Pre-Molten, Wet, and Dry Molten Globules: Comparison
Please note this is a comparison between Version 4 by Vladimir N. Uversky and Version 3 by Rita Xu.

Transitions between the unfolded and native states of the ordered globular proteins are accompanied by the accumulation of several intermediates, such as pre-molten globules, wet molten globules, and dry molten globules. Structurally equivalent conformations can serve as native functional states of intrinsically disordered proteins.

  • baroenzymology
  • cryoenzymology
  • intrinsically disordered proteins
  • macromolecular crowding

1. Introduction

In the classical picture of enzymology, the native structure of a protein is intimately correlated to its function [1], and the functional 3D structure of proteins is determined solely by their amino acid sequences [2][3][4][5]. A deviation from the native structure accompanied by the loss of biological activity was defined as protein denaturation. Hence, study of the process of the unfolding of a protein molecule (under various denaturing conditions) was as responsible for gaining crucial knowledge on the structure-function relationships in proteins as investigating protein refolding. In the three somewhat related phenomena, protein folding (i.e., spontaneous formation of a 3D structure by the nascent polypeptides in the cell), protein unfolding, and protein refolding from the unfolded state in the test tube, the transition between native structure and unfolded/denatured structure(s) was the common thread [6].
The transition between the native and denatured states of small globular proteins was initially considered a two-state process (two-state model of protein unfolding) [7][8][9]. Over the years, two generic folding intermediates were identified: the molten globule (MG) [10][11][12][13][14][15][16][17][18][19][20][21] and the pre-molten globule (PMG) [22][23][24][25][26]. Curiously, the existence of such folding intermediates was predicted in 1973 by Oleg B. Ptitsyn (1929–1999) based on the theoretical considerations of the potential mechanisms by which the hierarchical structure of a native globular protein can be rapidly formed, despite the astronomically large number of possibilities by which a polypeptide chain can be packed into a compact globule [27]. More recently, the concepts of “wet” and “dry” molten globules have emerged, where dry molten globular (DMG) intermediates represent an expanded form of the native protein with a dry core [28][29][30][31][32][33][34][35][36][37][38]. This is an interesting observation, as early studies indicated that the molten globule represents a highly hydrated state, with water inside the molten globule interior possessing characteristics of a highly associated liquid [21]. However, as early as in 1989, theoretical analysis of the denatured states of globular proteins suggested that since the compactness of a denatured protein may vary within a wide range, several denatured forms can be distinguished, such as coil, swollen globule, the “wet” molten globule (the compact state with pores occupied by solvent), and the “dry” molten globule where solvent does not penetrate inside the protein [39].

2. Molten Globule as an (Un)Folding Intermediate

Historically, protein denaturation and unfolding studies are based on the well-accepted and rather obvious (at least now) mantra stating, “Structure does exit since it can be broken”. These studies played crucial roles in establishing protein science in general and in understanding the basis of the correlation between protein amino acid sequence and function in particular. As early as in 1931, Hsien Wu (1893–1959) proposed the first theory of protein denaturation: the active structure is known to exist because it is destroyed by denaturation [40][41]. His paper published in the Chinese Journal of Physiology contained the first statement that protein function depends on prior structure [40][41]. However, even earlier, in 1925, Mortimer Louis Anson (1901–1968) and Alfred Ezra Mirsky (1900–1974) showed that intact hemoglobin can exist as such near the neutral point only, whereas dilute acid or alkali changed it to the denatured form, which could fold back to its native state upon restoration of native conditions, indicating that protein denaturation and unfolding are reversible processes [42]. In 1936, the first Western review on protein denaturation that represents the first modern theory of native and denatured proteins was published, where Alfred Mirsky and Linus Pauling (1901–1994) stated that the loss of certain highly specific properties constitutes the most significant change that occurs in the denaturation of a native protein [43]. By 1944, it became clear that native proteins have unique structures, that the denaturation processes are manifold in nature and magnitude, and that the addition of high concentrations of strong denaturants, such as guanidine hydrochloride (GdmHCl) or urea, to a protein causes a complete (or almost complete) disruption of all conformational interactions and, as consequence, to the transformation of a protein molecule into the highly disordered state of a random coil [44]. Furthermore, the authors of this seminal review stated: “The term denaturation has been used rather loosely and indiscriminately to denote ill-defined changes in the properties of proteins, caused by a variety of chemical, physical, and biological agents. The observation that many unrelated processes may cause similar changes in a protein early led to the belief that any single change, such as the formation of a coagulum, suffices to characterize a ‘denatured’ protein, and that all denaturing agents are alike in their action. Although proteins are now known to respond differently to various kinds of denaturation, the supposition of the singleness of the denaturation process has persisted” [44]. They also defined denaturation as “any non-proteolytic modification of the unique structure of a native protein, giving rise to definite changes in chemical, physical, or biological properties” [44]. It is obvious that a clear distinction should be made between the terms “denaturation” and “unfolding”. Here, as defined above, denaturation is a process leading to the elimination of protein functionality resulting from the disruption of functional 3D structure. This can be triggered by a wide range of conditions, with the resulting denatured forms possessing a wide spectrum of properties depending on the conditions in which they were achieved. On the contrary, protein unfolding is defined as a process leading to the complete elimination of all the conformational forces stabilizing the native protein structure, resulting, therefore, in the formation of a coil-like conformation. Retrospectively, finding partially folded species of globular proteins under a variety of denaturing conditions should not be surprising. This is because the unique 3-D structure of a protein molecule is stabilized by specific non-covalent interactions, such as hydrogen bonds, hydrophobic interactions, electrostatic interactions and salt bridges, and van der Waals interactions. Since these conformational forces have different physical natures, it is quite possible that they would react differently to the changes in the environment, where under specific conditions, some forces would decline and dissipate, whereas others would stay unchanged or even strengthen. In these cases, the protein molecule is obviously losing its biological activity; i.e., it is becoming denatured, but since not all the conformational forces are “shutdown”, denaturation is not necessarily accompanied by the complete unfolding of a protein, giving rise to the appearance of new conformations with properties halfway between those of native and completely unfolded states. Therefore, various degrees of denaturation/unfolding must exist, depending on the extent to which the structure of the protein has been modified under given conditions. Clearly, the fact that the extent of denaturation can be different is incompatible with the “all-or-none” hypothesis that a given protein can exist in only one of two states, the completely native or the completely denatured/unfolded [44]. These important considerations were rooted in the experimental evidence accumulated in the 1930s and 1940s, when the incomplete unfolding and existence of some intermediate stages of denaturation were recognized in several instances [45][46][47][48][49]. Furthermore, as follows from later studies, some denatured forms produced at milder denaturing conditions (e.g., heat- or pH-denatured proteins) can undergo additional structural alterations in the presence of strong denaturants, such as urea or GdmHCl [50]. Therefore, since the final denatured conformations of proteins are strongly dependent on the denaturing agent, not all denatured states are structurally similar, and under certain conditions the protein molecules are not completely unfolded. These very logical conclusions were formulated in a classical review by Charles Tanford (1921–2009) [9], which was one of the first papers providing in-depth analysis of the possibility that during the unfolding of globular proteins, accumulation of some equilibrium intermediate states might be expected. Unfortunately, since the results that were available at that time were too scanty, no serious generalization could be made. Furthermore, the vast majority of then reported studies suggested that accumulation of an intermediate during protein unfolding was regarded as an exception to the rule, whereas a conformational transition described by a two-state model represented the “normal” response of a protein to changes in its environment. Although proteins were shown to respond differently to various kinds of denaturation, the supposition of the singleness of the denaturation process persisted [44]. For the first time, an intermediate state accumulating during the unfolding process was identified as early as in 1973 by Tanford’s group while looking at the chemical unfolding of bovine carbonic anhydrase B (BCAB) by GdmHCl [51]. It is notable that the intermediate state identified by far- and near-UV circular dichroism (CD) spectroscopy was described as having the secondary structure of the native state but as having lost the tertiary structure [51]. A year later, Kin-Ping Wong and Larry M. Hamlin used circular dichroism, difference spectrophotometry, enzymatic activity, and viscosity to study acid denaturation of these proteins and showed that the denatured acid BCAB was enzymatically inactive and did not have a unique 3D structure as judged by near-UV CD; it also did not exist in the random-coiled state as indicated by viscosity and far-UV CD [52]. Around the same time, Pititsyn’s group [27] initiated their work, which eventually led to further early insights into the folding intermediate. It was suggested that the formation of a native-like secondary structure preceded the proteins acquiring their tertiary structures. The results of the analysis of acid- and temperature-induced denaturation from this group were found to support this notion [10][11][13]. It was Ohgushi and Wada who in 1983 coined the term “molten globule” to describe such folding intermediates [12]. The most defining characteristics of a “classic” MG are outlined below [14][15][16][25][53][54][55][56][57][58][59][60][61][62][63]. A protein molecule in the MG state is characterized by the presence of a significant secondary structure (which is often classified as native-like secondary structure) with no or little tertiary structure (tight packing of side chains of amino acid residues is absent). Furthermore, 2D-NMR coupled with a hydrogen-deuterium exchange showed that the protein molecule in the MG state is characterized not only by the native-like secondary structure content, but also by the native-like folding pattern [64][65][66][67][68][69][70][71][72][73]. Small-angle X-ray scattering (SAXS) analysis revealed that the molten globular proteins possess globular structure typical of native globular proteins [74][75][76][77][78][79]. In agreement with the preservation of globular structure, the protein molecule in this state is characterized by a high degree of compactness, as its expansion typically leads to a general increase of 10–20% in radius of gyration or a hydrodynamic radius (over the native state), which corresponds to the volume increase of ~50% [55][80]. A considerable increase in the accessibility of a protein molecule to proteases was noted as a specific property of the MG state [81][82][83][84][85][86][87]. There was also an increase in the solvent exposure of the hydrophobic core, which was now less compact than the core of a native globular protein. This was reflected in the characteristic capability of the MG to specifically bind a hydrophobic fluorescent probe 1-anilino-naphthalene-8-sulfonate (ANS) or 1,1′-Bis(4-anilino-5-naphthalenesulfonic acid) (bis-ANS) [88][89][90]. MGs can show substantial levels of structure in some cases [53]. Lynne Regan reported that one part of a protein can retain the native structure, whereas another part forms an MG [91]. That is expected as proteins in general are characterized by noticeable structural heterogeneity, and conformational stability/flexibility can vary across the protein regions [92][93][94]. The abundant existence of intrinsically disordered proteins (IDPs) with various levels of disorder, and the presence of intrinsically disordered protein regions (IDPRs) in numerous proteins serve as extreme examples of this phenomenon [81][92][93][94]. While earlier data on the denaturation/unfolding and refolding of small proteins were compatible with the two-state model comprised of N → D and D → N transitions, the fact that many proteins were shown to form MGs during their unfolding indicated that the reality was more complex, and one should consider protein unfolding as the sequential process N ↔ MG ↔ U. This clearly raised a question on the physical and thermodynamic nature of the corresponding N ↔ MG and MG ↔ U transitions. The answer to this question was retrieved first from the results of the multiparametric experimental analysis of equilibrium GdmHCl-induced unfolding of BCAB and S. aureus β-lactamase at 4 °C, which clearly showed that the molten globule was separated from the more unfolded states by the “all-or-none” transition (this was evidenced by the bimodal distribution function of the molecular dimensions within the transition from the molten globule to the unfolded state) [95]. Later, similar bimodal distribution in the HPLC gel-filtration profiles was observed within the unfolding pathways of the NAD+-dependent DNA ligase from the thermophile Thermus scotoductus [22][23]. In line with these observations, an analysis of then available data on the equilibrium urea- and GdmHCl-induced N → U, N → MG, and MG → U transitions of globular proteins revealed that the cooperativity of all these unfolding processes increased linearly with the increase of the molecular weight of the protein up to 25–30 kDa. This indicated that the solvent-induced transitions from the native to the unfolded state, from the native to the molten globule state, and from the molten globule to the unfolded state were characterized by an “all-or-none” nature, thereby suggesting that the molten globule represented a third thermodynamic state of a protein molecule [96][97]. The validity of this model was later supported by Vijay S. Pande and Daniel S. Rokhsar, who in 1998 analyzed the equilibrium properties of proteins with Monte Carlo simulations and showed that, in addition to a rigid native state and a nontrivial unfolded state, a generic phase diagram contained a thermodynamically distinct MG state, further supporting the idea that MG represented a third phase state of proteins [98].

3. Potential Functionality of Folding Intermediates

Even before the acknowledgement of the prevalence and biological importance of intrinsically disordered proteins with their considerable structural heterogeneity, it was recognized that folding intermediates, including MGs, might have biological relevance. One of the first notes about this scenario was a hypothesis that the MG state may be involved in the translocation of proteins across membranes [99]. This idea was successfully supported by experiments, and there is now enough evidence that translocation of proteins and their insertion into membranes involve the MG state [100][101][102][103][104]. Model systems with α-lactalbumin showed the binding of MG to lipid bilayers [105]. In general, globular proteins can be transformed into the MG states on interaction with the membrane surface [106]. Such N → MG transitions in the vicinity of a membrane can be induced by the action of the so-called “membrane field”, which is a combination of the local decrease in the effective dielectric constant of water near the organic surface with the effect of negative charges located on the membrane surface [107][108][109][110]. Release and loading of the large, tightly packed hydrophobic ligands from and to the globular proteins might be facilitated by the partial unfolding of the carrier (N → MG transition) resulting from the concerted action of the moderate local decrease of pH and of the dielectric constant in proximity to the target membranes [111]. Furthermore, many proteins responsible for the transport of large hydrophobic ligands might have MG properties in their preloaded apo-forms [112][113][114]. It was also shown that many carbohydrate- and amino acid-binding periplasmic protein in E. coli form molten globule, which bind to their respective ligands [115]. Chaperonins interact with MGs and prevent their aggregation [116]. Earlier, Martin et al. discussed how a chaperonin-mediated folding had an MG as an intermediate [117]. It was also pointed out that compact, MG-like intermediates are localized within a central cavity of the chaperonin GroEL [118][119][120]. Facilitated folding of actins and tubulins occurs via a nucleotide-dependent interaction between the cytoplasmic chaperonin and the distinctive folding intermediates [121]. The presence of MG during nascent peptide folding has been inferred [122]. Importantly, although aforementioned functionalities have been attributed to the MG-like conformations, the major emphasis of all these and similar studies was still focused on the assumption that these functional MGs were folding intermediates kinetically trapped by the chaperonins just after the protein biosynthesis but before proteins become completely folded [25][99][123] or appear as a result of point mutations preventing polypeptides from complete folding [25][124] or originate from the denaturing effects of the membrane field [99][100][101][102][103][104][105][106][107][108][109][110] or ligand binding or release [112][113][114]. However, the presence of MGs in the cells become an established fact. [125]. All these observations provided strong support to the validity and importance of the concept of MG as a folding intermediate of globular proteins in vivo.

4. How Can One Find Molten Globules, and Where Can They Be Found?

MGs of globular proteins are generally obtained by their mild denaturation that can be induced by acid, alkali, low to medium concentrations of chemical denaturants such as urea and GdmHCl, chaotropic salts, moderately high temperature, and, for some proteins, even by low temperature [126][127][128][129][130][131][132][133][134][135][136][137][138][139][140][141][142][143][144][145]. Later studies revealed that in some proteins, an MG can also be induced by various organic solvents [146][147][148][149]. However, it was also shown that fluorinated alcohols can preferentially stabilize α-helices leading to the formation of non-native helical structures in some all-β-sheet proteins. For example, such highly helical states were induced by 2,2,2-trifluoroethanol (TFE) in several all β-sheet proteins, such as cardiotoxin analogue II (CTX II), from the Taiwan cobra (Naja naja atra) [150], procerain, a cysteine protease from Calotropis procera [151], β-lactoglobulin [152][153][154][155] and mellitin, [152][155] to name a few. All β-sheets to mostly α-helical structure in β-lactoglobulin and mellitin were also induced by hexafluoroisopropanol (HFIP), as well as by non-fluorinated alcohols, isopropanol, ethanol, and methanol [152][155]. Curiously, it was pointed out that an alcohol-induced α-helical state of β-lactoglobulin structurally resembles a transiently populated folding intermediate with high levels of non-native α-helical structure, which is formed within a few milliseconds during the refolding of this protein [156], suggesting that an intermediate with the non-native α-helical structure can accumulate during the refolding process of β-lactoglobulin, emphasizing that the hierarchical model cannot correctly describe folding of some β-structural proteins, including β-lactoglobulin [154][156]. The secondary and tertiary structures were evaluated generally by far- and near-UVCD, respectively. Secondary structure can also be evaluated with Fourier-transform infrared spectroscopy (FTIR) or optical rotatory dispersion (ORD), whereas viscosity measurements, gel-filtration chromatography, dynamic light scattering (DLS), SAXS, and electron microscopy are used to track expansion of the molecular volume [61][62]. The decrease in the compactness accompanied by the increased solvent accessibility of the hydrophobic core is normally estimated by looking at the binding of the fluorescent dye ANS to a protein molecule [88][89][90]. However, it was also pointed out that since ANS and bis-ANS have a strong affinity to the partially folded MG state, they can shift the equilibrium from favoring the native state (N) to favoring the MG state [89]. As a result, the apparent destabilization of the native state is observed, as was shown for the nucleotide-binding chaperonin DnaK [89]. On the other hand, binding of ADP or ATP to the native state of this protein resulted in a shift of the equilibrium from the MG toward the N state [89]. Furthermore, as early as 1995, Anthony L. Fink (1943–2008) cautioned that “It is important to note that the presence of ANS tends to increase the propensity of molten globules and compact denatured states to aggregate, and that aggregation increases the ANS fluorescence emission” [62]. Some other techniques like hydrogen-deuterium exchange, NMR, X-ray, isothermal titration calorimetry (ITC), differential scanning calorimetry (DSC), and computational methods have also been increasingly applied in later years [71]. In general, all the techniques/methods applicable to looking at protein structure and stability can give valuable information about partially folded intermediates like MGs [157]. For example, various fluorescence techniques, such as analysis of the intrinsic and extrinsic fluorescence (both steady-state and time-resolved), fluorescence anisotropy, Förster resonance energy transfer (FRET), dynamic and static fluorescence quenching, and proteolytic susceptibility are also used quite often [158]. In additional to classical examples of α-lactabumin, BCAB, and β-lactamase, both equilibrium and kinetic (transient) MGs have been described for a number of proteins and their mutants [159][160][161]. One interesting comparison is between the MGs formed by α-amylases from a thermophile and those formed from a mesophile [162]. This analysis revealed that the MG of the thermophile was more stable, which is not surprising. The polyols were less effective in refolding of the MG of the mesophilic enzyme [162]. Another interesting class of proteins are from halophiles. These generally require >0.5 M KCl to be functional. In several cases, these proteins just like those from thermophiles are fairly stable towards unfolding. The mechanism of halo-adaption was investigated by Gloss et al. [163] by looking at the kinetics of folding of urea denatured dihydrofolate reductases (DHFR) from E. coli and a halophile. In both cases, after a burst intermediate, formation of two intermediates was detected. The data was consistent with salt ions destabilizing the unfolded states in both cases. The authors concluded that halo-adaption involves affecting the solvent via a hydrophobic effect, the Hofmeister effect, preferential hydration, and crowding. This is in line with the X-ray crystallography and structural data that showed extensive solvation but little salt binding in the case of many halophilic proteins [163]. Yet another example of complexity in halo-adoption by halophile proteins is the role of protein hydration [164]. Given their higher surface charge density, it is widely believed that these are highly hydrated even in their native forms. This excessive hydration was expected to be responsible for the exceptional stability of corresponding proteins under saline conditions. The results obtained with an engineered protein with a high number of acidic residues on its surface suggested that not only was the surface hydration of a halophilic protein not much larger than that of a mesophilic counterpart, but even its hydration dynamics during unfolding was not very different [164]. Study of the proteasome from the extremely halophilic archaeon Haloarcula marismortui revealed that while other enzymes unfolded under sub-saline conditions, the proteasome was more resistant [165]. The biological significance of this is that it underlines how proteasome degrades the damaged proteins under sub-saline conditions as the stress situation for the organisms [165]. Uversky compared the stabilities of proteins from mesophiles with those from halophiles, thermophiles, and barophiles while advancing a hypothesis about the role of protein dielectricity in affecting the solvent properties in the context of protein-protein interactions [166]. The research mentions the earlier work with β-lactoglobulin, in which it was reported that the molten globule formation by the protein in alcohol-cosolvent mixtures was directly dependent on the decrease in the dielectric constant of the water as a result of mixing the simple alcohols [109]. Interestingly enough, in an independent observation, Gupta et al. around the same time observed that for a number of proteins, the enzyme stability in aqueous-organic cosolvent mixtures was dictated by the polarity index of the organic solvent [167]. Solvents with a polarity index of 5.8 and above were good cosolvents, which did not destabilize the protein even when up to 50% (v/v) is added to the aqueous buffer [167]. Both dielectric constants and polarity indexes are measures of solvent polarity. Another interesting observation has been reported about MG formed by chymotrypsinogen [168]. A single cysteine reacts with glutathione at a very rapid rate. Such hyperactive cysteine residues are also present in serum albumin, lysozyme, and ribonuclease [168]. However, cysteine present in two proteins of a thermophile (in which glutathione is absent) did not display this hyper-reactivity. The authors infer that this unusually high reactivity of cysteine residues is relevant to the oxidative refolding of proteins in the organisms, which have oxidized glutathione-reduced glutathione system [168].

5. Molten Globules and Intrinsic Disorder in Proteins

Coming back to the hypothesis on the potential role of protein dielectricity in affecting the solvent properties mentioned earlier [166], in the context of functional relevance of partially unfolded protein intermediates, it was proposed that a protein lowers the dielectric constant of the local medium around its interface with the aqueous solvent/water rich medium. This facilitates the behavior of proteins acting as “unfoldases”. Many proteins, in order to be functional, have to be unfolded (then referred to as conditionally disordered proteins) [169][170]. In many cases, this conditional unfolding is initiated by the interacting protein, which acts as an unfoldase by lowering the local dielectric around it; this leads to the binding between the two as a part of a biologically relevant process. Examples include unfolding of BCL-xL while interacting with the intrinsically disordered PUMA, which in turn folds upon binding as entropic compensation [166]. Unfoldases also include ATP-dependent proteases (such as in proteomes) and molecular chaperonins. Early examples in which this unfoldase behavior was observed were pore-forming domains of some toxins and carrier proteins of large nonpolar ligands. The aggregation including where it leads to amyloid formations (and is responsible for many diseases) may also be initiated by protein lowering the dielectric around it. Few other examples relevant to this are available [170][171][172][173]. Therefore, this hypothesis provides a common thread running through diverse phenomena [166]. Interestingly enough, later work has confirmed that functionally relevant unfolded structures of many bacterial toxins are molten globules [174][175][176]. One should keep in mind that intrinsic disorder in proteins represent a highly heterogeneous phenomenon, and functional IDPs can be disordered to different degrees. In fact, the existence of native (i.e., functional) coils, PMGs, and MGs was reported [54][177][178][179][180]. Furthermore, different parts of a protein molecule can be disordered to different degrees, and a functional protein can contain ordered, molten globular, pre-molten globular, and coil-like domains. What’s more, IDPs/IDPRs (and, as a matter of fact, any protein molecule in general) can be structurally represented as a spatio-temporal combinations of foldons (independent foldable units of a protein), inducible foldons (disordered regions that can fold at least in part due to the interaction with their binding partners), inducible morphing foldons (disordered regions that can fold differently at interaction with different binding partners), semi-foldons (always semi-folded regions), non-foldons (non-foldable protein regions), and unfoldons (regions that undergo an order-to-disorder transition to become functional) [92][93][181][182]. Another important note is that these functional disordered elements (i.e., foldons, inducible foldons, inducible morphing foldons, semi-foldons, and non-foldons) can structurally exist as coils, PMGs, or MGs. There is another pointer to the complexity of the process. Bychkova et al. have discussed the differences between an MG and an IDP [125]. In the latter, there is a greater disruption of local structure; H-D exchange is higher. However, researchers do not have any data regarding a comparison between the two different forms of MGs (WMG and DMG) and IDPs. There is also an interesting observation that MG-like IDPs can drive liquid-liquid phase separation (LLPS) that leads to the formation of protein condensates [183]. It is reported that in the case of the replication transcription of respiratory syncytial virus that take place within the “viral factories”, which are liquid-like structures within the cytosol of infected cells, the phosphoprotein tetramer (which is involved in the process and has a highly disordered N-terminal domain and a molten globular C-terminal domain) displays LLPS during a thermal transition, which is accompanied by the folding of the MG domain [183]. When the phosphoprotein is mixed with a nucleoprotein, which is also a part of the viral replication complex, again a phase separation is observed. Based on their observations, the authors of this study concluded that for LLPS to take place in vitro and in the cell, a weak, MG-like structure must be present, and such a structure defines physicochemical grounds for the LLPS behind the viral replication factory [183]. This is an interesting observation, as more often, proteins driving LLPS are expected to be either native coils (as shown for many IDPs [184][185][186][187][188][189][190]) or native PMGs (see, e.g., data for the AB region of human retinoid X receptor subtype γ (hRXRγ) [191]).

References

  1. Lesk, A. Introduction to Protein Science: Architecture, Function, and Genomics; Oxford University Press: Oxford, UK, 2010.
  2. Anfinsen, C.B. Principles that govern the folding of protein chains. Science 1973, 181, 223–230.
  3. Anfinsen, C.B.; Haber, E.; Sela, M.; White, F.H., Jr. The kinetics of formation of native ribonuclease during oxidation of the reduced polypeptide chain. Proc. Natl. Acad. Sci. USA 1961, 47, 1309–1314.
  4. White, F.H., Jr. Regeneration of native secondary and tertiary structures by air oxidation of reduced ribonuclease. J. Biol. Chem. 1961, 236, 1353–1360.
  5. Anfinsen, C.B.; Haber, E. Studies on the reduction and re-formation of protein disulfide bonds. J. Biol. Chem. 1961, 236, 1361–1363.
  6. Seckler, R.; Jaenicke, R. Protein folding and protein refolding. FASEB J. 1992, 6, 2545–2552.
  7. Tsytlonok, M.; Itzhaki, L.S. The how’s and why’s of protein folding intermediates. Arch. Biochem. Biophys. 2013, 531, 14–23.
  8. Privalov, P.L. Stability of proteins: Small globular proteins. Adv. Protein Chem. 1979, 33, 167–241.
  9. Tanford, C. Protein denaturation. Adv. Protein Chem. 1968, 23, 121–282.
  10. Gil’manshin, R.I.; Dolgikh, D.A.; Ptitsyn, O.B.; Finkel’shtein, A.V.; Shakhnovich, E.I. Protein globule without the unique three-dimensional structure: Experimental data for alpha-lactalbumins and general model. Biofizika 1982, 27, 1005–1016.
  11. Dolgikh, D.A.; Gilmanshin, R.I.; Brazhnikov, E.V.; Bychkova, V.E.; Semisotnov, G.V.; Venyaminov, S.; Ptitsyn, O.B. Alpha-Lactalbumin: Compact state with fluctuating tertiary structure? FEBS Lett. 1981, 136, 311–315.
  12. Ohgushi, M.; Wada, A. ‘Molten-globule state’: A compact form of globular proteins with mobile side-chains. FEBS Lett. 1983, 164, 21–24.
  13. Ptitsyn, O.B.; Dolgikh, D.A.; Gil’manshin, R.I.; Shakhnovich, E.I.; Finkel’shtein, A.V. Fluctuating state of the protein globule. Mol. Biol. 1983, 17, 569–576.
  14. Kuwajima, K. The molten globule state as a clue for understanding the folding and cooperativity of globular-protein structure. Proteins 1989, 6, 87–103.
  15. Ptitsyn, O.B. Structures of folding intermediates. Curr. Opin. Struct. Biol. 1995, 5, 74–78.
  16. Ptitsyn, O.B. Molten globule and protein folding. Adv. Protein Chem. 1995, 47, 83–229.
  17. Arai, M.; Kuwajima, K. Role of the molten globule state in protein folding. Adv. Protein Chem. 2000, 53, 209–282.
  18. Bychkova, V.E.; Semisotnov, G.V.; Balobanov, V.A.; Finkelstein, A.V. The Molten Globule Concept: 45 Years Later. Biochemistry 2018, 83, S33–S47.
  19. Baldwin, R.L.; Rose, G.D. Molten globules, entropy-driven conformational change and protein folding. Curr. Opin. Struct. Biol. 2013, 23, 4–10.
  20. Balobanov, V.A.; Katina, N.S.; Finkelstein, A.V.; Bychkova, V.E. Intermediate States of Apomyoglobin: Are They Parts of the Same Area of Conformations Diagram? Biochemistry 2017, 82, 625–631.
  21. Kharakoz, D.P.; Bychkova, V.E. Molten globule of human alpha-lactalbumin: Hydration, density, and compressibility of the interior. Biochemistry 1997, 36, 1882–1890.
  22. Georlette, D.; Blaise, V.; Bouillenne, F.; Damien, B.; Thorbjarnardottir, S.H.; Depiereux, E.; Gerday, C.; Uversky, V.N.; Feller, G. Adenylation-dependent conformation and unfolding pathways of the NAD+-dependent DNA ligase from the thermophile Thermus scotoductus. Biophys. J. 2004, 86, 1089–1104.
  23. Georlette, D.; Blaise, V.; Dohmen, C.; Bouillenne, F.; Damien, B.; Depiereux, E.; Gerday, C.; Uversky, V.N.; Feller, G. Cofactor binding modulates the conformational stabilities and unfolding patterns of NAD(+)-dependent DNA ligases from Escherichia coli and Thermus scotoductus. J. Biol. Chem. 2003, 278, 49945–49953.
  24. Uversky, V.N.; Ptitsyn, O.B. Further evidence on the equilibrium “pre-molten globule state”: Four-state guanidinium chloride-induced unfolding of carbonic anhydrase B at low temperature. J. Mol. Biol. 1996, 255, 215–228.
  25. Ptitsyn, O.B.; Bychkova, V.E.; Uversky, V.N. Kinetic and equilibrium folding intermediates. Philos. Trans. R. Soc. Lond. B Biol. Sci. 1995, 348, 35–41.
  26. Uversky, V.N.; Ptitsyn, O.B. “Partly folded” state, a new equilibrium state of protein molecules: Four-state guanidinium chloride-induced unfolding of beta-lactamase at low temperature. Biochemistry 1994, 33, 2782–2791.
  27. Ptitsyn, O.B. Stages in the mechanism of self-organization of protein molecules. Dokl. Akad. Nauk. SSSR 1973, 210, 1213–1215.
  28. Sarkar, S.S.; Udgaonkar, J.B.; Krishnamoorthy, G. Unfolding of a small protein proceeds via dry and wet globules and a solvated transition state. Biophys. J. 2013, 105, 2392–2402.
  29. Acharya, N.; Jha, S.K. Dry Molten Globule-Like Intermediates in Protein Folding, Function, and Disease. J. Phys. Chem. B 2022, 126, 8614–8622.
  30. Acharya, N.; Mishra, P.; Jha, S.K. A dry molten globule-like intermediate during the base-induced unfolding of a multidomain protein. Phys. Chem. Chem. Phys. 2017, 19, 30207–30216.
  31. de Oliveira, G.A.P.; Silva, J.L. The push-and-pull hypothesis in protein unfolding, misfolding and aggregation. Biophys. Chem. 2017, 231, 20–26.
  32. Neumaier, S.; Kiefhaber, T. Redefining the dry molten globule state of proteins. J. Mol. Biol. 2014, 426, 2520–2528.
  33. Jha, S.K.; Marqusee, S. Kinetic evidence for a two-stage mechanism of protein denaturation by guanidinium chloride. Proc. Natl. Acad. Sci. USA 2014, 111, 4856–4861.
  34. Baldwin, R.L.; Frieden, C.; Rose, G.D. Dry molten globule intermediates and the mechanism of protein unfolding. Proteins 2010, 78, 2725–2737.
  35. Reiner, A.; Henklein, P.; Kiefhaber, T. An unlocking/relocking barrier in conformational fluctuations of villin headpiece subdomain. Proc. Natl. Acad. Sci. USA 2010, 107, 4955–4960.
  36. Jha, S.K.; Udgaonkar, J.B. Direct evidence for a dry molten globule intermediate during the unfolding of a small protein. Proc. Natl. Acad. Sci. USA 2009, 106, 12289–12294.
  37. Rami, B.R.; Udgaonkar, J.B. Mechanism of formation of a productive molten globule form of barstar. Biochemistry 2002, 41, 1710–1716.
  38. Kiefhaber, T.; Labhardt, A.M.; Baldwin, R.L. Direct NMR evidence for an intermediate preceding the rate-limiting step in the unfolding of ribonuclease A. Nature 1995, 375, 513–515.
  39. Finkelstein, A.V.; Shakhnovich, E.I. Theory of cooperative transitions in protein molecules. II. Phase diagram for a protein molecule in solution. Biopolymers 1989, 28, 1681–1694.
  40. Edsall, J.T. Hsien Wu and the First Theory of Protein Denaturation (1931). Adv. Protein Chem. 1995, 46, 1–5.
  41. Wu, H. Studies of Denaturation of Proteins XIII. A Theory of Denaturation. Chin. J. Physiol. 1931, 5, 321–344.
  42. Anson, M.L.; Mirsky, A.E. On Some General Properties of Proteins. J. Gen. Physiol. 1925, 9, 169–179.
  43. Mirsky, A.E.; Pauling, L. On the Structure of Native, Denatured, and Coagulated Proteins. Proc. Natl. Acad. Sci. USA 1936, 22, 439–447.
  44. Neurath, H.; Greenstein, J.P.; Putnam, F.W.; Erickson, J.A. The chemistry of protein denaturation. Chem. Rev. 1944, 34, 157–265.
  45. Greenstein, J.P. Sulfhydryl groups in proteins: I. Egg albumin in solutions of urea, guanidine, and their derivatives. J. Biol. Chem. 1938, 125, 501–513.
  46. Greenstein, J.P. Sulfhydryl groups in proteins: II. Edestin, Excelsin, and Globin in solutions of guanidine hydrochloride, urea, and their derivatives. J. Biol. Chem. 1939, 128, 233–240.
  47. Greenstein, J.P. Sulfhydryl groups in proteins: III. The effect on egg albumin of various salts of guanidine. J. Biol. Chem. 1939, 130, 519–526.
  48. Neurath, H.; Cooper, G.R.; Erickson, J.O. The denaturation of proteins and its apparent reversal: I. horse serum albumin. J. Biol. Chem. 1942, 142, 249–263.
  49. Neurath, H.; Cooper, G.R.; Erickson, J.O. The denaturation of proteins and its apparent reversal: II. horse serum pseudoglobulin. J. Biol. Chem. 1942, 142, 265–276.
  50. Aune, K.C.; Salahuddin, A.; Zarlengo, M.H.; Tanford, C. Evidence for residual structure in acid- and heat-denatured proteins. J. Biol. Chem. 1967, 242, 4486–4489.
  51. Wong, K.P.; Tanford, C. Denaturation of bovine carbonic anhydrase B by guanidine hydrochloride. A process involving separable sequential conformational transitions. J. Biol. Chem. 1973, 248, 8518–8523.
  52. Wong, K.P.; Hamlin, L.M. Acid denaturation of bovine carbonic anhydrase B. Biochemistry 1974, 13, 2678–2683.
  53. Dobson, C.M. Protein folding. Solid evidence for molten globules. Curr. Biol. 1994, 4, 636–640.
  54. Uversky, V.N. Protein folding revisited. A polypeptide chain at the folding-misfolding-nonfolding cross-roads: Which way to go? Cell. Mol. Life Sci. 2003, 60, 1852–1871.
  55. Tcherkasskaya, O.; Uversky, V.N. Polymeric aspects of protein folding: A brief overview. Protein Pept. Lett. 2003, 10, 239–245.
  56. Uverskii, V.N. How many molten globules states exist? Biofizika 1998, 43, 416–421.
  57. Ikeguchi, M.; Fujino, M.; Kato, M.; Kuwajima, K.; Sugai, S. Transition state in the folding of alpha-lactalbumin probed by the 6–120 disulfide bond. Protein Sci. 1998, 7, 1564–1574.
  58. Ptitsyn, O. How molten is the molten globule? Nat. Struct. Biol. 1996, 3, 488–490.
  59. Fink, A.L. Compact intermediates states in protein folding. Subcell. Biochem. 1995, 24, 27–53.
  60. Ewbank, J.J.; Creighton, T.E.; Hayer-Hartl, M.K.; Ulrich Hartl, F. What is the molten globule? Nat. Struct. Biol. 1995, 2, 10–11.
  61. Fink, A.L. Compact intermediate states in protein folding. Annu. Rev. Biophys. Biomol. Struct. 1995, 24, 495–522.
  62. Fink, A.L. Molten globules. Methods Mol. Biol. 1995, 40, 343–360.
  63. Vassilenko, K.S.; Uversky, V.N. Native-like secondary structure of molten globules. Biochim. Biophys. Acta 2002, 1594, 168–177.
  64. Redfield, C. Using nuclear magnetic resonance spectroscopy to study molten globule states of proteins. Methods 2004, 34, 121–132.
  65. Krishna, M.M.; Hoang, L.; Lin, Y.; Englander, S.W. Hydrogen exchange methods to study protein folding. Methods 2004, 34, 51–64.
  66. Bracken, C. NMR spin relaxation methods for characterization of disorder and folding in proteins. J. Mol. Graph. Model. 2001, 19, 3–12.
  67. Bose, H.S.; Whittal, R.M.; Baldwin, M.A.; Miller, W.L. The active form of the steroidogenic acute regulatory protein, StAR, appears to be a molten globule. Proc. Natl. Acad. Sci. USA 1999, 96, 7250–7255.
  68. Eliezer, D.; Yao, J.; Dyson, H.J.; Wright, P.E. Structural and dynamic characterization of partially folded states of apomyoglobin and implications for protein folding. Nat. Struct. Biol. 1998, 5, 148–155.
  69. Wu, L.C.; Laub, P.B.; Elove, G.A.; Carey, J.; Roder, H. A noncovalent peptide complex as a model for an early folding intermediate of cytochrome c. Biochemistry 1993, 32, 10271–10276.
  70. Chyan, C.L.; Wormald, C.; Dobson, C.M.; Evans, P.A.; Baum, J. Structure and stability of the molten globule state of guinea-pig alpha-lactalbumin: A hydrogen exchange study. Biochemistry 1993, 32, 5681–5691.
  71. Jeng, M.F.; Englander, S.W.; Elove, G.A.; Wand, A.J.; Roder, H. Structural description of acid-denatured cytochrome c by hydrogen exchange and 2D NMR. Biochemistry 1990, 29, 10433–10437.
  72. Bushnell, G.W.; Louie, G.V.; Brayer, G.D. High-resolution three-dimensional structure of horse heart cytochrome c. J. Mol. Biol. 1990, 214, 585–595.
  73. Baum, J.; Dobson, C.M.; Evans, P.A.; Hanley, C. Characterization of a partly folded protein by NMR methods: Studies on the molten globule state of guinea pig alpha-lactalbumin. Biochemistry 1989, 28, 7–13.
  74. Hsu, D.J.; Leshchev, D.; Kosheleva, I.; Kohlstedt, K.L.; Chen, L.X. Unfolding bovine alpha-lactalbumin with T-jump: Characterizing disordered intermediates via time-resolved x-ray solution scattering and molecular dynamics simulations. J. Chem. Phys. 2021, 154, 105101.
  75. Uversky, V.N.; Karnoup, A.S.; Segel, D.J.; Seshadri, S.; Doniach, S.; Fink, A.L. Anion-induced folding of Staphylococcal nuclease: Characterization of multiple equilibrium partially folded intermediates. J. Mol. Biol. 1998, 278, 879–894.
  76. Semisotnov, G.V.; Kihara, H.; Kotova, N.V.; Kimura, K.; Amemiya, Y.; Wakabayashi, K.; Serdyuk, I.N.; Timchenko, A.A.; Chiba, K.; Nikaido, K.; et al. Protein globularization during folding. A study by synchrotron small-angle X-ray scattering. J. Mol. Biol. 1996, 262, 559–574.
  77. Eliezer, D.; Chiba, K.; Tsuruta, H.; Doniach, S.; Hodgson, K.O.; Kihara, H. Evidence of an associative intermediate on the myoglobin refolding pathway. Biophys. J. 1993, 65, 912–917.
  78. Kataoka, M.; Hagihara, Y.; Mihara, K.; Goto, Y. Molten globule of cytochrome c studied by small angle X-ray scattering. J. Mol. Biol. 1993, 229, 591–596.
  79. Kataoka, M.; Kuwajima, K.; Tokunaga, F.; Goto, Y. Structural characterization of the molten globule of alpha-lactalbumin by solution X-ray scattering. Protein Sci. 1997, 6, 422–430.
  80. Uversky, V.N. Use of fast protein size-exclusion liquid chromatography to study the unfolding of proteins which denature through the molten globule. Biochemistry 1993, 32, 13288–13298.
  81. Uversky, V.N. Molten globular enzymes. In Structure and Intrinsic Disorder in Enzymology; Elsevier: Amsterdam, The Netherlands, 2023; pp. 303–325.
  82. Fontana, A.; de Laureto, P.P.; Spolaore, B.; Frare, E.; Picotti, P.; Zambonin, M. Probing protein structure by limited proteolysis. Acta Biochim. Pol. 2004, 51, 299–321.
  83. Fontana, A.; Polverino de Laureto, P.; De Filippis, V.; Scaramella, E.; Zambonin, M. Probing the partly folded states of proteins by limited proteolysis. Fold. Des. 1997, 2, R17–R26.
  84. Merrill, A.R.; Cohen, F.S.; Cramer, W.A. On the nature of the structural change of the colicin E1 channel peptide necessary for its translocation-competent state. Biochemistry 1990, 29, 5829–5836.
  85. Polverino de Laureto, P.; De Filippis, V.; Di Bello, M.; Zambonin, M.; Fontana, A. Probing the molten globule state of alpha-lactalbumin by limited proteolysis. Biochemistry 1995, 34, 12596–12604.
  86. Polverino de Laureto, P.; Frare, E.; Gottardo, R.; Fontana, A. Molten globule of bovine alpha-lactalbumin at neutral pH induced by heat, trifluoroethanol, and oleic acid: A comparative analysis by circular dichroism spectroscopy and limited proteolysis. Proteins 2002, 49, 385–397.
  87. Polverino de Laureto, P.; Frare, E.; Gottardo, R.; Van Dael, H.; Fontana, A. Partly folded states of members of the lysozyme/lactalbumin superfamily: A comparative study by circular dichroism spectroscopy and limited proteolysis. Protein Sci. 2002, 11, 2932–2946.
  88. Uversky, V.N.; Winter, S.; Lober, G. Use of fluorescence decay times of 8-ANS-protein complexes to study the conformational transitions in proteins which unfold through the molten globule state. Biophys. Chem. 1996, 60, 79–88.
  89. Shi, L.; Palleros, D.R.; Fink, A.L. Protein conformational changes induced by 1,1’-bis(4-anilino-5-naphthalenesulfonic acid): Preferential binding to the molten globule of DnaK. Biochemistry 1994, 33, 7536–7546.
  90. Semisotnov, G.V.; Rodionova, N.A.; Razgulyaev, O.I.; Uversky, V.N.; Gripas, A.F.; Gilmanshin, R.I. Study of the “molten globule” intermediate state in protein folding by a hydrophobic fluorescent probe. Biopolymers 1991, 31, 119–128.
  91. Regan, L. Molten globules move into action. Proc. Natl. Acad. Sci. USA 2003, 100, 3553–3554.
  92. Uversky, V.N. Protein intrinsic disorder and structure-function continuum. Prog. Mol. Biol. Transl. Sci. 2019, 166, 1–17.
  93. Uversky, V.N. Dancing Protein Clouds: The Strange Biology and Chaotic Physics of Intrinsically Disordered Proteins. J. Biol. Chem. 2016, 291, 6681–6688.
  94. Uversky, V.N. Unusual biophysics of intrinsically disordered proteins. Biochim. Biophys. Acta 2013, 1834, 932–951.
  95. Uversky, V.N.; Semisotnov, G.V.; Pain, R.H.; Ptitsyn, O.B. ‘All-or-none’ mechanism of the molten globule unfolding. FEBS Lett. 1992, 314, 89–92.
  96. Uversky, V.N.; Ptitsyn, O.B. All-or-none solvent-induced transitions between native, molten globule and unfolded states in globular proteins. Fold. Des. 1996, 1, 117–122.
  97. Ptitsyn, O.B.; Uversky, V.N. The molten globule is a third thermodynamical state of protein molecules. FEBS Lett. 1994, 341, 15–18.
  98. Pande, V.S.; Rokhsar, D.S. Is the molten globule a third phase of proteins? Proc. Natl. Acad. Sci. USA 1998, 95, 1490–1494.
  99. Bychkova, V.E.; Pain, R.H.; Ptitsyn, O.B. The ‘molten globule’ state is involved in the translocation of proteins across membranes? FEBS Lett. 1988, 238, 231–234.
  100. Song, M.; Shao, H.; Mujeeb, A.; James, T.L.; Miller, W.L. Molten-globule structure and membrane binding of the N-terminal protease-resistant domain (63-193) of the steroidogenic acute regulatory protein (StAR). Biochem. J. 2001, 356, 151–158.
  101. Bose, H.S.; Baldwin, M.A.; Miller, W.L. Evidence that StAR and MLN64 act on the outer mitochondrial membrane as molten globules. Endocr. Res. 2000, 26, 629–637.
  102. Ren, J.; Kachel, K.; Kim, H.; Malenbaum, S.E.; Collier, R.J.; London, E. Interaction of diphtheria toxin T domain with molten globule-like proteins and its implications for translocation. Science 1999, 284, 955–957.
  103. van der Goot, F.G.; Lakey, J.H.; Pattus, F. The molten globule intermediate for protein insertion or translocation through membranes. Trends Cell. Biol. 1992, 2, 343–348.
  104. van der Goot, F.G.; Gonzalez-Manas, J.M.; Lakey, J.H.; Pattus, F. A ‘molten-globule’ membrane-insertion intermediate of the pore-forming domain of colicin A. Nature 1991, 354, 408–410.
  105. Banuelos, S.; Muga, A. Binding of molten globule-like conformations to lipid bilayers. Structure of native and partially folded alpha-lactalbumin bound to model membranes. J. Biol. Chem. 1995, 270, 29910–29915.
  106. Watts, A. Biophysics of the membrane interface. Biochem. Soc. Trans. 1995, 23, 959–965.
  107. Bychkova, V.E.; Basova, L.V.; Balobanov, V.A. How membrane surface affects protein structure. Biochemistry 2014, 79, 1483–1514.
  108. Narizhneva, N.V.; Uversky, V.N. Decrease of dielectric constant transforms the protein molecule into the molten globule state. Biochemistry 1998, 63, 448–455.
  109. Uversky, V.N.; Narizhneva, N.V.; Kirschstein, S.O.; Winter, S.; Lober, G. Conformational transitions provoked by organic solvents in beta-lactoglobulin: Can a molten globule like intermediate be induced by the decrease in dielectric constant? Fold. Des. 1997, 2, 163–172.
  110. Bychkova, V.E.; Dujsekina, A.E.; Klenin, S.I.; Tiktopulo, E.I.; Uversky, V.N.; Ptitsyn, O.B. Molten globule-like state of cytochrome c under conditions simulating those near the membrane surface. Biochemistry 1996, 35, 6058–6063.
  111. Bychkova, V.E.; Dujsekina, A.E.; Fantuzzi, A.; Ptitsyn, O.B.; Rossi, G.L. Release of retinol and denaturation of its plasma carrier, retinol-binding protein. Fold. Des. 1998, 3, 285–291.
  112. Uversky, V.N.; Narizhneva, N.V. Effect of natural ligands on the structural properties and conformational stability of proteins. Biochemistry 1998, 63, 420–433.
  113. Uversky, V.N.; Narizhneva, N.V.; Ivanova, T.V.; Tomashevski, A.Y. Rigidity of human alpha-fetoprotein tertiary structure is under ligand control. Biochemistry 1997, 36, 13638–13645.
  114. Uversky, V.N.; Narizhneva, N.V.; Ivanova, T.V.; Kirkitadze, M.D.; Tomashevski, A. Ligand-free form of human alpha-fetoprotein: Evidence for the molten globule state. FEBS Lett. 1997, 410, 280–284.
  115. Prajapati, R.S.; Indu, S.; Varadarajan, R. Identification and thermodynamic characterization of molten globule states of periplasmic binding proteins. Biochemistry 2007, 46, 10339–10352.
  116. Rajaraman, K.; Raman, B.; Rao, C.M. Molten-globule state of carbonic anhydrase binds to the chaperone-like alpha-crystallin. J. Biol. Chem. 1996, 271, 27595–27600.
  117. Martin, J.; Langer, T.; Boteva, R.; Schramel, A.; Horwich, A.L.; Hartl, F.U. Chaperonin-mediated protein folding at the surface of groEL through a ‘molten globule’-like intermediate. Nature 1991, 352, 36–42.
  118. Li, S.; Bai, J.H.; Park, Y.D.; Zhou, H.M. Aggregation of creatine kinase during refolding and chaperonin-mediated folding of creatine kinase. Int. J. Biochem. Cell Biol. 2001, 33, 279–286.
  119. Hayer-Hartl, M.K.; Ewbank, J.J.; Creighton, T.E.; Hartl, F.U. Conformational specificity of the chaperonin GroEL for the compact folding intermediates of alpha-lactalbumin. EMBO J. 1994, 13, 3192–3202.
  120. Braig, K.; Simon, M.; Furuya, F.; Hainfeld, J.F.; Horwich, A.L. A polypeptide bound by the chaperonin groEL is localized within a central cavity. Proc. Natl. Acad. Sci. USA 1993, 90, 3978–3982.
  121. Melki, R.; Cowan, N.J. Facilitated folding of actins and tubulins occurs via a nucleotide-dependent interaction between cytoplasmic chaperonin and distinctive folding intermediates. Mol. Cell Biol. 1994, 14, 2895–2904.
  122. Zhou, B.; Tian, K.; Jing, G. An in vitro peptide folding model suggests the presence of the molten globule state during nascent peptide folding. Protein Eng. 2000, 13, 35–39.
  123. Bychkova, V.E.; Ptitsyn, O.B. The molten globule in vitro and in vivo. Chemtracts Biochem. Molec. Biol. 1993, 4, 133–163.
  124. Bychkova, V.E.; Ptitsyn, O.B. Folding intermediates are involved in genetic diseases? FEBS Lett. 1995, 359, 6–8.
  125. Bychkova, V.E.; Dolgikh, D.A.; Balobanov, V.A.; Finkelstein, A.V. The molten globule state of a globular protein in a cell is more or less frequent case rather than an exception. Molecules 2022, 27, 4361.
  126. Griko, Y.V.; Freire, E.; Privalov, P.L. Energetics of the alpha-lactalbumin states: A calorimetric and statistical thermodynamic study. Biochemistry 1994, 33, 1889–1899.
  127. Goto, Y.; Fink, A.L. Acid-induced folding of heme proteins. Methods Enzym. 1994, 232, 3–15.
  128. Kuroda, Y.; Kidokoro, S.; Wada, A. Thermodynamic characterization of cytochrome c at low pH. Observation of the molten globule state and of the cold denaturation process. J. Mol. Biol. 1992, 223, 1139–1153.
  129. Hughson, F.M.; Wright, P.E.; Baldwin, R.L. Structural characterization of a partly folded apomyoglobin intermediate. Science 1990, 249, 1544–1548.
  130. Goto, Y.; Takahashi, N.; Fink, A.L. Mechanism of acid-induced folding of proteins. Biochemistry 1990, 29, 3480–3488.
  131. Goto, Y.; Calciano, L.J.; Fink, A.L. Acid-induced folding of proteins. Proc. Natl. Acad. Sci. USA 1990, 87, 573–577.
  132. Maheshwari, D.; Yadav, R.; Rastogi, R.; Jain, A.; Tripathi, S.; Mukhopadhyay, A.; Arora, A. Structural and Biophysical Characterization of Rab5a from Leishmania Donovani. Biophys. J. 2018, 115, 1217–1230.
  133. Wahiduzzaman; Dar, M.A.; Haque, M.A.; Idrees, D.; Hassan, M.I.; Islam, A.; Ahmad, F. Characterization of folding intermediates during urea-induced denaturation of human carbonic anhydrase II. Int. J. Biol. Macromol. 2017, 95, 881–887.
  134. Khan, P.; Prakash, A.; Haque, M.A.; Islam, A.; Hassan, M.I.; Ahmad, F. Structural basis of urea-induced unfolding: Unraveling the folding pathway of hemochromatosis factor E. Int. J. Biol. Macromol. 2016, 91, 1051–1061.
  135. Ghosh, G.; Mandal, D.K. Differing structural characteristics of molten globule intermediate of peanut lectin in urea and guanidine-HCl. Int. J. Biol. Macromol. 2012, 51, 188–195.
  136. Mandal, A.K.; Samaddar, S.; Banerjee, R.; Lahiri, S.; Bhattacharyya, A.; Roy, S. Glutamate counteracts the denaturing effect of urea through its effect on the denatured state. J. Biol. Chem. 2003, 278, 36077–36084.
  137. Garcia, P.; Serrano, L.; Rico, M.; Bruix, M. An NMR view of the folding process of a CheY mutant at the residue level. Structure 2002, 10, 1173–1185.
  138. Cymes, G.D.; Grosman, C.; Delfino, J.M.; Wolfenstein-Todel, C. Detection and characterization of an ovine placental lactogen stable intermediate in the urea-induced unfolding process. Protein Sci. 1996, 5, 2074–2079.
  139. Das, B.K.; Bhattacharyya, T.; Roy, S. Characterization of a urea induced molten globule intermediate state of glutaminyl-tRNA synthetase from Escherichia coli. Biochemistry 1995, 34, 5242–5247.
  140. Rodionova, N.A.; Semisotnov, G.V.; Kutyshenko, V.P.; Uverskii, V.N.; Bolotina, I.A. Staged equilibrium of carbonic anhydrase unfolding in strong denaturants. Mol. Biol. 1989, 23, 683–692.
  141. Kuznetsova, I.M.; Stepanenko, O.V.; Turoverov, K.K.; Zhu, L.; Zhou, J.M.; Fink, A.L.; Uversky, V.N. Unraveling multistate unfolding of rabbit muscle creatine kinase. Biochim. Biophys. Acta 2002, 1596, 138–155.
  142. Zerovnik, E.; Jerala, R.; Kroon-Zitko, L.; Turk, V.; Pain, R.H. Denaturation of stefin B by GuHCl, pH and heat; evidence for molten globule intermediates. Biol. Chem. Hoppe Seyler 1992, 373, 453–458.
  143. Powell, L.M.; Pain, R.H. Effects of glycosylation on the folding and stability of human, recombinant and cleaved alpha 1-antitrypsin. J. Mol. Biol. 1992, 224, 241–252.
  144. Christensen, H.; Pain, R.H. Molten globule intermediates and protein folding. Eur. Biophys. J. 1991, 19, 221–229.
  145. Ptitsyn, O.B.; Pain, R.H.; Semisotnov, G.V.; Zerovnik, E.; Razgulyaev, O.I. Evidence for a molten globule state as a general intermediate in protein folding. FEBS Lett. 1990, 262, 20–24.
  146. Magsumov, T.; Ziying, L.; Sedov, I. Comparative study of the protein denaturing ability of different organic cosolvents. Int. J. Biol. Macromol. 2020, 160, 880–888.
  147. Prasanna Kumari, N.K.; Jagannadham, M.V. Deciphering the molecular structure of cryptolepain in organic solvents. Biochimie 2012, 94, 310–317.
  148. Santucci, R.; Polizio, F.; Desideri, A. Formation of a molten-globule-like state of cytochrome c induced by high concentrations of glycerol. Biochimie 1999, 81, 745–751.
  149. Wicar, S.; Mulkerrin, M.G.; Bathory, G.; Khundkar, L.H.; Karger, B.L. Conformational changes in the reversed phase liquid chromatography of recombinant human growth hormone as a function of organic solvent: The molten globule state. Anal. Chem. 1994, 66, 3908–3915.
  150. Jayaraman, G.; Kumar, T.K.; Arunkumar, A.I.; Yu, C. 2,2,2-Trifluoroethanol induces helical conformation in an all beta-sheet protein. Biochem. Biophys. Res. Commun. 1996, 222, 33–37.
  151. Dubey, V.K.; Shah, A.; Jagannadham, M.V.; Kayastha, A.M. Effect of organic solvents on the molten globule state of procerain: Beta-sheet to alpha-helix switchover in presence of trifluoroethanol. Protein Pept. Lett. 2006, 13, 545–547.
  152. Hirota-Nakaoka, N.; Goto, Y. Alcohol-induced denaturation of beta-lactoglobulin: A close correlation to the alcohol-induced alpha-helix formation of melittin. Bioorg. Med. Chem. 1999, 7, 67–73.
  153. Kuwata, K.; Hoshino, M.; Era, S.; Batt, C.A.; Goto, Y. alpha-->beta transition of beta-lactoglobulin as evidenced by heteronuclear NMR. J. Mol. Biol. 1998, 283, 731–739.
  154. Shiraki, K.; Nishikawa, K.; Goto, Y. Trifluoroethanol-induced stabilization of the alpha-helical structure of beta-lactoglobulin: Implication for non-hierarchical protein folding. J. Mol. Biol. 1995, 245, 180–194.
  155. Hirota, N.; Mizuno, K.; Goto, Y. Cooperative alpha-helix formation of beta-lactoglobulin and melittin induced by hexafluoroisopropanol. Protein Sci. 1997, 6, 416–421.
  156. Konuma, T.; Sakurai, K.; Yagi, M.; Goto, Y.; Fujisawa, T.; Takahashi, S. Highly Collapsed Conformation of the Initial Folding Intermediates of beta-Lactoglobulin with Non-Native alpha-Helix. J. Mol. Biol. 2015, 427, 3158–3165.
  157. Uversky, V.N. A multiparametric approach to studies of self-organization of globular proteins. Biochemistry 1999, 64, 250–266.
  158. Jacobs, M.D.; Fox, R.O. Staphylococcal nuclease folding intermediate characterized by hydrogen exchange and NMR spectroscopy. Proc. Natl. Acad. Sci. USA 1994, 91, 449–453.
  159. Alam Khan, M.K.; Das, U.; Rahaman, M.H.; Hassan, M.I.; Srinivasan, A.; Singh, T.P.; Ahmad, F. A single mutation induces molten globule formation and a drastic destabilization of wild-type cytochrome c at pH 6.0. J. Biol. Inorg. Chem. 2009, 14, 751–760.
  160. Jennings, P.A.; Wright, P.E. Formation of a molten globule intermediate early in the kinetic folding pathway of apomyoglobin. Science 1993, 262, 892–896.
  161. Radford, S.E.; Dobson, C.M.; Evans, P.A. The folding of hen lysozyme involves partially structured intermediates and multiple pathways. Nature 1992, 358, 302–307.
  162. Shokri, M.M.; Khajeh, K.; Alikhajeh, J.; Asoodeh, A.; Ranjbar, B.; Hosseinkhani, S.; Sadeghi, M. Comparison of the molten globule states of thermophilic and mesophilic alpha-amylases. Biophys. Chem. 2006, 122, 58–65.
  163. Gloss, L.M.; Topping, T.B.; Binder, A.K.; Lohman, J.R. Kinetic folding of Haloferax volcanii and Escherichia coli dihydrofolate reductases: Haloadaptation by unfolded state destabilization at high ionic strength. J. Mol. Biol. 2008, 376, 1451–1462.
  164. Qvist, J.; Ortega, G.; Tadeo, X.; Millet, O.; Halle, B. Hydration dynamics of a halophilic protein in folded and unfolded states. J. Phys. Chem. B 2012, 116, 3436–3444.
  165. Franzetti, B.; Schoehn, G.; Garcia, D.; Ruigrok, R.W.; Zaccai, G. Characterization of the proteasome from the extremely halophilic archaeon Haloarcula marismortui. Archaea 2002, 1, 53–61.
  166. Uversky, V.N. Hypothesis: The unfolding power of protein dielectricity. Intrinsically Disord. Proteins 2013, 1, e25725.
  167. Gupta, M.N.; Batra, R.; Tyagi, R.; Sharma, A. Polarity index: The guiding solvent parameter for enzyme stability in aqueous-organic cosolvent mixtures. Biotechnol. Prog. 1997, 13, 284–288.
  168. Bocedi, A.; Gambardella, G.; Cattani, G.; Bartolucci, S.; Limauro, D.; Pedone, E.; Iavarone, F.; Castagnola, M.; Ricci, G. Ultra-rapid glutathionylation of chymotrypsinogen in its molten globule-like conformation: A comparison to archaeal proteins. Sci. Rep. 2020, 10, 8943.
  169. Jakob, U.; Kriwacki, R.; Uversky, V.N. Conditionally and transiently disordered proteins: Awakening cryptic disorder to regulate protein function. Chem. Rev. 2014, 114, 6779–6805.
  170. Bardwell, J.C.; Jakob, U. Conditional disorder in chaperone action. Trends Biochem. Sci. 2012, 37, 517–525.
  171. Follis, A.V.; Chipuk, J.E.; Fisher, J.C.; Yun, M.K.; Grace, C.R.; Nourse, A.; Baran, K.; Ou, L.; Min, L.; White, S.W.; et al. PUMA binding induces partial unfolding within BCL-xL to disrupt p53 binding and promote apoptosis. Nat. Chem. Biol. 2013, 9, 163–168.
  172. Uversky, V.N.; Oldfield, C.J.; Dunker, A.K. Showing your ID: Intrinsic disorder as an ID for recognition, regulation and cell signaling. J. Mol. Recognit. 2005, 18, 343–384.
  173. Prakash, S.; Matouschek, A. Protein unfolding in the cell. Trends Biochem. Sci. 2004, 29, 593–600.
  174. Kukreja, R.; Singh, B. Biologically active novel conformational state of botulinum, the most poisonous poison. J. Biol. Chem. 2005, 280, 39346–39352.
  175. Cai, S.; Singh, B.R. Role of the disulfide cleavage induced molten globule state of type a botulinum neurotoxin in its endopeptidase activity. Biochemistry 2001, 40, 15327–15333.
  176. Kumar, R.; Chang, T.-W.; Singh, B.R. Evolutionary Traits of Toxins. In Biological Toxins and Bioterrorism. Toxinology; Gopalakrishnakone, P., Balali-Mood, M., Llewellyn, L., Singh, B.R., Eds.; Springer: Dordrecht, The Netherlands, 2015; pp. 527–557.
  177. Uversky, V.N. Natively unfolded proteins: A point where biology waits for physics. Protein Sci. 2002, 11, 739–756.
  178. Uversky, V.N. What does it mean to be natively unfolded? Eur. J. Biochem. 2002, 269, 2–12.
  179. Dunker, A.K.; Obradovic, Z. The protein trinity--linking function and disorder. Nat. Biotechnol. 2001, 19, 805–806.
  180. Dunker, A.K.; Lawson, J.D.; Brown, C.J.; Williams, R.M.; Romero, P.; Oh, J.S.; Oldfield, C.J.; Campen, A.M.; Ratliff, C.M.; Hipps, K.W.; et al. Intrinsically disordered protein. J. Mol. Graph. Model. 2001, 19, 26–59.
  181. Uversky, V.N. Recent Developments in the Field of Intrinsically Disordered Proteins: Intrinsic Disorder-Based Emergence in Cellular Biology in Light of the Physiological and Pathological Liquid-Liquid Phase Transitions. Annu. Rev. Biophys. 2021, 50, 135–156.
  182. Kulkarni, P.; Bhattacharya, S.; Achuthan, S.; Behal, A.; Jolly, M.K.; Kotnala, S.; Mohanty, A.; Rangarajan, G.; Salgia, R.; Uversky, V. Intrinsically Disordered Proteins: Critical Components of the Wetware. Chem. Rev. 2022, 122, 6614–6633.
  183. Salgueiro, M.; Camporeale, G.; Conci, J.; Sousa, B.; Visentin, A.; Corbat, A.; Grecco, H.; de Oliveira, G.A.; de Prat-Gay, G. Molten Globule Driven Liquid-Liquid Phase Separation at the Center of Viral Factory Assembly. Biophys. J. 2020, 118, 215a.
  184. Ambadipudi, S.; Biernat, J.; Riedel, D.; Mandelkow, E.; Zweckstetter, M. Liquid-liquid phase separation of the microtubule-binding repeats of the Alzheimer-related protein Tau. Nat. Commun. 2017, 8, 275.
  185. Darling, A.L.; Breydo, L.; Rivas, E.G.; Gebru, N.T.; Zheng, D.; Baker, J.D.; Blair, L.J.; Dickey, C.A.; Koren, J., 3rd; Uversky, V.N. Repeated repeat problems: Combinatorial effect of C9orf72-derived dipeptide repeat proteins. Int. J. Biol. Macromol. 2019, 127, 136–145.
  186. Ford, L.K.; Fioriti, L. Coiled-Coil Motifs of RNA-Binding Proteins: Dynamicity in RNA Regulation. Front. Cell Dev. Biol. 2020, 8, 607947.
  187. Franklin, J.M.; Guan, K.L. YAP/TAZ phase separation for transcription. Nat. Cell Biol. 2020, 22, 357–358.
  188. Ray, S.; Singh, N.; Kumar, R.; Patel, K.; Pandey, S.; Datta, D.; Mahato, J.; Panigrahi, R.; Navalkar, A.; Mehra, S.; et al. alpha-Synuclein aggregation nucleates through liquid-liquid phase separation. Nat. Chem. 2020, 12, 705–716.
  189. Hernandez-Sanchez, I.E.; Maruri-Lopez, I.; Martinez-Martinez, C.; Janis, B.; Jimenez-Bremont, J.F.; Covarrubias, A.A.; Menze, M.A.; Graether, S.P.; Thalhammer, A. LEAfing through literature: Late embryogenesis abundant proteins coming of age-achievements and perspectives. J. Exp. Bot. 2022, 73, 6525–6546.
  190. Shih, P.Y.; Fang, Y.L.; Shankar, S.; Lee, S.P.; Hu, H.T.; Chen, H.; Wang, T.F.; Hsia, K.C.; Hsueh, Y.P. Phase separation and zinc-induced transition modulate synaptic distribution and association of autism-linked CTTNBP2 and SHANK3. Nat. Commun. 2022, 13, 2664.
  191. Soltys, K.; Ozyhar, A. Ordered structure-forming properties of the intrinsically disordered AB region of hRXRgamma and its ability to promote liquid-liquid phase separation. J. Steroid Biochem. Mol. Biol. 2020, 198, 105571.
More
ScholarVision Creations