing
and thawing as to exposure to equivalent concentrations
of solute and redilution.8 When similar experiments
were carried out in the absence of glycerol, the correspondence between the effects of freezing and solution
exposure was good for cells in isotonic and 2 X isotonic
saline, but freezing was substantially more damaging
than solution exposure when the cells were suspended
in 0.6 X or 4 X saline.14 The mechanism of this effect is
unclear, but it is consistent with the proposition that
7
Downloaded by: Collections and Technical Services Department. Copyrighted material.
THE HISTORY AND PRINCIPLES OF CRYOPRESERVATION/PEGG
SEMINARS IN REPRODUCTIVE MEDICINE/VOLUME 20, NUMBER 1 2002
cells suspended in a range of tonicities of saline do not
behave as a uniform population. The matter remains
unresolved; the general view is that ice probably has no
direct role in slow freezing injury to erythrocytes, but
that possibility cannot be excluded.
In the foregoing discussion we have ignored the
fact that real biological systems are compartmentalized,
the intracellular spaces being separated from the extracellular space by semipermiable cell membranes. Therefore it is important to consider whether ice forms inside
or outside the cells or both. Freezing is a nucleationinduced event, and the probability of nucleation increases directly with the degree of supercooling and
with the volume. Thus, as our compartmentalized model
cools below the equilibrium freezing point of the saline,
it is inevitable that a nucleation event will occur in the
single, large extracellular compartment before very
many of the cells have nucleated. Once that has happened, further cooling will result in growth of that extracellular ice, concentration of the extracellular solution, and dehydration of the cells by osmosis through
their semipermeable membranes. Thus, providing the
cooling rate is sufficiently low and the water permeability of the cells sufficiently high, the intracellular spaces
will remain free of ice. This fundamental phenomenon
was discovered and elucidated by Mazur15 at Oak Ridge
National Laboratory. He provided a quantitative analysis of the effect of cooling rate on water transport during progressive cooling and correlated the predicted extent of intracellular supercooling with the known effect
of cooling rate on cell survival: the greater the degree of
supercooling, the greater the probability of freezing. It
was shown that cooling rates that produced significant
supercooling with each cell also caused the survival to
fall. The phenomenon is best illustrated by Figure 4,
compiled by Leibo,16 which shows, for three types of
cell with differing water permeabilities, an inverse correlation between intracellular freezing and survival. This
is the basis of the so-called “two-factor hypothesis” of
freezing injury: solute damage at low cooling rates
where extracellular ice is probably innocuous to cells in
suspension; intracellular freezing at high cooling rates,
which is generally lethal. Each cell/cryoprotectant combination is associated with an optimum cooling rate.
This discussion has dealt with a simplified model
of cells in suspension and the simplest of all mammalian
cells (the erythrocyte), together with a few examples of
more typical mammalian cells in suspension. But many
of us would like to be able to cryopreserve more complex systems—multicellular, organized tissues and even
organs. Several factors conspire to make this a vastly
more difficult task. Most obviously, tissues consist of
many types of cells that may differ in their requirements
for optimal preservation; the dimensions of the tissue or
organ will impose restrictions on the rates of cooling
and heating that can be used; and above all, the function
of an organ or tissue depends upon the interrelations
and interconnections between the cells and the preservation of an adequate three-dimensional arrangement
with intact nonliving intercellular structures such as
basement membranes, glycosaminoglycans, and collagen fibers. The point to be emphasized is that ice forming outside the cells may nevertheless be within the system we wish to preserve and may therefore produce
lethal injury. In fact, we have concluded that damage
due to extracellular ice is the single most serious obstacle to the extension of cryopreservation techniques to
structured multicellular systems.17,18 In the case of vascularized tissues we have obtained both experimental
and theoretical evidence for the crucial role of intravas-
Figure 4 The effect of cooling rate on the percentage of cells that were seen to freeze internally (solid line) and the percentage
that survived freezing and thawing (broken line). Mouse ova were frozen in 1 M DMSO, HeLa cells in tissue culture medium, and
erythrocytes in 1.5 M glycerol.16
Downloaded by: Collections and Technical Services Department. Copyrighted material.
8
cular freezing: rabbit kidneys, frozen after equilibration
with 2 M glycerol, were found to have ruptured glomerular capillaries19; mathematical modeling of the process of freezing in a Krogh cylinder model of vascularized tissue showed that the capillaries would have to
expand eight-fold to accommodate the required volume
of ice, which is far beyond their elastic limit.20
We conclude that ice is exceedingly damaging to
tissues, through a number of mechanisms, even in the
presence of concentrations of cryoprotectant that would
ordinarily be expected to provide a high degree of cell
protection. The route to effective cryopreservation of
such systems would therefore seem to be via ice-free
cooling, or vitrification. The term vitrification may be
defined for aqueous systems as the conversion of the system from a fluid to a solid solely by an increase in viscosity, without any crystallization of water and therefore in
the complete absence of ice. Before considering vitrification in more detail we must consider two other factors
that are important in conventional freeze-preservation,
but which acquire even greater importance in vitrification where the concentrations of added solute are much
greater: the chemical toxicity of the additives and the osmotic consequences of their addition and removal.
THE TOXICITY OF CRYOPROTECTANTS
Cryoprotectants are of necessity tolerated in high concentrations; glycerol is, in this sense, far less harmful
than NaCl, but at a sufficiently high concentration any
compound will be “toxic.” This has two important consequences: the highest concentration that the tissue will
tolerate prior to preservation is limited, and, during freezing, the concentration will increase as ice separates. In vitrification, as opposed to freezing, a much higher initial
concentration is needed but no further concentration occurs during cooling because freezing does not occur. In
both techniques one seeks the highest tolerable concentration, in freezing to reduce the salt concentration and in
vitrification to achieve the vitreous state without freezing.
In practice it is found that the maximum concentration
that can be achieved without impairment of viability is
dependent on the temperature and rate of addition and
removal; temperature dependence is due partly to chemical toxicity (which is reduced by reduction in temperature)
and partly to osmotic effects (which are increased by reduction in temperature).
Consider chemical toxicity first. Cryoprotectants
are frequently added at 0 to 4°C rather than at room
temperature to take advantage of the positive temperature coefficient of chemical toxicity, and this has assumed a far greater importance in attempts at vitrification. Thus, for the vitrification of mouse embryos, Rall
and Fahy used a two-stage incubation with their cryoprotectant mixture, first at 20°C with one quarter of
the final concentration and then at 4°C with the final
concentration.21 Much earlier than this, Elford and
Walter22 increased the concentration of dimethyl sulphoxide in smooth muscle to a final level of 50% w/v by
a series of step increases starting at 37°C (to 20%), then
after cooling to 7°C (to 30%), then at 14°C (to
40%), at 22°C (to 50%), and finally at 39°C (to
60%), following which the muscles were cooled to
79°C. Reversal of this stepwise process during warming permitted full recovery of contractile function, and
no ice was formed at 79°C. Thus, in this system, a
very high concentration of cryoprotectant was tolerated
when the temperature and rate of addition and removal
were appropriately optimized. Empirical experiments in
our laboratory have shown that rabbit kidneys will tolerate up to 4 M glycerol,23 and rabbit corneas will recover after exposure to 4.25 M dimethyl sulphoxide.24
Unfortunately, the concentrations of cryoprotectant needed to vitrify aqueous systems are usually in excess of 5.5 M and cannot be achieved without severe
toxicity. With some cryoprotectants, even using conventional freezing methods of cryopreservation, it seems
that toxic levels are reached. This was first pointed out
by Lovelock in his study of the cryopreservation of
erythrocytes with methanol, which, irrespective of the
initial concentration, caused recovery to fall to negligible levels when the temperature fell below 55°C.
With some systems a similar, if less severe, phenomenon
occurs with propan-1,2-diol (propylene glycol); for example with human platelets25 and rabbit kidneys,26 but
other systems, including rabbit cornea27 and human
embryos,28 seem to tolerate this cryoprotectant more
readily. It seems very likely that all cryoprotectants produce some damage at the concentrations used in practice. Certainly this is true of glycerol and human erythrocytes,8 but where the additive is effective this toxic
action is clearly less than the cryoprotective effect. The
crucial points are that there always is a toxic limit to the
concentration of cryoprotectant that can be used and
that the apparent toxic limit is strongly influenced by
the conditions under which the cryoprotectant is added
and removed.
OSMOTIC EFFECTS OF
CRYOPROTECTANTS
The most effective cryoprotectants penetrate cell membranes but they do so more slowly than water, which
means that some osmotic imbalance is inevitable during
the addition or removal of these compounds. Gross osmotic shock results in cell damage, ultimately in lysis,
and it is therefore both logical and of proven efficacy to
control changes in cell volume so that acceptable limits
are not transgressed. An important step in designing
cryopreservation methods is therefore to measure the
9
Downloaded by: Collections and Technical Services Department. Copyrighted material.
THE HISTORY AND PRINCIPLES OF CRYOPRESERVATION/PEGG
SEMINARS IN REPRODUCTIVE MEDICINE/VOLUME 20, NUMBER 1 2002
volume response to changes in external osmolality of a
nonpermeating solute and to correlate each volume
with subsequent structural and functional damage. The
next step is to measure or, failing that, to estimate the
water and solute permeability for the chosen cryoprotectant. Two stratagems for controlling volume during
the addition and removal of penetrating solutes are
available: the first is to use low rates of change in concentration or, what amounts to the same thing, to
change the concentration in several small steps; the
other approach, which is applicable only to the more
critical removal phase, is to incorporate solutes that do
not penetrate the cells and therefore function as “osmotic buffers” by restricting the inflow of water as
the concentration of cryoprotectant is reduced. These
highly effective procedures are illustrated for the addition and removal of dimethyl sulphoxide with human
oocytes29 in Figure 5. Such processes are even more important when concentrations of penetrating agents sufficient to achieve vitrification are used.
VITRIFICATION OF CELLS AND TISSUES
Vitrification occurs when the viscosity of the solution
reaches a sufficient value to inhibit the crystallization of
ice. Although the concentration of solute in the remaining liquid phase increases during progressive freezing, a
temperature will eventually be reached with many systems at which that residual liquid vitrifies in the presence of ice. It has already been pointed out that under
the cooling conditions that are usually used, ice does not
form inside the cells, and in particular because intracellular protein promotes vitrification, it is actually the case
that the cells in conventionally frozen material are vitrified. Polge, Smith, and Parkes were not, after all, in
error in the title of their paper,1 although the presence
of extracellular ice would render the term vitrification
inappropriate in today’s usage. Had it been possible to
reach the concentration of glycerol required to vitrify
before cooling was initiated, true vitrification of the
whole system would have occurred, but this would require about 80% w/w, far beyond the toxic limit at 0°C.
There are, however, more favorable solutes than
glycerol, perhaps the most encouraging recent additions
to the list being propan-1,2-diol and butan-2,3-diol,30
which will vitrify at a concentration less than 50%. This
approach may be termed the equilibrium approach to vitrification: the system will vitrify no matter how slowly it
is cooled. There is, however, another approach that is
best illustrated by Figure 6: the first stage in the process
of freezing is nucleation, and nucleation has an unusual
temperature dependence in that it becomes more active
rather than less active with reduction of temperature,
until it is limited by viscosity. However, the growth of
ice crystals has a more usual temperature dependence, in
that it is slowed and eventually arrested by cooling. The
interesting point is the manner in which these two processes interact: the rate of ice crystal growth in the solution illustrated in Figure 6 has reached low values before
the temperature zone for active nucleation is entered.
Consequently three possibilities exist: if cooled sufficiently rapidly the sample may escape both nucleation
and ice crystal growth; if cooling is more rapid the sample may be nucleated but without ice crystals; or, with
slower cooling, the sample may nucleate and ice then
may form. Critical cooling rates (that avoid the crystal-
Figure 5 Calculated changes in cell volume for mature human oocytes as they are exposed in sequence to 0.8, 1.5, 0.64, and 0 M
DMSO in isotonic media. This scheme will restrict the osmotically induced excursions in cell volume to 30%, during both the addition and the removal of DMSO.
Downloaded by: Collections and Technical Services Department. Copyrighted material.
10
Figure 6 A diagram to illustrate the dependence of the nucleation of ice and the growth of ice crystals on time and temperature.
The material is a thin film of 50% polyvinylpyrrolidone. The lines with arrows indicate cooling rates that will avoid nucleation
(300°C/s), will allow nucleation but avoid ice crystal growth (80°C/s), or will allow ice crystals both to form and to grow (20°C/s).
lization of ice) for a wide range of solutes have been
published by Sutton.31–33
Figure 6 carries another danger message and this
concerns warming: as the temperature is raised the sample traverses the nucleation zone first and then the zone
of ice crystal growth, so the stage is set for freezing to
occur during warming, even if it was avoided during
cooling. The obvious remedy for this situation is ultrarapid heating, and irradiation with microwaves is the
obvious technique to use because of its potential for depositing energy uniformly as well as rapidly, even in
bulky samples. However, the critical warming rates are
orders of magnitude greater than the corresponding
critical cooling rates. Considerable effort is now being
devoted to the definition of the problem of microwave
heating of appropriate systems from very low temperatures34,35 and to its solution. Other maneuvers that may
help include the use of extremely high hydrostatic pressures to depress both the freezing and the nucleation
temperature36 and the addition of materials that retard
the growth of ice crystals, such as the so-called “antifreeze” proteins found in some fishes and insects37,38
and certain synthetic so-called “ice-blockers.”39
THE ROLE OF MATHEMATICAL MODELING
Following the mathematical analysis by Mazur15 in
which he showed how intracellular freezing was controlled by the cooling rate and the permeability of the
cell membrane to water, there has been steady progress
in the modeling of cryopreservation in a variety of cells
and even in tissues. Such studies have been very useful as
a means of exploring mechanisms and testing ideas that
can then be subjected to experimental investigation. To
be useful as predictors, such models must be physically
realistic and be supplied with reliable numerical data:
this cannot always be done at the present time. The
most secure use of models is in predicting safe methods
for the addition and removal of cryoprotectants with
isolated cells at a fixed temperatures above 0°C. Here
one needs some standard chemical data on the cryoprotective solute and knowledge of the temperature, the size
of the cells, their water content, the time course of their
volume response to known concentrations of the cryoprotectant, and an impermeant solute and the limits
of their osmotic tolerance to that impermeant solute.
These data permit the calculation by means of standard
solute and solvent flux equations40 of osmotically active
volume, hydraulic conductivity, and solute permeability.
It is then possible to use the derived data to predict volume time courses for other regimes of addition and removal: this has been both useful and reliable.25,41
However, to predict the fluxes that occur during
cooling and their potential effects requires knowledge of
the temperature dependence, not only of the permeability coefficients but also of the effective osmotic volume
and the susceptibility of the cells to changes in volume.
Extrapolation of permeability constants to subzero temperatures from measurements made at suprazero temperatures seems hazardous, and data concerning the
11
Downloaded by: Collections and Technical Services Department. Copyrighted material.
THE HISTORY AND PRINCIPLES OF CRYOPRESERVATION/PEGG
SEMINARS IN REPRODUCTIVE MEDICINE/VOLUME 20, NUMBER 1 2002
other temperature coefficients is simply absent. In this
author’s opinion, there is no alternative at this time to
direct experimentation for the determination of optimal
cooling rates. The situation concerning rewarming is
even less satisfactory; the mechanisms are not even well
understood qualitatively, let alone quantitatively. Measurements of solute permeation into tissues have also
been published and are useful in guiding the design of
experimental protocols, but with less precision than
with cells.42,43
11.
12.
13.
14.
CONCLUSIONS
Cryopreservation does have a rational, scientific basis
but there are still many unanswered questions and some
cells and tissues remain refractory. Vitrification remains
an enticing vision but the problems of vitrifying organized systems remain formidable. Yet it is toward vitrification that the signposts for preservation of these refractory systems clearly point. At this time most if not
all of the systems that have been preserved by vitrification (such as early embryos, monocytes, pancreatic
islets, and blood vessels [see Refs. 21 and 44]) can also
be preserved by conventional freeze-preservation methods. The problem is to identify solutes that can be included in acceptable concentrations and that will vitrify
at realizable rates of cooling and warming, yet will also
be compatible with viability.
15.
16.
17.
18.
19.
20.
REFERENCES
1. Polge C, Smith AU, Parkes AS. Revival of spermatozoa after
vitrification and dehydration at low temperatures. Nature
(Lond) 1949;164:666
2. Parkes AS. Preservation of living cells and tissues at low temperatures. Proc III Internat Congress Animal Reproduction,
Cambridge, 1956:69
3. Smith AU. Biological Effects of Freezing and Supercooling.
London: Edward Arnold; 1961
4. Lovelock JE. The haemolysis of human red blood cells by
freezing and thawing. Biochim Biophys Acta (Amst) 1953;
10:414–426
5. Lovelock JE. The mechanism of the protective action of glycerol against haemolysis by freezing and thawing. Biochim
Biophys Acta (Amst) 1953;11:28–36
6. Lovelock JE. The protective action by natural solutes against
haemolysis by freezing and thawing. Biochem J 1954;56:
265–270
7. Lovelock JE. Haemolysis by thermal shock. Brit J Haemat
1955;1:17–29
8. Pegg DE, Diaper MP. On the mechanism of injury to slowly
frozen erythrocytes. Biophys J 1988;54:471–488
9. Mazur P, Rall WF, Rigopoulos N. The relative contributions
of the fraction of unfrozen water and of salt concentration to
the survival of slowly frozen human erythrocytes. Biophys J
1981;36:653–675
10. Mazur P, Rigopoulos N. Contributions of unfrozen fraction
and of salt concentration to the survival of slowly frozen
21.
22.
23.
24.
25.
26.
27.
28.
29.
human erythrocytes: influence of warming rate. Cryobiology
1983;20:274–289
Mazur P, Cole KW. Influence of cell concentration on the
contribution of unfrozen fraction and salt concentration to
the survival of slowly frozen human erythrocytes. Cryobiology 1985;22:505–536
Mazur P, Cole KW. Roles of unfrozen fraction, salt concentration and changes in cell volume in the survival of frozen
human erythrocytes. Cryobiology 1989;26:1–29
Pegg DE, Diaper MP. The “unfrozen fraction” hypothesis of
freezing injury to human erythrocytes: a critical examination
of the evidence. Cryobiology 1989;26:30–43
Pegg DE, Diaper MP. The effect of initial tonicity on freezethaw injury to human red cells suspended in solutions of sodium chloride. Cryobiology 1989;26:579–560
Mazur P. Kinetics of water loss from cells at subzero temperatures and the likelihood of intracellular freezing. J Gen Physiol 1963;47:347–369
Leibo SP. Preservation of mammalian cells and embryos by
freezing. In: Simatos D, Strong DM, Turk M, eds. Les Colloques de I’Institut National de la Sante et de la Recherche
Medicale: 62. Cryoimmunologie. Paris: Inserm; 1977:
311–334
Pegg DE, Jacobsen IA. Current status of cryopreservation of
whole organs with particular reference to the kidney. In: Marberger M, Dreikorn K, eds. Renal Preservation. Baltimore:
Williams and Wilkins; 1983:301–322
Pegg DE. Ice crystals in tissues and organs. In: Pegg DE,
Karow AM Jr, eds. The Biophysics of Organ Cryopreservation. New York and London: Plenum Press; 1987:117–140
Jacobsen IA, Pegg DE, Starklint H, et al. Effect of cooling
and warming rate on glycerolised rabbit kidneys. Cryobiology
1984;21:637–653
Rubinsky B, Pegg DE. A mathematical model for the freezing process in biological tissue. Proc R Soc Lond B Biol Sci
1988;234:343–358
Rall WF, Fahy GM. Ice-free cryopreservation of mouse embryos at 196°C by vitrification. Nature (Lond) 1985;313:
573–575
Elford BC, Walter CA. Effects of electrolyte composition and
pH on the structure and function of smooth muscle cooled to
79°C in unfrozen media. Cryobiology 1972;9:82–100
Jacobsen IA, Pegg, DE, Wusteman MC, Robinson SM.
Transplantation of rabbit kidneys perfused with glycerol solutions at 10°C. Cryobiology 1978;15:18–26
Taylor MJ. Equilibrium approach to ice-free cryopreservation
of corneas: tolerance of endothelium to high concentrations
of dimethyl sulphoxide. Cryobiology 1988;25:533
Arnaud FG, Pegg DE. Cryopreservation of human platelets
with propane-1,2diol. Cryobiology 1990;27:130–137
Jacobsen IA, Pegg DE, Starklint H, Hunt CJ, Diaper MP. Introduction and removal of cryoprotectants with rabbit kidneys: assessment by transplantation. Cryobiology 1988;25:
285–299
Rich SJ, Armitage WJ. Propane-1,2-diol as a potential component of a vitrification solution for corneas. Cryobiology
1990;27:42–54
Testart J. High pregnancy rate after early embryo freezing.
Fertil Steril 1986;46:268
Newton H, Pegg DE, Barrass R, Gosden RG. Osmotically
inactive volume, hydraulic conductivity and permeability to
dimethyl sulphoxide of human mature oocytes. J Reprod Fertil 1999;117:27–33
Downloaded by: Collections and Technical Services Department. Copyrighted material.
12
30. Boutron P. Non-equilibrium formation of ice in aqueous solutions: efficacy of polyalcohol solutions for vitrification. In: Pegg
DE, Karow AM Jr, eds. The Biophysics of Organ Cryopreservation. New York and London: Plenum Press; 1987:201–236
31. Sutton RL. Critical cooling rates to avoid ice crystallization
in solutions of cryoprotective agents. J Chem Soc Faraday
Trans 1991;87:101–105
32. Sutton RL. Critical cooling rates to avoid ice crystallization
in aqueous cryoprotectant solutions containing polymers. J
Chem Soc Faraday Trans 1991;87:3747–3751
33. Sutton RL. Critical cooling rates for aqueous cryoprotectants
in the presence of sugars and polysaccharides. Cryobiology
1992;29:585–598
34. Marsland TP, Evans S, Pegg DE. Dielectric measurements
for the design of an electromagnetic rewarming system. Cryobiology 1981;24:311–323
35. Robinson MP, Pegg DE. Rapid electromagnetic warming of
cells and tissues. IEEE Trans Biomed Eng 1999;46:1413–1425
36. Fahy GN, MacFarlane DR, Angell CA, Meryman HT. Vitrification as an approach to cryopreservation. Cryobiology
1984;21:407–426
37. De Vries AL. Biological antifreeze agents. Am Rev Physiol
1983;45:245
38. Sutton RL, Pegg DE. Devitrification in butane-2,3-diol solutions containing anti-freeze peptide. Cryo-Letters 1993;14:
13–20
39. Wowk B, Darwin M, Harris SB, Russell SR, Rasch CM. Effects of solute methoxylation on glass-forming ability and
stability of vitrification solutions. Cryobiology 1999;39:
215–227
40. Kleinhans FW. Membrane permeability modelling; KedemKatchalsky vs a two-parameter formalism. Cryobiology 1998;
37:271–289
41. Pfaff RT, Agca Y, Liu J, Woods EJ, Peter AT, Critser JK.
Cryobiology of rat embryos 1: determination of zygote membrane permeability coefficients for water and cryoprotectants, their activation energies and the development of
improved cryopreservation methods. Biol Reprod 2000;63:
1294–1302
42. Bateson EAJ, Busza AL, Pegg DE, Taylor MJ. Permeation of
rabbit common carotid arteries with dimethyl sulphoxide.
Cryobiology 1994;31:393–397
43. Newton H, Fisher J, Arnold JRP, Pegg DE, Faddy MJ, Gosden RG. Permeation of human ovarian tissue with cryoprotective agents in preparation for cryopreservation. Hum Reprod 1998;13:101–105
44. Song YC, Khirabadi BS, Lightfoot FG, Brockbank KGM,
Taylor MJ. Vitreous cryopreservation maintains the function
of vascular grafts. Nat Biotechnol 2000;18:296–299
13
Downloaded by: Collections and Technical Services Department. Copyrighted material.
THE HISTORY AND PRINCIPLES OF CRYOPRESERVATION/PEGG
Descargar