Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2024 Apr 18;124(9):5668–5694. doi: 10.1021/acs.chemrev.3c00752

Protecting Proteins from Desiccation Stress Using Molecular Glasses and Gels

Gil I Olgenblum §, Brent O Hutcheson , Gary J Pielak †,, Daniel Harries §,*
PMCID: PMC11082905  PMID: 38635951

Abstract

graphic file with name cr3c00752_0012.jpg

Faced with desiccation stress, many organisms deploy strategies to maintain the integrity of their cellular components. Amorphous glassy media composed of small molecular solutes or protein gels present general strategies for protecting against drying. We review these strategies and the proposed molecular mechanisms to explain protein protection in a vitreous matrix under conditions of low hydration. We also describe efforts to exploit similar strategies in technological applications for protecting proteins in dry or highly desiccated states. Finally, we outline open questions and possibilities for future explorations.

1. INTRODUCTION

1.1. Protein Stability and Desiccation Stress

Globular proteins are delicate. In solution they exist in a tight but facile ensemble of conformations centered around their native, biologically active state. Many fold via a two-state mechanism, with folding free energies in solution at room temperature and neutral pH of −5 to −15 kcal/mol, an energy comparable to that of a single hydrogen bond.1,2 Concomitantly, their melting temperatures — the temperature at which half the protein molecules unfold — range from ∼25 to ∼110 °C.3,4 This marginal stability has been selected via evolution to allow proteins to maintain the flexibility required for function while preserving their 3D structure under physiological conditions.

The sensitivity of proteins makes them vulnerable to changes in their surroundings. Living creatures must accommodate a variety of stresses brought on by deleterious high or low temperature,59 salinity,10 acidity or alkalinity,1113 radiation,14 oxidation,15 desiccation,16 and even hydrostatic pressure (for deep ocean dwellers).1,1720 Such stresses drive proteins toward or away from their native state, potentially impairing their biological function. Consequently, numerous strategies have evolved to combat these insults and enable proteins (and other biomacromolecules) structure to survive environmental changes.

We focus on desiccation stress. Proteins evolved in water, and their properties change when subjected to desiccation. In the dry state, globular proteins still exist in an ensemble of states, but changes among conformations are no longer facile,21 and the conformational ensemble, relative to solution, is “frozen”, or “stuck”. Moreover, removing water is expected to be, and usually is, destabilizing.

The fraction of proteins existing in non-native states under desiccation increases for several reasons. First, removing water negates the hydrophobic effect, the driving force keeping nonpolar side chains buried in the protein interior.2224 Specifically, the molecular underpinning of the hydrophobic effect — the increase in water entropy upon burial of the nonpolar moieties — is moderated with the onset of desiccation. Second, removing the distinction between inside and outside exposes more backbone, which increases the likelihood of non-native hydrogen bonding. The third reason involves hydration of charge. In solution, ionic interactions within and between proteins are weak because water has a dielectric constant that is about 80 times larger than air. Eliminating water increases the strength of native and non-native ionic interactions, both inter- and intra-protein, shifting the delicate balance that holds proteins in their functional form.

Cells are crowded with proteins, with typical concentrations of 100 to 300 g/L.2528 This range is likely the maximal concentration compatible with cellular function. Different proteins exhibit distinct tolerances before they are destabilized due to dehydration and subsequent increase in inter-protein interactions. Ubiquitin, a highly miscible protein, can be concentrated up to 100 g/L,29 whereas ribonuclease has a maximum solubility of only 2 g/L.30 As a rough estimate of the minimal average hydration of the solvated proteins, consider a radius of gyration of approximately 1.2 nm for ubiquitin31 and 1.4 nm for ribonuclease,32 and a water molecule with a diameter of about 0.3 nm. Then, the solubility of these proteins translates to an average spatial separation of ca. 13 water layers for ubiquitin pairs and 64 layers for ribonuclease pairs at their solubility limits.33 This simple estimate extends, for example, to DNA in plant nuclei, which also reaches concentrations ≈100 g/L,34 indicating a comparable minimal hydration level.

Under dehydrating conditions, water molecules evaporate, and solute concentrations increase, challenging protein stability.35,36 The contact of proteins with the air–water interface also increases. Proteins tend to adsorb and denature at the interface, to an extent comparable to their accumulation at water–oil interfaces (Figure 1).3739 This tendency is a result of favorable interactions of hydrophobic protein moieties with the nonaqueous environment presented by air. In aqueous solution, these hydrophobic parts are mostly buried inside the folded protein interior, but hydrophobic interfaces, such as air, tip the balance toward more favorable interactions of nonpolar protein residues with the environment. These hydrophobic interactions lead to exposure of the previously buried nonpolar moieties and promote protein denaturation.

Figure 1.

Figure 1

Schematic of proteins in water and at the water–air interface. Proteins tend to denature at the air–water interface, where they expose hydrophobic residues (in red) that are mostly buried in the compact native state. Aggregation may follow as neighboring proteins adhere nonspecifically to each other.

The elevated protein concentration on desiccation is in and of itself an additional challenge that can lead to protein aggregation. The problem arises because at high enough protein concentrations the probability of proteins sticking to each other grows due to the increased probability of protein–protein encounters.40,41 This problem is exacerbated by the exposure of hydrophobic parts of the protein that can interact with each other, leading to entangled aggregates (Figure 1). Aggregation is often irreversible, because the free energies required to disentangle the large aggregates exceed that provided by thermal energy.

The long human history of preserving foods by dehydration teaches us that dry proteins may become more susceptible to certain damaging chemical reactions. The main reason for this chemical sensitivity that can lead to irreversible protein unfolding is that, in the denatured state, proteins are vulnerable to interactions with harmful chemical agents. Processes leading to inactivation include oxidation, disulfide interchange catalyzed by reduction, β-elimination, and racemization.42 Irrespective of the reason or destabilizing factor, protein unfolding and aggregation are detrimental to the proper function of living organisms. Consequently, biology has evolved mechanisms to counter the negative effects of desiccation.

1.2. Biology’s Efforts to Overcome Desiccation Stress

Adaptation allows a wide range of organisms to counteract desiccation stress. Notable among these mechanisms is anhydrobiosis, defined as the ability to lose almost all cellular water and enter a state where metabolism stops but starts again upon rehydration.35 Anhydrobiosis allows organisms to avoid the damaging effects of desiccation on proteins, lipid membranes, and nucleic acids. Among the life forms capable of anhydrobiosis are brine shrimp, Aedes mosquitoes, and tardigrades. Brine shrimp are marketed to children as “The World’s only Instant Pets” because the dried organisms can be sent by mail and “brought to life” by simply adding water.43 Eggs of Aedes mosquitoes, associated with viral infections,44 can withstand 21-days of desiccation, readily hatching upon rehydration.45,46 Tardigrades, popularly known as water bears, are microscopic animals found throughout the world, various species of which are not only capable of anhydrobiosis46 but also, among other stresses capable of surviving high amounts ionizing radiation, pressures higher than those found in the deepest ocean trenches, and even the rigors of outer space.47 This adaptation allows them to virtually halt their metabolism and growth in dry or suboptimal conditions and later resume normal function when conditions improve by readsorbing water.

In the plant and fungi realm, desiccation tolerance is primarily, though not exclusively, observed in seeds, pollen, and spores. The ability of these reproductive elements to undergo reversible anhydrobiosis facilitates their efficient distribution, even under challenging environmental conditions, and their reactivation only when and where conditions are favorable.36

Our understanding of anhydrobiosis is incomplete but we know that cells respond to desiccation, as well as stresses such as freezing, in a variety of carefully regulated ways. One response is to accumulate or synthesize large amounts of additives.36 These additives or “protectants” include specialized proteins and small organic molecules, most often sugars and sugar alcohols. During dehydration, solute concentrations grow, eventually reaching the solubility limit of the additive solutes. If dehydrated fast enough, solutions may enter a metastable state and do not crystallize, but instead solidifies into an amorphous glass which shows molecular disorder but retains a solid-like rigidity (Section 2.1). Thus, under extreme desiccation and in the presence of these additives, a vitrified glassy or aerogel-like matrix forms. This matrix stabilizes the embedded proteins, allowing them to maintain their native structure, or preserving their ability to regain native structure, even as the cellular matrix loses nearly all its water.48 The vitrified matrix also combats chemical hazards by reducing diffusion of harmful reagents, including oxidative species.49 Upon rehydration, and in some cases even in the dry state, proteins remain active.50,51

The most prominent sugar accumulated in many organisms is trehalose,52 but many others are known (Table 1). In addition, many cells upregulate distinct proteins. Prominent classes are the late embryogenesis abundant (LEA), and cytosolically abundant heat soluble (CAHS) proteins.27,53,54 LEA proteins, first detected in plant seeds, have homologues in microorganisms and animals.35,55 Typically small hydrophilic and glycine-rich, LEA proteins are disordered when hydrated, but tend to gain structure when dried.56 CAHS proteins,57 a family found exclusively in tardigrades, are both sufficient and necessary for survival when dried.58 More specifically, they support desiccation survival not only in tardigrades but also when expressed recombinantly in yeast and bacteria.58,59 Unlike LEA proteins, CAHS protein form strong, reversible hydrogels and aerogels of known secondary structure.6063

Table 1. Common Additives in Anhydrobiosis.

System Additive Refs
bacteria trehalose (69)
yeast and fungi trehalose and glycerol (16,70)
plants sucrose, octulose in Craterostigma plantagineum, trehalose in Myrothamnus flabellifolia, raffinose in Xerophyta villosa, galactinol in X. viscosa, LEA proteins in most plants (7173)
lichen lichens that establish symbiosis with cyanobacteria accumulate sugars (glucose), but those containing green algae preferentially accumulate polyols (ribitol, arabitol, erythritol, sorbitol, etc.) (72,7476)
seeds sucrose, trehalose, nonreducing oligosaccharides of the raffinose group, LEA proteins (49,71,77,78)
insects glycerol, trehalose, LEA proteins (70,79,80)
tardigrades CAHS and LEA proteins, trehalose (81)
brine shrimp (Artemia) glycerol, trehalose, LEA proteins (82)
nematodes trehalose (8285)

The details of how LEA and CAHS proteins overcome desiccation stress are unclear, but their abundance and tendency toward disorder suggest that they work by actively participating in forming an aerogel or glassy matrix, possibly together with sugars that often accumulate simultaneously. Studies of LEA and CAHS proteins in vitro show that they typically protect other proteins and lipid membranes similarly (or even better) compared to protection that sugars provide, suggesting they act by similar mechanisms.56,58,6468

Although it may seem intuitive that immobilization in a dehydrated matrix stabilizes proteins, the molecular mechanisms are unresolved. Below, we describe some of the observations (Section 3) and the most prominent theories (Section 4) to explain stabilization in the additive-rich dry state.

1.3. Engineering Efforts to Stabilize Proteins under Desiccation Stress

The importance of protein stabilization in the pharma, cosmetics, and food industries cannot be overstated. Extending the catalytic lifetime of an enzyme, transporting proteins in their dry state, extending the shelf life of food or drugs, and storage around room temperature all demand methods to stabilize desiccated proteins.86 Moreover, some proteins must retain activity in nonaqueous environments, at nonphysiological pH or salinity, or when immobilized to a surface, which imposes similar demands on stabilization. As expected, strategies employed by industry are often similar to those that evolved in biology, including embedding the protein in a glassy matrix or aerogel (Table 2).

Table 2. Examples of Protein Desiccation in Glassy Matrices and Aerogels.

Matrix composition Matrix type Embedded proteins Refs
amorphous SiO2 sol–gels, sol–gels plus additives lysozyme, α-lactalbumin, hemoglobin, metmyoglobin, creatine kinase, hexokinase, heme proteins, glucose oxidase, peroxidases, catalase, tyrosinase, lactate dehydrogenase, urease, bovine carbonic anhydrase, alkaline phosphatase, acetylcholinesterase, butyrylcholinesterase, acid phosphatase, xylanase (8795,101103)
amorphous SiO2 aerogel glucose oxidase, acid phosphatase, xylanase (95)
Au-immobilized protein in SiO2 aerogel cytochrome c (104)
sucrose, trehalose, and leucine sprayed or lyophilized glass lysozyme, rubella vaccine, and influenza antigen (105108)
sucrose, glycerol, maltose, maltodextrin, sorbitol, 1,6-anhydroglucose, and trehalose freeze-dried glass igG1, monoclonal antibody, cytokines, pepsin, plasma components, trypsin lysozyme and catalase (109114)
trehalose convective air drying whey protein (115)
micelles/reverse micelles zwitterionic surfactants, block copolymers cytochrome c, laccase spore coat protein A (116,117)

Beyond vitrification in a glassy matrix of an additive such as sugar, methods that entrap or encapsulate proteins in polymer matrices are also used to stabilize proteins. During entrapment, the protein is fixed inside as the matrix that forms around it via polymerization. This process has been used to immobilize proteins at surfaces using sol–gels.8695 Encapsulation by an environment with cavities, such as reverse micelles, often similarly increase protein stability,96 although there are exceptions.97,98

From sugars in jam to salt in beef jerky and fish, food can be preserved by increasing the environment’s osmotic pressure, thereby dehydrating the food, impeding bacterial growth, and regulating volume, viscosity, and quinary interactions within cells.99 Osmotic pressure, Π, is a colligative property that describes the change in water chemical potential due to the presence of solutes, and is reported in units of the equivalent hydrostatic pressure required to exactly negate the change in water activity due to solute addition. This pressure is embodied in the van ’t Hoff equation (earning van ’t Hoff the first Nobel prize in chemistry), often generalized as Π = iϕCRT, where C is solute concentration, R is the gas constant, T is temperature, i is an index representing the extent of dissociation in solution (e.g., i = 2 for NaCl), and ϕ is the osmotic coefficient that describes deviation from ideal behavior, where ϕideal = 1.100 Molecular glasses that are almost devoid of water generate exceptionally high osmotic pressures that increase food and drugs shelf life due to their antibacterial effects and their potential to increase protein stability.

In summary, protein stabilization in glassy environments holds significant technological advantages and remains a central strategy. To design more efficient environments tailored for protein stabilization, it will be important to resolve the molecular mechanisms of stabilization. Specifically, it is key to link the properties of the glass-former and glass properties to effects of the glass on protein stability.

1.4. Outlook

We focus on recent efforts to resolve the mechanisms by which the glassy state protects dry proteins. We start by describing the properties of molecular glasses most relevant to protein protection. Next, we describe the main methods and proposed mechanisms for stabilizing proteins in glass. We end with open questions and a prospective. Glassy matrices also protect DNA and lipids, and similar strategies are employed in protection from other types of stress, including freezing. These aspects are not covered here, but excellent reviews are available.10,36,70,118120

2. NATURE OF MOLECULAR GLASSES

2.1. Defining a Glass

What distinguishes the glassy state from other states of matter? Unlike a solid crystal, molecular glasses are amorphous. They possess no long-range order, and atomic positions cannot be inferred from a repeating unit cell because no unit cell exists (Figure 2). Even though at the atomic level a glass resembles a supercooled liquid, unlike a simple liquid, glasses exhibit elastic deformation when subject to external forces. However, if an external force is applied slowly, a glass can display inelastic liquid-like deformation over extended periods of time.121 Thus, the fundamental nature of a glass depends on the time scale at which it is observed: a glass is essentially a supercooled liquid with such high viscosity that long-range diffusion is unobservable within relevant experimental times.122,123 Consequently, a glass behaves like a solid on short time scales and a liquid on (extremely) long time scales.

Figure 2.

Figure 2

Schematic of the disaccharide trehalose crystal structure (A) and in the glassy state (B). Trehalose’s crystal unit cell is shown as an orange rectangle in panel A, as resolved by Nagase et al.416 Trehalose glassy structure is resolved from molecular dynamics simulation.231

Many materials act as glass formers, including oxides and sulfides (such as SiO2,124 BO3,125 and As2S3126) salts (like ZnCl2,127 and BeF2128) alloys (including CuZr129) polymers, (including polystyrene,130 polyvinyl chloride,131 and polycarbonate132) and even pure water.133135 Notably, water can be more readily vitrified by adding large amounts of solutes, not only because solutes hinder crystallization of water into ice but also because many solutes are themselves glass formers.136 Solutes that facilitate glass formation include salts like LiCl137,138 and Ca(NO3)2;139,140 sugars like glucose,141 sucrose,142 and trehalose;143 and polyols such as glycerol,144,145 ethylene- and propylene glycols.146,147

The glass transition temperature, Tg, is an important property of a glass former. Tg characterizes the temperature at which the glass former transitions between supercooled liquid and amorphous glass and reflects a material’s intermolecular interactions. This temperature is readily understood by following a cooling process. In the absence of crystallization, as a glass former cools below its melting temperature, Tm, the volume, , and entropy, (where the macron stands for a molar quantity) of the supercooled liquid gradually decrease. However, once a certain temperature Tg is reached (Tg < Tm) the viscosity sharply increases (typically 1000-fold within a 10 K range), while diffusion sharply decreases (Figure 3D).136 This high viscosity corresponds to long molecular relaxation processes, making the equilibrium state virtually unattainable. This characteristic implies that the glass transition is a kinetic event, in contrast to crystallization and other first-order phase transitions that are equilibrium minima in free energies, making the glass state metastable. Moreover, the properties of glass, such as and , exhibit hysteresis,122 so that they depend on the method of preparation. For example, more rapid cooling leads to larger values of and compared to slow cooling because there is less time for relaxation. At the glass transition, the molar volume and entropy begin to deviate from their behavior in the liquid state, as seen in their weaker temperature dependencies in glass. This results in and values that are larger in the metastable glassy state compared to the thermodynamically stable crystal. This change in temperature dependence is also observed in the heat released or absorbed, , during cooling or heating.

Figure 3.

Figure 3

Methods for determining Tg. (A) Thermal expansion coefficient, αV, versus temperature. (B) Heat capacity, CP, versus temperature. Heating cycle is in black and cooling cycle in gray. (C) Gordon–Taylor curves for binary aqueous sugar mixtures showing Tg versus sugar weight fraction. For trehalose, sucrose, and glucose, the values of k are 4.76, 4.94, and, 4.52, respectively.417 (D) Natural logarithm of trehalose diffusion coefficient versus inverse temperature for different trehalose weight fractions as seen in molecular dynamics simulations.222 Blue background depicts the liquid regime and yellow depicts the glass.

Below Tg, the energy content is higher in a glass than it would be under equilibrium conditions because the amorphous structure is far from equilibrium. Values of , , and will decrease as the glass “ages” toward its equilibrium state through slow relaxation. Above Tg, the material has high molecular mobility and can reach thermal equilibrium. Enthalpic relaxation involves the reduction of the glass enthalpy, glass, toward its value in the crystal, crystal, below Tg. In contrast, enthalpic recovery is the enthalpy that the sample releases or dissipates as it relaxes toward equilibrium at the conclusion of a heating cycle, above Tg. Enthalpic recovery manifests as a peak in heat capacity, CP = (∂H/∂T)P (Figure 3B). This peak originates from surplus heat required to raise the temperature above Tg for a sample that aged (relaxed below Tg) toward its equilibrium state.148152 The extra heat can be attributed to the glass transitioning toward a more stable state, somewhat closer to the crystal. A proper dissection of the heat flow due to the heat capacity and enthalpy of relaxation as it passes through Tg requires careful analysis of calorimetric data.153,154

Quantifying Tg is challenging. First, the glass transition is a gradual process that can span a wide range of temperatures. Moreover, the range over which the transition occurs varies, depending on experimental methodology or experimental protocol. Finally, as described above, Tg depends on the rate of cooling or heating and the history of the sample, including its preparation and manipulation. Plus, there are important differences between cooling and heating cycles. For example, while cooling cycles exhibit hysteresis only in relation to cooling rates, heating cycles are also influenced by the properties of the preheated glass.

The value of Tg is typically determined by dilatometry155158 or calorimetry.136 In dilatometric measurements, Tg is identified by a distinct kink in plots of thermal expansion coefficient, αV = (∂ ln /∂T)P, versus temperature (Figure 3A). In calorimetric measurements, e.g., differential scanning calorimetry (DSC), Tg is determined from the change in the slope of CP with temperature (Figure 3B). Multiple distinct values of Tg can be gleaned from features in the CP versus T curves of the cooling or heating cycles, specifically the cooling or heating onset, offset, and midpoint, as denoted in Figure 3B.136,152,159,160 The midpoint Tg is the most frequently employed value due to the inherent difficulty in precisely pinpointing the onset and offset during the gradual process of the glass transition. Furthermore, two distinct methods, namely the inflection point and the half-step point, can be used to determine the CP midpoint. Notably, Tg values derived from the midpoints and the temperature of maximal enthalpic recovery are particularly valuable, as they facilitate the calculation of the activation energy associated with relaxation processes in the glassy state.148

For mixtures, such as those involving carbohydrate protectants and water, Tg varies considerably with composition. Typically, Tg increases with protectant content because of its higher viscosity, which arrests diffusion and reduces the likelihood of crystallization.161 In practical terms, the dependence of Tg on protectant concentration in a mixture is often described by the Fox,162eq 1, or Gordon–Taylor, eq 2, equations:163

2.1. 1
2.1. 2

where Tg,1 and Tg,2 are the glass transition temperatures of the pure components and w1, w2 are the components weight fractions, with w1 + w2 = 1. Figure 3C shows examples of Gordon–Taylor curves for sugar–water mixtures that use the additional empirical parameter k.

At elevated levels of hydration, Tg of an aqueous mixture can deviate significantly from the predictions made by the Fox and Gordon–Taylor equations due to the formation of crystalline ice. The formation of these ice particles is influenced by the rate of cooling.164,165 Slower cooling rates result in the development of larger ice crystals, effectively dehydrating the amorphous mixture as glass-forming components are typically expelled from the crystalline phase. Consequently, the mixture exhibits two distinct phases: crystalline ice and a concentrated amorphous glass. During a heating cycle, this mixture that contains ice and glass, will produce two calorimetric signals, one for the glass to liquid transition and second for the melting of crystalline ice.164 The glass transition temperature of the concentrated glass, Tg′, is larger than Tg of a pure glass mixture because the higher glass former concentration also leads to higher Tg. The melting temperature of ice crystals that are embedded in the glassy environment, Tm′, is lower than that of ice in pure water, Tm, because of the increased entropy associated with water–protectant mixing in the glass. Increasing the cooling rate reduces the size of the ice crystals so that when cooling is fast enough, the purely glassy behavior (in absence of any crystalline ice) described by eqs 1 and 2 is reestablished.

Glass formers are also categorized by their fragility, defined as the extent of deviation of viscosity from Arrhenius behavior.123,136 This fragility has little to do with the brittleness of the glass. Instead, fragility characterizes the viscosity of the liquid as it approaches its glass transition. The glass former fragility can be readily inferred from the Angell plots123,166 showing the logarithm of the viscosity, logη, versus Tg/T (Figure 4A). Scaling T by Tg in the Angell plot allows the comparison of glass formers that undergo the glass transformation at different temperatures. Below Tg, a “strong” liquid adheres to an Arrhenius relationship, while a “fragile” liquid exhibits super-Arrhenius behavior, resulting in a sharp upswing in viscosity as Tg is approached. The super-Arrhenius behavior of fragile glass formers implies that the energy barrier for molecular relaxation increases as the temperature is lowered toward Tg, and subsequently that the molecular motion becomes progressively more cooperative as temperature decreases.167

Figure 4.

Figure 4

Strong and fragile glasses. (A) Angell plots of SiO2, sorbitol, glycerol, sucrose, and trehalose aqueous mixtures from viscosity measurements. Adapted with permission from ref (418), copyright 1997, Springer Nature, and ref (136), copyright 2002, American Chemical Society. Black curve is for strong glass former, gray is for fragile. The curves of strong and fragile glass formers typically intersect at T = Tg because Tg is usually taken as the temperature where log η = 13 (in poise) or where τα = 100 s, which for many materials is close to Tg, as derived from dilatometry or calorimetry measurements.205 By contrast, β-relaxation is considerably faster, often on the order of 1 μs at Tg. (B) Angell plot of α- and β-relaxation for polyethylene terephthalate. Adapted with permission from ref (419), copyright 2006, John Wiley & Sons. T = Tg is marked by a vertical line.

Figure 4A shows examples of strong and fragile glass formers. For example, SiO2 is a strong glass former, whereas molecular mixtures and melts, such as those composed of polyols, sugars, and other organic compounds with or without water, tend to be fragile. It is useful to rank fragility by the steepness or fragility index, m, describing the change in viscosity as Tg is approached:168,169

2.1. 3

A larger value of m corresponds to a more fragile glass former. For example, the fragile anhydrous trehalose has m = 107, while for the less fragile (stronger) glycerol m = 47, and for the strong SiO2m = 20.170172

The fragility of a glass former is also related to its α-relaxation time, τα, which characterizes the continuous evolution of the structure of the liquid or metastable glass. There are different types of α-relaxation times that depend on the method of measurement, including shear (τS),173177 dielectric (τD),178,179 spin (τC),179181 and enthalpy (τH)182,183 relaxation times. These relaxation times are related to each other through the complex motions of the glassy matrix.

Like viscosity, each relaxation time serves as a means to assess liquid fragility based on its departure from Arrhenius behavior. These relaxation times are indicative of macroscopic motions that are progressively more constricted as the temperature drops and the supercooled liquid vitrifies. However, distinct relaxation times may manifest slightly different temperature-dependecies.184 Note that using τS in Angell plots is equivalent to using η since according to Maxwell’s relation the two are linearly dependent.185

For many glass formers, a second relaxation process, called the Johari–Goldstein or β-relaxation, emerges below Tg.186189 The corresponding β-relaxation time, τβ, in a fragile glass exhibits Arrhenius dependence below Tg (Figure 4B). α-relaxation is several orders of magnitude slower (Figure 4B). The difference in the two time scales increases with decreasing temperature. However, the extrapolation of α- and β-relaxation suggests they converge at high temperatures.190 Moreover, the activation energy for β-relaxation is smaller than that of α-relaxation. For example, for a 43.8 mol % toluene-pyridine glass at −140 °C, the activation energies are 50 and 5.4 kcal/mol for α- and β- relaxation, respectively.189

Although the precise origin of β-relaxation is unknown, it is generally considered to encompass the motion of an entire molecule or some larger repeat unit (e.g., for a polymer glass former).191 This process is typically associated with the reconfiguration of molecules “rattling” within the cages formed by neighboring molecules in the vitrified state, as well as shorter events involving the breaking of these cages.192 Because β-relaxation involves rapid and rigid movement of molecules (or repeat units), some argue that this shorter relaxation is a precursor to the α-relaxation and that these faster motions are essential for the viscous flow and diffusion responsible for the longer α-relaxation (Figure 5).193 Regardless of the argument’s validity, τα and τβ are correlated in terms of their pressure and aging dependencies, and tend to merge at elevated temperatures above Tg, suggesting that the processes are coupled.190,191,194

Figure 5.

Figure 5

Schematic of motions associated with α- and β-relaxations. β-relaxation involves limited “rattle” movements of molecules confined by neighboring molecules in the glass, while α-relaxation is related to larger motions that contribute to the ongoing changes in the glass structure. See text for details.

2.2. Theory of Glasses

So far, we have described the properties of a glass and the phenomenological parameters to characterize them. We now summarize some theoretical efforts to model the molecular origins of these properties. Initially, most of these efforts were directed to modeling the dependence of the viscosity, and consequently τ, on temperature in supercooled states. The simplest model, suggested by Andrade, describes an Arrhenius-like dependence of η on temperature,195

2.2. 4

where Ea is the activation energy and η the limiting viscosity value as T approaches infinity. Developments by Volgel,196 Flucher,197 and Tammann and Hesse198 led to the Volge–Flucher–Tammann (VFT) equation,199

2.2. 5

with T0 the temperature where the viscosity diverges, and B is an empirical parameter. B was later related to the fragility through B = DT0, with the empirical parameter D < 20 corresponding to a fragile glass former and D > 100 corresponding to strong glass former.136,200 Further improvement to the temperature dependence of the viscosity is given by the Williams–Landel–Ferry (WLF) equation.201 The dependence of viscosity on external pressure can be modeled by an analog of the VFT equation,

2.2. 6

where P0 is the idealized glass transition pressure, η0 is the extrapolated zero pressure viscosity and C is the empirical pressure fragility parameter. More complex models for the pressure dependence have also been proposed.200,202

Two important theoretical descriptors of mobility in glass and glass viscosity are the free volume and the excess configurational entropy. Under the free volume framework suggested by Doolittle,203,204 the viscosity increases upon cooling because of the reduced space afforded to each particle. The viscosity then depends on the free volume, Vf, defined as the volume that arises due to thermal expansion of the glass, so that Vf = VVoc with V the total volume and Voc the occupied volume. The dependence of viscosity on volume is then:175,176

2.2. 7

where V is a fitting parameter.

Excess configurational entropy, Sc, is associated with the diversity of configurations of the supercooled liquid particles with respect to the crystal. When extrapolated to below Tg, Sc approaches zero at the so-called Kauzmann temperature, TK, suggesting that below TK the entropy of the metastable state is less than the entropy of the crystal.205 In practice, this inversion of entropy is avoided by the purely kinetic glass transition, implying, paradoxically, that at some (long) time such an inversion could be attained.206 For many glass formers the value of the thermodynamic TK and kinetic T0 coincide, which might explain the apparent viscosity divergence of the VFT model (eq 5).207

Kauzmann attempted to resolve the paradox by suggesting the crystallization of the metastable state between Tg and TK.205 Despite its plausibility, Kauzmann’s resolution has not been corroborated experimentally, and subsequently, the “Kauzmann paradox” drove the development of new theories. In their work, Gibbs and DiMarzio (GD), developed a lattice theory suggesting the onset of a second-order phase transition upon cooling above or at TK.208 Building on GD theory and by using detailed balance arguments, DiMarzio and Yang developed a kinetic theory relating particle trapping and escaping from deep wells to glass properties. These properties include the viscosity, which unlike the VFT theory does not diverge at any temperature, circumventing the Kauzmann paradox.209 The dynamic nature of supercooled liquid and glass, including caging and transitions between metastable states, is further emphasized in the fluidized domain model,210,211 mode coupling theory,212215 random first order transition theory,207,216 and the self-consistent generalized Langevin equation.217

The influence of thermal history and aging on glass energetics can be analyzed using the Tool–Narayanaswamy–Moynihan (TNM) model. TNM is particularly valuable because it can be applied to fit differential scanning calorimetry (DSC) experiments, accommodating various scanning protocols, rates, aging times, and temperatures. Through TNM fits, insights are gained into the activation energy governing relaxation processes and the correlation between enthalpic recovery and sample history.148 To address the nonexponential nature of relaxation processes in glass, TNM introduces an empirical parameter 0 < β̃ ≤ 1, denoting the deviation from exponential behavior (the tilde distinguishes it from β-relaxation). The relaxation decay function is then expressed as218

2.2. 8

Here, τ0 represents the characteristic relaxation time that is given by

2.2. 9

where Ea is the activation energy, x is the nonlinearity parameter, and Tf is the fictive temperature signifying the temperature at which the measured property (e.g., volume or enthalpy) would reach equilibrium. To streamline the extraction of activation energies and pre-exponential factors for relaxation processes from constant rate DSC experiments, a minimal model that encompasses enthalpic relaxation and recovery has been proposed.219

An alternative molecular mechanism was suggested by Adam and Gibbs (AG).220 In their model, the properties of glass are determined by the equilibrium size distribution of cooperatively rearranging regions (CRRs), defined as the smallest region that can change configuration independent of its environment (Figure 6). CRRs and cooperative motion in glass are observed in simulations.221225 The idea is that the size of the CRRs increases upon cooling to comprise the entire glass at TK. By relating the size of the CRRs to Sc, the AG model relates the values of Tg and TK, and resolves the temperature dependence of viscosity through

2.2. 10

where L is a fit parameter.

Figure 6.

Figure 6

Schematic of cooperatively rearranging regions (CRRs) above Tg (A) and in a molecular glass (B). The glassy CRRs are shown in orange, liquid regions in blue.

CRRs have been further developed in Ginzburg’s “two-state, two (time) scale” model,190 which employs a Flory–Huggins type lattice model, where the sites are occupied by two types of elements, a liquid- and solid-like CRRs. Unlike the AG model, the size of these CRRs does not vary with temperature. Instead, the fraction of solid CRRs, ψ, increases upon cooling, with ψ = 1 corresponding to a lattice of only solid elements. The particle motions in the liquid and solid CRRs follow an Arrhenius behavior with a liquid activation energy, El, and solid activation energy, Es. By associating the fast β-relaxation with local motion in the liquid CRRs and the slow α-relaxation with the effective motion in both liquid and solid CRRs, the relaxation times are

2.2. 11

Here, ψ is determined for a given pressure and temperature by minimizing the lattice model free energy. This simple framework describes the pressure and temperature dependence of both relaxation time scales, including the super-Arrhenius behavior of τα, and the variation of Tg with degree of polymerization in polymeric glass formers. Further modification allows the model to capture the effect of cooling rate and thickness of a thin glass film on τα and Tg.226,227

To date, the predominant focus in theoretical research has been on simple glass formers, characterized as spherical particles that interact primarily through steric interactions,190,202,212,213,215,217 with occasional consideration of close-range interactions.214,226228 These models thus overlook the intricate structural characteristics of most real glass formers. While this simplification facilitates the generalization of properties concerning the α- and β-relaxation processes, including their temperature and pressure dependencies, it can hardly resolve the implications of more complex structure and intermolecular interactions (such as hydrogen bonding and van der Waals forces) to both relaxation processes.

3. PROTECTING PROTEINS UNDER CONDITIONS OF LIMITED HYDRATION

Anyone who has prepared glass candy knows molecular glasses are easy to make: sucrose and water are simply combined at high temperature and then cooled.229 Glucose is added to prevent sugar recrystallization, probably by enhancing the impact of mixing entropy or increasing packing frustration of the nonhomogeneous mixture. The result is a transparent and brittle glass. If exposed for too long, sorption of ambient water to this hygroscopic medium damages the glass. In the following sections we discuss the generalizability of the glass formation process. Are details of the glass forming process important in determining the properties of the glass? What are the key considerations when embedding proteins in glasses or aerogels?

3.1. Molecular Glasses

The glassy state is reached for molecular glass formers when a liquid is supercooled fast enough below Tm to prevent crystallization (Section 2.1). This quenching process typically involves simultaneous changes in temperature and hydration. Desiccation and temperature quenching are thus coupled in glass formation because the melting point of the mixture depends on hydration.

Glass is metastable. Therefore, its structure and properties depend on the method and conditions of preparation as well as the history of the sample once it is formed.123 Specifically, glasses may be quenched to states with different local structures and order, which may modify processes within the glass, including water sorption and crystallization, that directly impact properties including viscosity, glass transition temperature, and relaxation. This diversity complicates the technological application of preservation by glass in the pharma and food industries because glass can be unpredictable. Lack of predictability has led to a search for robust preparation methods. Four distinct methods have been reported for creating amorphous sugar glass with little or no water. These methods are freeze-drying, spray-drying, vacuum drying, and melt quenching (Figure 7),230 although many variations have been reported.231,232

Figure 7.

Figure 7

Schematic of methods for forming molecular glasses and gels. Sol–gels, aerogels, and xerogels have controllable distinct morphologies, while the differences in the structure of molecular glasses formed using different methods are not as well resolved (Section 3.1).

Freeze-drying, or lyophilization, involves freezing of an aqueous solution of sugar or other glass former (typically at concentrations of 0.25–1 M) well below the freezing point (typically −45 to −200°C), with subsequent drying at low pressure (less than ∼100 Torr). The initial solution may also include proteins or other molecules of interest for embedding in the glass. The fundamental idea is to exploit the slow protein dynamics in the quenched cold state and the low vapor pressure of water under vacuum, so that by reducing the pressure, water is removed as vapor by sublimation. This process unfolds in three steps.233,234 First, during the initial freezing, the degree of water crystallization and ice morphology is determined based on factors like cooling rate and glass-former concentration (Section 2.1). In the second step, referred to as primary drying, ice is removed through sublimation. The final step involves secondary drying, during which water trapped in the amorphous phase is evaporated, typically at higher temperatures (e.g., 40–60°C) to accelerate water evaporation.233 Evaporation rates in both drying stages are controlled by ice formation that occurs during the initial freezing.165 Since crystalline ice is a “lousy solvent” for many solutes, including proteins, freezing of pure water may involve these solutes’ exclusion resulting in effective dehydration,164 which may lead to protein aggregation. However, the concurrent increase in the concentration of the glass-forming agent, which in turn promotes protein stability in solution and within the glass, protects proteins from misfolding and aggregation.234 Importantly, a rapid freezing process impedes harmful water crystallization, affording the entrapment of proteins and other macromolecules in an innocuous matrix of supercooled water. The restricted motions at low temperatures protect the embedded protein (or other macromolecule) during vitrification.

Spray-drying involves spraying an aqueous solution of the glass former into air or nitrogen gas at elevated temperature, typically close to 100°C, leading to water evaporation. Spraying through a narrow nozzle produces a powder of finely tuned particle size distribution and morphology. However, the elevated temperature can damage many biomacromolecules.235 In the vacuum drying method, the solid glass former is dehydrated by heating under vacuum for several hours, at temperatures less than 100°C and pressures under 100 Torr. Nitrogen flow may be used to exclude oxygen and water vapor. Melt-quenching typically involves briefly heating the solid glass former well above the glass forming temperature under nitrogen gas flow, followed by rapidly cooling, e.g. by plunging into liquid nitrogen. The resulting glass in all these methodologies may be further milled to obtain flowing powders for technical applications.

Heating must be considered carefully because it promotes unwanted chemical reactions of both glass former and protein. In particular, sugars can undergo Maillard and caramelization reactions causing nonenzymatic browning.236 In the Maillard reaction, reducing sugars and amino acids react at moderate temperatures to produce melanoidins. Nonreducing sugars such as trehalose do not undergo this reaction. Caramelization typically proceeds at higher temperatures, and only sugar, reducing or nonreducing, reacts. The products are polymers, including caramelans, caramelens and caramelins.

Maillard and caramelization reactions are detrimental to both glass formation and embedded proteins. Thus, nonreducing sugars and ones that resist caramelizing will be easier to manipulate in technological applications, making trehalose an attractive choice.237 Its inertness may in part explain why trehalose is selected by so many organisms as the glass-former of choice in combating desiccation.238 Beyond chemical reactivity, heating can also induce protein unfolding and denaturation. Heating proteins close to and above their melting temperature should thus be avoided. However, the stabilization that sugars impart on proteins, even in the solution before the formation of glass, may circumvent the deleterious effect of solution heating en route to glass formation.5,239242

Sample preparation history can impact glass properties. In a study of trehalose glass formation, several properties were compared using different preparation methods.230 Interestingly, fragility, transition temperature, and the heat capacity change at the transition temperature varied only slightly between methods, but the onset temperature of crystallization and enthalpic relaxation were notably different. In addition, the method of preparation affected the rate and extent of water sorption, and sorption removed the effects of structural history in the amorphous dehydrated phase.

In another study, the preparation method influenced the amount of water in trehalose glass.232 Water content is important because it can affect the mode of protein protection, as well as many glass properties, including Tg (Figure 3C). In addition, the concentration of sugar in the initial solution from which the glass is prepared influenced the outcome of the preparation: high concentrations of sugar or large volumes led preferentially to formation of the trehalose dihydrate crystal, while lower concentration or volumes more often resulted in a glass.232 Trehalose glass may tolerate significant hydration while maintaining a high Tg (close to 100°C), perhaps because excess water tends to form the dihydrate crystal, which separates from the glass.243245 The removal of water from the glassy state by the dihydrate crystal may be another reason trehalose was selected by evolution for desiccation protection.

Adding components to the glassy matrix can improve the stability and function of embedded proteins. Mixing polyethylene glycol (PEG) with sugar alcohols or sugars (trehalose, lactose, or mannitol) in a freeze-dried matrix preserves lactate dehydrogenase or phosphofructokinase better than the respective single-component matrices.246 In other studies, adding glycerol to a sugar glass matrix improved its ability to protect proteins.247,248

3.2. Solgels, Aerogels, and Reverse Micelles

A sol–gel process involves an inorganic collodial suspension that forms a three-dimensional gel (Figure 7).249 In contrast to molecular glasses that form a matrix through noncovalent bonds, sol–gels processes entrap proteins by creating a chemically bonded network. Sol–gel syntheses immobilize proteins by encapsulating and adsorbing them into the gel’s porous structure.103 This strategy is considered “hard entrapment” in contrast to the “soft entrapment” by molecular glasses.250 For several decades, immobilization and entrapment of proteins and other biologically relevant molecules through the sol–gel processes has presented itself as an attractive methodology for enzyme applications and biosensor preparations.101,250252 Entrapping enzymes in these gels has evolved into one of the most popular ways to immobilize enzymes, with wide ranging applications in biotechnology and biomedicine, and for environmental, synthetic, and sensing uses.253259

The chemical process involved in forming the oxide matrix is simple, and is achieved under mild conditions, including low temperature and moderate pH. In addition, the porosity and chemical nature (such as hydrophobicity vs hydrophilicity) of the matrix can be tuned. Another advantage is that a significant amount of water can be maintained in the matrix, which facilitates protein activity.260,261

The sol–gel process is based on the ability to form silica, metaloxide, and organosiloxane matrices of defined porosity by the reaction of organic precursors at room temperature.228,262 Enzyme immobilization is typically achieved through a polymeric (or alkoxide) route. First, oxide precursors are suspended or dissolved, usually at acidic pH in the presence of water. For silica-based sol–gel syntheses, the matrix precursors are tetraalkoxysilanes or mono-, di-, or trialkyl alkoxysilanes, most commonly tetramethoxysilane (TMOS) and tetraethoxysilane (TEOS).250 In the next step, condensation reactions between silanol moieties are promoted by activating the hydrolyzed precursor with a base (e.g., KOH). Condensation results in a siloxane polymer matrix that grows until the onset of gelation. Proteins can be trapped in the resulting matrix as it forms around them.101,263265 Moreover, enzyme activity is often preserved in the matrix.

Sol–gel synthesis can be detrimental to protein stability. Alcohols (ethanol and methanol) released from TMOS and TEOS as byproducts of hydrolysis and condensation reactions can harm protein activity, although they are less damaging than long-chain alcohols.101,250,266,267 Moreover, sol–gel synthesis requires harsh conditions, such as hot concentrated KOH or HF, to release the proteins. Alternative sol–gel precursors such as poly(glyceryl silicate) help alleviate this problem.268 Recently, Potnis et al. reported a gentle release method that avoids strong acids and bases via their “Capture and Release Gels for Optimized Storage of Biospecimens System”,103 which avoids the damaging effects of acids, bases, and alcohols on proteins by using a short microwave treatment for hydrolysis of TMOS and then rotary evaporation of methanol from the solution of orthosilicic acid.103

Parameters for gel formation, including pore size and entrapment tightness, as well as pore-wall properties, can be controlled by adjusting the ratio of H2O to Si, pH, and solvent.269 A low pH and low H2O:Si ratio favor gel tightness, with smaller pores generated by a cluster–cluster polymerization, within a highly branched matrix. Under alkaline conditions, condensation is limited by the availability of hydrolyzed precursors, and polymer growth is driven by a monomer-cluster process, resulting in a looser silica matrix.262,269,270 Increasing gel tightness can block or slow conformational transitions necessary to achieve enzyme function, resulting in protein conformations with higher or lower activity. As for pore surface properties, NMR data show that even though the gel pores are hydrophobic, the presence of some free hydrophilic groups inside the pore, like hydroxyl groups from hybrid gel formed by TMOS/n-butyltrimethoxysilane precursors, has a role in slowing exchange of polar solvents, preserving enzyme activity.250

While gels formed by sol–gel synthesis often provide an excellent matrix for proteins, they are considerably hydrated, and developing ways to preserve proteins in the dry state are desirable. Indeed, the hydrated gels can be desiccated to produce solvent-free xerogels or aerogels that can hold proteins or other guest molecules. While both types of gels are formed by drying the hydrated gel, xerogels are realized by dehydrating liquid from the sol–gel, typically at, or close to, ambient conditions, while aerogels are achieved either by freezing with subsequent solvent extraction at low pressures through sublimation, or by removing the solvent at supercritical temperatures and pressures. The aerogel process is gentler, preventing capillary forces from acting on the sol–gel matrix as solvent is removed, leaving the matrix largely intact and highly voluminous.

Aerogels and xerogels differ in many ways. Perhaps most notably, aerogels shrink less upon drying, which leads to greater porosity, specific surface area, and a lower bulk density.95,271,272 In fact, aerogels hold the record for the lowest density solid.273 Protein can be trapped in silica aerogels by first forming the hydrated gel around the protein in solution and then replacing water with ethanol with subsequent critical-point drying in an atmosphere of CO2. Enzymes can be trapped in aerogels, typically at 35 to 40°C, without denaturation, exceeding the activity of proteins embedded in the more traditional xerogels.95 Thermal stabilization is also seen, and large substrates inaccessible to the enzymes in xerogels may be accessible in aerogels.

Both hydrogels and aerogels can also be realized using other matrix elements. For instance, proteins can be used as scaffolds for other protein or drug molecules in the aerogel formation process.274 Protein aerogels are advantageous due to the ability to control the surface area that can reach up to hundreds of square meters per gram, and to their exceptional porosity, enabling efficient drug intake, making them suitable for drug delivery. Furthermore, their low densities, typically between 0.003 and 0.5 g/cm3,275 are advantageous for drug delivery systems, particularly for medications administered mucosally, such as through oral or nasal routes.276 Milk or egg proteins are popular starting materials, as well as gelatin and elastin. Carbohydrates can also be used as matrix elements, including cellulose, chitosan, and hyaluronic acid.277 Proteins designed to combat desiccation can also form aerogels. CAHS D aerogels retain the structural units of their hydrogels, but the details depend on prelyophilization concentrations. Only at higher concentrations do slabs form that then comprise the walls of the aerogel pores. These changes in morphology are associated with a loss in disorder and an increase in large β sheets and a decrease in α helices and random coils.60 In summary, methods of gel preparation are important for determining gel properties, complicating the task of unraveling the mechanisms of protein stabilization in these complex matrices.

Nevertheless, encapsulation and protein restriction can even be detrimental to stability. In fact, in technological applications, entrapment in silica glass matrices often reduces the protein catalytic activity and may destabilize them relative to buffer.101 Moreover, in two studies, the interactions of proteins with the confining cavity of reverse micelles destabilized the protein.97,98 We discuss mechanisms of protein stabilization and destabilization caused by volume restriction in Section 4.4.

4. STABILIZATION MECHANISMS IN MOLECULAR GLASSES AND GELS

4.1. Crowding

To lay the groundwork, it is valuable to examine the impact of solute additives or cosolutes on protein stability in aqueous solutions in the liquid state. Perhaps, to a first approximation, the glassy state can be thought of as a highly concentrated solution. In aqueous solutions, protein stability is affected by forces facilitated by added crowders and their preferential inclusion or exclusion from the protein surface. Crowders can assume various forms, including macromolecules (such as proteins or DNA), large polymers, or smaller solutes, also referred to as cosolutes. Crowding, therefore, collectively refers here to the impact of a rising concentration of crowders on stability, structure, association, and aggregation of macromolecules, e.g., proteins.278282

Cosolute inclusion or exclusion is quantified using the preferential hydration coefficient, ΓS, which captures the average excess or deficit of solvent molecules adjacent to the protein compared to bulk solvent.231,283286 If a cosolute is excluded from a protein’s surface, excess water remains in the protein’s vicinity, that is, the protein is preferentially hydrated, and ΓS is positive. Conversely, preferentially included cosolutes result in negative ΓS. Preferential hydration can be determined from the difference between osmotic pressure of mixtures measured in the absence and presence of protein.283,287290

Importantly, the extent to which a cosolute stabilizes a protein is characterized by the change in ΓS upon folding, ΔΓS = Γnatives – Γdenatureds. This parameter directly correlates with the change in folding free energy as the protein is transferred from water to a cosolute-containing solution. If a cosolute is more excluded from the denatured state than it is from the native state, ΔΓS is negative and the protein is stabilized (lower folding free energy). By contrast, denaturants (e.g., urea) are preferentially included at the denatured protein state more than at the native state. ΔΓS is therefore positive and the protein is destabilized (higher folding free energy).283,291293 For example, for the β-hairpin model miniprotein MET16 ΔΓS is ca. −42 for the strongly stabilizing disaccharide trehalose over a wide range of concentrations, but only −23 for its moderately stabilizing monomer glucose.239,241 The same ΔΓS is directly proportional to (minus) the m-value, defined as the change in folding free energy with cosolute concentration.283,294296 For example, for MET16, the m-value is approximately 2.4 kJ mol–1 M–1 in the presence of trehalose and 1.1 kJ mol–1 M–1 in the presence of glucose.239,241 These differences correspond to a respective increase in the folded-to-unfolded ratio by a factor of approximately 2.6 and 1.6 for 1 M sugar. A similar trend in stabilizing effect has been noted for larger globular proteins, e.g., for equine cytochrome c the m-values are 10 kJ mol–1 M–1 for trehalose and 6.3 kJ mol–1 M–1 for glucose.297

The value of ΔΓS, whether negative (stabilizing) or positive (destabilizing), arises from the interplay of interactions between all solution components: protein, cosolute, and solvent. Protein–cosolute interactions include mutual steric excluded volume or hard-core repulsions, and soft (chemical) interactions. Excluded volume interactions are purely repulsive and therefore restrict the volume available to the protein in solution. As the cosolute volume fraction increases there is less space for the protein, forcing it into its most compact state, which, in the presence of water, is most often the native, biologically active, folded state.298 The stabilization exerted by excluded volume interactions is further modulated by solvent–cosolute nonideal interactions that impact solution osmotic pressure, Π. The presence of solvent–cosolute attractions (and correspondingly effective repulsive cosolute–cosolute interactions) increase Π for a given concentration. For cosolutes excluded from protein surfaces, this added osmotic pressure further destabilizes any protein exposed interface, thereby favoring compaction. Conversely, solvent–cosolute repulsions reduce Π and thus reduce the stabilizing effect of the cosolute.241,299

Protein–cosolute soft interactions can be stabilizing or destabilizing.5,282,300306 If the soft interactions are net repulsive the “effective” volume excluded by cosolutes increases, reinforcing the stabilizing effect. If they are attractive, the exposure of sites buried in the native state is favored, so that more cosolute–protein contacts can become available, making these attractive protein–cosolute interactions destabilizing. Soft interaction depends not only on the cosolute but also on the chemical character of the exposed protein interface. For instance, trehalose exhibits attractive soft interactions with the α-helical model protein AQ16 but repulsive interactions with the β-hairpin model protein MET16.241 These contrasting interactions are manifested in the contributions of soft interactions to the m-values: negative for AQ16 (−0.41 kJ mol–1 M–1) but positive for MET16 (0.22 kJ mol–1 M–1).241 Interestingly, attractive interactions might explain why dry proteins are often destabilized in the absence of a protectant, because the only chemical interactions available to a dry protein involve the protein itself. That is, proteins contain many attractive interactions, and unfolded proteins provide even more attractive interactions than do native proteins. Thus, adding molecules to block or prevent these interactions should lead to more proteins in the native state.

An often-neglected theme impacting stability is the nonideal cosolute–solvent chemical interaction,307 which stems from the difference between solution compositions at the protein interface and in bulk solution. Along with the release of cosolute and solvent on folding, the mixing of these liberated molecules with the differently concentrated bulk phase can impact the protein folding free energy.241 These nonideal mixing interactions generate heat that can show up as a stabilizing enthalpic contribution,299 and is observed for many proteins, particularly in the presence of low molecular weight cosolutes.239,240,282,308312

Much effort has been invested to model and predict the effect of cosolute-protein–water interactions. Excluded volume interactions were first modeled by Asakura and Oosawa313,314 and later by using scaled particle theory.301,315320 Soft interactions are usually addressed using weak protein–cosolute binding terms,302,306,321 yet newer models also consider repulsive soft interactions242,307,322 and even explicitly incorporate cosolute–solvent intractions.299,307,323

How far does the analogy of crowding carry to proteins in a glassy matrix? The first hurdle is that there is little or no water in the glassy state, so a picture of a cosolute solvated within a solvent is blurred or even completely inverted (i.e., the cosolute becomes the solvent). Another difference is that the glassy state is amazingly viscous (Section 2.1), so the system may not sample all states on the time scale of the observation. Can we even speak of a well-defined folding free energy? Nevertheless, the image of a protein crowded by cosolute molecules in a desiccated glassy matrix inspires several mechanisms, as described next and in Table 3.

Table 3. Examples of Mechanisms for Protecting Proteins in Glasses and Gels.

Protein Protectant(s) Drying Method(s) Mechanism/notes Refs
plasma trehalose, glucose, sucrose freeze-drying DSC, FTIR, LS, NMR, OD NS (331)
MET16 trehalose, glucose evaporation CD, SRCD NS (231)
GB1 trehalose, glucose, sucrose, maltose, fucose, rhamnose, l-galactose, sorbitol, hexanediol freeze-drying DSC, NMR, TGA water replacement (238)
lysozyme trehalose, sucrose, dextran spray drying, lyophilization DSC, FTIR, SAXS, WAXS water replacement and anchorage (105,106)
lysozyme trehalose freeze-drying FTIR, Raman water entrapment (66,238,332335)
GB1 and CI2 CAHS proteins freeze-drying DSC, NMR, TGA volume restriction and electrostatic interactions  
lactate dehydrogenase and lipoprotein lipase CAHS, globular proteins, Ficoll, and trehalose evaporation, lyophilization enzyme activity NS (67)
myoglobin trehalose evaporation DSC, absorption water entrapment (312,326)
lysozyme, α-lactalbumin, metmyoglobin, RNase A sol–gel matrix - CD, DSC volume restriction, water entrapment (102)
α3W, ubiquitin, cytochrome c reverse micelle - NMR volume restriction (96)
frataxins, titin domain I27 polyacrylamide gel - fluorescence volume restriction (336)
H+-ATPase trehalose - enzyme activity assay, viscometry increased viscosity (337)
myoglobin trehalose evaporation absorbance increased viscosity (338)
myoglobin trehalose evaporation FTIR anchorage (339)
lysozyme glycerol, trehalose freeze-drying Raman, neutron scattering anchorage  
lysozyme trehalose, lactose, myoinositol lyophilization FTIR water replacement (340)
alkaline phosphatase inulin, trehalose spray drying DSC, DVS water replacement (341)
lysozyme sol–gel matrix - DSC, TGA-DTA-MS, CD volume restriction (342)

4.2. Water Retention

One way to extend the ideas from stabilization by excluded crowders in solution to the glassy state is to consider a crowder’s ability to hold on to water upon desiccation. In the water retention hypothesis, a layer of water is maintained by protectant molecules (Figure 8). This residual water interacts with the target protein surface, preserving its stability.324 This mechanism requires that such protectants adsorb large amounts of water and therefore be more hygroscopic than nonstabilizing cosolutes and proteins. However, this hypothesis has tested negative for protein-based protectants from desiccation tolerant organisms because these protector proteins retain no more water than typical globular proteins.324,325 This mechanism is also unlikely for molecular glasses since their stabilizing efficacy usually, but not always,326 increases as water content decreases.327329 Furthermore, glass-embedded proteins and lipid membranes likely retain equivalent amount of water as dry macromolecules on their own.238,330 In contrast to most other theories that explain protein stabilization within glass matrices (Sections 4.3 to 4.7), water retention diverges from the theoretical models of glass outlined in section 2.2. It does not invoke the impact of viscosity, reduced free volume, or glass cooperativity on protein stability.

Figure 8.

Figure 8

Schematics of hypotheses for protein stabilization by glassy matrices. (A) Water molecules are retained by the molecular glass, leading to stabilizing hydration of the protein. Water molecules are present both within the bulk glass and in proximity of the protein. (B) Water molecules are trapped near the protein interface by a glass that is dry compared with the protein surroundings. Entrapped water is strongly included in the protein environment, while the glass is dehydrated. (C) Steric exclusion restricts the protein to a smaller space (shown in white), supporting the stability of its compact native state. (D) The high viscosity in glass slows protein unfolding. Red arrows represent friction forces. (E) Water molecules anchor the protein to the glass matrix (illustrated by small arrows), restricting unfolding. (F) Water–protein hydrogen bonds are replaced by protectant–protein hydrogen bonds (shown by dashed lines), stabilizing the native state. Hypotheses are detailed in Sections 4.2 to 4.7.

4.3. Water Entrapment

A variation on water retention, water entrapment involves water trapped by the glass matrix as the protein “fights” with the matrix for water, as opposed to the available water being retained by the protectant molecules. The protein thus acts like a sponge that sequesters water from the glassy matrix, resulting in protein hydration (Figure 8).326,343 This idea resonates with the notion of CRRs introduced by Adam and Gibbs (Section 2.2), so that the protein possibly inhabits the higher mobility regions that engulf the more rigid CRRs. These more liquid-like regions may be more hydrated because of the inclusion of water molecules around the protein. Thus, the potential stabilizing effect of protein crowding (i.e., stabilization in dense liquid solutions by cosolute exclusion and preferential hydration) directly extends to vitrified glasses.231,332,344

The concept of water entrapment was introduced by Belton and Gil332 in the context of lysozyme’s interactions with a marginally hydrated trehalose glass, where the interaction was probed using FTIR and Raman spectrophotometry. Subsequent evidence from experiments utilizing calorimetry and viscometry reveal that embedded myoglobin draws in water from the adjacent glassy matrix.326 Additionally, molecular dynamic simulations of glass showcase how water molecules become ensnared between the primarily sugar-rich layer and the protein.345347 Notably, protein–sugar interactions within this context are limited312,344,346,347 and weaker compared to stronger protein–water hydrogen bonds.345

4.4. Volume Restriction

In this model, the reduction in water content during vitrification and the concomitant accumulation of larger protectant molecules results in additional steric constraints. The augmented excluded volume interactions arising from higher protectant concentration reduce the available space for proteins in addition to reducing protectant free volume, as described by Doolittle’s model, eq 7. Confining a protein to a smaller volume stabilizes its compact native state.348,349 The stabilization results from the reduced conformational entropy of the denatured states, which increases their free energy with respect to the less affected native state.349,350

The influence of glass confinement on protein structural stability was demonstrated by Eggers and Valentine, who found that the thermal stability of proteins generally increases when encapsulated in a silica glass matrix.102 Since then, various experimental,96,312,336,351356 theoretical,349,350,357 and computational358,359 efforts show that confinement enhances protein stability. However, as discussed (Section 3.2), the conditions used to remove the protein from confinement and competing interactions within the confined space can also lead to destabilization.

The increase in protein stability due to volume restriction coincides with an acceleration of folding rates,350 driven by the restricted search for the folded, native state in configurational space.358 However, when the cavity size becomes too small, folding slows due to the strong entrapment of the protein, preventing the necessary partial unfolding required to correct misfolded states.350 Consequently, to maintain protein stability, theoretical considerations require that the dimensions of the cavities within the glassy matrix that hold the protein must not fall below some critical size, beyond which folding is compromised.

4.5. Increased Viscosity

Conformational changes are required for protein unfolding. Protection by vitrification is often assumed to be the result of the large increase in viscosity below Tg (Figure 4A). As water content decreases, the viscosity of the mixture increases (in concert with Tg that also increases upon dehydration as described by the Fox equation, eq 1, and Gordon–Taylor equation, eq 2, Section 2.1), impeding movements large enough to trigger unfolding.360 In this purely kinetic mechanism, the increased viscosity causes greater friction between the protein and its environment,337 reducing diffusion.338 Viscosity depends on temperature as described by Andrade’s equation, eq 4, or the VFT models, eq 5. Slower diffusion also means that encounters between proteins become less frequent, diminishing the chance of their aggregation.

The rate constant for a reaction that relies only on diffusion, such as folding, decreases with viscosity according to Kramer’s theory, k = η–1eΔE/(RT)eΔG/(RT), where η is viscosity, ΔE is the activation energy, and ΔG is the activation free energy.361,362 The link to activation free energy is correct only if the viscosity follows the Arrhenius relation, η = ηeE/(RT)eG/(RT), where η is a pre-exponential constant and E and G are the viscosity activation energy and free energy, respectively. Given that the viscosity of the fragile glass formers (e.g., sugars) becomes exceedingly high below Tg,120,136,363 it is tempting to conclude that friction between the protein and surrounding glassy matrix plays a substantial role in protein stabilization.

The relationship between protection and increased viscosity is satisfying, easy to grasp, and has been demonstrated for several proteins in crowded solutions,360,362,364366 yet the presumed link between low water content and a high melting temperature in a protein-protectant mixture does not always hold.326 Even in cases where this relationship holds, it is not clear that viscosity alone is responsible for the increased stabilization. The fundamental objection is that at thermal equilibrium, higher viscosity (at least in its most simple definition) should slow both the folding and unfolding rates to the same extent, so that the equilibrium constant for folding is independent of viscosity, KeqDN = kDN/kND = eΔG0DN/(RT) (where ΔG0DN = ΔGN – ΔGD), and protein stability (in terms of free energy) is unchanged (Figure 9). Of course, cosolutes often do change stability, but that is because they affect the free energy of the folded state, unfolded state, or both.282

Figure 9.

Figure 9

Schematic energy landscape for protein folding in media of high and low viscosity. In this example the denatured state, D, is unstable compared to the native state, N, by ΔG0D → N, which is in principle unaltered by viscosity, unlike the folding and unfolding activation energies (Section 4.5).

4.6. Anchorage

The anchorage hypothesis combines elements from the entrapment and viscosity hypotheses. Here, protein motion becomes entwined with the glass matrix, or at least with adjacent CRRs (Section 2.2), through water molecules trapped between the protein and protectant (Figure 8).344,345 This coupling effectively “slaves” protein motion to that of the cooperative glassy matrix, linking the structural fluctuations associated with α-relaxation processes (Section 2.1) to the dynamics of the protein, up to 300 K for the CO molecule bound to carboxymyoglobin in trehalose glass339 and even 350 K for lysozyme in a glass composed of trehalose and glycerol.367 Although the correlation between protein stability and α-relaxation for a specific glass former at different compositions has been demonstrated, the degree of correlation can differ among various glass formers.368,369 By contrast, the relationship between the rate of protein unfolding and β-relaxation appears to be more consistent.370 This observation implies that protein stability in molecular glass is more closely associated with the local fluctuations characteristic of β-relaxation than the large-scale structural alterations of α-relaxation.

As already noted, reducing water content increases both the size of CRRs (as in the AG and Ginzburg’s models, Section 2.2) and glass viscosity (as in Andrade’s and VFT models), slowing both matrix- and protein dynamics.371 Adding small quantities of water or an alternative molecular plasticizer—substances that reduce the glass viscosity (e.g., glycerol)—to the matrix can also enhance protein stability by reinforcing the rigidity of the matrix, facilitated by robust hydrogen bonds between the glass-forming agent and the plasticizer.21,345,372,373

Within the framework of anchorage, differences in the efficacy of sugars in stabilizing proteins in the glassy state can be attributed to either increased matrix rigidity or increased protein-glass coupling. The rigidity idea suggests that the efficacy of fragile glass-forming materials surpasses that of stronger glass formers below Tg because fragile liquids experience a steeper increase in viscosity with decreasing temperature (Figure 4A).367,374376 By contrast, it has been suggested that more fragile glass formers allow faster conformational changes with higher probability, which is detrimental to protein stability.367 The coupling notion implies that the number of water molecules shared by the protein and glass depends on the sugar and that stronger coupling stabilizes the protein. Nonetheless, both considerations explain the discrepancies in sugar stability. For instance, trehalose is more fragile than many other disacharides376 but also shares more water molecules than sucrose,346 likely due to the higher affinity of trehalose for water.377 This duality makes it challenging to determine which property leads to trehalose’s efficacy compared to other sugars. The inconsistent relationships between glass composition, including hydration level, sugar, and polyol content and the extent of glass-induced stability increase highlight some of the unresolved questions concerning glasses and dry protein stability.247

4.7. Water Replacement

Water replacement is the only hypothesis that emphasizes direct protein–protectant interactions. Specifically, it posits that protectants function by substituting water–protein hydrogen bonds with protectant-protein hydrogen bonds upon desiccation.120,340,378,379 This hypothesis goes beyond current glass theories, which focus on properties of the glass former, by emphasizing the coupling between protein and protectant through direct molecular interactions.

The idea was proposed by Crowe et al.,380382 who demonstrated that stabilization of lipid membranes by sugars results from the establishment of hydrogen bonds between sugar hydroxyls and the phospholipid head groups of lipids.118,330,382391 The notion of stabilization via direct hydrogen bonding was expanded to encompass proteins.119 For instance, changes in the amide II band of lysozyme upon dehydration are mitigated in the presence of trehalose.340 Further evidence comes from the NMR-based observation that degree of protection is related to the number of hydroxyl groups in the protectant.238 Likewise, CD data reveal that the glass matrix influences protein stability through intermolecular chemical interactions.231

The proposed direct interactions are often invoked to rationalize variation in protein stabilization observed among molecular glasses with similar Tg values.341 Nevertheless, it remains challenging to dismiss the influence of the protectant’s fragility (and consequently, viscosity and relaxation dynamics),374,376 as well as potential contributions from the other hypotheses.

5. EMBEDDED IN GLASS — PROTEIN PERSPECTIVE

5.1. Alternate Protein Structures

The solvating environment plays a critical role not only in protein stability, shifting the equilibrium toward or away from the native state, but also in the structure of the compact native state and, even more so, the structures in the extended denatured ensemble. By analogy to compaction of intrinsically disordered proteins under crowded conditions,322 the denatured state confined within a glassy matrix experiences a substantial reduction in conformational entropy (Section 4.4), which results in compaction.348,350 Interestingly, simulations of a protein with a β-hairpin native fold, derived from the C-terminal of protein G, suggest that volume restriction on its own is sufficient to impact the denatured state by favoring structural elements in the ensemble with native-like structure.358

Confirmation of the retention of native-like structure in the denatured state is provided by circular dichroism (CD) data on another β-hairpin model protein, MET16, embedded in trehalose and glucose glasses.231 These experiments reveal that the denatured state in the glass matrix closely resembles the native state and differs from the denatured state in aqueous solution (Figure 10). Similarly, novel partially unfolded trapped states were spectroscopically observed for myoglobin in sol–gels.392 This type of compaction may reduce the amount of harmful, aggregation-prone misfolded conformations (Section 5.2). In addition, compaction of the denatured state may reduce the folding time because the conformational search is more confined.

Figure 10.

Figure 10

Scheme of the β-hairpin model protein MET16 solvated in aqueous solution and embedded in a sugar glass. The native state (center) is conserved between solution and glass environments. In solution, the denatured state is extended (right), while within the glass matrix, it adopts a more compact structure (left) that more closely resembles the native state structure.

Finally, the compaction of MET16’s denatured state within the glassy matrix changes the folding mechanism, from predominantly entropically driven in solution to enthalpy driven in the glass. This mechanistic change reverses the temperature dependence of the folding free energy and leads to increased protein stabilization in the glass at ambient temperatures.231

Although the compact denatured states of MET16 in trehalose and glucose glass are indistinguishable, trehalose affords more stabilization, which is strongly enthalpy driven. This sugar-specific shift in equilibrium thermodynamics suggests that the roles played in solution by hydrophobic contacts and water release upon folding are replaced in the glass by specific interactions between the glass and the restricted folded and denatured states. The differences between sugars in the extent of stabilization suggest that the matrix is not simply inert. Instead, nonsteric intermolecular glass–protein interactions play a role in protein stabilization.

Secondary structure content from FTIR data on dry samples shows substantial structural differences between aqueous solutions and sugar glass for larger globular proteins, including bovine serum albumin, myoglobin, and lysozyme.393396 When these proteins are combined with sugar and polyol glass formers like sucrose, trehalose, maltose, and glycerol in their desiccated form, a marked increase in α-helix content is observed compared to the same proteins desiccated in the absence of protectant. Furthermore, changes in the content of α-helix, β-sheet, and various types of unordered structures are noted between aqueous solutions and sugar glass.393 These structural changes also impact the temperature dependence of folding, generally extending the stability of α-helix and β-sheet structures to higher temperatures.

5.2. Intra-Protein Interactions

Beyond modifying protein structure and interactions with their environment, protein confinement by dense and crowded solutions, as well as integration into a protective glass matrix, also impact interactions between protein residues within the protein and among neighboring proteins. As a result, the amorphous glass matrix modulates protein stability and protects against unfolding and irreversible aggregation, processes that are usually detrimental to function.

The increase in number and strength of intra-protein interactions by protectants is evident in simulations of highly crowded solutions, even at concentrations below the glass transition. A study combining MD simulation and NMR demonstrates that polyols and sugars reduce the length, and hence increase the strength, of intra-protein hydrogen bonds (Figure 11A).397 In another simulation study, an osmotic pressure of 3.9 Osmolal exerted by the sugar alcohol sorbitol was shown to considerably strengthen intra-protein hydrogen bonds, leading to stabilization.398 Notably, this increase in hydrogen bonding strength seems to have a more pronounced impact on intrabackbone bonds compared to those involving side chains.397,398

Figure 11.

Figure 11

Schematics of the impact of molecular glass on protein–protein interactions. (A) Compaction by a sugar glass leads to stronger intra-protein hydrogen bonds in the matrix (left panel, dashed lines), whereas hydrogen bonds are weaker in aqueous solution (right). (B) Protein dilution by glassy protectant (top) versus aggregation of dry protein in the absence of protectants (bottom).

Furthermore, by disproportionately weakening the protein’s hydrogen bonds in the native and denatured states with the surrounding solution, polyols and sugars can effectively mitigate the loss of hydrogen bonds between the protein and its environment upon folding.241 The implication is that even in an environment that affords the protein fewer and weaker hydrogen bonds (as is typically the case for glass formers), native state stability can be enhanced simply by destabilizing the unfolded state more than the folded state.

The protection provided by sugar glass against unfolding is particularly notable for residues engaged in internal protein hydrogen bonds, especially residues fostering two or more hydrogen bonds.238 DSC and TGA, together with liquid-observed vapor exchange NMR spectroscopy reveal that protection afforded to residues by the sugar glass is linked to the number of intramolecular hydrogen bonds within the protein.238 This protection implies an increase in native state stability within the glass, driven by the reinforcement of intra-protein hydrogen bonds, compared to the liquid state. This strengthening of intra-protein hydrogen bonds appears to be a consistent trend among sugar glass formers.

5.3. Inter-Protein Interactions

In dilute solution, proteins tend to aggregate when interactions between neighboring proteins become more favorable than their interactions with the solvent.399 Most cells maintain an extremely concentrated environment at their normal hydration levels,27,400 causing proteins to be practically supersaturated.401 Even before desiccation, a small decrease in water content often leads to aggregation. Under extreme desiccation, where most of the water is removed, proteins are left with a minimal hydration layer that barely separates one protein from another. Consequently, desiccation can exacerbate aggregation by bringing proteins into close proximity, increasing intermolecular interactions and the likelihood of sticking to each other.

Molecular glass matrices modulate the inter-protein interactions that lead to aggregation at elevated protein concentrations. Molecular glasses have been shown to mitigate aggregation of a diverse range of proteins.51,402406 A factor contributing to this protection is the physical separation facilitated by the matrix (Figure 11B).399 In essence, diluting the protein with the protectant maintains some distance between proteins, thus protecting them from aggregation.407 For instance, simulations demonstrate that trehalose in its glassy state weakens inter-protein hydrogen bonds, thereby limiting protein–protein contacts.345,408

In molecular glasses, protection from aggregation relies not only on protein dilution but also on enhancement of protein stability. Because inter-protein interactions often involve nonspecific hydrophobic forces,40 these short-range interactions contribute and can even dominate aggregation at high protein concentrations.409 Conversely, aggregation is mitigated by additives that exhibit limited binding to the protein interface, because they promote protein compaction and thus reduce interactions between exposed hydrophobic protein moieties.410 Specific interactions, such as electrostatic forces and hydrogen bonds, may also contribute to protein aggregation, but their impact is less predictable because they strongly depends on the specific protein.238,399,411

Aggregation, including the formation of amyloids, generally proceeds from partially unfolded protein states.412,413 Consequently, increasing the stability of the native state, which reduces the frequency of unfolding events, also alleviates aggregation.240,414 Notably, during desiccation, including in processes such as lyophilization, unfolding can lead to aggregation, underscoring the advantage of inhibiting conformational changes.406 Inhibition of conformational changes can be achieved by adding sugars that stabilize proteins.238 For instance, SAXS measurements of lysozyme dried in sucrose suggest the preservation of its native state.106 In contrast, lysozyme dried in the absence of sugar displays significant structural distortions compared to the native state in solution.415 This observation suggests that, indeed, diluting the protein with sugars in the glassy state achieves both protection from aggregation and increased protein stability by mitigating destabilizing protein–protein interactions.

6. CONCLUSIONS

Desiccation protection is essential for the survival of organisms experiencing harsh conditions and imperative in technological applications in the food, pharma, and biomedical industries. Organisms have evolved various strategies to protect their macromolecules from the detrimental effect of drying. One of the most common invokes the formation of amorphous glassy matrices made from molecularly small metabolites (most notably sugars) alone or in combination with intrinsically disordered proteins (e.g., LEA and CAHS proteins). The idea that nonvolatile additives that form amorphous vitrified media can overcome the detrimental impacts of drying is shown by their implementation in industry. A host of molecular glassy matrices, silica glass, and aerogels have been suggested as protective media. Due to their simplicity of application, the use of glassy and amorphous media of nonvolatile solutes has become an attractive solution, and research in this area is poised to result in the discovery of robust, hypothesis-driven strategies of protection.

Despite advances in developing glasses for desiccation protection, the molecular mechanisms of protection remain controversial and comprise an active area of research. Many of the mechanisms described in this review highlight the difficulty in pinpointing their details. The extent of involvement of the matrix and residual water as well as the details of the “fight over water” between protein and matrix elements are difficult to deconvolute. One problem is the separation of time and length scales. In solution, proteins fold much more slowly than the solvent relaxes around them, whereas in glasses, relaxation slows, and proteins may even seem immobile, yet experiments show that ensembles of protein configurations correspond, at least in part, to native and denatured states.

Another persistent question is the quality and degree of protection and preservation that glasses afford. It has been known for years that trehalose has been selected as a protectant by many organisms, yet what makes this sugar superior to others is not completely understood. More generally, it is unclear which glass properties most closely correspond to the protection efficacy. It is tempting to choose easily measured properties, such as the glass transition temperature or molecular inertness, but unfortunately these properties do not always correspond to the most protective glass formers.

Beyond understanding how the glass impacts the embedded proteins, a relevant aspect that has not been widely explored is the way that proteins modify the properties of the matrix. This facet is important because in the vitrified state, the ratio of protein to glass-forming molecules is large, suggesting that almost every glass former is in close contact with a protein. It is hard to imagine how the glass properties would not then be impacted by the protein. Can these effects teach us about the properties that make a particularly good protectant?

Another important aspect that is only beginning to be explored is the advantages afforded by mixtures of glass formers. Specifically, mixtures of sugars and plasticizers such as glycerol are known to improve the protective efficacy of the glass, but the mechanism is unresolved. It is also known that some protective proteins form gels that may act in combination with other additives such as trehalose. Could the cohabitation of gel forming proteins and molecularly small glass formers lead to better protective properties?

The glassy matrix is complex and poses many challenges, both theoretical and experimental, yet the importance of its protective propensity indicates that a better molecular-level understanding of how the matrix stabilizes proteins will lead to a better grasp of biology and thus pave the way to the development of improved glasses with an even wider range of applications.

Acknowledgments

G.I.O. thanks the Harvey M. Krueger Family Center for Nanoscience & Nanotechnology for their fellowship. G.J.P. and D.H. thank the United States-Israel Binational Science Foundation for support (BSF-2017063 and BSF-2023649). Support from the Israel science foundation to D.H. (ISF Grant No. 1207/21) is gratefully acknowledged. G.J.P. thanks the US National Science Foundation for support (CHE-2203505 and MCB-2335137). The Fritz Haber Research Center is supported by the Minerva Foundation.

Glossary

LIST OF ABBREVIATIONS

AG

Adam and Gibbs (model)

CAHS

cytosolic abundant heat soluble

CD

circular dichroism (spectroscopy)

CRR

cooperatively rearranging region

GD

Gibbs and DiMarzio (model)

DSC

differential scanning calorimetry

DVS

dynamic vapor sorption

FTIR

Fourier-transform infrared (spectroscopy)

GB1

B1 domain of streptococcal protein G

GBP

Escherichia coli glucose binding protein

LEA

late embryogenesis abundant

LS

light scattering

MD

molecular dynamics

NMR

nuclear magnetic resonance (spectroscopy)

NS

not stated

OD

oxidative damage

PEG

polyethylene glycol

SAXS

small-angle X-ray scattering

SRCD

synchrotron radiation CD

TEOS

tetraethoxysilane

TGA

thermogravimetric analysis

TMOS

tetramethoxysilane

TNM

Tool–Narayanaswamy–Moynihan (model)

VFT

Volgel–Flucher–Tammann (model)

WAXS

wide-angle X-ray scattering

WLF

Williams–Landel–Ferry (model)

Biographies

Gil I. Olgenblum earned his B.Sc. and M.Sc. from the Hebrew University of Jerusalem, Israel, researching protein stability and preservation under desiccation. He is pursuing his Ph.D. in Chemistry at the Hebrew University, focusing on protein crowding in concentrated solution and in dry media.

Brent O. Hutcheson earned a B.S. in Biochemistry from American University in Washington, D.C., and is pursuing a Ph.D. in Chemistry from UNC Chapel Hill. His dissertation research focuses on the preservation of proteins in the dry state.

Gary J. Pielak earned a B.A. in Chemistry from Bradley University in Peoria, Illinois, and a Ph.D. in Biochemistry from Washington State University in Pullman, Washington. He was a postdoctoral fellow in the laboratory of Michael Smith at the University of British Columbia in Vancouver, Canada, and then the laboratory of Robert J. P. Williams at the University of Oxford. Gary is Kenan Distinguished Professor of Chemistry, Biochemistry and Biophysics at the University of North Carolina at Chapel Hill. His research focuses on protein chemistry in living cells and the protection of dry proteins.

Daniel Harries received a B.Sc. in Chemistry and a Ph.D. in Theoretical Chemistry from the Hebrew University of Jerusalem, Israel. He was a postdoctoral fellow in the laboratory of V. Adrian Parsegian at the National Institutes of Health in Bethesda, Maryland. Daniel currently holds the Dr. & Mrs. Philip Gotlieb Chair in Physical Chemistry at the Institute of Chemistry and the Fritz Haber Research Center at Hebrew University. His main research interests involve interactions between macromolecules and how these are modified in biologically realistic solutions that are stressed, confined, or crowded by numerous additional cosolutes.

The authors declare no competing financial interest.

Special Issue

Published as part of Chemical Reviewsvirtual special issue “Molecular Crowding”.

References

  1. Cheung L. K. Y.; Sanders A. D.; Pratap-Singh A.; Dee D. R.; Dupuis J. H.; Baldelli A.; Yada R. Y. Effects of High Pressure on Protein Stability, Structure, and Function—Theory and Applications. Effect of High-Pressure Technologies on Enzymes: Science and Applications 2023, 19–48. 10.1016/B978-0-323-98386-0.00005-1. [DOI] [Google Scholar]
  2. Huyghues-Despointes B. M. P.; Scholtz J. M.; Pace C. N. Protein Conformational Stabilities Can Be Determined from Hydrogen Exchange Rates. Nat. Struct. Biol. 1999, 6, 910–912. 10.1038/13273. [DOI] [PubMed] [Google Scholar]
  3. Gorania M.; Seker H.; Haris P. I. Predicting a Protein’s Melting Temperature from its Amino Acid Sequence. Conf. Proc. IEEE. Eng. Med. Biol. Soc. 2010, 1820–1823. [DOI] [PubMed] [Google Scholar]
  4. Stellwagen E.; Wilgus H. Relationship of Protein Thermostability to Accessible Surface Area. Nature 1978, 275, 342–343. 10.1038/275342a0. [DOI] [PubMed] [Google Scholar]
  5. Xie G.; Timasheff S. N. The Thermodynamic Mechanism of Protein Stabilization by Trehalose. Biophys. Chem. 1997, 64, 25–43. 10.1016/S0301-4622(96)02222-3. [DOI] [PubMed] [Google Scholar]
  6. Xie G.; Timasheff S. N. Temperature Dependence of the Preferential Interactions of Ribonuclease A in Aqueous Co-solvent Systems: Thermodynamic Analysis. Protein Sci. 1997, 6, 222–232. 10.1002/pro.5560060124. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Becktel W. J.; Schellman J. A. Protein Stability Curves. Biopolymers 1987, 26, 1859–1877. 10.1002/bip.360261104. [DOI] [PubMed] [Google Scholar]
  8. Aune K. C.; Tanford C. Thermodynamics of the Denaturation of Lysozyme by Guanidine Hydrochloride. I. Dependence on pH at 25°. Biochem. 1969, 8, 4579–4585. 10.1021/bi00839a052. [DOI] [PubMed] [Google Scholar]
  9. De Virgilo C.; Hottiger T.; Dominguez J.; Boller T.; Wiemken A. The Role of Trehalose Synthesis for the Acquisition of Thermotolerance in Yeast. Eur. J. Biochem. 1994, 219, 179–186. 10.1111/j.1432-1033.1994.tb19928.x. [DOI] [PubMed] [Google Scholar]
  10. Yancey P.; Organic H. Osmolytes as Compatible, Metabolic and Counteracting Cytoprotectants in High Osmolarity and Other Stresses. J. Exp. Biol. 2005, 208, 2819–2830. 10.1242/jeb.01730. [DOI] [PubMed] [Google Scholar]
  11. Isom D. G.; Castañeda C. A.; Cannon B. R.; Velu P. D.; García-Moreno E B. Charges in the Hydrophobic Interior of Proteins. Proc. Natl. Acad. Sci. U.S.A. 2010, 107, 16096–16100. 10.1073/pnas.1004213107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Matthew J. B.; Gurd F. R. N.; Garcia-Moreno B. E.; Flanagan M. A.; March K. L.; Shire S. J. pH-Dependent Processes in Protein. Crit. Rev. Biochem. Mol. Biol. 1985, 18, 91–197. 10.3109/10409238509085133. [DOI] [PubMed] [Google Scholar]
  13. Tanford C. The Interpretation of Hydrogen Ion Titration Curves of Proteins. Adv. Protein Chem. 1963, 17, 69–165. 10.1016/S0065-3233(08)60052-2. [DOI] [Google Scholar]
  14. Yoshinaga K.; Yoshioka H.; Kurosaki H.; Hirasawa M.; Uritani M.; Hasegawa K. Protection by Trehalose of DNA from Radiation Damage. Biosci., Biotechnol., Biochem. 1997, 61, 160–161. 10.1271/bbb.61.160. [DOI] [PubMed] [Google Scholar]
  15. Hochachka P. W.; Somero G. N.. Biochemical Adaptation: Mechanism and Process in Physiological Evolution. Biochem. Mol. Biol. Educ. 2002, 30, 215–216. 10.1002/bmb.2002.494030030071 [DOI] [Google Scholar]
  16. Wiemken A. Trehalose in Yeast, Stress Protectant Rather than Reserve Carbohydrate. Antonie Van Leeuwenhoek 1990, 58, 209–217. 10.1007/BF00548935. [DOI] [PubMed] [Google Scholar]
  17. Chen T.; Dave K.; Gruebele M. Pressure- and Heat-Induced Protein Unfolding in Bacterial Cells: Crowding vs Sticking. FEBS Lett. 2018, 592, 1357–1365. 10.1002/1873-3468.13025. [DOI] [PubMed] [Google Scholar]
  18. Luong T. Q.; Kapoor S.; Winter R. Pressure—A Gateway to Fundamental Insights into Protein Solvation, Dynamics, and Function. ChemPhysChem 2015, 16, 3555–3571. 10.1002/cphc.201500669. [DOI] [PubMed] [Google Scholar]
  19. Meersman F.; Daniel I.; Bartlett D. H.; Winter R.; Hazael R.; McMillan P. F. High-Pressure Biochemistry and Biophysics. Rev. Mineral. Geochem. 2013, 75, 607–648. 10.2138/rmg.2013.75.19. [DOI] [Google Scholar]
  20. Ravindra R.; Winter R. On the Temperature–Pressure Free-Energy Landscape of Proteins. ChemPhysChem 2003, 4, 359–365. 10.1002/cphc.200390062. [DOI] [PubMed] [Google Scholar]
  21. Sakai V. G.; Khodadadi S.; Cicerone M. T.; Curtis J. E.; Sokolov A. P.; Roh J. H. Solvent Effects on Protein Fast Dynamics: Implications for Biopreservation. Soft Matter 2013, 9, 5336–5340. 10.1039/c3sm50492a. [DOI] [Google Scholar]
  22. Tanford C. Contribution of Hydrophobic Interactions to the Stability of the Globular Conformation of Proteins. J. Am. Chem. Soc. 1962, 84, 4240–4247. 10.1021/ja00881a009. [DOI] [Google Scholar]
  23. Graziano G. Contrasting the Hydration Thermodynamics of Methane and Methanol. Phys. Chem. Chem. Phys. 2019, 21, 21418–21430. 10.1039/C9CP03213D. [DOI] [PubMed] [Google Scholar]
  24. Camilloni C.; Bonetti D.; Morrone A.; Giri R.; Dobson C. M.; Brunori M.; Gianni S.; Vendruscolo M. Towards a Structural Biology of the Hydrophobic Effect in Protein Folding. Sci. Rep. 2016, 6, 28285. 10.1038/srep28285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Wiśniewski J. R.; Hein M. Y.; Cox J.; Mann M. A “Proteomic Ruler” for Protein Copy Number and Concentration Estimation without Spike-in Standards. Mol. Cell. Proteomics 2014, 13, 3497–3506. 10.1074/mcp.M113.037309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Brown G. C. Total Cell Protein Concentration as an Evolutionary Constraint on the Metabolic Control Distribution in Cells. J. Theor. Biol. 1991, 153, 195–203. 10.1016/S0022-5193(05)80422-9. [DOI] [PubMed] [Google Scholar]
  27. Theillet F. X.; Binolfi A.; Frembgen-Kesner T.; Hingorani K.; Sarkar M.; Kyne C.; Li C.; Crowley P. B.; Gierasch L.; Pielak G. J.; et al. Physicochemical Properties of Cells and their Effects on Intrinsically Disordered Proteins (IDPs). Chem. Rev. 2014, 114, 6661–6714. 10.1021/cr400695p. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Oh S.; Lee C. H.; Yang W.; Li A.; Mukherjee A.; Basan M.; Ran C.; Yin W.; Tabin C. J.; Fu D.;et al. Protein and Lipid Mass Concentration Measurement in Tissues by Stimulated Raman Scattering Microscopy. Proc. Natl. Acad. Sci. U.S.A. 2022, 119, e2117938119. 10.1073/pnas.2117938119 [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Roscoe B. P.; Thayer K. M.; Zeldovich K. B.; Fushman D.; Bolon D. N. A. Analyses of the Effects of All Ubiquitin Point Mutants on Yeast Growth Rate. J. Mol. Biol. 2013, 425, 1363. 10.1016/j.jmb.2013.01.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Pace C. N.; Treviñ S.; Scholtz J. M. Protein Structure, Stability and Solubility in Water and Other Solvents. Philos.Trans. R. Soc. B 2004, 359, 1225–1235. 10.1098/rstb.2004.1500. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Marianayagam N. J.; Jackson S. E. Native-State Dynamics of the Ubiquitin Family: Implications for Function and Evolution. J. R Soc. Interface 2005, 2, 47. 10.1098/rsif.2004.0025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Zhou Y. Salt Effects on Protein Titration and Binding. J. Phys. Chem. B 1998, 102, 10615–10621. 10.1021/jp982542x. [DOI] [Google Scholar]
  33. Erickson H. P. Size and Shape of Protein Molecules at the Nanometer Level Determined by Sedimentation, Gel Filtration, and Electron Microscopy. Biol. Proced Online 2009, 11, 32–51. 10.1007/s12575-009-9008-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Grierson D.; Covey S. N.. Organization of Nuclear DNA. Plant Molecular Biology. Tertiary Level Biology; Springer: Boston, MA, 1988; Chapter 1, pp 1–21. 10.1007/978-1-4615-3666-6_1 [DOI] [Google Scholar]
  35. Grzyb T.; Skłodowska A. Introduction to Bacterial Anhydrobiosis: A General Perspective and the Mechanisms of Desiccation-Associated Damage. Microorganisms 2022, 10, 432. 10.3390/microorganisms10020432. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Romero-Perez P. S.; Dorone Y.; Flores E.; Sukenik S.; Boeynaems S.. When Phased without Water: Biophysics of Cellular Desiccation, from Biomolecules to Condensates. Chem. Rev. 2023, 123, 9010. 10.1021/acs.chemrev.2c00659 [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Bergfreund J.; Bertsch P.; Fischer P. Adsorption of Proteins to Fluid Interfaces: Role of the Hydrophobic Subphase. J. Colloid Interface Sci. 2021, 584, 411–417. 10.1016/j.jcis.2020.09.118. [DOI] [PubMed] [Google Scholar]
  38. Zhao Y.; Cieplak M. Proteins at Air–Water and Oil–Water Interfaces in an All-Atom Model. Phys. Chem. Chem. Phys. 2017, 19, 25197–25206. 10.1039/C7CP03829A. [DOI] [PubMed] [Google Scholar]
  39. D’Imprima E.; Floris D.; Joppe M.; Sánchez R.; Grininger M.; Kühlbrandt W. Protein Denaturation at the Air-Water Interface and How to Prevent It. Elife 2019, 8, e42747. 10.7554/eLife.42747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Arakawa T.; Ejima D.; Akuta T. Protein Aggregation Under High Concentration/Density State During Chromatographic and Ultrafiltration Processes. Int. J. Biol. Macromol. 2017, 95, 1153–1158. 10.1016/j.ijbiomac.2016.11.005. [DOI] [PubMed] [Google Scholar]
  41. Hofmann M.; Winzer M.; Weber C.; Gieseler H. Prediction of Protein Aggregation in High Concentration Protein Solutions Utilizing Protein-Protein Interactions Determined by Low Volume Static Light Scattering. J. Pharm. Sci. 2016, 105, 1819–1828. 10.1016/j.xphs.2016.03.022. [DOI] [PubMed] [Google Scholar]
  42. Frank F.Freeze-drying of Pharmaceuticals and Biopharmaceuticals RSC Publishing: Cambridge, UK, 2007, Chapter 1, pp 1–12.. [Google Scholar]
  43. Morris J. E. Hydration, Its Reversibility, and the Beginning of Development in the Brine Shrimp, Artemia salina. Comp Biochem Physiol A Physiol 1971, 39, 843–857. 10.1016/0300-9629(71)90205-2. [DOI] [Google Scholar]
  44. Simmons C. P.; Farrar J. J.; van Vinh Chau N.; Wills B. Dengue. N Engl J. Med. 2012, 366, 1423–1455. 10.1056/NEJMra1110265. [DOI] [PubMed] [Google Scholar]
  45. Prasad A.; Sreedharan S.; Bakthavachalu B.; Laxman S. Aedes aegypti Eggs Use Rewired Polyamine and Lipid Metabolism to Survive Extreme Desiccation. bioRxiv 2022, 12 (30), 522323. 10.1101/2022.12.30.522323. [DOI] [Google Scholar]
  46. Westh P.; Ramløv H.. Trehalose Accumulation in the Tardigrade Adorybiotus coronifer During Anhydrobiosis. J. Exp. Zool. 1991, 258, 303. 10.1002/jez.1402580305 [DOI] [Google Scholar]
  47. Schill R. O., Ed. Water Bears: The Biology of Tardigrades; Zoological Monographs Vol. 2; Springer, 2018. [Google Scholar]
  48. Fox K. C. Putting Proteins Under Glass. Science (1979) 1995, 267, 1922–1923. 10.1126/science.7701317. [DOI] [PubMed] [Google Scholar]
  49. Smolikova G.; Leonova T.; Vashurina N.; Frolov A.; Medvedev S. Desiccation Tolerance as the Basis of Long-Term Seed Viability. Int. J. Mol. Sci. 2021, 22, 101. 10.3390/ijms22010101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Léonard-Nevers V.; Marton Z.; Lamare S.; Hult K.; Graber M. Understanding Water Effect on Candida antarctica Lipase B Activity and Enantioselectivity Towards Secondary Alcohols. J. Mol. Catal. B Enzym 2009, 59, 90–95. 10.1016/j.molcatb.2009.01.008. [DOI] [Google Scholar]
  51. Costantino H. R.; Griebenow K.; Mishra P.; Langer R.; Klibanov A. M. Fourier-transform infrared spectroscopic investigation of protein stability in the lyophilized form. Biochim. Biophys. Acta, Prot. Struc. Mol. Enz. 1995, 1253, 69–74. 10.1016/0167-4838(95)00156-O. [DOI] [PubMed] [Google Scholar]
  52. Onwe R. O.; Onwosi C. O.; Ezugworie F. N.; Ekwealor C. C.; Okonkwo C. C. Microbial Trehalose Boosts the Ecological Fitness of Biocontrol Agents, the Viability of Probiotics During Long-Term Storage and Plants Tolerance to Environmental-Driven Abiotic Stress. Sci. Total Environ. 2022, 806, 150432. 10.1016/j.scitotenv.2021.150432. [DOI] [PubMed] [Google Scholar]
  53. Battaglia M.; Covarrubias A. A. Late Embryogenesis Abundant (LEA) Proteins in Legumes. Front. Plant Sci. 2013, 4, 45960. 10.3389/fpls.2013.00190. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Gao J.; Lan T. Functional Characterization of the Late Embryogenesis Abundant (LEA) Protein Gene Family From Pinus tabuliformis (Pinaceae) in Escherichia coli. Sci. Rep. 2016, 6, 19467. 10.1038/srep19467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Leprince O.; Buitink J. Introduction to Desiccation Biology: From Old Borders to New Frontiers. Planta 2015, 242, 369–378. 10.1007/s00425-015-2357-6. [DOI] [PubMed] [Google Scholar]
  56. Caramelo J. J.; Iusem N. D. When Cells Lose Water: Lessons From Biophysics and Molecular Biology. Prog. Biophys. Mol. Biol. 2009, 99, 1–6. 10.1016/j.pbiomolbio.2008.10.001. [DOI] [PubMed] [Google Scholar]
  57. Yamaguchi A.; Tanaka S.; Yamaguchi S.; Kuwahara H.; Takamura C.; Imajoh-Ohmi S.; Horikawa D. D.; Toyoda A.; Katayama T.; Arakawa K.; et al. Two Novel Heat-Soluble Protein Families Abundantly Expressed in an Anhydrobiotic Tardigrade. PLoS One 2012, 7, e44209. 10.1371/journal.pone.0044209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Boothby T. C.; Tapia H.; Brozena A. H.; Piszkiewicz S.; Smith A. E.; Giovannini I.; Rebecchi L.; Pielak G. J.; Koshland D.; Goldstein B. Tardigrades Use Intrinsically Disordered Proteins to Survive Desiccation. Mol. Cell 2017, 65, 975–984. 10.1016/j.molcel.2017.02.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Boothby T. C.; Pielak G. J. Intrinsically Disordered Proteins and Desiccation Tolerance: Elucidating Functional and Mechanistic Underpinnings of Anhydrobiosis. Bioessays 2017, 39, 1700119. 10.1002/bies.201700119. [DOI] [PubMed] [Google Scholar]
  60. Eicher J.; Hutcheson B. O.; Pielak G. J. Properties of a Tardigrade Desiccation-Tolerance Protein Aerogel. Biophys. J. 2023, 122, 2500–2505. 10.1016/j.bpj.2023.05.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Eicher J. E.; Brom J. A.; Wang S.; Sheiko S. S.; Atkin J. M.; Pielak G. J. Secondary Structure and Stability of a Gel-Forming Tardigrade Desiccation-Tolerance Protein. Protein Sci. 2022, 31, e4495. 10.1002/pro.4495. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Wang S.; Eicher J.; Pielak G. J. Trifluoroethanol and the Behavior of a Tardigrade Desiccation-Tolerance Protein. Protein Sci. 2023, 32, e4716. 10.1002/pro.4716. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Malki A.; Teulon J. M.; Camacho-Zarco A. R.; Chen S.-w. W.; Adamski W.; Maurin D.; Salvi N.; Pellequer J. L.; Blackledge M. Intrinsically Disordered Tardigrade Proteins Self-Assemble into Fibrous Gels in Response to Environmental Stress. Angew. Chem., Int. Ed. 2022, 61, e202109961. 10.1002/anie.202109961. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Hoekstra F. A.; Golovina E. A.; Buitink J. Mechanism of Plant Desiccation Tolerance. Trends Plant Sci. 2001, 6, 431–438. 10.1016/S1360-1385(01)02052-0. [DOI] [PubMed] [Google Scholar]
  65. Dong A.; Prestrelski S. J.; Allison S. D.; Carpenter J. F. Infrared Spectroscopic Studies of Lyophilization- and Temperature-Induced Protein Aggregation. J. Pharm. Sci. 1995, 84, 415–424. 10.1002/jps.2600840407. [DOI] [PubMed] [Google Scholar]
  66. Crilly C. J.; Brom J. A.; Warmuth O.; Esterly H. J.; Pielak G. J. Protection by Desiccation-Tolerance Proteins Probed at the Residue Level. Protein Sci. 2022, 31, 396–406. 10.1002/pro.4231. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Piszkiewicz S.; Gunn K. H.; Warmuth O.; Propst A.; Mehta A.; Nguyen K. H.; Kuhlman E.; Guseman A. J.; Stadmiller S. S.; Boothby T. C.; et al. Protecting Activity of Desiccated Enzymes. Protein Sci. 2019, 28, 941–951. 10.1002/pro.3604. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Piszkiewicz S.; Pielak G. J. Protecting Enzymes from Stress-Induced Inactivation. Biochemistry 2019, 58, 3825–3833. 10.1021/acs.biochem.9b00675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Welsh D. T.; Herbert R. A. Osmotically Induced Intracellular Trehalose, but not Glycine Betaine Accumulation Promotes Desiccation Tolerance in Escherichia coli. FEMS Microbiol. Lett. 1999, 174, 57–63. 10.1111/j.1574-6968.1999.tb13549.x. [DOI] [PubMed] [Google Scholar]
  70. Yancey P. H.; Clark M. E.; Hand S. C.; Bowlus R. D.; Somero G. N. Living with Water Stress: Evolution of Osmolyte Systems. Science (1979) 1982, 217, 1214–1222. 10.1126/science.7112124. [DOI] [PubMed] [Google Scholar]
  71. Farrant J. M.; Moore J. P. Programming Desiccation-Tolerance: From Plants to Seeds to Resurrection Plants. Curr. Opin. Plant Biol. 2011, 14, 340–345. 10.1016/j.pbi.2011.03.018. [DOI] [PubMed] [Google Scholar]
  72. Gasulla F.; Del Campo E. M.; Casano L. M.; Guéra A. Advances in Understanding of Desiccation Tolerance of Lichens and Lichen-Forming Algae. Plants 2021, 10, 807. 10.3390/plants10040807. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Zhang Q.; Song X.; Bartels D.; Oliveira M.; Abreu I. A.; Wisniewski J. R. Enzymes and Metabolites in Carbohydrate Metabolism of Desiccation Tolerant Plants. Proteomes 2016, 4, 40. 10.3390/proteomes4040040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Palmqvist K. Tansley Review No. 117 Carbon Economy in Lichens. New Phytol 2000, 148, 11–36. 10.1046/j.1469-8137.2000.00732.x. [DOI] [PubMed] [Google Scholar]
  75. Sadowsky A.; Mettler-Altmann T.; Ott S. Metabolic Response to Desiccation Stress in Strains of Green Algal Photobionts (Trebouxia) From Two Antarctic Lichens of Southern Habitats. Phycologia 2016, 55, 703–714. 10.2216/15-127.1. [DOI] [Google Scholar]
  76. Hell A. F.; Gasulla F.; González-Houcarde M.; Pelegrino M. T.; Seabra A. B.; del Campo E. M.; Casano L. M.; Centeno D. C. Polyols-Related Gene Expression Is Affected by Cyclic Desiccation in Lichen Microalgae. Environ. Exp. Bot. 2021, 185, 104397. 10.1016/j.envexpbot.2021.104397. [DOI] [Google Scholar]
  77. Buitink J.; Leprince O. Intracellular Glasses and Seed Survival in the Dry State. C.R. Biol. 2008, 331, 788–795. 10.1016/j.crvi.2008.08.002. [DOI] [PubMed] [Google Scholar]
  78. Koster K. L. Glass Formation and Desiccation Tolerance in Seeds. Plant Physiol 1991, 96, 302–304. 10.1104/pp.96.1.302. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Thorat L. J.; Gaikwad S. M.; Nath B. B. Trehalose as an Indicator of Desiccation Stress in Drosophila melanogaster Larvae: A Potential Marker of Anhydrobiosis. Biochem. Biophys. Res. Commun. 2012, 419, 638–642. 10.1016/j.bbrc.2012.02.065. [DOI] [PubMed] [Google Scholar]
  80. Morano K. A. Anhydrobiosis: Drying Out with Sugar. Curr. Biol. 2014, 24, R1121–R1123. 10.1016/j.cub.2014.10.022. [DOI] [PubMed] [Google Scholar]
  81. Kasianchuk N.; Rzymski P.; Kaczmarek Ł. The Biomedical Potential of Tardigrade Proteins: A Review. Biomed. Pharmacother 2023, 158, 114063. 10.1016/j.biopha.2022.114063. [DOI] [PubMed] [Google Scholar]
  82. Hibshman J. D.; Clegg J. S.; Goldstein B. Mechanisms of Desiccation Tolerance: Themes and Variations in Brine Shrimp, Roundworms, and Tardigrades. Front Physiol 2020, 11, 592016. 10.3389/fphys.2020.592016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Crowe J. H.; Madin K. A. C.; Loomis S. H. Anhydrobiosis in Nematodes: Metabolism During Resumption of Activity. J. Exp. Zool. 1977, 201, 57–64. 10.1002/jez.1402010107. [DOI] [Google Scholar]
  84. Cassada R. C.; Russell R. L. The Dauerlarva, a Post-Embryonic Developmental Variant of the Nematode Caenorhabditis elegans. Dev. Biol. 1975, 46, 326–342. 10.1016/0012-1606(75)90109-8. [DOI] [PubMed] [Google Scholar]
  85. Erkut C.; Kurzchalia T. V. The C. elegans dauer Larva as a Paradigm to Study Metabolic Suppression and Desiccation Tolerance. Planta 2015, 242, 389–396. 10.1007/s00425-015-2300-x. [DOI] [PubMed] [Google Scholar]
  86. Moehlenbrock M. J.; Minteer S. D.. Introduction to the Field of Enzyme Immobilization and Stabilization. Methods Mol. Biol.; Humana Press Inc., 2017; Vol. 1504. [DOI] [PubMed] [Google Scholar]
  87. Coche-Guérente L.; Cosnier S.; Labbé P. Sol–Gel Derived Composite Materials for the Construction of Oxidase/Peroxidase Mediatorless Biosensors. Chem. Mater. 1997, 9, 1348–1352. 10.1021/cm960498x. [DOI] [Google Scholar]
  88. Lim J.; Malati P.; Bonet F.; Dunn B. Nanostructured Sol-Gel Electrodes for Biofuel Cells. J. Electrochem. Soc. 2007, 154, A140. 10.1149/1.2404904. [DOI] [Google Scholar]
  89. Nguyen D. T.; Smit M.; Dunn B.; Zink J. I. Stabilization of Creatine Kinase Encapsulated in Silicate Sol-Gel Materials and Unusual Temperature Effects on its Activity. Chem. Mater. 2002, 14, 4300–4306. 10.1021/cm020398t. [DOI] [Google Scholar]
  90. Hussain F.; Birch D. J. S.; Pickup J. C. Glucose Sensing Based on the Intrinsic Fluorescence of Sol-Gel Immobilized Yeast Hexokinase. Anal. Biochem. 2005, 339, 137–143. 10.1016/j.ab.2005.01.016. [DOI] [PubMed] [Google Scholar]
  91. Vinogradov V. V.; Avnir D.. Enzyme Renaturation to Higher Activity Driven by the Sol-Gel Transition: Carbonic Anhydrase. Sci. Rep, 2015, 5, 14411. 10.1038/srep14411 [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Braun S.; Rappoport S.; Zusman R.; Avnir D.; Ottolenghi M. Biochemically Active Sol-Gel Glasses: the Trapping of Enzymes. Mater. Lett. 1990, 10, 1–5. 10.1016/0167-577X(90)90002-4. [DOI] [Google Scholar]
  93. Braun S.; Shtelzer S.; Rappoport S.; Avnir D.; Ottolenghi M. Biocatalysis by Sol-Gel Entrapped Enzymes. J. Non Cryst. Solids 1992, 147–148, 739–743. 10.1016/S0022-3093(05)80708-2. [DOI] [Google Scholar]
  94. Altstein M.; Segev G.; Aharonson N.; Ben-Aziz O.; Turniansky A.; Avnir D. Sol-Gel-Entrapped Cholinesterases: A Microtiter Plate Method for Monitoring Anti-cholinesterase Compounds. J. Agric. Food Chem. 1998, 46, 3318–3324. 10.1021/jf9801015. [DOI] [Google Scholar]
  95. Ganonyan N.; Benmelech N.; Bar G.; Gvishi R.; Avnir D. Entrapment of Enzymes in Silica Aerogels. Mater. Today 2020, 33, 24–35. 10.1016/j.mattod.2019.09.021. [DOI] [Google Scholar]
  96. Peterson R. W.; Anbalagan K.; Tommos C.; Wand A. J. Forced Folding and Structural Analysis of Metastable Proteins. J. Am. Chem. Soc. 2004, 126, 9498–9499. 10.1021/ja047900q. [DOI] [PubMed] [Google Scholar]
  97. Senske M.; Smith A. E.; Pielak G. J. Protein Stability in Reverse Micelles. Angew. Chem. 2016, 128, 3650–3653. 10.1002/ange.201508981. [DOI] [PubMed] [Google Scholar]
  98. Cheng K.; Wu Q.; Zhang Z.; Pielak G. J.; Liu M.; Li C. Crowding and Confinement Can Oppositely Affect Protein Stability. ChemPhysChem 2018, 19, 3350–3355. 10.1002/cphc.201800857. [DOI] [PubMed] [Google Scholar]
  99. Sukenik S.; Ren P.; Gruebele M. Weak Protein-Protein Interactions in Live Cells are Quantified by Cell-Volume Modulation. Proc. Natl. Acad. Sci. U. S. A. 2017, 114, 6776–6781. 10.1073/pnas.1700818114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Harries D.; Rösgen J. A Practical Guide on How Osmolytes Modulate Macromolecular Properties. Methods Cell Biol. 2008, 84, 679–735. 10.1016/S0091-679X(07)84022-2. [DOI] [PubMed] [Google Scholar]
  101. Kandimalla V.; Tripathi V. S.; Ju H. Immobilization of Biomolecules in Sol–Gels: Biological and Analytical Applications. Crit. Rev. Anal. Chem. 2006, 36, 73–106. 10.1080/10408340600713652. [DOI] [Google Scholar]
  102. Eggers D. K.; Valentine J. S. Molecular Confinement Influences Protein Structure and Enhances Thermal Protein Stability. Protein Sci. 2001, 10, 250–261. 10.1110/ps.36201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Potnis C. S.; Grapperhaus C. A.; Gupta G. Investigating BioCaRGOS, a Sol-Gel Matrix for the Stability of Heme Proteins under Enzymatic Degradation and Low pH. ACS Omega 2023, 8, 32053–32059. 10.1021/acsomega.3c04012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Harper-Leatherman A. S.; Wallace J. M.; Rolison D. R.. Cytochrome c Stabilization and Immobilization in Aerogels. Methods Mol. Biol. (Clifton, N.J.); Humana Press: Totowa, NJ, 2011; Vol. 679. [DOI] [PubMed] [Google Scholar]
  105. Liao Y. H.; Brown M. B.; Nazir T.; Quader A.; Martin G. P. Effects of Sucrose and Trehalose on the Preservation of the Native Structure of Spray-Dried Lysozyme. Pharm. Res. 2002, 19, 1847–1853. 10.1023/A:1021445608807. [DOI] [PubMed] [Google Scholar]
  106. Bogdanova E.; Lages S.; Phan-Xuan T.; Kamal M. A.; Terry A.; Millqvist Fureby A.; Kocherbitov V. Lysozyme-Sucrose Interactions in the Solid State: Glass Transition, Denaturation, and the Effect of Residual Water. Mol. Pharmaceutics 2023, 20, 4664–4675. 10.1021/acs.molpharmaceut.3c00403. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Shokri S.; Shahkarami M. K.; Shafyi A.; Mohammadi A.; Esna-ashari F.; Hamta A. Evaluation of the Thermal Stability of Live-Attenuated Rubella Vaccine (Takahashi strain) Formulated and Lyophilized in Different Stabilizers. J. Virol. Methods 2019, 264, 18–22. 10.1016/j.jviromet.2018.08.013. [DOI] [PubMed] [Google Scholar]
  108. Sou T.; Morton D. A. V.; Williamson M.; Meeusen E. N.; Kaminskas L. M.; McIntosh M. P. Spray-Dried Influenza Antigen with Trehalose and Leucine Produces an Aerosolizable Powder Vaccine Formulation that Induces Strong Systemic and Mucosal Immunity after Pulmonary Administration. J. Aerosol Med. Pulm. Drug Delivery 2015, 28, 361–371. 10.1089/jamp.2014.1176. [DOI] [PubMed] [Google Scholar]
  109. Liao Y. H.; Brown M. B.; Quader A.; Martin G. P. Protective Mechanism of Stabilizing Excipients Against Dehydration in the Freeze-Drying of Proteins. Pharm. Res. 2002, 19, 1854–1861. 10.1023/A:1021497625645. [DOI] [PubMed] [Google Scholar]
  110. Wang B.; Tchessalov S.; Warne N. W.; Pikal M. J. Impact of Sucrose Level on Storage Stability of Proteins in Freeze-Dried Solids: I. Correlation of Protein-Sugar Interaction with Native Structure Preservation. J. Pharm. Sci. 2009, 98, 3131–3144. 10.1002/jps.21621. [DOI] [PubMed] [Google Scholar]
  111. Brogna R.; Oldenhof H.; Sieme H.; Figueiredo C.; Kerrinnes T.; Wolkers W. F. Increasing Storage Stability of Freeze-Dried Plasma Using Trehalose. PLoS One 2020, 15, e0234502. 10.1371/journal.pone.0234502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Garidel P.; Pevestorf B.; Bahrenburg S. Stability of Buffer-Free Freeze-Dried Formulations: A Feasibility Study of a Monoclonal Antibody at High Protein Concentrations. Eur. J. Pharm. Biopharm 2015, 97, 125–139. 10.1016/j.ejpb.2015.09.017. [DOI] [PubMed] [Google Scholar]
  113. López-Díez E. C.; Bone S. The Interaction of Trypsin with Trehalose: an Investigation of Protein Preservation Mechanisms. Biochim Biophys Acta Gen Subj 2004, 1673, 139–148. 10.1016/j.bbagen.2004.04.010. [DOI] [PubMed] [Google Scholar]
  114. Carpenter J. F.; Crowe L. M.; Crowe J. H. Stabilization of Phosphofructokinase with Sugars During Freeze-Drying: Characterization of Enhanced Protection in the Presence of Divalent Cations. Biochim Biophys Acta Gen Subj 1987, 923, 109–115. 10.1016/0304-4165(87)90133-4. [DOI] [PubMed] [Google Scholar]
  115. Haque M. A.; Chen J.; Aldred P.; Adhikari B. Drying and Denaturation Characteristics of Whey Protein Isolate in the Presence of Lactose and Trehalose. Food Chem. 2015, 177, 8–16. 10.1016/j.foodchem.2014.12.064. [DOI] [PubMed] [Google Scholar]
  116. Kembaren R.; Fokkink R.; Westphal A. H.; Kamperman M.; Kleijn J. M.; Borst J. W. Balancing Enzyme Encapsulation Efficiency and Stability in Complex Coacervate Core Micelles. Langmuir 2020, 36, 8494–8502. 10.1021/acs.langmuir.0c01073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Doussin S.; Birlirakis N.; Georgin D.; Taran F.; Berthault P. Novel Zwitterionic Reverse Micelles for Encapsulation of Proteins in Low-Viscosity Media. Chem.—Eur. J. 2006, 12, 4170–4175. 10.1002/chem.200501422. [DOI] [PubMed] [Google Scholar]
  118. Crowe J. H.; Crowe L. M.; Chapman D. Preservation of Membranes in Anhydrobiotic Organisms: The Role of Trehalose. Science (1979) 1984, 223, 701–703. 10.1126/science.223.4637.701. [DOI] [PubMed] [Google Scholar]
  119. Crowe J. H.; Hoekstra F. A.; Crowe L. M. Anhydrobiosis. Annu. Rev. Physiol. 1992, 54, 579–599. 10.1146/annurev.ph.54.030192.003051. [DOI] [PubMed] [Google Scholar]
  120. Crowe J. H.; Carpenter J. F.; Crowe L. M. The Role of Vitrification in Anhydrobiosis. Annu. Rev. Physiol. 1998, 60, 73–103. 10.1146/annurev.physiol.60.1.73. [DOI] [PubMed] [Google Scholar]
  121. Pfaender H. G.Schott Guide to Glass; Chapman & Hall, 1996. [Google Scholar]
  122. Elliott S. R.Amorphous Solids: An Introduction. InDefects and Disorder in Crystalline and Amorphous Solids; Catlow C. R. A., Eds.; NATO ASI Book Series Vol. 418; Springer: Dordrecht, 1994; pp 73–86. 10.1007/978-94-011-1942-9_4 [DOI] [Google Scholar]
  123. Angell C. A. Formation of Glasses from Liquids and Biopolymers. Science (1979) 1995, 267, 1924–1935. 10.1126/science.267.5206.1924. [DOI] [PubMed] [Google Scholar]
  124. Philipp H. R.Silicon Dioxide (SiO2) (Glass). Handbook of Optical Constants of Solids; Palik E. D., Ed.; Adademic Press, 1997; pp 749–763. 10.1016/B978-012544415-6.50038-8 [DOI] [Google Scholar]
  125. Soppe W.; van der Marel C.; van Gunsteren W. F.; den Hartog H. W. New Insights Into the Structure of B2O3 glass. J. Non-Cryst. Solids 1988, 103, 201–209. 10.1016/0022-3093(88)90199-8. [DOI] [Google Scholar]
  126. Zitkovsky I.; Boolchand P. Molecular Structure of As2S3 glass. J. Non-Cryst. Solids 1989, 114, 70–72. 10.1016/0022-3093(89)90071-9. [DOI] [Google Scholar]
  127. Van Uitert L. G.; Wemple S. H. ZnCl2 glass: A Potential Ultralow-Loss Optical Fiber Material. Appl. Phys. Lett. 1978, 33, 57–59. 10.1063/1.90189. [DOI] [Google Scholar]
  128. Baldwin C. M.; Almeida R. M.; Mackenzie J. D. Halide Glasses. J. Non-Cryst. Solids 1981, 43, 309–344. 10.1016/0022-3093(81)90101-0. [DOI] [Google Scholar]
  129. Qiao J.; Jia H.; Liaw P. K. Metallic Glass Matrix Composites. Mater. Sci. Eng.: R: Rep. 2016, 100, 1–69. 10.1016/j.mser.2015.12.001. [DOI] [Google Scholar]
  130. Fox T. G.; Flory P. J. The Glass Temperature and Related Properties of Polystyrene Influence of Molecular Weight. J. Polym. Sci. 1954, 14, 315–319. 10.1002/pol.1954.120147514. [DOI] [Google Scholar]
  131. Alford S.; Dole M. Specific Heat of Synthetic High Polymers. VI. A Study of the Glass Transition in Polyvinyl Chloride. J. Am. Chem. Soc. 1955, 77, 4774–4777. 10.1021/ja01623a024. [DOI] [Google Scholar]
  132. Cangialosi D.; Wübbenhorst M.; Groenewold J.; Mendes E.; Schut H.; Van Veen A.; Picken S. J. Physical aging of polycarbonate far below the glass transition temperature: Evidence for the diffusion mechanism. Phys. Rev. B Condens. Matter Mater. Phys. 2004, 70, 224213. 10.1103/PhysRevB.70.224213. [DOI] [Google Scholar]
  133. Debenedetti P. G.; Stanley H. E. Supercooled and Glassy Water. Phys. Today 2003, 56, 40–46. 10.1063/1.1595053. [DOI] [Google Scholar]
  134. Mishima O.; Stanley H. E. The Relationship Between Liquid, Supercooled and Glassy Water. Nature 1998, 396, 329–335. 10.1038/24540. [DOI] [Google Scholar]
  135. Handle P. H.; Loerting T.; Sciortino F. Supercooled and Glassy Water: Metastable Liquid(s), Amorphous Solid(s), and a No-Man’s Land. Proc. Natl. Acad. Sci. U.S.A. 2017, 114, 13336–13344. 10.1073/pnas.1700103114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Angell C. A. Liquid Fragility and the Glass Transition in Water and Aqueous Solutions. Chem. Rev. 2002, 102, 2627–2650. 10.1021/cr000689q. [DOI] [PubMed] [Google Scholar]
  137. Moynihan C. T.; Bressel R. D.; Angell C. A. Conductivity and Dielectric Relaxation in Concentrated Aqueous Lithium Chloride Solutions. J. Chem. Phys. 1971, 55, 4414–4424. 10.1063/1.1676768. [DOI] [Google Scholar]
  138. Moynihan C. T.; Balitactac N.; Boone L.; Litovitz T. A. Comparison of Shear and Conductivity Relaxation Times for Concentrated Lithium Chloride Solutions. J. Chem. Phys. 1971, 55, 3013–3019. 10.1063/1.1676531. [DOI] [Google Scholar]
  139. Angell C. A.; Bressel R. D. Fluidity and Conductance in Aqueous Electrolyte Solutions. An Approach from the Glassy State and High-Concentration Limit. I. Ca(NO3)2 Solutions. J. Phys. Chem. 1972, 76, 3244–3253. 10.1021/j100666a023. [DOI] [Google Scholar]
  140. Angell C. A.; Donnella J. Mechanical Collapse vs Ideal Glass Formation in Slowly Vitrified Solutions: A Plausibility Test. J. Chem. Phys. 1977, 67, 4560–4563. 10.1063/1.434597. [DOI] [Google Scholar]
  141. Shirke S.; Ludescher R. D. Molecular Mobility and the Glass Transition in Amorphous Glucose, Maltose, and Maltotriose. Carbohydr. Res. 2005, 340, 2654–2660. 10.1016/j.carres.2005.08.016. [DOI] [PubMed] [Google Scholar]
  142. Vanhal I.; Blond G. Impact of Melting Conditions of Sucrose on its Glass Transition Temperature. J. Agric. Food. Chem. 1999, 47, 4285–4290. 10.1021/jf981411q. [DOI] [PubMed] [Google Scholar]
  143. Chen T.; Fowler A.; Toner M. Literature Review: Supplemented Phase Diagram of the Trehalose–Water Binary Mixture. Cryobiology 2000, 40, 277–282. 10.1006/cryo.2000.2244. [DOI] [PubMed] [Google Scholar]
  144. Schröter K.; Donth E.; Schrö K.; Donth E. Viscosity and Shear Response at the Dynamic Glass Transition of Glycerol. J. Chem. Phys. 2000, 113, 9101–9108. 10.1063/1.1319616. [DOI] [Google Scholar]
  145. Win K. Z.; Menon N. Glass Transition of Glycerol in the Volume-Temperature Plane. Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 2006, 73, 040501 10.1103/PhysRevE.73.040501. [DOI] [PubMed] [Google Scholar]
  146. Gao C.; Zhou G. Y.; Xu Y.; Hua T. C. Glass Transition and Enthalpy Relaxation of Ethylene Glycol and its Aqueous Solution. Thermochim. Acta 2005, 435, 38–43. 10.1016/j.tca.2005.03.024. [DOI] [Google Scholar]
  147. Birge N. O. Specific-Heat Spectroscopy of Glycerol and Propylene Glycol Near the Glass Transition. Phys. Rev. B 1986, 34, 1631. 10.1103/PhysRevB.34.1631. [DOI] [PubMed] [Google Scholar]
  148. Bogdanova E.; Kocherbitov V.. Assessment of Activation Energy of Enthalpy Relaxation in Sucrose-Water System: Effects of DSC Cycle Type and Sample Thermal History. J. Thermal Anal. Calorim. 2022, 147, 9695–9709. [Google Scholar]
  149. Kawai K.; Hagiwara T.; Takai R.; Suzuki T. Comparative Investigation by Two Analytical Approaches of Enthalpy Relaxation for Glassy Glucose, Sucrose, Maltose, and Trehalose. Pharm. Res. 2005, 22, 490–495. 10.1007/s11095-004-1887-6. [DOI] [PubMed] [Google Scholar]
  150. Hancock B. C.; Shamblin S. L.; Zografi G. Molecular Mobility of Amorphous Pharmaceutical Solids Below Their Glass Transition Temperatures. Pharm. Res. 1995, 12, 799–806. 10.1023/A:1016292416526. [DOI] [PubMed] [Google Scholar]
  151. Shamblin S. L.; Zografi G. Enthalpy Relaxation in Binary Amorphous Mixtures Containing Sucrose. Pharm. Res. 1998, 15, 1828–1834. 10.1023/A:1011997721086. [DOI] [PubMed] [Google Scholar]
  152. Hodge I. M. Enthalpy Relaxation and Recovery in Amorphous Materials. J. Non Cryst. Solids 1994, 169, 211–266. 10.1016/0022-3093(94)90321-2. [DOI] [Google Scholar]
  153. Thomas L. C.Modulated DSC® Paper# 5 Measurement of Glass Transitions and Enthalpic Recovery. TA Instruments: New Castle, DE, 2005. [Google Scholar]
  154. Guigo N.; Sbirrazzuoli N.. Thermal Analysis of Biobased Polymers and Composites. In Handbook of Thermal Analysis and Calorimetry; Vyazovkin S., Koga N., Schick C., Eds.; Elsevier Science B.V., 2018; Vol. 6, Chapter 10, pp 399–429. 10.1016/B978-0-444-64062-8.00002-4 [DOI] [Google Scholar]
  155. Miller G. W. Thermal analyses of polymers. VII. Calorimetric and Dilatometric Aspects of the Glass Transition. J. Appl. Polym. Sci. 1971, 15, 2335–2348. 10.1002/app.1971.070151002. [DOI] [Google Scholar]
  156. Lofaj F.; Satet R.; Hoffmann M. J.; de Arellano López A. R. Thermal Expansion and Glass Transition Temperature of the Rare-Earth Doped Oxynitride Glasses. J. Eur. Ceram. Soc. 2004, 24, 3377–3385. 10.1016/j.jeurceramsoc.2003.10.012. [DOI] [Google Scholar]
  157. Lunkenheimer P.; Loidl A.; Riechers B.; Zaccone A.; Samwer K. Thermal Expansion and the Glass Transition. Nat. Phys. 2023, 19, 694–699. 10.1038/s41567-022-01920-5. [DOI] [Google Scholar]
  158. Ediger M. D.; Angell C. A.; Nagel S. R. Supercooled Liquids and Glasses. J. Phys. Chem. 1996, 100, 13200–13212. 10.1021/jp953538d. [DOI] [Google Scholar]
  159. Hancock B. C.; Dalton C. R.; Pikal M. J.; Shamblin S. L. A Pragmatic Test of a Simple Calorimetric Method for Determining the Fragility of Some Amorphous Pharmaceutical Materials. Pharm. Res. 1998, 15, 762–767. 10.1023/A:1011931305755. [DOI] [PubMed] [Google Scholar]
  160. Weng L.; Elliott G. D. Dynamic and Thermodynamic Characteristics Associated with the Glass Transition of Amorphous Trehalose–Water Mixtures. Phys. Chem. Chem. Phys. 2014, 16, 11555–11565. 10.1039/C3CP55418J. [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. MacFarlane D. R.; Angell C. A. Nonexistent Glass Transition for Amorphous Solid Water. J. Phys. Chem. 1984, 88, 759–762. 10.1021/j150648a029. [DOI] [Google Scholar]
  162. Fox T. G. Influence of Diluent and of Copolymer Composition on the Glass Temperature of a Polymer System. Bull. Am. Phys. Soc. 1956, 1, 123. [Google Scholar]
  163. Gordon M.; Taylor J. S. Ideal Copolymers and the Second-Order Transitions of Synthetic Rubbers. i. Non-Crystalline Copolymers. J. Appl. Chem. 1952, 2, 493–500. 10.1002/jctb.5010020901. [DOI] [Google Scholar]
  164. Roos Y.; Karel M. Amorphous State and Delayed Ice Formation in Sucrose Solutions. Int. J. Food Sci. Technol. 1991, 26, 553–566. 10.1111/j.1365-2621.1991.tb02001.x. [DOI] [Google Scholar]
  165. Bogdanova E.; Fureby A. M.; Kocherbitov V. Influence of Cooling Rate on Ice Crystallization and Melting in Sucrose-Water System. J. Pharm. Sci. 2022, 111, 2030–2037. 10.1016/j.xphs.2022.01.027. [DOI] [PubMed] [Google Scholar]
  166. Zhang D. M.; Sun D. Y.; Gong X. G. Angell Plot from the Potential Energy Landscape Perspective. Phys. Rev. E 2022, 106, 064129. 10.1103/PhysRevE.106.064129. [DOI] [PubMed] [Google Scholar]
  167. Berthier L.; Biroli G. Theoretical Perspective on the Glass Transition and Amorphous Materials. Rev. Mod. Phys. 2011, 83, 587–642. 10.1103/RevModPhys.83.587. [DOI] [Google Scholar]
  168. Hodge I. M. Strong and Fragile Liquids — a Brief Critique. J. Non-Cryst. Solids 1996, 202, 164–172. 10.1016/0022-3093(96)00151-2. [DOI] [Google Scholar]
  169. Branca C.; Magazù S.; Maisano G.; Migliardo P.; Villari V.; Sokolov A. P. The Fragile Character and Structure-Breaker Role of α,α-Trehalose: Viscosity and Raman Scattering Findings. J. Phys.: Condens. Matter 1999, 11, 3823. 10.1088/0953-8984/11/19/305. [DOI] [Google Scholar]
  170. Rössler E.; Hess K. U.; Novikov V. N. Universal Representation of Viscosity in Glass Forming Liquids. J. Non-Cryst. Solids 1998, 223, 207–222. 10.1016/S0022-3093(97)00365-7. [DOI] [Google Scholar]
  171. Ferreira A. G. M.; Egas A. P. V.; Fonseca I. M. A.; Costa A. C.; Abreu D. C.; Lobo L. Q. The Viscosity of Glycerol. J. Chem. Thermodyn 2017, 113, 162–182. 10.1016/j.jct.2017.05.042. [DOI] [Google Scholar]
  172. De Gusseme A.; Carpentier L.; Willart J. F.; Descamps M. Molecular Mobility in Supercooled Trehalose. J. Phys. Chem. B 2003, 107, 10879–10886. 10.1021/jp0343234. [DOI] [Google Scholar]
  173. Wendt H.; Richert R. Purely Mechanical Solvation Dynamics in Supercooled Liquids: The S0 ← T1 (0–0) Transition of Naphthalene. J. Phys. Chem. A 1998, 102, 5775–5781. 10.1021/jp981613p. [DOI] [Google Scholar]
  174. Mezei F.; Knaak W.; Farago B.. Neutron Spin Echo Study of Dynamic Correlations Near Liquid-Glass Transition. Phys. Scr. 1987, T19B, 363. 10.1088/0031-8949/1987/T19B/007 [DOI] [PubMed] [Google Scholar]
  175. Cummins H. Z.; Li G.; Du W. M.; Hernandez J. Relaxational Dynamics in Supercooled Liquids: Experimental Tests of the Mode Coupling Theory. Physica A 1994, 204, 169–201. 10.1016/0378-4371(94)90424-3. [DOI] [Google Scholar]
  176. Li G.; Du W. M.; Sakai A.; Cummins H. Z. Light-Scattering Investigation of α and β Relaxation Near the Liquid-Glass Transition of the Molecular Glass Salol. Phys. Rev. A 1992, 46, 3343. 10.1103/PhysRevA.46.3343. [DOI] [PubMed] [Google Scholar]
  177. Sidebottom D.; Bergman R.; Börjesson L.; Torell L. M. Two-Step Relaxation Decay in a Strong Glass Former. Phys. Rev. Lett. 1993, 71, 2260. 10.1103/PhysRevLett.71.2260. [DOI] [PubMed] [Google Scholar]
  178. Stickel F.; Fischer E. W.; Richert R. Dynamics of Glass-Forming-Liquids. II. Detailed Comparison of Dielectric Relaxation, DC-Conductivity, and Viscosity Data. J. Chem. Phys. 1996, 104, 2043–2055. 10.1063/1.470961. [DOI] [Google Scholar]
  179. Dixon P. K.; Wu L.; Nagel S. R.; Williams B. D.; Carini J. P. Scaling in the Relaxation of Supercooled Liquids. Phys. Rev. Lett. 1990, 65, 1108. 10.1103/PhysRevLett.65.1108. [DOI] [PubMed] [Google Scholar]
  180. Farnan I.; Stebbins J. F. The Nature of the Glass Transition in a Silica-Rich Oxide Melt. Science (1979) 1994, 265, 1206–1209. 10.1126/science.265.5176.1206. [DOI] [PubMed] [Google Scholar]
  181. Stebbins J. F.; Farnan I.; Xue X. The Structure and Dynamics of Alkali Silicate Liquids: A View from NMR Spectroscopy. Chem. Geol. 1992, 96, 371–385. 10.1016/0009-2541(92)90066-E. [DOI] [Google Scholar]
  182. Oguni M.; Hikawa H.; Suga H. Enthalpy Relaxation in Vapor-Deposited Butyronitrile. Thermochim. Acta 1990, 158, 143–156. 10.1016/0040-6031(90)80061-3. [DOI] [Google Scholar]
  183. Birge N. O.; Nagel S. R. Specific-Heat Spectroscopy of the Glass Transition. Phys. Rev. Lett. 1985, 54, 2674. 10.1103/PhysRevLett.54.2674. [DOI] [PubMed] [Google Scholar]
  184. Sasabe H.; Moynihan C. T. Structural Relaxation in Poly(Vinyl Acetate). J. Polym. Sci. 1978, 16, 1447–1457. 10.1002/pol.1978.180160810. [DOI] [Google Scholar]
  185. Hansen J.-P.; McDonald I. R.. Hydrodynamics and Transport Coefficients. In Theory of Simple Liquids, 3rd ed.; Hansen J.-P., McDonald I. R., Eds.; Academic Press, 2006; Chapter 8, pp 219–254. 10.1016/B978-012370535-8/50010-0 [DOI] [Google Scholar]
  186. Johari G. P. Glass Transition and Secondary Relaxations in Molecular Liquids and Crystals. Ann. N.Y. Acad. Sci. 1976, 279, 117–140. 10.1111/j.1749-6632.1976.tb39701.x. [DOI] [Google Scholar]
  187. Johari G. P.; Goldstein M.; Soc J. A.; Kubert B.; Szabo J.; Giulieri S. Viscous Liquids and the Glass Transition. III. Secondary Relaxations in Aliphatic Alcohols and Other Nonrigid Molecules. J. Chem. Phys. 1971, 55, 4245–4252. 10.1063/1.1676742. [DOI] [Google Scholar]
  188. Johari G. P.; Goldstein M. Molecular Mobility in Simple Glasses. J. Phys. Chem. 1970, 74, 2034–2035. 10.1021/j100704a038. [DOI] [Google Scholar]
  189. Johari G. P.; Goldstein M. Viscous Liquids and the Glass Transition. II. Secondary Relaxations in Glasses of Rigid Molecules. J. Chem. Phys. 1970, 53, 2372–2388. 10.1063/1.1674335. [DOI] [Google Scholar]
  190. Ginzburg V. V. A Simple Mean-Field Model of Glassy Dynamics and Glass Transition. Soft Matter 2020, 16, 810–825. 10.1039/C9SM01575B. [DOI] [PubMed] [Google Scholar]
  191. Ngai K. L.; Paluch M. Classification of Secondary Relaxation in Glass-Formers Based on Dynamic Properties. J. Chem. Phys. 2004, 120, 857–873. 10.1063/1.1630295. [DOI] [PubMed] [Google Scholar]
  192. Caporaletti F.; Capaccioli S.; Valenti S.; Mikolasek M.; Chumakov A. I.; Monaco G. A Microscopic Look at the Johari-Goldstein Relaxation in a Hydrogen-Bonded Glass-Former. Sci. Rep. 2019, 9, 14319. 10.1038/s41598-019-50824-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Capaccioli S.; Paluch M.; Prevosto D.; Wang L.-M.; Ngai K. L. Many-Body Nature of Relaxation Processes in Glass-Forming Systems. J. Phys. Chem. Lett. 2012, 3, 735–743. 10.1021/jz201634p. [DOI] [PubMed] [Google Scholar]
  194. Roland C. M.; Casalini R. Density Scaling of the Dynamics of Vitrifying Liquids and its Relationship to the Dynamic Crossover. J. Non-Cryst. Solids 2005, 351, 2581–2587. 10.1016/j.jnoncrysol.2005.03.056. [DOI] [Google Scholar]
  195. Andrade E. N. D. C. The Viscosity of Liquids. Nature 1930, 125, 309–310. 10.1038/125309b0. [DOI] [Google Scholar]
  196. Vogel H. Das Temperaturabhangigkeitsgesetz Der Viskositat von Flussigkeiten. Phys. Z. 1921, 22, 645–646. [Google Scholar]
  197. Fulcher G. S. Analysis of Recent Measurements of the Viscosity of Glasses. J. Am. Ceram. Soc. 1925, 8, 339–355. 10.1111/j.1151-2916.1925.tb16731.x. [DOI] [Google Scholar]
  198. Tammann G.; Hesse W. Die Abhängigkeit der Viscosität von der Temperatur bie unterkühlten Flüssigkeiten. Z. Anorg. Allg. Chem. 1926, 156, 245–257. 10.1002/zaac.19261560121. [DOI] [Google Scholar]
  199. Garca-Coln L. S.; Del Castillo L. F.; Goldstein P. Theoretical Basis for the Vogel-Fulcher-Tammann Equation. Phys. Rev. B 1989, 40, 7040. 10.1103/PhysRevB.40.7040. [DOI] [PubMed] [Google Scholar]
  200. Bair S.The Pressure and Temperature Dependence of the Low-Shear Viscosity. In High Pressure Rheology for Quantitative Elastohydrodynamics, 2nd ed.; Bair S., Ed.; Elsevier, 2019; Chapter 5, pp 97–133. 10.1016/B978-0-444-64156-4.00005-2 [DOI] [Google Scholar]
  201. Williams M. L.; Landel R. F.; Ferry J. D. The Temperature Dependence of Relaxation Mechanisms in Amorphous Polymers and Other Glass-forming Liquids. J. Am. Chem. Soc. 1955, 77, 3701–3707. 10.1021/ja01619a008. [DOI] [Google Scholar]
  202. Phan A. D.; Wakabayashi K.. Theory of Structural and Secondary Relaxation in Amorphous Drugs under Compression. Pharmaceutics 2020, 12, 177. 10.3390/pharmaceutics12020177 [DOI] [PMC free article] [PubMed] [Google Scholar]
  203. Doolittle A. K. Studies in Newtonian Flow. III. The Dependence of the Viscosity of Liquids on Molecular Weight and Free Space (in Homologous Series). J. Appl. Phys. 1952, 23, 236–239. 10.1063/1.1702182. [DOI] [Google Scholar]
  204. Doolittle A. K.; Angke Jakarta M. Studies in Newtonian Flow. II. The Dependence of the Viscosity of Liquids on Free-Space. J. Appl. Phys. 1951, 22, 1471–1475. 10.1063/1.1699894. [DOI] [Google Scholar]
  205. Kauzmann W. The Nature of the Glassy State and the Behavior of Liquids at Low Temperatures. Chem. Rev. 1948, 43, 219–256. 10.1021/cr60135a002. [DOI] [Google Scholar]
  206. Angell C. A. Entropy and Fragility in Supercooling Liquids. J. Res. Natl. Inst. Stand. Technol. 1997, 102, 171. 10.6028/jres.102.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  207. Lubchenko V.; Wolynes P. G. Theory of Structural Glasses and Supercooled Liquids. Annu. Rev. Phys. Chem. 2007, 58, 235–266. 10.1146/annurev.physchem.58.032806.104653. [DOI] [PubMed] [Google Scholar]
  208. Gibbs J. H.; DiMarzio E. A. Nature of the Glass Transition and the Glassy State. J. Chem. Phys. 1958, 28, 373–383. 10.1063/1.1744141. [DOI] [Google Scholar]
  209. Di Marzio E. A.; Yang A. J. M. Configurational Entropy Approach to the Kinetics of Glasses. J. Res. Natl. Inst. Stand. Technol. 1997, 102, 135. 10.6028/jres.102.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  210. Stillinger F. H.; Hodgdon J. A. Translation-Rotation Paradox for Diffusion in Fragile Glass-Forming Liquids. Phys. Rev. E 1994, 50, 2064. 10.1103/PhysRevE.50.2064. [DOI] [PubMed] [Google Scholar]
  211. Sillescu H. Heterogeneity at the Glass Transition: a Review. J. Non Cryst. Solids 1999, 243, 81–108. 10.1016/S0022-3093(98)00831-X. [DOI] [Google Scholar]
  212. Bengtzelius U.; Gotze W.; Sjolander A. Dynamics of Supercooled Liquids and the Glass Transition. J. Phys. C: Solid State Phys. 1984, 17, 5915. 10.1088/0022-3719/17/33/005. [DOI] [Google Scholar]
  213. Leutheusser E. Dynamical Model of the Liquid-Glass Transition. Phys. Rev. A 1984, 29, 2765. 10.1103/PhysRevA.29.2765. [DOI] [Google Scholar]
  214. Das S. P. Mode-Coupling Theory and the Glass Transition in Supercooled Liquids. Rev. Mod. Phys. 2004, 76, 785–851. 10.1103/RevModPhys.76.785. [DOI] [Google Scholar]
  215. Janssen L. M. C. Mode-Coupling Theory of the Glass Transition: A Primer. Front. Phys. 2018, 6, 403911. 10.3389/fphy.2018.00097. [DOI] [Google Scholar]
  216. Kirkpatrick T. R.; Thirumalai D.. Colloquium: Random First Order Transition Theory Concepts in Biology and Physics. Rev. Mod. Phys. 2015, 87, 183. 10.1103/RevModPhys.87.183 [DOI] [Google Scholar]
  217. Yeomans-Reyna L.; Chávez-Rojo M. A.; Ramírez-González P. E.; Juárez-Maldonado R.; Chávez-Páez M.; Medina-Noyola M.. Dynamic Arrest Within the Self-Consistent Generalized Langevin Equation of Colloid Dynamics. Phys. Rev. E: Stat. Nonlinear Soft Matter Phys. 2007, 76, 041504. 10.1103/PhysRevE.76.041504 [DOI] [PubMed] [Google Scholar]
  218. Hodge I. M.; Berens A. R. Effects of Annealing and Prior History on Enthalpy Relaxation in Glassy Polymers. 2. Mathematical Modeling. Macromolecules 1982, 15, 762–770. 10.1021/ma00231a016. [DOI] [Google Scholar]
  219. Argatov I.; Kocherbitov V. Analysis of the Minimal Model for the Enthalpy Relaxation and Recovery in Glass Transition: Application to Constant-Rate Differential Scanning Calorimetry. Continuum Mechanics and Thermodynamics 2021, 33, 107–123. 10.1007/s00161-020-00891-3. [DOI] [Google Scholar]
  220. Adam G.; Gibbs J. H. On the Temperature Dependence of Cooperative Relaxation Properties in Glass-Forming Liquids. J. Chem. Phys. 1965, 43, 139–146. 10.1063/1.1696442. [DOI] [Google Scholar]
  221. Lacevic N.; Starr F. W.; Schrøder T. B.; Glotzer S. C. Spatially Heterogeneous Dynamics Investigated via a Time-Dependent Four-Point Density Correlation Function. J. Chem. Phys. 2003, 119, 7372–7387. 10.1063/1.1605094. [DOI] [Google Scholar]
  222. Olgenblum G. I.; Sapir L.; Harries D.. Properties of Aqueous Trehalose Mixtures: Glass Transition and Hydrogen Bonding. J. Chem. Theory Comput. 2020, 16, 1249. 10.1021/acs.jctc.9b01071 [DOI] [PMC free article] [PubMed] [Google Scholar]
  223. Simmons D. S.; Douglas J. F. Nature and Interrelations of Fast Dynamic Properties in a Coarse-Grained Glass-Forming Polymer Melt. Soft Matter 2011, 7, 11010–11020. 10.1039/c1sm06189e. [DOI] [Google Scholar]
  224. Lang R. J.; Simmons D. S. Interfacial Dynamic Length Scales in the Glass Transition of a Model Freestanding Polymer Film and Their Connection to Cooperative Motion. Macromolecules 2013, 46, 9818–9825. 10.1021/ma401525q. [DOI] [Google Scholar]
  225. Hung J. H.; Patra T. K.; Meenakshisundaram V.; Mangalara J. H.; Simmons D. S. Universal Localization Transition Accompanying Glass Formation: Insights from Efficient Molecular Dynamics Simulations of Diverse Supercooled Liquids. Soft Matter 2019, 15, 1223–1242. 10.1039/C8SM02051E. [DOI] [PubMed] [Google Scholar]
  226. Ginzburg V. V. Modeling the Glass Transition of Free-Standing Polymer Thin Films Using the ‘sL-TS2’ Mean-Field Approach. Macromolecules 2022, 55, 873–882. 10.1021/acs.macromol.1c02370. [DOI] [Google Scholar]
  227. Ginzburg V. V. Combined Description of Polymer PVT and Relaxation Data Using a Dynamic “SL-TS2” Mean-Field Lattice Model. Soft Matter 2021, 17, 9094–9106. 10.1039/D1SM00953B. [DOI] [PubMed] [Google Scholar]
  228. Nguyen H. H.; Kim M. An Overview of Techniques in Enzyme Immobilization. Appl. Sci. Converg. Technol. 2017, 26, 157–163. 10.5757/ASCT.2017.26.6.157. [DOI] [Google Scholar]
  229. Barham P.; Edwards J. S. A.; Schafheitle J. M.. The Science of Cooking; Springer, 2001. [Google Scholar]
  230. Surana R.; Pyne A.; Suryanarayanan R. Effect of Preparation Method on Physical Properties of Amorphous Trehalose. Pharm. Res. 2004, 21, 1167–1176. 10.1023/B:PHAM.0000033003.17251.c3. [DOI] [PubMed] [Google Scholar]
  231. Olgenblum G. I.; Wien F.; Sapir L.; Harries D.. β-Hairpin Miniprotein Stabilization in Trehalose Glass Is Facilitated by an Emergent Compact Non-Native State. J. Phys. Chem. Lett. 2021, 12, 7659. 10.1021/acs.jpclett.1c02379 [DOI] [PubMed] [Google Scholar]
  232. Crowe L. M.; Reid D. S.; Crowe J. H. Is Trehalose Special for Preserving Dry Biomaterials?. Biophys. J. 1996, 71, 2087–2093. 10.1016/S0006-3495(96)79407-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  233. Craig D. Q. M.; Royall P. G.; Kett V. L.; Hopton M. L. The Relevance of the Amorphous State to Pharmaceutical Dosage Forms: Glassy Drugs and Freeze Dried Systems. Int. J. Pharm. 1999, 179, 179–207. 10.1016/S0378-5173(98)00338-X. [DOI] [PubMed] [Google Scholar]
  234. Roy I.; Gupta M. N. Freeze-Drying of Proteins: Some Emerging Concerns. Biotechnol Appl. Biochem 2004, 39, 165–177. 10.1042/BA20030133. [DOI] [PubMed] [Google Scholar]
  235. Ameri M.; Maa Y. F. Spray Drying of Biopharmaceuticals: Stability and Process Considerations. Drying Technology 2006, 24, 763–768. 10.1080/03602550600685275. [DOI] [Google Scholar]
  236. Amaya-Farfan J.; Rodriguez-Amaya D. B.. The Maillard Reactions. In Chemical Changes During Processing and Storage of Foods; Rodriguez-Amaya D. B., Amaya-Farfan J., Eds.; Academic Press; 2021; Chapter 6, pp 215–263. 10.1016/B978-0-12-817380-0.00006-3 [DOI] [Google Scholar]
  237. Teramoto N.; Sachinvala N. D.; Shibata M. Trehalose and Trehalose-based Polymers for Environmentally Benign, Biocompatible and Bioactive Materials. Molecules 2008, 13, 1773–1816. 10.3390/molecules13081773. [DOI] [PMC free article] [PubMed] [Google Scholar]
  238. Brom J. A.; Petrikis R. G.; Pielak G. J. How Sugars Protect Dry Protein Structure. Biochemistry 2023, 62, 1044–1052. 10.1021/acs.biochem.2c00692. [DOI] [PMC free article] [PubMed] [Google Scholar]
  239. Politi R.; Harries D. Enthalpically Driven Peptide Stabilization by Protective Osmolytes. Chem. Commun. 2010, 46, 6449–6451. 10.1039/c0cc01763a. [DOI] [PubMed] [Google Scholar]
  240. Sukenik S.; Politi R.; Ziserman L.; Danino D.; Friedler A.; Harries D. Crowding Alone Cannot Account for Cosolute Effect on Amyloid Aggregation. PLoS One 2011, 6, e15608. 10.1371/journal.pone.0015608. [DOI] [PMC free article] [PubMed] [Google Scholar]
  241. Olgenblum G. I.; Carmon N.; Harries D. Not always sticky: Specificity of Protein Stabilization by Sugars is Conferred by Protein-Water Hydrogen Bonds. J. Am. Chem. Soc. 2023, 145, 23308–23320. 10.1021/jacs.3c08702. [DOI] [PMC free article] [PubMed] [Google Scholar]
  242. Arsiccio A.; Sarter T.; Polidori I.; Winter G.; Pisano R.; Shea J. E. Thermodynamic Modeling and Experimental Data Reveal That Sugars Stabilize Proteins According to an Excluded Volume Mechanism. J. Am. Chem. Soc. 2023, 145, 16678–16690. 10.1021/jacs.3c04293. [DOI] [PubMed] [Google Scholar]
  243. Ding S. P.; Fan J.; Green J. L.; Lu Q.; Sanchez E.; Angell C. A. Vitrification of Trehalose by Water Loss From its Crystalline Dihydrate. J. Therm. Anal. 1996, 47, 1391–1405. 10.1007/BF01992835. [DOI] [Google Scholar]
  244. Cesàro A.; De Giacomo O.; Sussich F. Water Interplay in Trehalose Polymorphism. Food Chem. 2008, 106, 1318–1328. 10.1016/j.foodchem.2007.01.082. [DOI] [Google Scholar]
  245. Sussich F.; Skopec C.; Brady J.; Cesàro A. Reversible Dehydration of Trehalose and Anhydrobiosis: from Solution State to an Exotic Crystal?. Carbohydr. Res. 2001, 334, 165–176. 10.1016/S0008-6215(01)00189-6. [DOI] [PubMed] [Google Scholar]
  246. Srirangsan P.; Kawai K.; Hamada-Sato N.; Watanabe M.; Suzuki T. Stabilizing Effects of Sucrose–Polymer Formulations on the Activities of Freeze-Dried Enzyme Mixtures of Alkaline Phosphatase, Nucleoside Phosphorylase and Xanthine Oxidase. Food Chem. 2011, 125, 1188–1193. 10.1016/j.foodchem.2010.10.029. [DOI] [Google Scholar]
  247. Ramirez J. F.; Kumara U. G. V. S. S.; Arulsamy N.; Boothby T. C. Water Content, Transition Temperature and Fragility Influence Protection and Anhydrobiotic Capacity. bioRxiv 2023, 6, 547256. 10.1101/2023.06.30.547256. [DOI] [PMC free article] [PubMed] [Google Scholar]
  248. Weng L.; Elliott G. D. Local Minimum in Fragility for Trehalose/Glycerol Mixtures: Implications for Biopharmaceutical Stabilization. J. Phys. Chem. B 2015, 119, 6820–6827. 10.1021/acs.jpcb.5b01675. [DOI] [PubMed] [Google Scholar]
  249. Asmatulu R. Nanocoatings for Corrosion Protection of Aerospace Alloys. Corrosion Protection and Control Using Nanomaterials 2012, 357–374. 10.1533/9780857095800.2.357. [DOI] [Google Scholar]
  250. Cottone G.; Giuffrida S.; Bettati S.; Bruno S.; Campanini B.; Marchetti M.; Abbruzzetti S.; Viappiani C.; Cupane A.; Mozzarelli A.; et al. More than a Confinement: “Soft” and “Hard” Enzyme Entrapment Modulates Biological Catalyst Function. Catal. 2019, 9, 1024. 10.3390/catal9121024. [DOI] [Google Scholar]
  251. Deshmukh K.; Kovarik T.; Krenek T.; Docheva D.; Stich T.; Pola J.; et al. Recent Advances and Future Perspectives of Sol–Gel Derived Porous Bioactive Glasses: a Review. RSC Adv. 2020, 10, 33782–33835. 10.1039/D0RA04287K. [DOI] [PMC free article] [PubMed] [Google Scholar]
  252. Betancor L.; Luckarift H. R. Bioinspired Enzyme Encapsulation for Biocatalysis. Trends Biotechnol. 2008, 26, 566–572. 10.1016/j.tibtech.2008.06.009. [DOI] [PubMed] [Google Scholar]
  253. Vinogradov V. V.; Avnir D. Exceptional Thermal Stability of Therapeutical Enzymes Entrapped in Alumina Sol–Gel Matrices. J. Mater. Chem. B 2014, 2, 2868–2873. 10.1039/c3tb21643h. [DOI] [PubMed] [Google Scholar]
  254. Monton M. R. N.; Forsberg E. M.; Brennan J. D. Tailoring Sol-Gel-Derived Silica Materials for Optical Biosensing. Chem. Mater. 2012, 24, 796–811. 10.1021/cm202798e. [DOI] [Google Scholar]
  255. Hou J.; Dong G.; Ye Y.; Chen V. Enzymatic Degradation of Bisphenol-A with Immobilized Laccase on TiO2 Sol–Gel Coated PVDF Membrane. J. Membr. Sci. 2014, 469, 19–30. 10.1016/j.memsci.2014.06.027. [DOI] [Google Scholar]
  256. Reetz M. T.; Tielmann P.; Wiesenhöfer W.; Könen W.; Zonta A. Second Generation Sol-Gel Encapsulated Lipases: Robust Heterogeneous Biocatalysts. Adv. Synth. Catal. 2003, 345, 717–728. 10.1002/adsc.200303016. [DOI] [Google Scholar]
  257. David A. E.; Yang A. J.; Wang N. S. Enzyme Stabilization and Immobilization by Sol-Gel Entrapment. Methods Mol. Biol. 2011, 679, 49–66. 10.1007/978-1-60761-895-9_6. [DOI] [PubMed] [Google Scholar]
  258. Buthe A. Entrapment of Enzymes in Nanoporous Sol–Gels. Methods Mol. Biol. 2011, 743, 223–237. 10.1007/978-1-61779-132-1_18. [DOI] [PubMed] [Google Scholar]
  259. Badawy M. E. I.; El-Aswad A. F.. Bioactive Paper Sensor Based on the Acetylcholinesterase for the Rapid Detection of Organophosphate and Carbamate Pesticides. Int. J. Anal. Chem. 2014, 2014, 536823. 10.1155/2014/536823 [DOI] [PMC free article] [PubMed] [Google Scholar]
  260. Avnir D.; Braun S.; Lev O.; Ottolenghi M. Reviews Enzymes and Other Proteins Entrapped in Sol-Gel Materials. Chem. Mater. 1994, 6, 1605–1614. 10.1021/cm00046a008. [DOI] [Google Scholar]
  261. Smith K.; Silvernail N. J.; Rodgers K. R.; Elgren T. E.; Castro M.; Parker R. M. Sol-Gel Encapsulated Horseradish peroxidase: A catalytic material for peroxidation. J. Am. Chem. Soc. 2002, 124, 4247–4252. 10.1021/ja012215u. [DOI] [PubMed] [Google Scholar]
  262. Hench L. L.; West J. K. The Sol-Gel Process. Chem. Rev. 1990, 90, 33–72. 10.1021/cr00099a003. [DOI] [Google Scholar]
  263. Campás M.; Marty J.-L.. Encapsulation of Enzymes Using Polymers and Sol-Gel Techniques. In Methods in Biotechnology: Immobilization of Enzymes and Cells, 2nd ed.; Humana Press, 2006; pp 77–85 . [Google Scholar]
  264. Gupta R.; Chaudhury N. K. Entrapment of Biomolecules in Sol–Gel Matrix for Applications in Biosensors: Problems and Future Prospects. Biosens Bioelectron 2007, 22, 2387–2399. 10.1016/j.bios.2006.12.025. [DOI] [PubMed] [Google Scholar]
  265. Jerónimo P. C. A.; Araújo A. N.; Montenegro M. C. B. S. M. Optical Sensors and Biosensors Based on Sol–Gel Films. Talanta 2007, 72, 13–27. 10.1016/j.talanta.2006.09.029. [DOI] [PubMed] [Google Scholar]
  266. Jin W.; Brennan J. D. Properties and Applications of Proteins Encapsulated within Sol–Gel Derived Materials. Anal. Chim. Acta 2002, 461, 1–36. 10.1016/S0003-2670(02)00229-5. [DOI] [Google Scholar]
  267. Liu D. M.; Chen I. W. Encapsulation of Protein Molecules in Transparent Porous Silica Matrices via an Aqueous Colloidal Sol–Gel Process. Acta Mater. 1999, 47, 4535–4544. 10.1016/S1359-6454(99)00335-3. [DOI] [Google Scholar]
  268. Gill I.; Ballesteros A. Encapsulation of Biologicals within Silicate, Siloxane, and Hybrid Sol-Gel polymers: An Efficient and Generic Approach. J. Am. Chem. Soc. 1998, 120, 8587–8598. 10.1021/ja9814568. [DOI] [Google Scholar]
  269. Brikner C. J.; Scherer G. W.. Sol-Gel Science: The Physics and Chemistry of Sol-Gel Processing; Academic Press, 1990. [Google Scholar]
  270. Buckley A. M.; Greenblatt M. The Sol-Gel Preparation of Silica Gels. J. Chem. Educ. 1994, 71, 599. 10.1021/ed071p599. [DOI] [Google Scholar]
  271. Chong S.; Riley B. J.; Peterson J. A.; Olszta M. J.; Nelson Z. J. Gaseous Iodine Sorbents: A Comparison between Ag-Loaded Aerogel and Xerogel Scaffolds. ACS Appl. Mater. Interfaces. 2020, 12, 26127–26136. 10.1021/acsami.0c02396. [DOI] [PubMed] [Google Scholar]
  272. Khalil H. P. S. A.; Yahya E. B.; Tajarudin H. A.; Balakrishnan V.; Nasution H. Insights into the Role of Biopolymer-Based Xerogels in Biomedical Applications. Gels 2022, 8, 334. 10.3390/gels8060334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  273. Smirnova I.; Gurikov P. Aerogel production: Current Status, Research Directions, and Future Opportunities. J. Supercrit. Fluids 2018, 134, 228–233. 10.1016/j.supflu.2017.12.037. [DOI] [Google Scholar]
  274. Lovskaya D.; Bezchasnyuk A.; Mochalova M.; Tsygankov P.; Lebedev A.; Zorkina Y.; Zubkov E.; Ochneva A.; Gurina O.; Silantyev A.; et al. Preparation of Protein Aerogel Particles for the Development of Innovative Drug Delivery Systems. Gels 2022, 8, 765. 10.3390/gels8120765. [DOI] [PMC free article] [PubMed] [Google Scholar]
  275. Thomas S., Pothan L. A., Mavelil-Sam R., Eds. Biobased Aerogels: Polysaccharide and Protein-based Materials; Royal Society of Chemistry, 2018. 10.1039/9781782629979 [DOI] [Google Scholar]
  276. García-González C. A.; Sosnik A.; Kalmár J.; De Marco I.; Erkey C.; Concheiro A.; Alvarez-Lorenzo C. Aerogels in Drug Delivery: From Design to Application. J. Controlled Release 2021, 332, 40–63. 10.1016/j.jconrel.2021.02.012. [DOI] [PubMed] [Google Scholar]
  277. Cheng J.; Liu J.; Li M.; Liu Z.; Wang X.; Zhang L.; Wang Z. Hydrogel-Based Biomaterials Engineered from Natural-Derived Polysaccharides and Proteins for Hemostasis and Wound Healing. Front. Bioeng. Biotechnol. 2021, 9, 780187. 10.3389/fbioe.2021.780187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  278. Lee J. C.; Timasheff S. N. The Stabilization of Proteins by Sucrose. J. Biol. Chem. 1981, 256, 7193–7201. 10.1016/S0021-9258(19)68947-7. [DOI] [PubMed] [Google Scholar]
  279. Bhat R.; Timasheff S. N. Steric Exclusion is the Principal Source of the Preferential Hydration of Proteins in the Presence of Polyethylene Glycols. Protein Sci. 1992, 1, 1133–1143. 10.1002/pro.5560010907. [DOI] [PMC free article] [PubMed] [Google Scholar]
  280. Bolen D. W.; Baskakov I. V. The Osmophobic Effect: Natural Selection of a Thermodynamic Force in Protein Folding. J. Mol. Biol. 2001, 310, 955–963. 10.1006/jmbi.2001.4819. [DOI] [PubMed] [Google Scholar]
  281. Shimizu S.; Matubayasi N. Preferential solvation: Dividing surface vs excess numbers. J. Phys. Chem. B 2014, 118, 3922–3930. 10.1021/jp410567c. [DOI] [PubMed] [Google Scholar]
  282. Speer S. L.; Stewart C. J.; Sapir L.; Harries D.; Pielak G. J. Macromolecular Crowding Is More than Hard-Core Repulsions. Annu. Rev. Biophys. 2022, 51, 267–300. 10.1146/annurev-biophys-091321-071829. [DOI] [PubMed] [Google Scholar]
  283. Sapir L.; Harries D. Wisdom of the Crowd. Bunsen-Magazin 2017, 19, 152–162. [Google Scholar]
  284. Shimizu S. Estimating Hydration Changes Upon Biomolecular Reactions from Osmotic Stress, High Pressure, and Preferential Hydration Experiments. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 1195–1199. 10.1073/pnas.0305836101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  285. Parsegian V. A.; Rand R. P.; Rau D. C. Osmotic Stress, Crowding, Preferential Hydration, and Binding: A Comparison of Perspectives. Proc. Natl. Acad. Sci. U.S.A. 2000, 97, 3987–3992. 10.1073/pnas.97.8.3987. [DOI] [PMC free article] [PubMed] [Google Scholar]
  286. Olsson C.; Genheden S.; García Sakai V.; Swenson J. Mechanism of Trehalose-Induced Protein Stabilization from Neutron Scattering and Modeling. J. Phys. Chem. B 2019, 123, 3679–3687. 10.1021/acs.jpcb.9b01856. [DOI] [PubMed] [Google Scholar]
  287. Courtenay E. S.; Capp M. W.; Anderson C. F.; Record M. T. Vapor Pressure Osmometry Studies of Osmolyte–Protein Interactions: Implications for the Action of Osmoprotectants in Vivo and for the Interpretation of “Osmotic Stress” Experiments in Vitro. Biochemistry 2000, 39, 4455–4471. 10.1021/bi992887l. [DOI] [PubMed] [Google Scholar]
  288. Timasheff S. N. Protein Hydration, Thermodynamic Binding, and Preferential Hydration. Biochemistry 2002, 41, 13473–13482. 10.1021/bi020316e. [DOI] [PubMed] [Google Scholar]
  289. Schellman J. A. Selective Binding and Solvent Denaturation. Biopolymers 1987, 26, 549–559. 10.1002/bip.360260408. [DOI] [PubMed] [Google Scholar]
  290. Shumilin I.; Harries D. Enhanced Solubilization in Multi-Component Mixtures: Mechanism of Synergistic Amplification of Cyclodextrin Solubility by Urea and Inorganic Salts. J. Mol. Liq. 2023, 380, 121760. 10.1016/j.molliq.2023.121760. [DOI] [Google Scholar]
  291. Badasyan A.; Tonoyan S.; Giacometti A.; Podgornik R.; Parsegian V. A.; Mamasakhlisov Y.; Morozov V. Osmotic Pressure Induced Coupling Between Cooperativity and Stability of a Helix-Coil Transition. Phys. Rev. Lett. 2012, 109, 068101 10.1103/PhysRevLett.109.068101. [DOI] [PubMed] [Google Scholar]
  292. Gnutt D.; Gao M.; Brylski O.; Heyden M.; Ebbinghaus S. Excluded-Volume Effects in Living Cells. Angew. Chem., Int. Ed. 2015, 54, 2548–2551. 10.1002/anie.201409847. [DOI] [PMC free article] [PubMed] [Google Scholar]
  293. Gruebele M.; Dave K.; Sukenik S. Globular Protein Folding In Vitro and In Vivo. Annu. Rev. Biophys. 2016, 45, 233–251. 10.1146/annurev-biophys-062215-011236. [DOI] [PubMed] [Google Scholar]
  294. Wyman J. Linked Functions and Reciprocal Effects in Hemoglobin: A Second Look. Adv. Protein Chem. 1964, 19, 223–286. 10.1016/S0065-3233(08)60190-4. [DOI] [PubMed] [Google Scholar]
  295. Greene R. F.; Pace C. N. Urea and Guanidine Hydrochloride Denaturation of Ribonuclease, Lysozyme, α Chymotrypsin, and β Lactoglobulin. J. Biol. Chem. 1974, 249, 5388–5393. 10.1016/S0021-9258(20)79739-5. [DOI] [PubMed] [Google Scholar]
  296. Rosgen J.; Pettitt B. M.; Bolen D. W.. Protein Folding, Stability, and Solvation Structure in Osmolyte Solutions. Biophys. J. 2005, 89, 2988–2997. 10.1529/biophysj.105.067330 [DOI] [PMC free article] [PubMed] [Google Scholar]
  297. Davis-Searles P. R.; Morar A. S.; Saunders A. J.; Erie D. A.; Pielak G. J. Sugar-Induced Molten-Globule Model. Biochem 1998, 37, 17048–17053. 10.1021/bi981364v. [DOI] [PubMed] [Google Scholar]
  298. Richards F.Areas, Volumes, Packing, and Protein Structure. Annu. Rev. Biophys. 1977, 6, 151–176. 10.1146/annurev.bb.06.060177.001055 [DOI] [PubMed] [Google Scholar]
  299. Stewart C. J.; Olgenblum G. I.; Propst A.; Harries D.; Pielak G. J. Resolving the Enthalpy of Protein Stabilization by Macromolecular Crowding. Protein Sci. 2023, 32, e4573. 10.1002/pro.4573. [DOI] [PMC free article] [PubMed] [Google Scholar]
  300. Gnutt D.; Timr S.; Ahlers J.; König B.; Manderfeld E.; Heyden M.; Sterpone F.; Ebbinghaus S. Stability Effect of Quinary Interactions Reversed by Single Point Mutations. J. Am. Chem. Soc. 2019, 141, 4660–4669. 10.1021/jacs.8b13025. [DOI] [PubMed] [Google Scholar]
  301. Zhou H. X.; Rivas G.; Minton A. P. Macromolecular Crowding and Confinement: Biochemical, Biophysical, and Potential Physiological Consequences. Annu. Rev. Biophys. 2008, 37, 375–397. 10.1146/annurev.biophys.37.032807.125817. [DOI] [PMC free article] [PubMed] [Google Scholar]
  302. Schellman J.; Protein A. Stability in Mixed Solvents: A Balance of Contact Interaction and Excluded Volume. Biophys. J. 2003, 85, 108–125. 10.1016/S0006-3495(03)74459-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  303. Timr S.; Madern D.; Sterpone F. Protein Thermal Stability. Prog. Mol. Biol. Transl Sci. 2020, 170, 239–272. 10.1016/bs.pmbts.2019.12.007. [DOI] [PubMed] [Google Scholar]
  304. Guseman A. J.; Goncalves G. M. P.; Speer S. L.; Young G. B.; Pielak G. J. Protein Shape Modulates Crowding Effects. Proc. Natl. Acad. Sci. U. S. A. 2018, 115, 10965–10970. 10.1073/pnas.1810054115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  305. Timasheff S. N. Control of Protein Stability and Reactions by Weakly Interacting Cosolvents: the Simplicity of the Complicated. Adv. Protein Chem. 1998, 51, 355–428. 10.1016/S0065-3233(08)60656-7. [DOI] [PubMed] [Google Scholar]
  306. Minton A. P. Quantitative Assessment of the Relative Contributions of Steric Repulsion and Chemical Interactions to Macromolecular Crowding. Biopolymers 2013, 99, 239–244. 10.1002/bip.22163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  307. Sapir L.; Harries D. Macromolecular Stabilization by Excluded Cosolutes: Mean Field Theory of Crowded Solutions. J. Chem. Theory Comput. 2015, 11, 3478–3490. 10.1021/acs.jctc.5b00258. [DOI] [PubMed] [Google Scholar]
  308. Pace C. N. Determination and Analysis of Urea and Guanidine Hydrochloride Denaturation Curves. Methods Enzymol. 1986, 131, 266–280. 10.1016/0076-6879(86)31045-0. [DOI] [PubMed] [Google Scholar]
  309. Canchi D. R.; Paschek D.; Garcia A. E. Equilibrium Study of Protein Denaturation by Urea. J. Am. Chem. Soc. 2010, 132, 2338–2344. 10.1021/ja909348c. [DOI] [PubMed] [Google Scholar]
  310. Canchi D. R.; García A. E.. Cosolvent Effects on Protein Stability. Annu. Rev. Phys. Chem. 2013, 64, 273–293. 10.1146/annurev-physchem-040412-110156 [DOI] [PubMed] [Google Scholar]
  311. Zosel F.; Soranno A.; Buholzer K. J.; Nettels D.; Schuler B. Depletion Interactions Modulate the Binding Between Disordered Proteins in Crowded Environments. Proc. Natl. Acad. Sci. U. S. A. 2020, 117, 13480–13489. 10.1073/pnas.1921617117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  312. Semeraro E. F.; Giuffrida S.; Cottone G.; Cupane A. Biopreservation of Myoglobin in Crowded Environment: A Comparison between Gelatin and Trehalose Matrixes. J. Phys. Chem. B 2017, 121, 8731–8741. 10.1021/acs.jpcb.7b07266. [DOI] [PubMed] [Google Scholar]
  313. Asakura S.; Oosawa F. Interaction Between Particles Suspended in Solutions of Macromolecules. J. Polym. Sci. 1958, 33, 183–192. 10.1002/pol.1958.1203312618. [DOI] [Google Scholar]
  314. Asakura S.; Oosawa F. On Interaction between Two Bodies Immersed in a Solution of Macromolecules. J. Chem. Phys. 1954, 22, 1255–1256. 10.1063/1.1740347. [DOI] [Google Scholar]
  315. Sharp K. A. Analysis of the Size Dependence of Macromolecular Crowding Shows that Smaller is Better. Proc. Natl. Acad. Sci. U. S. A. 2015, 112, 7990–7995. 10.1073/pnas.1505396112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  316. Gibbons R. M. The Scaled Particle Theory for Particles of Arbitrary Shape. Mol. Phys. 1969, 17, 81–86. 10.1080/00268976900100811. [DOI] [Google Scholar]
  317. Lebowitz J. L.; Helfand E.; Praestgaard E. Scaled Particle Theory of Fluid Mixtures. J. Chem. Phys. 1965, 43, 774–779. 10.1063/1.1696842. [DOI] [Google Scholar]
  318. Reiss H.; Frisch H. L.; Helfand E.; Lebowitz J. L. Aspects of the Statistical Thermodynamics of Real Fluids. J. Chem. Phys. 1960, 32, 119–124. 10.1063/1.1700883. [DOI] [Google Scholar]
  319. Saunders A. J.; Davis-Searles P. R.; Allen D. L.; Pielak G. J.; Erie D. A. Osmolyte-Induced Changes in Protein Conformational Equilibria. Biopolymers 2000, 53, 293–307. . [DOI] [PubMed] [Google Scholar]
  320. Davis-Searles P. R.; Saunders A. J.; Erie D. A.; Winzor D. J.; Pielak G. J. Interpreting the Effects of Small Uncharged Solutes on Protein-Folding Equilibria. Annu. Rev. Biophys. 2001, 30, 271–306. 10.1146/annurev.biophys.30.1.271. [DOI] [PubMed] [Google Scholar]
  321. Courtenay E. S.; Capp M. W.; Saecker R. M.; Record M. T. Thermodynamic Analysis of Interactions Between Denaturants and Protein Surface Exposed on Unfolding: Interpretation of Urea and Guanidinium Chloride m-Values and Their Correlation With Changes in Accessible Surface Area (ASA) Using Preferential Interaction Coefficients and the Local-Bulk Domain Model. Proteins Struct. Funct. Bioinf. 2000, 41, 72–85. . [DOI] [PubMed] [Google Scholar]
  322. Soranno A.; Hofmann H.; Nettels D. Single-Molecule Spectroscopy Reveals Polymer Effects of Disordered Proteins in Crowded Environments. Proc. Natl. Acad. Sci. U.S.A. 2014, 111, 4874–4879. 10.1073/pnas.1322611111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  323. Sapir L.; Harries D. Macromolecular Compaction by Mixed Solutions: Bridging versus Depletion Attraction. Curr. Opin. Colloid Interface Sci. 2016, 22, 80–87. 10.1016/j.cocis.2016.02.010. [DOI] [Google Scholar]
  324. Brom J. A.; Pielak G. J. Desiccation-Tolerance and Globular Proteins Adsorb Similar Amounts of Water. Protein Sci. 2022, 31, e4288. 10.1002/pro.4288. [DOI] [PMC free article] [PubMed] [Google Scholar]
  325. Sanchez-Martinez S.; Ramirez J. F.; Meese E. K.; Childs C. A.; Boothby T. C. The Tardigrade Protein CAHS D Interacts with, but does not Retain, Water in Hydrated and Desiccated Systems. Sci. Rep. 2023, 13, 10449. 10.1038/s41598-023-37485-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  326. Olsson C.; Jansson H.; Swenson J. The Role of Trehalose for the Stabilization of Proteins. J. Phys. Chem. B 2016, 120, 4723–4731. 10.1021/acs.jpcb.6b02517. [DOI] [PubMed] [Google Scholar]
  327. Bellavia G.; Giuffrida S.; Cottone G.; Cupane A.; Cordone L. Protein Thermal Denaturation and Matrix Glass Transition in Different Protein-Trehalose-Water Systems. J. Phys. Chem. B 2011, 115, 6340–6346. 10.1021/jp201378y. [DOI] [PubMed] [Google Scholar]
  328. Bellavia G.; Cottone G.; Giuffrida S.; Cupane A.; Cordone L. Thermal Denaturation of Myoglobin in Water-Disaccharide Matrixes: Relation with the Glass Transition of the System. J. Phys. Chem. B 2009, 113, 11543–11549. 10.1021/jp9041342. [DOI] [PubMed] [Google Scholar]
  329. Bellavia G.; Cordone L.; Cupane A. Calorimetric Study of Myoglobin Embedded in Trehalose-Water Matrixes. J. Therm. Anal. Calorim. 2009, 95, 699–702. 10.1007/s10973-008-9490-4. [DOI] [Google Scholar]
  330. Crowe J. H.; Spargo B. J.; Crowe L. M. Preservation of Dry Liposomes does not Require Retention of Residual Water. Proc. Natl. Acad. Sci. U. S. A. 1987, 84, 1537–1540. 10.1073/pnas.84.6.1537. [DOI] [PMC free article] [PubMed] [Google Scholar]
  331. Stadmiller S. S.; Pielak G. J. Protein-Complex Stability in Cells and in Vitro Under Crowded Conditions. Curr. Opin. Struct. Biol. 2021, 66, 183–192. 10.1016/j.sbi.2020.10.024. [DOI] [PubMed] [Google Scholar]
  332. Belton P. S.; Gil A. M. IR and Raman Spectroscopic Studies of the Interaction of Trehalose with Hen Egg White Lysozyme. Biopolymers 1994, 34, 957–961. 10.1002/bip.360340713. [DOI] [PubMed] [Google Scholar]
  333. Crilly C. J.; Brom J. A.; Kowalewski M. E.; Piszkiewicz S.; Pielak G. J. Dried Protein Structure Revealed at the Residue Level by Liquid-Observed Vapor Exchange NMR. Biochemistry 2021, 60, 152–159. 10.1021/acs.biochem.0c00863. [DOI] [PMC free article] [PubMed] [Google Scholar]
  334. Crilly C. J.; Eicher J. E.; Warmuth O.; Atkin J. M.; Pielak G. J. Water’s Variable Role in Protein Stability Uncovered by Liquid-Observed Vapor Exchange NMR. Biochemistry 2021, 60, 3041–3045. 10.1021/acs.biochem.1c00552. [DOI] [PMC free article] [PubMed] [Google Scholar]
  335. Crilly C. J.; Noonan Brom J. A.; Rockcliffe D. A.; Pielak G. J.. Liquid-Observed Vapor Exchange (LOVE) NMR Reveals Residue-Level Effects of Protectants on a Dried Protein. Biophys. J. 2020, 118, 36a. 10.1016/j.bpj.2019.11.37931839262 [DOI] [Google Scholar]
  336. Bolis D.; Politou A. S.; Kelly G.; Pastore A.; Andrea Temussi P. Protein Stability in Nanocages: A Novel Approach for Influencing Protein Stability by Molecular Confinement. J. Mol. Biol. 2004, 336, 203–212. 10.1016/j.jmb.2003.11.056. [DOI] [PubMed] [Google Scholar]
  337. Sampedro J. G.; Uribe S. Trehalose-Enzyme Interactions Result in Structure Stabilization and Activity Inihibition. The Role of Viscosity. Mol. Cell. Biochem. 2004, 256–257, 319–327. 10.1023/B:MCBI.0000009878.21929.eb. [DOI] [PubMed] [Google Scholar]
  338. Hagen S. J.; Hofrichter J.; Eaton W. A. Protein Reaction Kinetics in a Room-Temperature Glass. Science 1995, 269, 959–962. 10.1126/science.7638618. [DOI] [PubMed] [Google Scholar]
  339. Giuffrida S.; Cottone G.; Librizzi F.; Cordone L. Coupling between the Thermal Evolution of the Heme Pocket and the External Matrix Structure in Trehalose Coated Carboxymyoglobin. J. Phys. Chem. B 2003, 107, 13211–13217. 10.1021/jp036194x. [DOI] [Google Scholar]
  340. Carpenter J. F.; Crowe J. H. An Infrared Spectroscopic Study of the Interactions of Carbohydrates with Dried Proteins. Biochemistry 1989, 28, 3916–3922. 10.1021/bi00435a044. [DOI] [PubMed] [Google Scholar]
  341. Grasmeijer N.; Stankovic M.; De Waard H.; Frijlink H. W.; Hinrichs W. L. J. Unraveling Protein Stabilization Mechanisms: Vitrification and Water Replacement in a Glass Transition Temperature Controlled System. Biochimica et Biophysica Acta (BBA) - Proteins and Proteomics 2013, 1834, 763–769. 10.1016/j.bbapap.2013.01.020. [DOI] [PubMed] [Google Scholar]
  342. Doekhie A.; Slade M. N.; Cliff L.; Weaver L.; Castaing R.; Paulin J.; Chen Y. C.; Edler K. J.; Koumanov F.; Marchbank K. J.; et al. Thermal Resilience of Ensilicated Lysozyme via Calorimetric and in vivo Analysis. RSC Adv. 2020, 10, 29789–29796. 10.1039/D0RA06412B. [DOI] [PMC free article] [PubMed] [Google Scholar]
  343. Killian M. S.; Taylor A. J.; Castner D. G.. Stabilization of Dry Protein Coatings with Compatible Solutes. Biointerphases 2018, 13, 06E401. 10.1116/1.5031189 [DOI] [PMC free article] [PubMed] [Google Scholar]
  344. Cottone G.; Ciccotti G.; Cordone L. Protein–Trehalose–Water Structures in Trehalose Coated Carboxy-Myoglobin. J. Chem. Phys. 2002, 117, 9862–9866. 10.1063/1.1518960. [DOI] [Google Scholar]
  345. Lerbret A.; Affouard F. Molecular Packing, Hydrogen Bonding, and Fast Dynamics in Lysozyme/Trehalose/Glycerol and Trehalose/Glycerol Glasses at Low Hydration. J. Phys. Chem. B 2017, 121, 9437–9451. 10.1021/acs.jpcb.7b07082. [DOI] [PubMed] [Google Scholar]
  346. Cottone G.; Giuffrida S.; Ciccotti G.; Cordone L. Molecular Dynamics Simulation of Sucrose- and Trehalose-Coated Carboxy-Myoglobin. Proteins Struct. Funct. Bioinf. 2005, 59, 291–302. 10.1002/prot.20414. [DOI] [PubMed] [Google Scholar]
  347. Corradini D.; Strekalova E. G.; Eugene Stanley H.; Gallo P. Microscopic Mechanism of Protein Cryopreservation in an Aqueous Solution with Trehalose. Sci. Rep. 2013, 3, 1218. 10.1038/srep01218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  348. Thirumalai D.; Klimov D. K.; Lorimer G. H. Caging Helps Proteins Fold. Proc. Natl. Acad. Sci. U.S.A. 2003, 100, 11195–11197. 10.1073/pnas.2035072100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  349. Zhou H. X. Protein Folding in Confined and Crowded Environments. Arch. Biochem. Biophys. 2008, 469, 76–82. 10.1016/j.abb.2007.07.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  350. Takagi F.; Koga N.; Takada S. How Protein Thermodynamics and Folding Mechanisms are Altered by the Chaperonin Cage: Molecular simulations. Proc. Natl. Acad. Sci. U.S.A. 2003, 100, 11367–11372. 10.1073/pnas.1831920100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  351. Ravindra R.; Zhao S.; Gies H.; Winter R. Protein Encapsulation in Mesoporous Silicate: The Effects of Confinement on Protein Stability, Hydration, and Volumetric Properties. J. Am. Chem. Soc. 2004, 126, 12224–12225. 10.1021/ja046900n. [DOI] [PubMed] [Google Scholar]
  352. Campanini B.; Bologna S.; Cannone F.; Chirico G.; Mozzarelli A.; Bettati S. Unfolding of Green Fluorescent Protein mut2 in wet nanoporous silica gels. Protein Sci. 2005, 14, 1125–1133. 10.1110/ps.041190805. [DOI] [PMC free article] [PubMed] [Google Scholar]
  353. Wheeler K. E.; Nocek J. M.; Hoffman B. M. NMR Spectroscopy Can Characterize Proteins Encapsulated in a Sol-Gel Matrix. J. Am. Chem. Soc. 2006, 128, 14782–14783. 10.1021/ja066244m. [DOI] [PubMed] [Google Scholar]
  354. Zhao S.; Gies H.; Winter R. Stability of Proteins Confined in MCM-48 Mesoporous Molecular Sieves - The Effects of pH, Temperature and Co-Solvents. Zeitschrift fur Physikalische Chemie 2007, 221, 139–154. 10.1524/zpch.2007.221.1.139. [DOI] [Google Scholar]
  355. Simorellis A. K.; Flynn P. F. Fast Local Backbone Dynamics of Encapsulated Ubiquitin. J. Am. Chem. Soc. 2006, 128, 9580–9581. 10.1021/ja061705p. [DOI] [PubMed] [Google Scholar]
  356. Shi Z.; Peterson R. W.; Wand A. J. New Reverse Micelle Surfactant Systems Optimized for High-Resolution NMR Spectroscopy of Encapsulated Proteins. Langmuir 2005, 21, 10632–10637. 10.1021/la051409a. [DOI] [PubMed] [Google Scholar]
  357. Denesyuk N. D.; Thirumalai D.. Theory and Simulations for Crowding-Induced Changes in Stability of Proteins with Applications to λ Repressor. arXiv 2020, 2012.11111. 10.48550/arXiv.2012.11111 [DOI] [Google Scholar]
  358. Klimov D. K.; Newfield D.; Thirumalai D. Simulations of β-Hairpin Folding Confined to Spherical Pores Using Distributed Computing. Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 8019–8024. 10.1073/pnas.072220699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  359. Friedel M.; Sheeler D. J.; Shea J. E. Effects of Confinement and Crowding on the Thermodynamics and Kinetics of Folding of a Minimalist β-Barrel Protein. J. Chem. Phys. 2003, 118, 8106–8113. 10.1063/1.1564048. [DOI] [Google Scholar]
  360. Butler S. L.; Falke J. J. Effects of Protein Stabilizing Agents on Thermal Backbone Motions: A Disulfide Trapping Study. Biochemistry 1996, 35, 10595–10600. 10.1021/bi961107v. [DOI] [PMC free article] [PubMed] [Google Scholar]
  361. Kramers H. A. Brownian Motion in a Field of Force and the Diffusion Model of Chemical Reactions. Physica 1940, 7, 284–304. 10.1016/S0031-8914(40)90098-2. [DOI] [Google Scholar]
  362. Jacob M.; Schmid F. X. Protein Folding as a Diffusional Process. Biochemistry 1999, 38, 13773–13779. 10.1021/bi991503o. [DOI] [PubMed] [Google Scholar]
  363. Slade L.; Levine H. Glass Transitions and Water-Food Structure Interactions. Adv. Food Nutr. Res. 1995, 38, 103–269. 10.1016/S1043-4526(08)60084-4. [DOI] [PubMed] [Google Scholar]
  364. Goldberg J. M.; Baldwin R. L. Kinetic Mechanism of a Partial Folding Reaction. 2. Nature of the Transition State. Biochem 1998, 37, 2556–2563. 10.1021/bi972403q. [DOI] [PubMed] [Google Scholar]
  365. Chrunyk B. A.; Matthews C. R. Role of Diffusion in the Folding of the a Subunit of Tryptophan Synthase from Escherichia coli. Biochem 1990, 29, 2149–2154. 10.1021/bi00460a027. [DOI] [PubMed] [Google Scholar]
  366. Vaucheret H.; Signon L.; Le Bras G.; Garel J. R.; et al. Mechanism of Renaturation of a Large Protein, Aspartokinase-Homoserine Dehydrogenase. Biochemistry 1987, 26, 2785–2790. 10.1021/bi00384a020. [DOI] [PubMed] [Google Scholar]
  367. Caliskan G.; Mechtani D.; Roh J. H.; Kisliuk A.; Sokolov A. P.; Azzam S.; Cicerone M. T.; Lin-Gibson S.; Peral I. Protein and solvent dynamics: How Strongly are They Coupled?. J. Chem. Phys. 2004, 121, 1978–1983. 10.1063/1.1764491. [DOI] [PubMed] [Google Scholar]
  368. Duddu S. P.; Zhang G.; Dal Monte P. R. The Relationship between Protein Aggregation and Molecular Mobility Below the Glass Transition Temperature of Lyophilized Formulations Containing a Monoclonal Antibody. Pharm. Res. 1997, 14, 596–600. 10.1023/A:1012196826905. [DOI] [PubMed] [Google Scholar]
  369. Pikal M. J.; Rigsbee D.; Roy M. L.; Galreath D.; Kovach K. J.; Wang B.; Carpenter J. F.; Cicerone M. T. Solid State Chemistry of Proteins: II. The Correlation of Storage Stability of Freeze-Dried Human Growth Hormone (hGH) with Structure and Dynamics in the Glassy Solid. J. Pharm. Sci. 2008, 97, 5106–5121. 10.1002/jps.21374. [DOI] [PubMed] [Google Scholar]
  370. Cicerone M. T.; Douglas J. F. β-Relaxation Governs Protein Stability in Sugar-Glass Matrices. Soft Matter 2012, 8, 2983–2991. 10.1039/c2sm06979b. [DOI] [Google Scholar]
  371. Cordone L.; Cottone G.; Giuffrida S.; Palazzo G.; Venturoli G.; Viappiani C. Internal Dynamics and Protein–Matrix Coupling in Trehalose-Coated Proteins. Biochim. Biophys. Acta 2005, 1749, 252–281. 10.1016/j.bbapap.2005.03.004. [DOI] [PubMed] [Google Scholar]
  372. Anopchenko A.; Psurek T.; Vanderhart D.; Douglas J. F.; Obrzut J.. Dielectric Study of the Antiplasticization of Trehalose by Glycerol. Phys. Rev. E: Stat. Nonlinear Soft Matter Phys. 2006, 74, 031501. 10.1103/PhysRevE.74.031501 [DOI] [PubMed] [Google Scholar]
  373. Obrzut J.; Anopchenko A.; Douglas J. F.; Rust B. W. Relaxation and Antiplasticization Measurements in Trehalose–Glycerol Mixtures – A Model Formulation for Protein Preservation. J. Non-Cryst. Solids 2010, 356, 777–781. 10.1016/j.jnoncrysol.2009.07.045. [DOI] [Google Scholar]
  374. Hatley R. H. M.; Blair J. A. Stabilisation and Delivery of Labile Materials by Amorphous Carbohydrates and their Derivatives. J. Mol. Catal. B Enzym 1999, 7, 11–19. 10.1016/S1381-1177(99)00018-1. [DOI] [Google Scholar]
  375. Hancock B. C.; Shamblin S. L.; Zografi G. Molecular Mobility of Amorphous Pharmaceutical Solids Below Their Glass Transition Temperatures. Pharm. Res. 1995, 12, 799–806. 10.1023/A:1016292416526. [DOI] [PubMed] [Google Scholar]
  376. Hatley R. H.; Glass M. Fragility and the Stability of Pharmaceutical Preparations—Excipient Selection. Pharm. Dev. Technol. 1997, 2, 257–264. 10.3109/10837459709031445. [DOI] [PubMed] [Google Scholar]
  377. Ekdawi-Sever N. C.; Conrad P. B.; De Pablo J. J. Molecular Simulation of Sucrose Solutions Near the Glass Transition Temperature. J. Phys. Chem. A 2001, 105, 734–742. 10.1021/jp002722i. [DOI] [Google Scholar]
  378. Allison S. D.; Chang B.; Randolph T. W.; Carpenter J. F. Hydrogen Bonding between Sugar and Protein Is Responsible for Inhibition of Dehydration-Induced Protein Unfolding. Arch. Biochem. Biophys. 1999, 365, 289–298. 10.1006/abbi.1999.1175. [DOI] [PubMed] [Google Scholar]
  379. Chang L. L.; Pikal M. J. Mechanisms of Protein Stabilization in the Solid State. J. Pharm. Sci. 2009, 98, 2886–2908. 10.1002/jps.21825. [DOI] [PubMed] [Google Scholar]
  380. Quinn P. J.; Koynova R. D.; Lis L. J.; Tenchov B. G. Lamellar Gel-Lamellar Liquid Crystal Phase Transition of Dipalmitoylphosphatidylcholine Multilayers Freeze-Dried from Aqueous Trehalose Solutions. A Real-Time X-Ray Diffraction Study. Biochim. Biophys. Acta 1988, 942, 315–323. 10.1016/0005-2736(88)90033-8. [DOI] [PubMed] [Google Scholar]
  381. Madden T. D.; Bally M. B.; Hope M. J.; Cullis P. R.; Schieren H. P.; Janoff A. S. Protection of Large Unilamellar Vesicles by Trehalose During Dehydration: Retention of Vesicle Contents. Biochim. Biophys. Acta 1985, 817, 67–74. 10.1016/0005-2736(85)90069-0. [DOI] [PubMed] [Google Scholar]
  382. Crowe L. M.; Crowe J. H.; Rudolph A.; Womersley C.; Appel L. Preservation of Freeze-Dried Liposomes by Trehalose. Arch. Biochem. Biophys. 1985, 242, 240–247. 10.1016/0003-9861(85)90498-9. [DOI] [PubMed] [Google Scholar]
  383. Strauss G.; Hauser H. Stabilization of Lipid Bilayer Vesicles by Sucrose During Freezing. Proc. Natl. Acad. Sci. U. S. A. 1986, 83, 2422–2426. 10.1073/pnas.83.8.2422. [DOI] [PMC free article] [PubMed] [Google Scholar]
  384. Quinn P. J. Effect of Sugars on the Phase Behaviour of Phospholipid Model Membranes. Biochem. Soc. Trans. 1989, 17, 953–957. 10.1042/bst0170953. [DOI] [PubMed] [Google Scholar]
  385. Harrigan P. R.; Madden T. D.; Cullis P. R. Protection of Liposomes During Dehydration or Freezing. Chem. Phys. Lipids 1990, 52, 139–149. 10.1016/0009-3084(90)90157-M. [DOI] [PubMed] [Google Scholar]
  386. Crowe L. M.; Womersley C.; Crowe J. H.; Reid D.; Appel L.; Rudolph A. Prevention of Fusion and Leakage in Freeze-Dried Liposomes by Carbohydrates. Biochim. Biophys. Acta, Proteins Proteomics 1986, 861, 131–140. 10.1016/0005-2736(86)90411-6. [DOI] [Google Scholar]
  387. Crowe L. M.; Crowe J. H.; Chapman D. Interaction of Carbohydrates with Dry Dipalmitoylphosphatidylcholine. Arch. Biochem. Biophys. 1985, 236, 289–296. 10.1016/0003-9861(85)90628-9. [DOI] [PubMed] [Google Scholar]
  388. Crowe J. H.; Crowe L. M.; Carpenter J. F.; Rudolph A. S.; Wistrom C. A.; Spargo B. J.; Anchordoguy T. J. Interactions of Sugars with Membranes. Biochimica et Biophysica Acta (BBA) - Reviews on Biomembranes 1988, 947, 367–384. 10.1016/0304-4157(88)90015-9. [DOI] [PubMed] [Google Scholar]
  389. Crowe L. M.; Crowe J. H. Trehalose and Dry Dipalmitoylphosphatidylcholine Revisited. Biochimica et Biophysica Acta (BBA) - Biomembranes 1988, 946, 193–201. 10.1016/0005-2736(88)90392-6. [DOI] [PubMed] [Google Scholar]
  390. Crowe J. H.; Crowe L. M.; Hoekstra F. A. Phase Transitions and Permeability Changes in Dry Membranes During Rehydration. J. Bioenerg Biomembr 1989, 21, 77–91. 10.1007/BF00762213. [DOI] [PubMed] [Google Scholar]
  391. Crowe J. H.; Crowe L. M.; Carpenter J. F.; Aurell Wistrom C. Stabilization of Dry Phospholipid Bilayers and Proteins by Sugars. Biochem. J. 1987, 242, 1. 10.1042/bj2420001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  392. Samuni U.; Navati M. S.; Juszczak L. J.; Dantsker D.; Yang M.; Friedman J. M. Unfolding and Refolding of Sol-Gel Encapsulated Carbonmonoxymyoglobin: An Orchestrated Spectroscopic Study of Intermediates and Kinetics. J. Phys. Chem. B 2000, 104, 10802–10813. 10.1021/jp000802g. [DOI] [Google Scholar]
  393. Imamura K.; Ohyama K. I.; Yokoyama T.; Maruyama Y.; Kazuhiro N. Temperature Scanning FTIR Analysis of Secondary Structures of Proteins Embedded in Amorphous Sugar Matrix. J. Pharm. Sci. 2009, 98, 3088–3098. 10.1002/jps.21568. [DOI] [PubMed] [Google Scholar]
  394. Liao Y. H.; Brown M. B.; Martin G. P. Investigation of the Stabilisation of Freeze-Dried Lysozyme and the Physical Properties of the Formulations. Eur. J. Pharm. Biopharm. 2004, 58, 15–24. 10.1016/j.ejpb.2004.03.020. [DOI] [PubMed] [Google Scholar]
  395. Wolkers W. F.; Hoekstra F. A. Heat Stability of Proteins in Desiccation-Tolerant Cattail (Typha latifolia L.) Pollen: a Fourier Transform Infrared Spectroscopic Study. Comp. Biochem. Physiol. A: Physiol. 1997, 117, 349–355. 10.1016/S0300-9629(96)00274-5. [DOI] [Google Scholar]
  396. Wolkers W. F.; Bochicchio A.; Selvaggi G.; Hoekstra F. A. Fourier Transform Infrared Microspectroscopy Detects Changes in Protein Secondary Structure Associated with Desiccation Tolerance in Developing Maize Embryos. Plant Physiol. 1998, 116, 1169–1177. 10.1104/pp.116.3.1169. [DOI] [PMC free article] [PubMed] [Google Scholar]
  397. Li J.; Chen J.; An L.; Yuan X.; Yao L.. Polyol and Sugar Osmolytes can Shorten Protein Hydrogen Bonds to Modulate Function. Commun. Biol. 2020, 3, 528. 10.1038/s42003-020-01260-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  398. Gilman-Politi R.; Harries D.. Unraveling the Molecular Mechanism of Enthalpy Driven Peptide Folding by Polyol Osmolytes. J. Chem. Theory Comput. 2011, 7, 3816–3828. 10.1021/ct200455n [DOI] [PubMed] [Google Scholar]
  399. Wang W.; Nema S.; Teagarden D. Protein aggregation—Pathways and Influencing Factors. Int. J. Pharm. 2010, 390, 89–99. 10.1016/j.ijpharm.2010.02.025. [DOI] [PubMed] [Google Scholar]
  400. Zimmerman S. B.; Trach S. O. Estimation of Macromolecule Concentrations and Excluded Volume Effects for the Cytoplasm of Escherichia coli. J. Mol. Biol. 1991, 222, 599–620. 10.1016/0022-2836(91)90499-V. [DOI] [PubMed] [Google Scholar]
  401. Vecchi G.; Sormanni P.; Mannini B.; Vandelli A.; Tartaglia G. G.; Dobson C. M.; Hartl F. U.; Vendruscolo M. Proteome-Wide Observation of the Phenomenon of Life on the Edge of Solubility. Proc. Natl. Acad. Sci. U.S.A. 2020, 117, 1015–1020. 10.1073/pnas.1910444117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  402. Lueckel B.; Helk B.; Bodmer D.; Leuenberger H. Effects of Formulation and Process Variables on the Aggregation of Freeze-Dried Interleukin-6 (IL-6) After Lyophilization and on Storage. Pharm. Dev. Technol. 1998, 3, 337–346. 10.3109/10837459809009861. [DOI] [PubMed] [Google Scholar]
  403. Garzon-Rodriguez W.; Koval R. L.; Chongprasert S.; Krishnan S.; Randolph T. W.; Warne N. W.; Carpenter J. F. Optimizing Storage Stability of Lyophilized Recombinant Human Interleukin-11 with Disaccharide/Hydroxyethyl Starch Mixtures. J. Pharm. Sci. 2004, 93, 684–696. 10.1002/jps.10587. [DOI] [PubMed] [Google Scholar]
  404. Strickley R. G.; Anderson B. D. Solid-state Stability of Human Insulin II. Effect of Water on Reactive Intermediate Partitioning in Lyophiles from pH 2–5 Solutions: Stabilization Against Covalent Dimer Formation. J. Pharm. Sci. 1997, 86, 645–653. 10.1021/js9700311. [DOI] [PubMed] [Google Scholar]
  405. Sane S. U.; Wong R.; Hsu C. C. Raman Spectroscopic Characterization of Drying-Induced Structural Changes in a Therapeutic Antibody: Correlating Structural Changes with Long-Term Stability. J. Pharm. Sci. 2004, 93, 1005–1018. 10.1002/jps.20014. [DOI] [PubMed] [Google Scholar]
  406. Chang B. S.; Beauvais R. M.; Dong A.; Carpenter J. F. Physical Factors Affecting the Storage Stability of Freeze-Dried Interleukin-1 Receptor Antagonist: Glass Transition and Protein Conformation. Arch. Biochem. Biophys. 1996, 331, 249–258. 10.1006/abbi.1996.0305. [DOI] [PubMed] [Google Scholar]
  407. Perez-Moral N.; Adnet C.; Noel T. R.; Parker R. Characterization of the Rate of Thermally-Induced Aggregation of β-Lactoglobulin and Its Trehalose Mixtures in the Glass State. Biomacromolecules 2010, 11, 2985–2992. 10.1021/bm100785d. [DOI] [PubMed] [Google Scholar]
  408. Lerbret A.; Affouard F.; Hédoux A.; Krenzlin S.; Siepmann J.; Bellissent-Funel M.-C.; Descamps M. How Strongly Does Trehalose Interact with Lysozyme in the Solid State? Insights from Molecular Dynamics Simulation and Inelastic Neutron Scattering. J. Phys. Chem. B 2012, 116, 11103–11116. 10.1021/jp3058096. [DOI] [PubMed] [Google Scholar]
  409. Kumar V.; Dixit N.; Zhou L.; Fraunhofer W. Impact of Short Range Hydrophobic Interactions and Long Range Electrostatic Forces on the Aggregation Kinetics of a Monoclonal Antibody and a Dual-Variable Domain Immunoglobulin at Low and High Concentrations. Int. J. Pharm. 2011, 421, 82–93. 10.1016/j.ijpharm.2011.09.017. [DOI] [PubMed] [Google Scholar]
  410. Arakawa T.; Tsumoto K.; Nagase K.; Ejima D. The Effects of Arginine on Protein Binding and Elution in Hydrophobic Interaction and Ion-Exchange Chromatography. Protein Expr Purif 2007, 54, 110–116. 10.1016/j.pep.2007.02.010. [DOI] [PubMed] [Google Scholar]
  411. Durbin S. D.; Feher G. Protein Crystallization. Annu. Rev. Phys. Chem. 1996, 47, 171–204. 10.1146/annurev.physchem.47.1.171. [DOI] [PubMed] [Google Scholar]
  412. Calamai M.; Chiti F.; Dobson C. M. Amyloid Fibril Formation can Proceed from Different Conformations of a Partially Unfolded Protein. Biophys. J. 2005, 89, 4201–4210. 10.1529/biophysj.105.068726. [DOI] [PMC free article] [PubMed] [Google Scholar]
  413. Dobson C. M. Protein Folding and Misfolding. Nature 2003 426:6968 2003, 426, 884–890. 10.1038/nature02261. [DOI] [PubMed] [Google Scholar]
  414. Sukenik S.; Sapir L.; Harries D.. Osmolyte Induced Changes in Peptide Conformational Ensemble Correlate with Slower Amyloid Aggregation: A Coarse-Grained Simulation Study. J. Chem. Theory Comput. 2015, 11, 5918–5928. 10.1021/acs.jctc.5b00657 [DOI] [PubMed] [Google Scholar]
  415. Phan-Xuan T.; Bogdanova E.; Millqvist Fureby A.; Fransson J.; Terry A. E.; Kocherbitov V. Hydration-Induced Structural Changes in the Solid State of Protein: A SAXS/WAXS Study on Lysozyme. Mol. Pharmaceutics 2020, 17, 3246–3258. 10.1021/acs.molpharmaceut.0c00351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  416. Nagase H.; Ogawa N.; Endo T.; Shiro M.; Ueda H.; Sakurai M. Crystal Structure of an Anhydrous Form of Trehalose: Structure of Water Channels of Trehalose Polymorphism. J. Phys. Chem. B 2008, 112, 9105–9111. 10.1021/jp800936z. [DOI] [PubMed] [Google Scholar]
  417. Gelman Constantin J.; Schneider M.; Corti H. R. Glass Transition Temperature of Saccharide Aqueous Solutions Estimated with the Free Volume/Percolation Model. J. Phys. Chem. B 2016, 120, 5047–5055. 10.1021/acs.jpcb.6b01841. [DOI] [PubMed] [Google Scholar]
  418. Miller D. P.; De Pablo J. J.; Corti H. Thermophysical Properties of Trehalose and Its Concentrated Aqueous Solutions. Pharm. Res. 1997, 14, 578–590. 10.1023/A:1012192725996. [DOI] [PubMed] [Google Scholar]
  419. Ngai K. L.; Casalini R., Capaccioli S., Paluch M.; Roland C. M.. Dispersion of the Structural Relaxation and the Vitrification of Liquids. In Fractals, Diffusion, and Relaxation in Disordered Complex Systems: Advances in Chemical Physics, Part B; Coffey W. T., Kalmykov Y. P., Eds.; Wiley, 2006; Chapter 10, pp 497–593. 10.1002/0470037148.ch10 [DOI] [Google Scholar]

Articles from Chemical Reviews are provided here courtesy of American Chemical Society

RESOURCES