A cryoprotectant is a chemical substance used to protect biological tissues, cells, and organs from damage caused by ice crystal formation during cryopreservation, enabling storage at cryogenic temperatures while preserving functionality upon thawing.[1] These agents work by lowering the freezing point of water, increasing solution viscosity, and stabilizing cellular structures through mechanisms such as vitrification, where the solution transitions to a glass-like state without ice crystals.[2] Cryoprotectants are essential in fields like medicine, biotechnology, and reproductive science, where they facilitate the long-term preservation of sperm, embryos, stem cells, and tissues for transplantation or research.[3]The primary mechanisms of cryoprotectants involve colligative properties that reduce intracellular ice formation by dehydrating cells and moderating electrolyte concentrations, as well as direct stabilization of biomembranes via hydrogen bonding and water replacement.[1] Penetrating cryoprotectants, such as dimethyl sulfoxide (DMSO) and glycerol, diffuse across cell membranes to protect intracellular compartments, while non-penetrating types like sucrose and polyvinylpyrrolidone (PVP) act extracellularly to inhibit ice recrystallization and osmotic stress.[2] Natural cryoprotectants, including antifreeze proteins from cold-adapted organisms and sugars like trehalose, mimic these effects by binding to ice surfaces or stabilizing lipid bilayers, often reducing oxidative damage from reactive oxygen species during freeze-thaw cycles.[3]Common applications include cryopreservation of hematopoietic stem cells for cancer therapy, where combinations of 5-15% DMSO with sugars improve post-thaw viability to over 80%, and vitrification techniques for oocytes and embryos in assisted reproduction, minimizing chilling injury.[1] However, challenges persist due to potential toxicity at high concentrations, such as DMSO-induced cytotoxicity affecting membraneintegrity, prompting research into less toxic natural alternatives like melatonin and deep eutectic solvents for enhanced cytoprotection.[2] Ongoing advancements focus on optimizing CPA formulations to balance efficacy and safety across diverse biological systems.[3]
Definition and Overview
Definition
Cryoprotectants are chemical compounds or mixtures added to aqueous solutions to protect biological structures, such as cells, tissues, and organs, from damage caused by ice formation during freezing processes in cryopreservation.[4] These agents are essential for enabling the long-term storage of viable biological materials at ultra-low temperatures, typically below -130°C, by mitigating the physical and chemical stresses associated with phase transitions of water.[5]The core functions of cryoprotectants include minimizing intracellular ice crystal growth, which can puncture cell membranes and disrupt cellular architecture; reducing the effects of solute concentration increases during freezing, which otherwise lead to osmotic dehydration and toxicity; and stabilizing biomolecular structures like proteins and lipid membranes through interactions such as hydrogen bonding with water molecules.[4] By performing these roles, cryoprotectants help preserve cellular integrity and functionality upon thawing.[6]Cryoprotectants are broadly classified into two categories based on their ability to penetrate cell membranes: permeable cryoprotectants, which cross cellular barriers to act intracellularly (e.g., glycerol and dimethyl sulfoxide), and non-permeable cryoprotectants, which remain extracellular and primarily influence the surrounding environment (e.g., sugars like sucrose and polymers like polyethylene glycol).[5] This classification guides their application in cryopreservation protocols to balance protection and potential toxicity.[4]A key physical principle underlying their efficacy is the colligative property of lowering the freezing point of solutions to delay ice nucleation while elevating the glass transition temperature (Tg), which promotes the formation of a stable, amorphous glass state instead of crystalline ice during rapid cooling—a process known as vitrification.[6]
Historical Development
The concept of cryoprotection emerged from early 20th-century investigations into the effects of low temperatures on biological materials. In the 1930s, Basile J. Luyet and Paul M. Gehenio conducted pioneering studies on vitrification, demonstrating that rapid cooling could transform water in plant tissues into a glass-like state without ice crystal formation, potentially preserving cellular integrity.[7] Their work, detailed in publications such as Life and Death at Low Temperatures (1940), highlighted the detrimental role of intracellular ice and laid foundational ideas for avoiding freezing damage through non-crystalline solidification.[8]A major breakthrough occurred in 1949 when Christopher Polge, Audrey U. Smith, and Alan S. Parkes accidentally discovered glycerol's protective effects during experiments on fowl spermatozoa. By adding glycerol to samples before freezing and thawing, they achieved revival rates exceeding 50% in multiple species, marking the first reliable use of a permeable cryoprotectant to mitigate ice-induced cell death. This serendipitous finding, stemming from contamination in a lab batch, revolutionized gamete preservation and spurred broader applications in cell banking.[9]The 1950s and 1960s saw expansion with the introduction of dimethyl sulfoxide (DMSO) by James E. Lovelock and Mervyn W. H. Bishop in 1959, who demonstrated its efficacy in protecting human and bovine red blood cells from freezing injury by reducing electrolyte concentration effects.[10] Unlike glycerol, DMSO's high permeability allowed rapid penetration into cells impermeable to larger molecules, enabling survival rates of up to 80% post-thaw in erythrocytes.[11] Concurrently, slow-freezing protocols were refined, optimizing cooling rates to 1°C per minute to minimize intracellular ice while leveraging cryoprotectants for extracellular stabilization.[7]From the 1980s onward, vitrification techniques gained prominence, with William F. Rall and Gregory M. Fahy reporting in 1985 the successful ice-free cryopreservation of mouseembryos at -196°C using high-concentration mixtures of permeable cryoprotectants like DMSO, acetamide, and propylene glycol.[12] This approach achieved over 90% embryo survival by inducing a glassy state, significantly reducing cryoprotectant toxicity through minimized exposure times compared to slow freezing.[13] In the 2000s, efforts shifted toward organ cryopreservation, where challenges like cryoprotectant toxicity, osmotic stress, and uneven distribution persisted, limiting success to small tissues despite advances in formulation screening.[14] By the 2020s, research emphasized bio-based alternatives, such as trehalose and antifreeze proteins from natural sources, which improved cell viability in oocytes and sperm by inhibiting ice recrystallization with reduced toxicity—e.g., L-proline oligomers boosting mouseoocyte survival to 99% alongside low-dose DMSO.[15]
Mechanisms of Action
Vitrification and Glass Transition
Vitrification refers to the process of rapidly cooling a solution to form a stable amorphous solid, or glass, rather than crystalline ice, thereby avoiding the mechanical damage caused by ice crystal formation during cryopreservation.[16] This biophysical mechanism is central to the protective action of cryoprotectants, which enable the transition to a vitreous state at practical cooling rates, preserving the structural integrity of biological samples.[12] By suppressing nucleation and growth of ice, vitrification maintains the sample in a non-crystalline, supercooled liquid-like state that solidifies without phase separation.[17]The glass transition temperature (Tg) is the critical point below which the supercooled liquid acquires the properties of a viscous, glassy solid, with dramatically reduced molecular mobility that inhibits ice formation.[18] Cryoprotectants elevate the Tg of aqueous solutions, allowing stabilization at temperatures above the homogeneous nucleation temperature of pure water, thus facilitating vitrification under controlled cooling conditions.[16] For instance, permeable cryoprotectants like DMSO increase Tg by disrupting water's hydrogen bonding network and forming a more rigid matrix.[19]The Gordon-Taylor equation is commonly used to predict Tg in binary mixtures of cryoprotectants and water:T_g = \frac{w_1 T_{g1} k + w_2 T_{g2}}{w_1 k + w_2}where w_1 and w_2 are the weight fractions of the cryoprotectant and water, respectively, T_{g1} and T_{g2} are their respective glass transition temperatures, and k is an empirical constant reflecting intermolecular interactions.[20] This model accurately describes how cryoprotectant addition shifts Tg, aiding the design of vitrification solutions.[21]High concentrations of cryoprotectants, typically 40-50% w/v, are essential to achieve vitrification while preventing devitrification—recrystallization—during rewarming, as they sufficiently raise Tg to suppress icenucleation throughout the thermal cycle.[16] Factors such as molecular weight, with higher values generally increasing Tg due to reduced chain mobility; hydrogen bonding capacity, which strengthens the glassy matrix; and interactions with water, where cryoprotectants act as plasticizers or stabilizers, profoundly influence this elevation.[22]
Ice Crystal Inhibition and Stabilization
Cryoprotectants inhibit ice crystal formation primarily through colligative effects that reduce the amount of free water available for nucleation and growth. By increasing the solute concentration in solution, these agents lower the freezing point via freezing point depression, described by the equation \Delta T_f = K_f \cdot m, where \Delta T_f is the change in freezing temperature, K_f is the cryoscopic constant of the solvent, and m is the molality of the solute.[23] This colligative property, rooted in Raoult's law, disrupts the equilibrium between ice and liquid water, thereby decreasing the temperature at which ice crystals initiate and expand.[8] For instance, high concentrations of permeable cryoprotectants like dimethyl sulfoxide (DMSO) form hydrogen bonds with water molecules, further elevating the fraction of unfrozen water and suppressing ice propagation.[23]In addition to colligative mechanisms, cryoprotectants can adsorb directly onto nascent ice nuclei or crystal surfaces, sterically hindering their growth. This adsorption-inhibition process limits the incorporation of water molecules into the icelattice, effectively capping crystal size and preventing the formation of sharp, damaging edges.[23] Specialized ice-binding proteins, such as antifreeze proteins (AFPs), exemplify this through their flat ice-binding faces that anchor to specific planes (e.g., basal or prism) on icecrystals, curving the ice-water interface and inhibiting further advancement.[24]Cryoprotectants also stabilize biomolecules against freeze-induced damage by preserving native structures and mitigating stress during phase changes. They maintain hydration shells around proteins, preventing denaturation by counteracting the loss of bound water that occurs as ice forms and concentrates extracellular solutes.[25] Similarly, these agents reduce osmotic stress and inhibit deleterious membranephase transitions from fluid to gel states, which could otherwise lead to leakage or rupture of lipid bilayers.[26] Non-permeating sugars like trehalose, for example, form a protective matrix that supports protein folding and membrane integrity during dehydration.[23]A critical distinction in ice formation lies between extracellular and intracellular locations, with cryoprotectants preferentially minimizing the latter to avoid mechanical disruption. Extracellular ice, which forms during slow cooling, induces cell shrinkage via osmosis but generally spares organelles; however, intracellular ice—arising from rapid freezing—produces sharp crystals that can puncture membranes and organelles, leading to lysis.[23] Permeable cryoprotectants penetrate cells to equilibrate osmotic gradients, thereby reducing the propensity for intracellular nucleation and growth while promoting controlled extracellular ice formation.[27]Antifreeze proteins serve as potent, naturally evolved inhibitors of ice crystal dynamics, exhibiting thermal hysteresis and recrystallization inhibition. Thermal hysteresis creates a gap between the melting point and the non-equilibrium freezing point, where AFPs bind to ice surfaces and generate pinning sites that resist ice front propagation at temperatures below the melting point.[28] During thawing, these proteins further prevent recrystallization—the merging of small crystals into larger, more destructive ones—by adsorbing to grain boundaries and maintaining crystal multiplicity.[24] This dual action, observed in polar fish and insects, has inspired biomimetic designs for enhanced cryopreservation.[29]During freezing, cryoprotectant concentration gradients emerge as ice forms extracellularly, progressively concentrating solutes in the unfrozen channels and exacerbating the solute effect—increased ionic strength and pH shifts that destabilize cells.[5] To mitigate this, combinations of permeable and non-permeable cryoprotectants are employed, allowing intracellular equilibration to buffer osmotic imbalances and limit extreme solute buildup, often reducing required concentrations by up to 50% while preserving viability.[23]
Toxicity Considerations
Cryoprotectants can induce various forms of toxicity during cryopreservation, primarily arising from their chemical interactions with cellular components, osmotic imbalances, and exposure to subzero temperatures. These adverse effects limit the concentrations usable for effective ice prevention, particularly in vitrification protocols requiring high CPA levels.[30][5]Chemical toxicity manifests through direct disruption of cellular structures, such as dimethyl sulfoxide (DMSO) extracting lipids from phospholipid bilayers, which decreases membrane thickness, forms transient pores at 10-20% concentrations, and destroys bilayer integrity at 25-30%. This solvent-like action compromises membrane permeability and function, leading to leakage of intracellular contents. Glycerol, while less disruptive, can depolymerize actin filaments in sperm at concentrations above 1.5%, impairing cytoskeletal stability.[30]Osmotic toxicity occurs due to rapid volume changes in cells during CPA addition or removal; hypertonic solutions cause shrinkage, while hypotonic conditions lead to swelling and potential lysis, as seen with glycerol's low permeability exacerbating extracellular dehydration. Chilling injury below 0°C further compounds this, as supercooling without ice formation triggers phase transitions in membranes and proteins, promoting instability even in the presence of CPAs.[30][31][32]Toxicity is highly dose-dependent, with concentrations exceeding 10% v/v often inducing apoptosis via caspase activation or necrosis through membrane blebbing and ultrastructural damage; for instance, DMSO above 1.41 M (about 10% v/v) reduces myocardial contractility in rats and triggers mitochondrial permeability transition. Exposure time and temperature modulate these effects, as prolonged contact at warmer temperatures accelerates chemical reactions, while lower temperatures slow permeation but prolong osmotic stress.[30][33]Species-specific sensitivities influence vulnerability, with mammalian cells generally more susceptible than plant cells owing to differences in membrane fluidity; mammalian plasma membranes, rich in unsaturated lipids, undergo greater perturbations from CPAs like DMSO, whereas plant cells maintain rigidity through sterols and natural solutes, enhancing tolerance. For example, flounder embryos exhibit higher glycerol toxicity compared to bacterial cells, highlighting phylogenetic variations in CPA metabolism.[30][34][35]Basic mitigation involves stepwise addition and removal of CPAs to minimize osmotic shock, allowing gradual equilibration and reducing peak intracellular concentrations. Antioxidants, such as ascorbic acid, counteract reactive oxygen species (ROS) generated by CPAs like DMSO during freezing, which oxidize lipids and proteins, thereby preserving cell viability.[30][36]Key toxicity metrics include LD50 values; for glycerol, the oral LD50 in mice is approximately 4.1 g/kg, indicating relatively low acute systemic risk compared to DMSO's 17.4 g/kg in rats, though cellular effects dominate in cryopreservation contexts.[37]
Types of Cryoprotectants
Permeable Cryoprotectants
Permeable cryoprotectants are small organic molecules with molecular weights typically below 500 Da that can diffuse across cell membranes to provide intracellular protection during cryopreservation.[38] These agents primarily function by colligatively depressing the freezing point of water, inhibiting intracellular ice formation, and stabilizing biomolecules through interactions like hydrogen bonding.[4] Key properties include high aqueous solubility, relatively low viscosity to facilitate mixing and penetration, and the ability to form hydrogen bonds with water molecules, which disrupts ice crystal lattice formation.[39] However, they often exhibit drawbacks such as potential cytotoxicity at higher concentrations, arising from osmotic stress or chemical interactions with cellular components.[40]Among the most widely used permeable cryoprotectants is glycerol (1,2,3-propanetriol), a polyol that penetrates cell membranes slowly but effectively protects against freezing damage. Discovered as a cryoprotectant in 1949 by Polge, Smith, and Parkes during experiments on fowl spermatozoa, glycerol has since become a standard for slow-freezing protocols in reproductive biology.[41] Its efficacy stems from its kosmotropic nature, enabling strong hydrogen bonding with water to prevent ice nucleation.[39]Dimethyl sulfoxide (DMSO) is another prominent example, known for its rapid membrane penetration due to its amphiphilic structure. Introduced by Lovelock and Bishop in 1959 as a protective agent for red blood cells during slow cooling, DMSO is commonly employed at concentrations of 10-15% v/v for mammalian cells and tissues.[10] It excels in vitrification applications by quickly equilibrating intracellularly, but its higher toxicity compared to glycerol necessitates careful post-thaw removal to avoid adverse effects like hemolysis.[42]Ethylene glycol serves as a less toxic alternative to DMSO, particularly in oocyte and embryo vitrification, where it penetrates efficiently and minimizes osmotic injury. With a molecular weight of 62 Da, it allows for higher concentrations without excessive cytotoxicity, often used in mixtures to balance efficacy and safety.[43] In usage protocols, permeable cryoprotectants like these are typically added stepwise to achieve equilibrium distribution across cell membranes during slow freezing at rates of 0.3-1°C/min, with final concentrations ranging from 5-20% v/v to optimize dehydration and solute penetration while limiting toxicity.[44][1]Comparatively, DMSO offers superior performance for rapid cooling scenarios due to its faster diffusionkinetics, achieving better post-thaw viability in sensitive cells than glycerol, which requires longer equilibration times.[4] However, glycerol generally exhibits lower cytotoxicity, making it preferable for applications where prolonged exposure is unavoidable, such as in spermcryopreservation.[45]
Non-Permeable Cryoprotectants
Non-permeable cryoprotectants are solutes that do not cross cell membranes, including both small compounds lacking membrane transporters and large polymers, thereby acting primarily in the extracellular space to create osmotic gradients that promote cellular dehydration and prevent intracellular ice formation during freezing.[4] These agents remain confined to the extracellular compartment, aiding in vitrification by stabilizing the external environment.[46]Key examples include polyvinylpyrrolidone (PVP), a synthetic polymer used as a cryoprotectant in blood plasma and islet cell preservation due to its ability to inhibit ice recrystallization extracellularly.[47]Sucrose and trehalose, non-permeable disaccharides, stabilize biomolecules by replacing water molecules in hydrogen bonding networks, with trehalose exhibiting superior glass formation owing to its unique α-1,1-glycosidic linkage that yields a higher glass transition temperature compared to sucrose.[48]Hydroxyethyl starch (HES), a modified polysaccharide, serves as an effective extracellular stabilizer in organ perfusion and red blood cell cryopreservation, enhancing viability by promoting dehydration and reducing ice crystal growth.[49]These cryoprotectants are typically employed at concentrations of 10-30% w/v, such as 2-6% HES to supplement permeable agents or 0.3-1 M (approximately 10-34%) for sucrose and trehalose, balancing efficacy with practicality in handling viscous solutions.[50]Trehalose's di-saccharide structure contributes to its effectiveness at these levels by forming a more stable amorphous glass state during cooling.[51]Non-permeable cryoprotectants offer advantages including lower toxicity relative to permeable counterparts, as they avoid intracellular accumulation, and are frequently combined with penetrating agents in protocols to optimize osmotic balance and overall cryopreservation outcomes.[6]
Natural Cryoprotectants
In Insects
Insects employ cryoprotectants to either avoid freezing by supercooling body fluids or tolerate extracellular ice formation while protecting intracellular compartments. Freeze-avoiding species accumulate polyhydric alcohols like glycerol and sorbitol to depress the freezing point and enhance supercooling capacity.[52][53]Freeze-tolerant insects, such as the larvae of the gall fly Eurosta solidaginis, rely on polyhydric alcohols like glycerol and sorbitol (up to 0.5 M glycerol in northern populations), alongside disaccharides like trehalose and amino acids such as proline at elevated levels (up to several hundred mM) to stabilize membranes and proteins during controlled extracellular freezing. These species often produce ice nucleators to initiate freezing at higher subzero temperatures (supercooling points around -10°C to -20°C), allowing controlled ice formation while cryoprotectants limit damage. Some freeze-tolerant insects also have ice-binding proteins that inhibit ice recrystallization, thereby limiting intracellular ice damage.[54][55][56][57][58]The synthesis of these cryoprotectants occurs via rapid enzymatic conversion from stored glycogen reserves during cold acclimation, often triggered by decreasing temperatures in autumn. This process involves activation of glycogen phosphorylase and the pentose phosphate pathway, leading to seasonal shifts in hemolymph composition where polyols and sugars dominate in winter.[59][60]Representative examples illustrate the diversity of insect adaptations. The Antarctic midge Belgica antarctica accumulates erythritol (along with glucose, sucrose, and trehalose) to up to 100 mM or more, supporting freeze tolerance in larvae exposed to -15°C or lower during the austral summer. In bark beetles like Cucujus clavipes, alanine serves as a non-toxic cryoprotectant, accumulating to high levels (often >100 mM) to stabilize proteins and contribute to supercooling without the toxicity risks of polyols.[61][62]These strategies provide adaptive benefits, with glycerol particularly effective in lowering the hemolymph freezing point colligatively (approximately 1.86°C per mole) while also stabilizing molecular structures, allowing survival at -20°C to -30°C without toxicity in many species.[63][64]
In Amphibians and Other Vertebrates
Amphibians and certain other vertebrates have evolved natural cryoprotective strategies to survive subzero temperatures, primarily through the accumulation of endogenous cryoprotective agents (CPAs) that enable either freeze tolerance or avoidance. In temperate zone amphibians, freeze tolerance is a prominent adaptation, allowing partial extracellular freezing while protecting intracellular fluids. This contrasts with polar species, such as certain fish, which employ freeze-avoidance mechanisms to maintain all body fluids in a liquid state below the freezing point. These strategies reflect evolutionary divergences: temperate species favor tolerance to endure episodic freezes during hibernation, whereas polar species prioritize avoidance to prevent any ice formation in perpetually cold environments.[65][66]The wood frog (Rana sylvatica), a classic example of freeze tolerance in amphibians, accumulates massive amounts of glucose as its primary CPA during winter hibernation. Triggered by extracellular ice nucleation, the liver rapidly mobilizes glucose from glycogen stores, reaching plasma concentrations exceeding 200 mM and up to 300 mM in vital organs like the heart and liver. This hyperosmotic response draws water out of cells via cryoprotective dehydration, minimizing intracellular ice formation and stabilizing cellular structures during freezing episodes that can reach -16°C in subarctic populations. Glucose mobilization is organ-specific, with high levels prioritized in the brain, heart, and liver to preserve essential functions, while peripheral tissues rely on lower concentrations.[66][67][68]In some Australian burrowing frogs, such as species in the genus Cyclorana, urea serves as a key permeable CPA, accumulating to levels of 50-100 mM or higher during periods of environmental stress like aestivation and hibernation. As a small-molecule stabilizer, urea permeates cell membranes to counteract dehydration and ionic imbalances, enhancing survival in both dry and cold conditions by stabilizing proteins and membranes without the osmotic extremes of glucose. This adaptation is particularly vital in arid-temperate regions where frogs face combined desiccation and frost risks.[69][70]Among other vertebrates, polar fish like Antarctic notothenioids and Arctic cod produce antifreeze glycoproteins (AFGPs), which are repetitive glycopeptides secreted into the blood to inhibit ice crystal growth and lower the freezing point of bodily fluids by 1-2°C. These AFGPs bind to ice surfaces, preventing recrystallization and enabling freeze avoidance in seawater that is -1.9°C. In contrast, freeze-tolerant reptiles such as the wood turtle (Glyptemys insculpta) employ glucose accumulation similar to amphibians, combined with upregulated antioxidant defenses like superoxide dismutase and catalase to mitigate oxidative damage from ischemia-reperfusion during thaw.[71][72][73]Mechanisms underlying these adaptations include organ-specific mobilization of CPAs, often regulated by hormonal signals like norepinephrine in amphibians, and cryoprotective dehydration facilitated by aquaporins—water channel proteins that selectively permit extracellular ice-driven water efflux while retaining intracellular solutes. In wood frogs, aquaporin expression in skin and bladder aids controlled dehydration, limiting body water loss to 65% during freezing and preventing lethal intracellular icing. This aquaporin-mediated process underscores the integration of membrane transport with osmotic protection across freeze-tolerant vertebrates.[66][74]
Applications
Cryopreservation of Cells and Tissues
Cryopreservation of cells and tissues relies on cryoprotectants to mitigate damage from ice formation, osmotic stress, and chilling injury during cooling to subzero temperatures and subsequent thawing. In biomedical applications, these agents enable the long-term storage of viable biological materials for clinical use, such as fertility preservation and regenerative medicine. Protocols typically involve either slow freezing, which allows controlled ice nucleation outside cells, or vitrification, which achieves an amorphous glass state to prevent ice entirely. Success depends on optimizing cryoprotectant concentration, cooling/warming rates, and post-thaw recovery to maintain cellular function.Slow freezing protocols for cells like hematopoietic stem cells commonly use permeable cryoprotectants such as 10% dimethyl sulfoxide (DMSO) in a stepwise manner, with cooling rates of 1-2°C/min down to -80°C before transfer to liquid nitrogen. This method promotes extracellular ice formation while minimizing intracellular ice, achieving post-thaw viabilities exceeding 80% for mesenchymal stem cells when combined with appropriate serum or polymers. For instance, rat mesenchymal stem cells cryopreserved with 10% DMSO at 1°C/min exhibited immediate post-thaw viability around 85%, with sustained proliferation over days.[75]Vitrification, in contrast, employs flash-freezing with high-concentration mixtures of 20-40% permeable and non-permeable cryoprotectants, such as ethylene glycol combined with sucrose, to induce a glass-like state without ice crystals. For human oocytes, a multi-step protocol equilibrates cells in increasing concentrations of ethylene glycol (up to 12.5 mol/kg) plus 0.75 mol/kg sucrose, followed by plunging into liquid nitrogen, yielding high survival rates by avoiding ice-induced damage. This approach has become standard for delicate structures like oocytes, where post-thaw integrity is critical for fertilization potential.[76]Key applications include sperm and egg banking, established since the 1950s for fertility preservation, and cryopreservation of hematopoietic stem cells for bone marrow transplants, where 10% DMSO enables engraftment rates comparable to fresh cells. In tissue engineering, cryoprotectants facilitate storage of skin grafts, maintaining viability for up to years in allogeneic models to support burn treatment and wound healing. These methods ensure off-the-shelf availability of cells and tissues for therapies. Recent advances as of 2025 include strategies for cryopreserving 3D biofabricated tissues using biomaterial matrices as cryoprotective supports.[77][78][79][80]Standard protocols encompass cryoprotectant agent (CPA) loading through stepwise exposure to minimize toxicity, controlled cooling or vitrification, rapid warming to prevent recrystallization, and gradual unloading to avoid osmotic swelling. For example, loading occurs at 0-4°C in incremental concentrations, followed by warming at 37°C in a water bath and dilution with isotonic media to remove CPAs safely. These steps are essential to preserve membrane integrity and metabolic activity across cell types.[81][82]Post-thaw recovery rates serve as primary success metrics, often exceeding 80% for isolated cells but declining for thicker tissues due to uneven CPA penetration and thermal gradients. Challenges in scaling to whole organs persist, as seen in kidney trials using VMP vitrification solution, where rat models achieved functional recovery after 100-day storage via nanowarming, though initial dysfunction and ice formation limit human translation. Ongoing refinements aim to enhance uniformity for clinical viability.[83][84]
Food Preservation and Processing
Cryoprotectants play a crucial role in food preservation by mitigating the adverse effects of freezing, such as ice crystal formation that leads to cellular damage, drip loss, and protein denaturation. These additives stabilize food structures during freeze-thaw cycles, preserving texture, flavor, and nutritional quality in products like meats, seafood, dairy, and fruits. By lowering the freezing point and inhibiting large ice crystal growth, cryoprotectants enable efficient industrial freezing processes while minimizing quality degradation over extended storage periods. Emerging bio-based cryoprotectants, such as natural sugars and proteins from plants or microbes, are gaining traction for sustainable applications as of 2025.[26]In seafood processing, sugars such as trehalose, a non-permeable cryoprotectant, are widely used at concentrations of 1-5% to prevent drip loss and protein denaturation in fish fillets. For instance, in frozen tilapia fillets treated with 5% trehalose, thawing loss was reduced from 9.18% (control) to 6.60% by maintaining cell membrane integrity and limiting ice recrystallization during storage. This stabilization is particularly effective in surimi and fillets, where trehalose at 5-8% concentrations enhances water-holding capacity and retards myofibrillar protein changes, ensuring better gel-forming ability upon thawing.[26][85][86]Glycerol serves as a common permeable cryoprotectant in frozen dairy products like ice cream to promote smoothness by reducing ice crystal size and preventing coarsening during storage. This addition lowers the freezing point of the mixture, resulting in a creamier texture and decreased hardness, which enhances sensory appeal without significantly altering flavor. In fruit preservation, polymers are employed to further reduce ice crystal formation, protecting cellular structures and minimizing structural collapse in frozen berries and other produce.[87][88]Industrial techniques like air-blast freezing, which circulates cold air at -30°C to -40°C over seafood, are often combined with cryoprotective agents to accelerate freezing rates and limit dehydration in products such as fish fillets. For higher-quality preservation, cryogenic methods using liquid nitrogen achieve ultra-rapid freezing, and when enhanced by polymers, they form smaller ice crystals that preserve the integrity of delicate foods like shrimp and vegetables. These approaches reduce freezer burn and maintain product weight, with air-blast methods yielding typical losses of 1-2% in unpacked seafood.[89][90][91]The use of cryoprotectants extends shelf life for frozen meats and seafood to 6-12 months at -18°C to -25°C by inhibiting microbial growth and oxidative changes, as seen in treated fish mince that retains functional properties far longer than untreated samples. Nutritionally, they support retention of key vitamins, such as vitamin C in frozen vegetables, where rapid freezing with protectants inactivates degradative enzymes and preserves substantial levels of initial content after months of storage.[92][93][94]Regulatory oversight ensures safe application, with the U.S. Food and Drug Administration (FDA) approving cryoprotectants like glycerol, trehalose, and polyvinyl alcohol at specified levels to meet safety standards without health risks. The food industry is shifting toward bio-based cryoprotectants, such as natural sugars and antifreeze proteins derived from plants or microbes, to replace synthetics and align with consumer demand for cleaner labels and sustainability.[95][96][26]
Challenges and Future Directions
Toxicity Mitigation Strategies
One effective strategy to mitigate cryoprotectant toxicity involves stepwise equilibration, where cryoprotective agents (CPAs) are added gradually to cells or tissues to minimize osmotic stress and intracellular concentration spikes. For instance, a two-step procedure for glycerol addition begins with exposure to 0.87 mol/kg glycerol combined with 0.08 Osm/kg nonpermeating solute for 12 minutes at 37°C, followed by a rapid shift to 20 mol/kg glycerol with 0.2 Osm/kg nonpermeating solute for 1 minute at 4°C.[97] This approach keeps cell volume within tolerable limits (0.2–2 times isotonic volume), reducing toxicity exposure compared to direct high-concentration loading.[97]CPA cocktails, combining multiple agents at reduced individual concentrations, further lower toxicity while maintaining cryoprotective efficacy. A common formulation pairs dimethyl sulfoxide (DMSO) with trehalose; for example, 10% DMSO supplemented with 50 mM trehalose enhances post-thaw viability of murine spermatogonial stem cells to 90%, compared to 76% with 10% DMSO alone, by stabilizing membranes and mitigating oxidative damage.[48] Similarly, 5% DMSO with 146 mM trehalose supports proliferation of human umbilical cord blood stem cells equivalent to non-frozen controls, allowing a 50% reduction in DMSO dosage without compromising protection.[48]Incorporating antioxidants and buffers into CPA solutions addresses reactive oxygen species (ROS) generation and pH fluctuations during exposure. Catalase, at 200 U/ml, scavenges hydrogen peroxide in cryoprotective media containing 10% dimethylacetamide or methanol, preserving membrane integrity in zebrafish sperm at 95% during 10-minute pre-freeze equilibration (versus 70–75% without catalase).[98]Alternative delivery methods enhance precision and reduce systemic exposure to toxic CPAs. Polymeric nanoparticles facilitate targeted intracellular release of agents like trehalose, bypassing membrane impermeability and minimizing extracellular high concentrations of permeable CPAs such as DMSO, thereby lowering overall toxicity while improving low-temperature viability.[99]Genetic engineering induces endogenous production of protective defenses; for example, inactivating class-I PI3K in C. elegans via the age-1(mg44) mutation boosts survival to 94% after 5-minute exposure to 75% M22 CPA at 20°C, compared to 7.5% in wild-type, by upregulating cellular stress responses.[100]These strategies collectively enable 20–50% reductions in effective CPA concentrations, such as halving DMSO from 10% to 5% in trehalose combinations, while improving post-thaw viability from as low as 10% to over 80% in endothelial cells.[97][48]
Emerging Bio-Based and Synthetic Advances
Recent advancements in bio-based cryoprotectants have focused on trehalose derivatives and analogs derived from bacterial sources to enhance organ preservation outcomes. For instance, sulfoxide-functionalized trehalose has been developed as a DMSO-free alternative, enabling cryopreservation of peripheral blood mononuclear cells (PBMCs) with preserved expansion potential, phenotype, and cytotoxicity comparable to traditional methods.[101] Plant-derived options, such as oat peptides, have emerged as effective cryoprotectants by mimicking antifreeze proteins (AFPs), inhibiting ice recrystallization and maintaining high post-thaw viability of Lactobacillus bulgaricus.[102] These bio-based agents address toxicity concerns while promoting sustainability, with polyphenols from plants like barley exhibiting AFP-like thermal hysteresis to stabilize gluten networks in frozen doughs.[26]Synthetic innovations include peptide-based ice blockers and fluorinated compounds designed for reduced toxicity and enhanced penetration. Peptide antifreeze agents, such as those derived from bovine gelatin, provide dual cryoprotection by adsorbing to ice surfaces and preserving rheological properties.[103] In 2024 research, spontaneous peptide coatings on surfaces delayed freezing by 5°C, offering potential for scalable tissue protection without genetic modification.[104] These synthetics enable vitrification with minimized ice formation, as seen in polyampholyte formulations that boost post-thaw recovery in mammalian cells by 20-30%.[105]A pivotal development in 2024 involved the vitrification of human-scale organ volumes using the M22 solution, achieving successful glass formation in 3-liter bags without cracking during nanowarming, a step toward clinical organ banking.[106] Computational models have optimized CPA mixtures, such as those guided by mass transport simulations, reducing exposure times by approximately 20% while maintaining tissue viability above 85% in rat kidney models.[107] In September 2025, researchers at Texas A&M University reported advancements in achieving a glass-like state for cryopreservation, potentially improving viability in organ transplantation and biological sample storage without relying on toxic CPAs.[108] These advances tackle scalability challenges, with bio-based options mitigating environmental impacts from synthetic fluorocarbons. Looking ahead, integration of such CPAs with 3D bioprinting promises enhanced recovery of engineered tissues, potentially securing regulatory approvals for routine use by 2030 as perfusion protocols mature.[109]