Abstract
The effects of crowding in biological environments on biomolecular structure, dynamics, and function remain not well understood. Computer simulations of atomistic models of concentrated peptide and protein systems at different levels of complexity are beginning to provide new insights. Crowding, weak interactions with other macromolecules and metabolites, and altered solvent properties within cellular environments appear to remodel the energy landscape of peptides and proteins in significant ways including the possibility of native state destabilization. Crowding is also seen to affect dynamic properties, both conformational dynamics and diffusional properties of macromolecules. Recent simulations that address these questions are reviewed here and discussed in the context of relevant experiments.
Introduction
Realistic biological environments involve a high degree of complexity and crowding at the molecular level that is still only at the beginning of being fully understood.1−5 Increasing evidence suggests that crowding can markedly affect biomolecular structure, dynamics, and thus possibly their function.6 The most obvious aspect of cellular crowding, a greatly diminished free volume with mostly entropic consequences, has long been studied in various ways using experiments, theory, and simulation.7−12 However, the degree to which artificial crowders approximate real cellular environments or even reproduce just excluded-volume effects without any other artifacts is unclear.6,13−15 At the same time, increasing attention is now focused on the effects of interactions between the cellular components16−24 and altered solvent properties,25−30 often contributing enthalpically,15,31,32 as well as the consequences of a substantial compositional heterogeneity present in biological cells.33,34
The concentration of biological macromolecules in cells is between 100 and 450 g/L, occupying 5–40% of the cytoplasmic volume.35 In addition, structural elements such as the cytoskeleton, compartments, and membrane surfaces are present, whereas genomic DNA fills out eukaryotic nuclei and a good fraction of prokaryotic cells. A variety of metabolites, small molecules that are intermediates of metabolic reactions such as phosphates, carbohydrates, amino acids, alcohols, vitamins, and other cofactors, as well as ions at significant concentrations further contribute to the rich physicochemical environment inside living cells. Experiments that are carried out under in vivo conditions or in the presence of cell extracts face substantial technical challenges.1,2,25,36 Often, it is difficult to achieve sufficient resolution to provide detailed insights, although recent progress especially with in-cell nuclear magnetic resonance (NMR)24,37−55 and in-cell fluorescence spectroscopy56−58 techniques is encouraging. An alternative is computational modeling, which can describe the structure and dynamics of biomolecules with high resolution in both space and time within model and resource limitations.59 In particular, molecular dynamics (MD) simulations using atomistic force fields are a popular avenue for studying biological systems. However, MD studies have so far largely focused on single molecules or complexes without considering crowded environments, even as multimillion atom systems are now becoming feasible,60−62 whereas questions in regard to crowding have been addressed mostly with simplified models.10 This may reflect in part the computational costs involved with fully atomistic detailed models of crowded environments, but the more important constraint is that sufficient experimental data for building high-resolution models of biological environments has not been readily available. There has also been a lack of suitable experimental data that predictions from simulating such models could be compared to. However, there is a recent influx of data from high-throughput experiments that probe the composition of a cell,63−65 and new experiments can directly probe macromolecule structure and dynamics in cellular environments.16−22,25,32,37,40,41,44−46,49,54,56,66−85 This now enables and, we would argue, requires detailed computational studies of realistic models of crowded biological environments to connect with the newly emerging experimental data and generate new hypotheses.86
Key Questions
The questions to be addressed here revolve around the structure, interactions, and dynamics of biomolecules, proteins, nucleic acids, and metabolites in cellular environments. Essentially, the key point is to determine to what extent the vast amount of knowledge about the biomolecular structure–dynamics–function paradigm generated during decades of structural biology and mechanistic analyses of biological function under in vitro conditions translates to real biological environments and ultimately what makes up life.87 The end point is a fully integrated view of biology that scales from single molecules to whole-cell models88,89 to explain life based on physical and chemical principles.90
Structure and Stability
Most of the knowledge about protein structure has been derived from crystallography where the environment, although relatively dense in terms of packing, lacks the heterogeneity of biological environments and where packing contacts are more likely related to artifacts than being representative of what would be seen in crowded biological environments. Structures based on nuclear magnetic resonance (NMR), on the other hand, are traditionally obtained under dilute conditions that are far from crowded cellular environments. Thus, it remains largely unclear whether and how the structure and stability of proteins, and more broadly the conformational energy landscape, is affected by crowding in the cell.91 Simple arguments based on a volume exclusion effect due to crowding focus on a shift in populations toward compact states on an otherwise largely unchanged landscape,11 but this view conflicts with recent experiments.17,22,23,73,92,93 Largely unaltered conformations in vivo vs in vitro are seen with in-cell solution NMR40,44 and solid-state (SS) NMR,54 but other evidence from experiments and simulations indicates that cellular environments may remodel the folded or unfolded ensemble or both16,17,49,71,94−100 with potentially varying effects in different cellular compartments.82 This suggests a more complicated picture where it is unclear whether remodeling of the ensemble is again driven mostly by entropic volume exclusion effects or by more enthalpic effects due to interactions with protein crowders and other components of the biological environment.16,17 Significant crowding effects are also expected for intrinsically disordered peptides (IDPs) where extended and dynamic ensembles have to be accommodated under crowded environments.37,69,75,76,83,100,101
Although proteins have received most of the attention, similar, less well-explored questions have been raised about how nucleic acid structures may act as crowders43 or be affected by crowded cellular environments compared to what we know about DNA and RNA from crystallography and solution NMR studies.102−109
Dynamics
Biomolecular dynamics links structure with function, and it is equally unclear how it may be affected in the context of cellular environments. Conformational dynamics is a key mechanistic factor in most biological macromolecules, and any effect of cellular crowding on such dynamics would directly impact biological function. There is some insight into altered folding kinetics under crowded conditions,48,56,80,82 and a few studies suggest that the internal backbone and side chain dynamics on sub-microsecond time scales remain largely unaltered.78,110−112 On the other hand, only a few studies have examined how functionally relevant motions in the native state, e.g., dynamics related to allosteric mechanisms, ligand binding, or catalysis, and dynamics of intrinsically disordered peptides may change upon crowding.113−115 This leaves many open questions yet to be addressed.
Diffusional motions of biomolecules in cellular environments, on the other hand, are somewhat better understood. Diffusion rates are a key determinant for the kinetics of many biological processes. Ligand diffusion is rate-limiting in many enzyme reactions, and protein diffusion is an important factor in most protein–protein and protein–nucleic acid interactions.116−120 There is a general understanding that diffusion of proteins is slowed down by about 1 order of magnitude under crowded conditions due to increased viscosity and interactions with crowder molecules,77,121−124 although this may not be generally true for natively disordered proteins.122 Anomalous diffusion results in diffusion rates varying on different time scales due to confinement within cellular compartments and caging effects due to temporary encapsulation by larger crowders.34,123,125−128 However, details of how diffusion is modulated by different cellular microenvironments and by the presence of DNA or membrane surfaces have remained unclear.77 There is also an incomplete understanding of how smaller molecules, metabolites and other ligands navigate the dense cellular environments where available space is not just limited and solvent viscosity may be altered but where the various macromolecular surfaces also provide ample opportunities for distractions that could, for example, slow down diffusion toward an enzyme active site.129
Interactions
Cellular crowding undoubtedly increases the opportunities for frequent encounters of biomolecules due to weak interactions.79,130,131 It is not clear, however, whether such soft interactions simply hinder diffusion and alter the conformational energy landscape as discussed above or whether interactions between biomolecules are altered with possible functional implications.16,66,70,74,84,132 Crowding has been found to affect native complex formation31,68,133−135 and aggregation,136−138 while other evidence indicates enhanced formation of nonspecific complexes between proteins in the cellular milieu.14 Weak interactions under highly crowded conditions may further lead to protein-enriched regions, forming distinct phases within the cell such as droplets with gel-like characteristics.139,140
Metabolite–protein interactions are also subject to crowding effects. Studies of substrate binding and product release during enzyme reactions under crowded conditions have suggested crowding effects.129 Moreover, effective metabolite concentrations in vivo may be reduced as a result of metabolites being sequestered from bulk solvent by nonspecific binding to biomolecular surfaces,141 whereas competition by promiscuous interactions with the environment, on the other hand, may modulate specific binding of ligands to target binding sites.132,142
Solvent and Membrane Environments
Finally, a fundamental question is how the solvent environment is altered in the presence of biomolecules at very high concentrations. Simple geometric considerations suggest that, although water still makes up about 70% of the cytoplasm, only a relatively small fraction is actually bulk-like, as defined by water molecules being several layers away from the closest macromolecule. This has been shown to impact water dynamics,29,143 viscosity,66,144,145 and dielectric properties,26,146 which then in turn may further exacerbate crowding effects on the macromolecules compared to dilute environments.29 While experiments have provided varying insights,28,147 a more detailed analysis of the properties of aqueous solvent inside biological cells is desirable. There are also many unanswered questions about how membrane bilayers respond to crowding,148 both inside the membrane by membrane proteins and outside by a crowded cytoplasm and how membrane surfaces modulate crowding effects in the cytoplasms.
A full review of all experiments, simulations, and theoretical studies that have looked at these questions is not possible here. Instead, the specific focus of this article is to highlight recent advances based on computer simulations of dense macromolecular systems in atomistic detail, and relate those findings to experimental studies, both past and future.
Cellular Environments in Computer Simulations
Computer simulations have examined biological systems at a wide range of scales from full atomistic detail to very coarse representations. Most simulations of crowded cellular environments have involved simplified models where macromolecules, crowder molecules, and the solvent environment are represented with different degrees of coarse-graining.95−98,127,149−152 For example, coarse-grained models of proteins in the presence of spherical crowders without any explicit representation of solvent are one avenue for studying the effect of crowding on protein folding or association equilibria,30,153 whereas Brownian, Stokesian, or other stochastic dynamics simulations of rigid protein models at different levels of detail have offered insights into how crowding modulates diffusive properties.127,151,154−158 Another possibility is postprocessing of trajectories using particle-insertion methods.159 All of these approaches involve simplifications that are attractive for saving computational costs and have allowed the exploration of relatively long time scales, in the range of micro- to milliseconds. However, there are significant trade-offs as intermolecular interactions and solvent effects often do not fully reflect cellular environments and intramolecular dynamics may be neglected or not fully accounted for.
As with any modeling, the degree of detail should match the questions that are being asked and an in silico model of a complete cell does not necessarily have to include every atom that is present. However, we argue that the central questions that are being asked here, conformational sampling, stability, and dynamics of biomolecules in the cellular environment, require, at the minimum, a model with the following features: (1) Proteins, nucleic acids, and metabolites should be fully flexible with an interaction potential that is able to maintain native states under dilute conditions but that can also sample non-native states in response to interactions with cellular components. (2) Weak interactions between the cellular components are physically accurate and predictive without requiring previous knowledge about specific complexes, since nonspecific interactions dominate in cellular environments. (3) The essential characteristics of the solvent environment are captured including an ability to reflect altered solvent behavior under crowded conditions such as reduced solvent dynamics and hydrodynamics interactions.
These requirements are a tall order for most coarse-grained approaches that typically focus only on certain aspects while failing to capture other features. For example, Go models can describe the thermodynamics of protein folding and complex association assuming that the structures of the end states are known and may offer some insight into how crowding shifts populations between known states.160−162 However, perturbations of the energy landscape toward unknown conformational states that may be induced as a result of crowding are difficult to capture with structure-based models. On the other hand, simulations of rigid or spherical biomolecules via Brownian and Stokesian dynamics schemes can describe diffusive properties well,155,157,163,164 but these approaches neglect intramolecular dynamics and conformational sampling. More sophisticated physics-based coarse-grained models165 such as PRIMO,166 OPEP,150 or COFFDROP167 that are combined with implicit solvent and/or models for hydrodynamic interactions156 are in a better position to satisfy these criteria, as the increased level of model detail requires fewer constraints and provides more transferability and ability to capture the various features outlined above. Such models can at least at a general qualitative level capture the main features of biomolecular structure and dynamics cellular environments.149,168,169 Atomistic simulations, with explicit26,143,170−176 or implicit solvent,27,97,98 or multiscale models, where atomistic and coarse-grained resolutions are mixed,169 provide even greater levels of detail and can, at least in principle, satisfy all of the requirements for modeling crowded cellular environments outlined above. However, because the balance between molecular stability, weak interactions, and solvent interactions in crowded environments depends on subtle shifts between enthalpic and entropic energy terms, the major challenge is an accurate interaction potential, both at the coarse-grained and atomistic level that can accurately reproduce both intra- and intermolecular interactions.177−179 For example, just a slight unbalance between protein–protein and protein–water interactions is enough to lead to aggregation177 and overcompaction artifacts180,181 with significantly mispredicted diffusive properties being one consequence. The choice of the water model is also a concern, since the most widely used models (SPC/E182 and TIP3P183) significantly mispredict viscosity and dielectric properties26,184,185 which in turn impacts macromolecular properties.185 Therefore, future efforts should continue the improvement of force fields186−192 to better balance the interactions between the various cellular components.
Another issue is the high computational cost of the most detailed models that limits the time scales that can be accessed. Current computer hardware, especially GPUs and other many-core coprocessors such as Intel Xeon Phi and high-performance simulation software,193−198 are now allowing for atomistic simulations of biomolecules in crowded cellular environments on time scales well into the microsecond regime with millisecond dynamics being within reach now199 and probably becoming routine within a decade. While such time scales are short compared to most functional dynamics of biological processes, they cover the majority of diffusive processes within the cell for all but the very largest particles such as ribosomes based on the extrapolation of sub-microsecond diffusion rates in crowded environments estimated from simulation.173 Therefore, the development of more accurate coarse-grained models that can describe biomolecular dynamics on millisecond to second time scales remains an attractive goal in the foreseeable future. The length of the simulations also determines the degree of sampling and the statistical significance of properties determined via averaging. However, this is a minor issue, since parallel computing architectures facilitate replicate simulations and crowded systems with multiple copies of a given system of interest allow ensemble averaging in addition to time averaging.
One strategy for overcoming limitations in simulation times is the use of enhanced sampling techniques200 such as umbrella sampling,201 generalized-ensemble methods,202,203 metadynamics,204 λ-dynamics,205 or accelerated MD,206 just to name a few. These methods generally focus on specific processes where significant kinetic barriers hinder sampling, and they are usually used to extract thermodynamic quantities such as ligand binding affinities. So far, enhanced sampling methods have not seen extensive use, in part because there is still very limited knowledge about what states may exist under crowded conditions and enhanced sampling methods are less advantageous for exploratory simulations where little is known about a system’s behavior. However, as more information is becoming available, the energetics of protein–protein or protein–ligand interactions or conformational dynamics between different states seen under crowded conditions are good problems for using enhanced sampling. Enhanced sampling methods are also less effective for accelerating processes that are limited by slow diffusion, a major factor in crowded cellular environments. Diffusion-limited processes can be accelerated, however, by using mean-field models of cellular environments,27,103,152 although the implicit models have other drawbacks, as they may oversimplify cellular environments169 and the extraction of kinetic properties is hindered. Developing enhanced sampling methods that target the sampling of diffusion-limited processes while fully capturing all components of cellular environments and allowing the extraction of thermodynamic and kinetic properties remains a significant challenge that we hope can be addressed in the future with new methods.
While it is clear that the simulation of crowded cellular environments is still in its infancy, the aim of this Feature Article is to review what has been reported so far from molecular simulations of proteins in dense cosolvents,97,98,112,143,207,208 of dense peptide and protein solutions,26,29,143,159,170−172,174,175,209−213 and of more complex models of cellular environments.112,163,168,214
Atomistic Simulations of Cellular Crowding
Table 1 gives an overview of crowding simulations studied via atomistic models in recent years. We selected studies of peptides and proteins where at least the solute for which crowding effects are studied is represented in atomistic detail but allowed for crowders and/or solvent environments to be modeled in a coarse-grained or implicit fashion. We included simulations of dense amino acid and peptide systems, as they offer insight into protein amino acid interactions under crowded conditions. Despite some overlap, we did not consider simulations involving amyloidogenic peptides.216 The concentrations of solutes in those studies are generally low compared to crowded environments, whereas their analysis primarily focuses on their aggregation propensities offering less insights into cellular crowding. In addition to studies of peptides and proteins, there are also a few studies involving nucleotides and nucleic acids under crowded and concentrated conditions,103,217−219 but since general insights have remained limited, these studies will also not be discussed here.
Table 1. Simulations of Crowded Peptide and Protein Systems with Atomistic Detail.
solutesa | solvent | sampling | year | groupref |
---|---|---|---|---|
Gly/Val/Phe/Asn (50–300 g/L) | expl. water | 1 μs MD | 2013 | Elcock178 |
Gly/Ala/Val/Phe (50–300 g/L) | expl. water | 10 ns MD | 2015 | Price175 |
(GlySer)2 (9–560 g/L) | TIP3P water | 8 μs MD | 2014 | Rao29 |
27 × (Xxxl)4 (40 mM; 10–30 g/L) | expl. water | 100 ns MD | 2016 | Taiji213 |
Gly5 (0.03/0.3 M; 10/100 g/L) | TIP3P water | 100 ns MD | 2014 | Pettitt176 |
1 × Ala/Ala15/melittin | GBk (ε = 2–80) | 44–83 ns ReXe-MD | 2008 | Feig27 |
1 × Trp-cage/melittin + 8 × protein G (CG) (40% vol; 540 g/L) | GBk (ε = 80) | 20 ns ReXe-MD | 2012 | Feig169 |
1 × chignolin, 0–8 × HSc (0–40% vol) | TIP4P water | 1 μs MD | 2012 | Rajagopalan215 |
1 × Trp-cage + 8 × BPTI (7% vol; 95 g/L), 8 × HSc (7%/20% vol) | implicit | 107 ReXe-MC moves | 2015 | Irbäck98 |
1 × Trp-cage/GB1m3 + 8 × BPTI (7% vol; 95 g/L), 8 × GB1d (7% vol; 95 g/L), 8 × HSc (20% vol) | implicit | 107 ReXe-MC moves | 2016 | Irbäck97 |
4/8 × villinb (6/9.2 mM; 25/40 g/L) | SPC wat./NaCl | 50–200 ns MD | 2014 | Zagrovic177 |
8 × GB1d (10–30% vol; 135–405 g/L), 4 × GB1d/8 × villinb (10–43% vol; 135–580 g/L) | TIP3P water | 300 ns MD | 2012; 2013 | Sugita/Feig26,171 |
1 × CI2,f 8 × lysozyme (100 g/L), 8 × BSAg (100 g/L) | TIP3P water | 100–250 ns MD | 2012 | Sugita/Feig170 |
1/4 × ubiquitin (16% vol; 216 g/L), 19/157 metabolites | TIP3P water, KCl, MgCl2 | 200 ns MD | 2011 | Guallar207 |
1 × ubiquitin, glucose (325 g/L) | TIP3P water | 500 ns MD | 2013 | Dal Peraro112 |
1–3 × ubiquitin, glucose (108/325 g/L) | TIP3P water | 600 ns MD, 30–70 ns ReXe-MD | 2014 | Dal Peraro208 |
3 × ubiquitin, Glu/Arg/glucose (300 g/L) | TIP3P water | 600–800 ns MD | 2016 | Dal Peraro143 |
1 × ACTRh/6 × NCBDi/1 × IRF-3j (175–296 g/L), 4 × ACTRh/24 × NCBDi/4 × IRF-3j (183 g/L) | TIP3P water | 100 ns–3 μs MD | 2016 | Orozco174 |
8 × BSAg (400 g/L), 88 × GB1d (400 g/L), 132 × villinb (400 g/L) | TIP3P water | 100 ns MD | 2017 | Sugita/Feig172 |
100s–1000s of proteins/RNA (300 g/L), 5000–40,000s of metabolites | TIP3P water | 20–140 ns MD | 2016 | Sugita/Feig173 |
Reported concentrations are underlined.
Villin headpiece.
Hard sphere.
B1 subunit of protein G.
Replica exchange sampling; given times are per replica.
Chymotrypsin inhibitor 2.
Bovine serum albumin.
Activator for thyroid hormone and retinoid receptor.
Nuclear coactivator-binding domain of CREB.
Interferon regulatory transcription factor.
Generalized Born implicit solvent.
Any amino acid.
As a representative overview of the kind of systems that are now being studied in atomistic detail via computer simulations, Figure 1 shows models that have been studied recently in our group. The systems range from mixtures of a few proteins at different concentrations to models of cytoplasmic environments with thousands of macromolecules. The number of atoms in these systems ranges from 50K to 100M atoms, and the time scales that were reached varied between tens of nanoseconds and microseconds. In all of these cases, all-atom force fields were used to describe the solutes, aqueous solvent, and any other molecules such as metabolites that were present, allowing for unrestrained dynamics of all components. On the basis of simulations of such systems, exciting new ideas have been found for how the stability and dynamics of proteins may be modulated by interactions with other molecules in the cell.
Figure 1.
Crowded and cellular systems studied via molecular dynamics simulations: Concentrated solution of villin in explicit water and ions (top left); chymotrypsin inhibitor 2 (CI2; blue) in the presence of bovine serum albumin (BSA) crowders with explicit water (top right); model of a bacterial cytoplasm consisting of proteins, RNA, metabolites, ions, and water shown only for part of the system in cyan (bottom).
Solute Interactions
A key feature of crowded environments is the opportunity for frequent encounters between biomolecules. The well-understood volume-exclusion effect of crowding typically assumes crowder molecules that are repulsive or at least not attractive so that such interactions would be minimized to just unavoidable random collisions. However, atomistic simulations of crowded environments have shown many examples of attractive interactions between proteins97,98,170,171,173,174,177 and between proteins and other molecules such as metabolites.143,173,207 Such interactions are generally nonspecific; i.e., the formed complexes cover a broad range of arrangements and do not correspond to specific functional complexes, but they have the potential to significantly perturb protein structure and dynamics over solutes in dilute environments that are free from such contacts. Weak transient interactions, commonly termed quinary interactions,2 nevertheless, do have the potential to be functionally relevant, for example, to bring enzymes involved in a common metabolic pathway into proximity to facilitate substrate channeling.173 Whether such close interactions are formed depends greatly on the involved partners. For example, chymotrypsin inhibitor 2 (CI2) interacts extensively with lysozyme but hardly at all with bovine serum albumin (BSA), whereas BSA prefers to interact with other BSA molecules.170 A similar finding was reported for Trp-cage in the presence of protein G vs bovine pancreatic trypsin inhibitor (BPTI) crowders, with BPTI more strongly self-interacting with other BPTI molecules.97 Differential protein interaction propensities can be traced to the nature of the amino acids decorating the surface of different proteins that lead to charge differences98 and hydrophobic patches97,171 along with sufficiently good shape complementarity to allow close interactions. Simulations of dense solutions of amino acids and small peptides29,143,175,176,178,213 have offered additional insights into which amino acids are more favorable to interact178 and how such interactions are formed via hydrogen bonding,176,213 aromatic ring interactions,213 dipole–dipole interactions,176 or mediation via water molecules.29 Generally, the degree to which amino acid interactions were found to be favorable reflects established hydrophobicity profiles;213 e.g., isoleucine, leucine, and valine readily interact due to their hydrophobicity, whereas basic and acidic residues strongly prefer to remain solvated.143,213 There are differences that can be at least in part be explained by strong intermolecular self-interactions, e.g., between asparagine and glutamine side chains,213 but variations were also seen as a function of the force field that was used.178,213
A concern that has been raised is whether the overall strength of peptide–peptide and protein–protein interactions in dense solutions is correctly calibrated in the current generation of force fields. On the basis of simulations of villin, it has been suggested that protein–protein interactions may be overall too strong relative to protein–water interactions.177 This would also be generally consistent with the observation that intrinsically disordered peptides are too compact with most force fields220−222 unless amino acid–water interactions are enhanced.180,191 However, simulations of concentrated amino acids and peptides seem to come to the opposite conclusion that solubilities are significantly overestimated compared to experimental values.176,178 This is clearly an area that will require further attention, but significant challenges exist in how to effectively compare with experimental measurements. For example, the extraction of solubilities that are comparable with macroscopic measurements, while possible in principle,176 requires a system with a large number of solutes that are simulated over long enough times to overcome kinetic bottlenecks of aggregate formation. Another possibility is to match rotational and/or translational diffusion rates for concentrated solutions as the slow-down in diffusion upon crowding is expected to depend upon the extent of protein interactions170 while deviations from nonideal behavior vary with the crowder type.123 A meaningful comparison between experiments and simulations also requires long simulations as well as reliable force field parameters for solutes and solvent and faces potential experimental difficulties in analyzing highly retarded and likely complex dynamics in concentrated solutions, e.g., via nuclear magnetic resonance (NMR) spectroscopy.81,175
Apart from interactions between peptides and proteins, some studies have examined interactions with high concentrations of smaller cosolvents such as metabolites173,207 or glucose112,143,208 or osmolytes such as trimethylamine N-oxide (TMAO).223 The general finding is that such molecules may interact extensively with protein solutes172,207 or among themselves,207 displace water molecules,143 and further affect the structure and/or dynamics of proteins as a consequence of such interactions.173 The simulations can be compared with experiments224−227 by quantifying preferential solvation via the integration of MD-derived density fluctuation of water and cosolvents to obtain Kirkwood–Buff integrals.228−232 While this approach can be readily applied for proteins in two-component, or at the most, three-component solvents with small cosolvent molecules, the extension to cellular environments where a large number of components are present and density distributions of larger molecules such as proteins do not converge well is not as straightforward.
Protein Structure Stability
A central question is whether crowded cellular environments affect the stability of protein native states and, more broadly, whether the conformational landscape is modulated over dilute conditions. As mentioned already above, one aspect of crowding, the reduction of accessible volume, is well understood and generally believed to stabilize the native state as the most compact state a given protein will assume.233 Unfortunately, such a simple view is not borne out in the atomistic simulations of peptides and proteins under crowded conditions. Instead, the simulations suggest that crowded cellular environments may perturb the native state, either directly due to interactions with other solutes170,173 or indirectly due to altered solvent properties (see example in Figure 2).26,27,29,172 There is also evidence from simulation that crowding may not just affect the native state but also folding intermediates215 or the unfolded ensemble based on the analysis of an intrinsically disordered peptide.174
Figure 2.
Villin native state destabilization under crowded conditions.
One theme for such native state perturbations is the competition between inter- and intramolecular interactions that can lead to a shift toward more extended states. Examples range from concentrated (GlySer)2 solutions,29 Trp-cage in the presence of protein G169 or BPTI,97,98 villin in villin/protein G mixtures,171 pyruvate dehydrogenase α subunit in a cytoplasmic environment,173 and ordered and natively disordered proteins in a heterogeneous environment.174 The presumed general mechanism in all of these cases is that favorable intermolecular interactions with peptide or protein crowders combined with an overall reduction of solvent exposure due to complex formation can drive partial unfolding. However, specific interactions with crowders such as edge-to-edge β-sheet interactions may also have the opposite effect of stabilizing the native state, as demonstrated for the GB1m3 hairpin.97
A second theme is the modulation of secondary structure propensities as a result of crowder interactions. While this has been systematically analyzed using coarse-grained models,95,96 atomistic simulations have provided evidence for an apparent stabilization of secondary structure elements, especially helices, as a result of crowding.27,97,98,174 One way to understand that is via crowder amino acid side chain interactions that stabilize helices95 or β-hairpin turns.97,98 Another interpretation is based on a reduced dielectric response of solvent in crowded environments26,27 where intramolecular hydrogen bonding would be strengthened as a result of reduced electrostatic screening.27
A third theme is a disruption of the hydrophobic core that would lead to a general destabilization of natively folded peptides.27,29,169,174 This could simply be a corollary of a shift toward extended states driven by protein–protein interactions, but it may also be understood in the context of a reduced dielectric response of the solvent that would imply a weakened hydrophobic effect.27
Finally, a newly emerging theme is the role of metabolites in modulating protein structure. Functionally relevant ligand-induced conformational changes are well documented,234 and it is easy to imagine that nonspecific metabolite interactions could have similar effects. Indeed, in a recent study of a model for a bacterial cytoplasm, nonspecific binding of negatively charged metabolites was found to modulate the structure of phosphoglucokinase and induce structural compaction via electrostatic screening.173
Experimental validation of the specific effects seen in the different simulations is largely missing so far and is therefore difficult to tell how real the specific effects seen in the simulations are. We imagine that it would be relatively straightforward to test the predicted effects of metabolites on protein structures, whereas in-cell NMR24,38,41,42,44−46,115 and fluorescence techniques56,58,235 should be able to test how real the crowding-induced perturbations of the conformational ensembles are. Specifically, the hope would be that the experiments could identify the presence of the predicted non-native subpopulations in the presence of certain protein crowders, but it may be challenging to discern minor populations of heterogeneous non-native conformations induced by crowding in vivo.(38) Other avenues for comparing the simulations with experiments would be to identify key residues based on the simulations that appear to play a role in facilitating crowding-induced structural changes and propose experimentally testable mutations that should be resistant to crowding effects on structural stability.
While the results from specific simulations are questionable without experimental validation, the overall picture from the simulations taken together is that the conformational landscape of proteins may become perturbed as a result of nonspecific weak interactions in cellular environments and a resulting competition between intra- and intermolecular interactions that can stabilize non-native states. The possibility of native-state destabilization in vivo is generally consistent with experiments,22,23 but the simulations, furthermore, suggest specific mechanisms by which the energy landscape may be altered such as direct contacts with nearby crowder protein surfaces that shift the balance toward unfolded states and a role of altered solvent properties and metabolites. Nevertheless, much remains to be learned about more general principles that would, for example, allow predictions of how the structure and stability of a specific protein may be affected by cellular environments. Such insight is desirable to advance the fundamental understanding of biology, but it is also practically relevant in the context of protein design and therapeutic developments involving protein targets.
Protein Dynamics
Dynamics is what connects structure to function for all but a few proteins. The dynamics of proteins spans a wide range of time scales and covers internal, conformational rearrangements as well as diffusive motions. Internal dynamics may be further distinguished into fluctuations around the native state and conformational transitions between different states including the folding–unfolding transition. The available simulations suggest that fluctuations around the native state, as measured by S2 order parameters or Debye–Waller B-factors, are only moderately affected by crowding.97,98,143,174,208 Other results point at increased fluctuations as a result of protein–protein contacts in CI2 when interacting with lysozyme,170 whereas somewhat reduced backbone dynamics were observed in other studies as a result of crowding.112,174 Experiments also indicate moderate effects of crowding on fast internal motions but a stronger influence on slower conformational dynamics.41,110,236 A more detailed understanding of how crowding may affect internal dynamics beyond these general observations can be gained from direct comparisons between MD and NMR,143 but further studies are needed to address this question. A somewhat clearer picture exists for the kinetics of conformational transitions between different states. Crowding with either proteins174 or glucose143,208 led to a significant slowdown in conformational sampling. A similar conclusion was also reached for (GlySer)2.29 The slowdown in conformational transitions has been correlated with crowders, such as glucose, or water molecules becoming trapped between solutes and exhibiting slow diffusion and long residence times.29,208 It remains difficult, however, to clearly separate cause and effect with respect to retarded solvation causing slow conformational dynamics or vice versa. In any case, as conformational transitions are essential in many biological mechanisms, understanding how much such kinetic processes are slowed down in cellular environments is critical for fully understanding biological processes.
Translational and rotational diffusion is well-known to be significantly slower under crowded conditions (by about a factor of 10).237−240 Furthermore, experiments show varying degrees of deviations from nonideality with respect to how the observed diffusion slows down with increased viscosity as a function of crowder type (proteins vs nonproteins and also between different proteins) that are taken to indicate different degrees of protein–crowder interactions.121−123 Diffusion rates from atomistic simulations under crowded conditions are generally consistent with these observations,170,173,175,208 and simulations clearly show a slowdown in diffusion that strongly depends on protein–protein interactions.170,173 This means in the context of a diverse cytoplasmic environment that different copies of the same protein may experience very different diffusion rates depending on the local environment. Or for a given molecule, diffusion may vary significantly on μs–ms time scales as different local environments are sampled as the molecule diffuses across a cell.173 In addition, there is evidence for anomalous diffusion as a result of caging by crowders, especially when interactions with those crowders are weak.170 One may take that argument a step further to postulate that a diffusing protein molecule in a heterogeneous cellular environment should experience a spectrum of diffusion time scales as crowders of different sizes and with different levels of attraction are encountered. This analysis informs dynamic models of processes at cellular levels that require diffusion rates as input.241
Taken together, new insight into the dynamics of biomolecules under crowded conditions from the simulations is a significant degree of heterogeneity in the diffusion on microsecond time scales, while other findings so far are largely confirming more general observations that are known already from experiments. We believe that simulations are especially valuable for describing how dynamics varies over different time scales in the presence of crowding. This will require that more efforts are made to extend future simulations of crowded systems well into the tens of microseconds regime which is technically becoming possible with specialized hardware such as the Anton 2 supercomputer.242
Solvent Properties
Most of the simulations discussed here involve explicit solvent. Therefore, the analysis of how crowding and cellular environments affects solvent properties is straightforward. The first realization from such analysis is that bulk-like water, defined here as water molecules beyond the second solvation layer from the closest macromolecular solute, only represents a small fraction under cellular conditions with macromolecular concentrations of 30–40% vol.26 For concentrations beyond 40% vol, even water molecules in the second solvation layer become rare. This means that under crowded conditions most of the water is either in the first solvation shell in direct contact with a macromolecular solute or sufficiently close to still experience the electric field of a solute (see Figure 3). While this has only a minor impact on water structure, analyzed in terms of pairwise radial distribution and hydrogen bonding,26,29 water dynamics has been found to be more strongly altered. According to the simulations, self-diffusion of water is reduced and residence times are increased significantly upon crowding.26,29 Further analysis suggest that the crowding effects vary with the size of the protein crowders.172 Smaller proteins led to a more pronounced slowdown in dynamics than large proteins at the same volume fraction, essentially as a result of larger protein surface areas for the smaller crowders.
Figure 3.
Interfacial water under crowded conditions with one (1) or two (2) layers of water between noninteracting proteins.
As a further indication of reduced dynamics, the dielectric response of water is also lowered to values between 20 and 60,26 in agreement with experimental measurements.146 The dielectric response of the entire cellular environment is more difficult to assess. Polar and charged metabolites and ions may increase the polarizability,207 while proteins with interior dielectric constants lower than pure water243,244 would be expected to reduce the overall dielectric response. A net decrease in the dielectric response in cellular environment would also be consistent with a tendency toward native-state destabilization with a partial loss of the hydrophobic core and an increase in secondary structure formation as discussed above.
Taken together, the simulations suggest that solvent exhibits significantly altered properties in crowded environments which in turn is expected to affect biomolecular structure, dynamics, and function. The general idea is that crowding reduces water dynamics and lowers the dielectric response which in turn suggests a reduced hydrophobic effect in crowded cellular environments. There is limited experimental data consistent with this picture,146 but further experimental probes that quantify solvent dynamics and describe electrostatic properties under crowded conditions are needed to validate the theoretical predictions.32
Limitations and Failures with Current Simulations
Since most simulations on crowded cellular systems lack direct experimental validation so far, it is difficult to assess how realistic the predictions made by the simulations really are. We expect that the overall picture of altered conformational landscapes and reduced diffusional dynamics as a result of intermolecular interactions with crowders is largely correct but there are indications that protein–protein interactions could be too strong177 and that there is a bias toward the sampling of compact states vs more disordered states due to force field artifacts180,191,220−222 limiting the ability to accurately describe conformational ensembles of peptides and proteins under crowded conditions. Ongoing force field improvements are aimed at addressing such problems.186 The other major issue are the relatively short time scales in the atomistic simulations published to date that have largely limited the study of dynamics to sub-microsecond time scales. Although longer time scales can be reached with coarse-grained models,154,155 such models do not provide as much detailed insight and a major challenge going forward is how to extend atomistic simulations of crowded systems further into the μs–ms regime. Quantitative studies of dynamics are also affected by the venerable three-site TIP3P and SPC/E water models in common use today that do not describe viscosity and dielectric properties of water as well as newer, more expensive four- and five-site models.245,246 The use of the newer water models may appear straightforward, but the main challenge is that the current force fields will need to be reparametrized to be compatible with these water models while the higher computational costs due to the additional sites further limit the time scales that can be reached in simulations.
Connections with Experiment
The recent detailed atomistic simulations of crowded cellular environments have been motivated and enabled by experiments in the growing field of cellular structural biology.24,38,39,41,42,44−46,54,56,85,247 There are increasing opportunities now to advance the field by complementing the experiments with simulations,143,170,171 where the role of simulations is ideally to provide more detailed insights and make predictions that can then be validated via additional experiments. Especially new NMR and fluorescence experiments5,6,15,17,19,20,22,44,48−50,52,54,56,71,73,75,82,84,92,105,108,132,248 are well suited to connect simulations with experiments as the experimental observables can be extracted easily from simulations and the high resolution in both spatial and temporal scales, especially with NMR, matches the level of detail in MD results well. As the experiments offer insights into biomolecular structures under concentrated conditions in vitro and in vivo, simulations can provide additional details about the exact mechanisms by which the structure and dynamics of a given system may be altered as a result of quinary interactions with cellular environments. Simulations are in principle able to combine dynamic and structural analyses for a given protein in the presence of crowders and make a direct connection with how crowding interactions lead to a modified structural ensemble. One advantage of simulations is that even minor populations can be observed that could be more difficult to detect experimentally and simulations can, at least in principle, pinpoint and quantify which interactions with other molecules in a more complex cell-like environment are most relevant for explaining crowding effects observed, for example, via in-cell NMR experiments. It is also possible to consider a variety of mutants for both the protein that is studied and the crowder proteins to obtain a more detailed view of how individual residues contribute to crowding effects. The results of such simulations would then lead directly back to experimentally testable hypotheses with respect to mutant proteins predicted to exhibit altered susceptibility to crowding effects.
As new simulations and experiments of cell-like environments have just recently begun to emerge in parallel, there are not many examples of direct comparisons yet as different systems have been studied via experiment and simulation. Focusing both sides, experiments and simulations, on the same systems in the same environments will be necessary to lead to more productive and meaningful interactions. However, coordinating experiments and simulations is not without challenges. Experimental systems may involve large crowders such as bovine serum albumin (BSA) that are computationally costly to study via simulation or involve complex in vivo environments that are basically impossible to reproduce exactly in silico. On the other hand, some computational studies involve systems that are difficult to prepare or analyze experimentally such as highly concentrated protein solutions171 or involve probe molecules at low concentrations in highly complex environments.173 Nevertheless, there are questions raised by recent simulations that could be tested experimentally, such as, for example, the idea that the two domains of phosphoglucokinase are closer in cellular environments because of electrostatic effects,173 and variations in secondary structure propensities as a function of crowder interactions.95,96 On the other hand, recent experiments of altered protein stabilities of small proteins like CI2,22 protein L,73 protein G,71 and SH348 in the presence of protein crowders as well as cosolute and crowding effects on the binding energetics of a dimer-forming protein G variant,68 just to give some examples, are a strong motivation for simulations to determine the exact mechanisms by which destabilization occurs.
Finally, direct imaging methods are rapidly advancing in recent years and super-resolution microscopy249 and high-resolution cryo-electron microscopy250 are primed to study the structure and dynamics of cellular environments, especially if the resolution both in space and time can be further increased.
Next Steps
Detailed simulations of crowded cellular environments have so far largely focused on proteins and protein crowders. There is also some insight into the role of metabolites as cosolvents under crowded conditions, but there is clearly room for additional studies to explore how metabolites may modulate biomolecular structure and dynamics in crowded environments. A related and possibly more important question is how metabolites behave in the presence of high concentrations of biomolecules where there are ample opportunities for nonspecific interactions that could reduce effective concentrations and distract from finding enzyme active sites. Insight into such questions would be especially relevant for understanding the dynamics of drug molecules where the kinetics of reaching a given target site is a key determinant of efficacy and important for understanding the mechanisms of unfavorable potential side effects.
Other largely unexplored questions revolve around interactions with nucleic acids, membranes, and cytoskeletal components. Some studies are hinting at crowding effects altering the structure and dynamics of DNA and RNA,103,104 but how the presence of nucleic acids in the cellular interior affects proteins that do not specifically bind DNA or RNA is unclear. Chromosomal DNA fills a large part of bacterial cells251−253 and most cytoplasmic proteins will not be able to avoid nonspecific interactions with the DNA, yet there is almost no insight, especially from simulation, how that may affect protein diffusion and/or stability. The major challenge in addressing such questions is the lack of realistic high-resolution molecular models of chromosomal DNA.
Membrane protein crowding has received some attention with limited insights from simulations.148,254,255 However, interactions between membranes and crowded cellular interiors are unclear. For example, one may wonder how nonmembrane binding proteins that are pushed to interact with membranes simply by virtue of cellular crowding maintain their stability and whether they become kinetically trapped or gain fluidity by facing a membrane bilayer instead of a crowded cytoplasm. It is also possible that the structure and dynamics of membranes themselves are affected by the presence of a crowded cytoplasm, little of which has been explored so far. Clearly, this is a wide open area that is ripe for detailed atomistic simulations to generate insights and guide experiments.
The ultimate goal is a molecular-level understanding of entire cells. Fully atomistic whole-cell simulations, at least for bacterial cells, are not too far from reality. However, a larger impact will probably be realized by using the detailed insights from atomistic simulations of cellular environments to build more simplified models that capture the essential physics but scale to longer time scales and allow the exploration of biological questions by the broader community without requiring access to extreme high-performance computing environments. Computationally tractable, yet physically realistic in silico whole-cell models also offer much potential for transforming rational drug design protocols that up to now largely rely on single molecules and highly empirical approaches.256,257
Acknowledgments
Funding was provided by the National Institute of Health Grants R01 GM084953 and GM103695 (to M.F.), the National Science Foundation Grant MCB 1330560 (to M.F.), the Fund from the High Performance Computing Infrastructure (HPCI) Strategic Program (hp120309, hp130003, hp140229, hp150233) and HPCI general trial use project (hp150145, hp160120) and FLAGSHIP 2020 project focused area 1 “Innovative drug discovery infrastructure through functional control of biomolecular systems” (hp160207, hp170254) of the Ministry of Education, Culture, Sports, Science and Technology (MEXT) (to Y.S.), a Grant-in-Aid for Scientific Research on Innovative Areas “Novel measurement techniques for visualizing “live” protein molecules at work” (No. 26119006) (to Y.S.), a grant from JST CREST on “Structural Life Science and Advanced Core Technologies for Innovative Life Science Research” (to Y.S.), RIKEN QBiC, iTHES, and Dynamic Structural Biology (to Y.S.), a Grant-in-Aid for Scientific Research (C) from MEXT (No. 25410025) (to I.Y.). Computer time was used at NSF XSEDE facilities (TG-MCB090003).
Biographies
Michael Feig is a Professor of Biochemistry & Molecular Biology at Michigan State University, USA. He obtained a Diplom degree in Physics from the Technical University of Berlin, Germany, continued the study of Computational Chemistry resulting in a Ph.D. degree at the University of Houston, Texas, USA, and completed postdoctoral training at The Scripps Research Institute in La Jolla, California, USA, before becoming faculty at MSU in 2003. His primary interests are in the development and application of molecular modeling and molecular dynamics techniques to biological questions, especially to connect biomolecular structure and dynamics of single molecules to biological function at the cellular level.
Isseki Yu is a Research Scientist at Interdisciplinary Theoretical Science Research Group (iTHES), RIKEN, in Japan. He also works as a member of Theoretical Molecular Science Laboratory, RIKEN. He obtained his Ph.D. in engineering at Yokohama National University in 2003. He was a postdoctoral fellow in Nagoya University (2003–2008) and an assistant professor in Aoyama Gakuin University (2008–2013). In 2013, he joined the Theoretical Molecular Science Laboratory, RIKEN, and started to investigate the cellular environment with large-scale molecular dynamics simulations.
Po-hung Wang is a Postdoctoral Researcher at the Theoretical Molecular Science Laboratory, RIKEN, Japan. In 2013, he obtained his Ph.D. in Physics in the University College London, UK. During his Ph.D. study, his primary interest was small molecule transport in proteins. After joining RIKEN, he shifted his focus to the study of various interactions between macromolecules in crowded environments using molecular simulations.
Grzegorz Nawrocki completed his M.S. degree in Biophysics at Jagiellonian University, Poland. He earned his Ph.D. in Biophysics at Polish Academy of Sciences, Poland, where he worked with Professor Marek Cieplak on characterizing interactions of biomolecules with the surfaces of solid bodies. Currently, he is a Postdoctoral Research Associate at Michigan State University working on developing force fields and analysis methods for molecular dynamics simulations. He studies diffusion, stability, and interactions of proteins and lipid bilayers in the crowded cellular environment.
Yuji Sugita is a Chief Scientist at Theoretical Molecular Science Laboratory, RIKEN, in Japan. He also works as a team leader at Laboratory of Molecular Function Simulation, RIKEN, and at Computational Biophysics Research Team, RIKEN. He obtained his Ph.D. in chemistry at Kyoto University in 1998, working with Nobuhiro Go. He was a postdoctoral fellow in RIKEN (1998), a research associate in IMS (1998–2002), and a lecturer in the University of Tokyo (2002–2007). In 2007, he joined RIKEN as a PI (Associate Chief Scientist) and was promoted to Chief Scientist in 2012. His group is primarily working on the development of a high-performance molecular dynamics program for investigating the structure–dynamics–function relationship of biomolecules in cellular environments.
The authors declare no competing financial interest.
References
- Gershenson A.; Gierasch L. M. Protein Folding in the Cell: Challenges and Progress. Curr. Opin. Struct. Biol. 2011, 21, 32–41. 10.1016/j.sbi.2010.11.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wirth A. J.; Gruebele M. Quinary Protein Structure and the Consequences of Crowding in Living Cells: Leaving the Test-Tube Behind. BioEssays 2013, 35, 984–993. 10.1002/bies.201300080. [DOI] [PubMed] [Google Scholar]
- Mittal S.; Chowhan R. K.; Singh L. R. Macromolecular Crowding: Macromolecules Friend or Foe. Biochim. Biophys. Acta, Gen. Subj. 2015, 1850, 1822–1831. 10.1016/j.bbagen.2015.05.002. [DOI] [PubMed] [Google Scholar]
- Spitzer J.; Pielak G. J.; Poolman B. Emergence of Life: Physical Chemistry Changes the Paradigm. Biol. Direct 2015, 10, 33. 10.1186/s13062-015-0060-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heyden M.; Ebbinghaus S.; Winter R. Das Innere der Zelle: Ein komplexes Lösungsmittel. Chem. Unserer Zeit 2017, 51, 26–33. 10.1002/ciuz.201700777. [DOI] [Google Scholar]
- Gnutt D.; Ebbinghaus S. The Macromolecular Crowding Effect – From in vitro into the Cell. Biol. Chem. 2016, 397, 37–44. 10.1515/hsz-2015-0161. [DOI] [PubMed] [Google Scholar]
- Shahid S.; Hassan M. I.; Islam A.; Ahmad F. Size-Dependent Studies of Macromolecular Crowding on the Thermodynamic Stability, Structure and Functional Activity of Proteins: In vitro and in silico Approaches. Biochim. Biophys. Acta, Gen. Subj. 2017, 1861, 178–197. 10.1016/j.bbagen.2016.11.014. [DOI] [PubMed] [Google Scholar]
- Minton A. P. The Influence of Macromolecular Crowding and Macromolecular Confinement on Biochemical Reactions in Physiological Media. J. Biol. Chem. 2001, 276, 10577–10580. 10.1074/jbc.R100005200. [DOI] [PubMed] [Google Scholar]
- Minton A. P. Models for Excluded Volume Interaction Between an Unfolded Protein and Rigid Macromolecular Cosolutes: Macromolecular Crowding and Protein Stability Revisited. Biophys. J. 2005, 88, 971–985. 10.1529/biophysj.104.050351. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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]
- Cheung M. S.; Klimov D.; Thirumalai D. Molecular Crowding Enhances Native State Stability and Refolding Rates of Globular Proteins. Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 4753–4758. 10.1073/pnas.0409630102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stagg L.; Zhang S. Q.; Cheung M. S.; Wittung-Stafshede P. Molecular Crowding Enhances Native Structure and Stability of α/β Protein Flavodoxin. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 18976–18981. 10.1073/pnas.0705127104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mukherjee S. K.; Gautam S.; Biswas S.; Kundu J.; Chowdhury P. K. Do Macromolecular Crowding Agents Exert only an Excluded Volume Effect? A Protein Solvation Study. J. Phys. Chem. B 2015, 119, 14145–14156. 10.1021/acs.jpcb.5b09446. [DOI] [PubMed] [Google Scholar]
- Luh L. M.; Hänsel R.; Löhr F.; Kirchner D. K.; Krauskopf K.; Pitzius S.; Schäfer B.; Tufar P.; Corbeski I.; Güntert P.; et al. Molecular Crowding Drives Active Pin1 into Nonspecific Complexes with Endogenous Proteins Prior to Substrate Recognition. J. Am. Chem. Soc. 2013, 135, 13796–13803. 10.1021/ja405244v. [DOI] [PubMed] [Google Scholar]
- Senske M.; Tork L.; Born B.; Havenith M.; Herrmann C.; Ebbinghaus S. Protein Stabilization by Macromolecular Crowding through Enthalpy rather than Entropy. J. Am. Chem. Soc. 2014, 136, 9036–9041. 10.1021/ja503205y. [DOI] [PubMed] [Google Scholar]
- Cohen R. D.; Pielak G. J. Electrostatic Contributions to Protein Quinary Structure. J. Am. Chem. Soc. 2016, 138, 13139–13142. 10.1021/jacs.6b07323. [DOI] [PubMed] [Google Scholar]
- Monteith W. B.; Cohen R. D.; Smith A. E.; Guzman-Cisneros E.; Pielak G. J. Quinary Structure Modulates Protein Stability in Cells. Proc. Natl. Acad. Sci. U. S. A. 2015, 112, 1739–1742. 10.1073/pnas.1417415112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sarkar M.; Pielak G. J. An Osmolyte Mitigates the Destabilizing Effect of Protein Crowding. Protein Sci. 2014, 23, 1161–1164. 10.1002/pro.2510. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sarkar M.; Lu J.; Pielak G. J. Protein Crowder Charge and Protein Stability. Biochemistry 2014, 53, 1601–1606. 10.1021/bi4016346. [DOI] [PubMed] [Google Scholar]
- Wang Y.; Sarkar M.; Smith A. E.; Krois A. S.; Pielak G. J. Macromolecular Crowding and Protein Stability. J. Am. Chem. Soc. 2012, 134, 16614–16618. 10.1021/ja305300m. [DOI] [PubMed] [Google Scholar]
- Benton L. A.; Smith A. E.; Young G. B.; Pielak G. J. Unexpected Effects of Macromolecular Crowding on Protein Stability. Biochemistry 2012, 51, 9773–9775. 10.1021/bi300909q. [DOI] [PubMed] [Google Scholar]
- Miklos A. C.; Sarkar M.; Wang Y.; Pielak G. J. Protein Crowding Tunes Protein Stability. J. Am. Chem. Soc. 2011, 133, 7116–7120. 10.1021/ja200067p. [DOI] [PubMed] [Google Scholar]
- Inomata K.; Ohno A.; Tochio H.; Isogai S.; Tenno T.; Nakase I.; Takeuchi T.; Futaki S.; Ito Y.; Hiroaki H.; et al. High-Resolution Multi-Dimensional NMR Spectroscopy of Proteins in Human Cells. Nature 2009, 458, 106–111. 10.1038/nature07839. [DOI] [PubMed] [Google Scholar]
- Sakakibara D.; Sasaki A.; Ikeya T.; Hamatsu J.; Hanashima T.; Mishima M.; Yoshimasu M.; Hayashi N.; Mikawa T.; Walchli M.; et al. Protein Structure Determination in Living Cells by In-Cell NMR Spectroscopy. Nature 2009, 458, 102–105. 10.1038/nature07814. [DOI] [PubMed] [Google Scholar]
- 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]
- Harada R.; Sugita Y.; Feig M. Protein Crowding Affects Hydration Structure and Dynamics. J. Am. Chem. Soc. 2012, 134, 4842–4849. 10.1021/ja211115q. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tanizaki S.; Clifford J.; Connelly B. D.; Feig M. Conformational Sampling of Peptides in Cellular Environments. Biophys. J. 2008, 94, 747–759. 10.1529/biophysj.107.116236. [DOI] [PMC free article] [PubMed] [Google Scholar]
- King J. T.; Arthur E. J.; Brooks C. L.; Kubarych K. J. Crowding Induced Collective Hydration of Biological Macromolecules over Extended Distances. J. Am. Chem. Soc. 2014, 136, 188–194. 10.1021/ja407858c. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lu C.; Prada-Gracia D.; Rao F. Structure and Dynamics of Water in Crowded Environments Slows Down Peptide Conformational Changes. J. Chem. Phys. 2014, 141, 045101. 10.1063/1.4891465. [DOI] [PubMed] [Google Scholar]
- Qi H. W.; Nakka P.; Chen C.; Radhakrishnan M. L. The Effect of Macromolecular Crowding on the Electrostatic Component of Barnase–Barstar Binding: A Computational, Implicit Solvent-Based Study. PLoS One 2014, 9, e98618. 10.1371/journal.pone.0098618. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim Y. C.; Mittal J. Crowding Induced Entropy-Enthalpy Compensation in Protein Association Equilibria. Phys. Rev. Lett. 2013, 110, 208102. 10.1103/PhysRevLett.110.208102. [DOI] [PubMed] [Google Scholar]
- Zhang N.; An L.; Li J.; Liu Z.; Yao L. Quinary Interactions Weaken the Electric Field Generated by Protein Side-Chain Charges in the Cell-like Environment. J. Am. Chem. Soc. 2017, 139, 647–654. 10.1021/jacs.6b11058. [DOI] [PubMed] [Google Scholar]
- Ando T.; Yu I.; Feig M.; Sugita Y. Thermodynamics of Macromolecular Association in Heterogeneous Crowding Environments: Theoretical and Simulation Studies with a Simplified Model. J. Phys. Chem. B 2016, 120, 11856–11865. 10.1021/acs.jpcb.6b06243. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kondrat S.; Zimmermann O.; Wiechert W.; von Lieres E. The Effect of Composition on Diffusion of Macromolecules in a Crowded Environment. Phys. Biol. 2015, 12, 046003. 10.1088/1478-3975/12/4/046003. [DOI] [PubMed] [Google Scholar]
- Ellis R. J.; Minton A. P. Cell Biology - Join the Crowd. Nature 2003, 425, 27–28. 10.1038/425027a. [DOI] [PubMed] [Google Scholar]
- Gershenson A. Deciphering Protein Stability in Cells. J. Mol. Biol. 2014, 426, 4–6. 10.1016/j.jmb.2013.10.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cedeño C.; Pauwels K.; Tompa P. Protein Delivery into Plant Cells: Toward in vivo Structural Biology. Front. Plant Sci. 2017, 8, 519. 10.3389/fpls.2017.00519. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pastore A.; Temussi P. A.. The Emperor’s New Clothes: Myths and Truths of In-Cell NMR. Arch. Biochem. Biophys. 2017, in press. DOI: 10.1016/j.abb.2017.02.008. [DOI] [PubMed] [Google Scholar]
- Hänsel R.; Luh L. M.; Corbeski I.; Trantirek L.; Dötsch V. In-Cell NMR and EPR Spectroscopy of Biomacromolecules. Angew. Chem., Int. Ed. 2014, 53, 10300–10314. 10.1002/anie.201311320. [DOI] [PubMed] [Google Scholar]
- Ikeya T.; Hanashima T.; Hosoya S.; Shimazaki M.; Ikeda S.; Mishima M.; Güntert P.; Ito Y. Improved In-Cell Structure Determination of Proteins at Near-Physiological Concentration. Sci. Rep. 2016, 6, 38312. 10.1038/srep38312. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li C.; Liu M. Protein Dynamics in Living Cells Studied by In-Cell NMR Spectroscopy. FEBS Lett. 2013, 587, 1008–1011. 10.1016/j.febslet.2012.12.023. [DOI] [PubMed] [Google Scholar]
- Luchinat E.; Banci L. In-Cell NMR: A Topical Review. IUCrJ 2017, 4, 108–118. 10.1107/S2052252516020625. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Majumder S.; Xue J.; DeMott C. M.; Reverdatto S.; Burz D. S.; Shekhtman A. Probing Protein Quinary Interactions by In-Cell Nuclear Magnetic Resonance Spectroscopy. Biochemistry 2015, 54, 2727–2738. 10.1021/acs.biochem.5b00036. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Müntener T.; Häussinger D.; Selenko P.; Theillet F.-X. In-Cell Protein Structures from 2D NMR Experiments. J. Phys. Chem. Lett. 2016, 7, 2821–2825. 10.1021/acs.jpclett.6b01074. [DOI] [PubMed] [Google Scholar]
- Ohno A.; Inomata K.; Tochio H.; Shirakawa M. In-Cell NMR Spectroscopy in Protein Chemistry and Drug Discovery. Curr. Top. Med. Chem. 2011, 11, 68–73. 10.2174/156802611793611878. [DOI] [PubMed] [Google Scholar]
- Rahman S.; Byun Y.; Hassan M. I.; Kim J.; Kumar V. Towards Understanding Cellular Structure Biology: In-cell NMR. Biochim. Biophys. Acta, Proteins Proteomics 2017, 1865, 547–557. 10.1016/j.bbapap.2017.02.018. [DOI] [PubMed] [Google Scholar]
- Sharaf N. G.; Barnes C. O.; Charlton L. M.; Young G. B.; Pielak G. J. A Bioreactor for In-Cell Protein NMR. J. Magn. Reson. 2010, 202, 140–146. 10.1016/j.jmr.2009.10.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smith A. E.; Zhou L. Z.; Gorensek A. H.; Senske M.; Pielak G. J. In-Cell Thermodynamics and a New Role for Protein Surfaces. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, 1725–1730. 10.1073/pnas.1518620113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Danielsson J.; Mu X.; Lang L.; Wang H.; Binolfi A.; Theillet F.-X.; Bekei B.; Logan D. T.; Selenko P.; Wennerström H.; et al. Thermodynamics of Protein Destabilization in Live Cells. Proc. Natl. Acad. Sci. U. S. A. 2015, 112, 12402–12407. 10.1073/pnas.1511308112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smith M. J.; Marshall C. B.; Theillet F.-X.; Binolfi A.; Selenko P.; Ikura M. Real-Time NMR Monitoring of Biological Activities in Complex Physiological Environments. Curr. Opin. Struct. Biol. 2015, 32, 39–47. 10.1016/j.sbi.2015.02.003. [DOI] [PubMed] [Google Scholar]
- Li C.; Zhao J.; Cheng K.; Ge Y.; Wu Q.; Ye Y.; Xu G.; Zhang Z.; Zheng W.; Zhang X.; et al. Magnetic Resonance Spectroscopy As a Tool for Assessing Macromolecular Structure and Function in Living Cells. Annu. Rev. Anal. Chem. 2017, 10, 157–182. 10.1146/annurev-anchem-061516-045237. [DOI] [PubMed] [Google Scholar]
- Smith A. E.; Zhang Z.; Pielak G. J.; Li C. NMR Studies of Protein Folding and Binding in Cells and Cell-Like Environments. Curr. Opin. Struct. Biol. 2015, 30, 7–16. 10.1016/j.sbi.2014.10.004. [DOI] [PubMed] [Google Scholar]
- Xu G.; Ye Y.; Liu X.; Cao S.; Wu Q.; Cheng K.; Liu M.; Pielak G. J.; Li C. Strategies for Protein NMR in Escherichia coli. Biochemistry 2014, 53, 1971–1981. 10.1021/bi500079u. [DOI] [PubMed] [Google Scholar]
- Kaplan M.; Cukkemane A.; van Zundert G. C. P.; Narasimhan S.; Daniels M.; Mance D.; Waksman G.; Bonvin A. M. J. J.; Fronzes R.; Folkers G. E.; et al. Probing a Cell-Embedded Megadalton Protein Complex by DNP-Supported Solid-State NMR. Nat. Methods 2015, 12, 649–652. 10.1038/nmeth.3406. [DOI] [PubMed] [Google Scholar]
- Kay L. E. New Views of Functionally Dynamic Proteins by Solution NMR Spectroscopy. J. Mol. Biol. 2016, 428, 323–331. 10.1016/j.jmb.2015.11.028. [DOI] [PubMed] [Google Scholar]
- König I.; Zarrine-Afsar A.; Aznauryan M.; Soranno A.; Wunderlich B.; Dingfelder F.; Stuber J. C.; Plückthun A.; Nettels D.; Schuler B. Single-Molecule Spectroscopy of Protein Conformational Dynamics in Live Eukaryotic Cells. Nat. Methods 2015, 12, 773–779. 10.1038/nmeth.3475. [DOI] [PubMed] [Google Scholar]
- Sustarsic M.; Kapanidis A. N. Taking the Ruler to the Jungle: Single-Molecule FRET for Understanding Biomolecular Structure and Dynamics in Live Cells. Curr. Opin. Struct. Biol. 2015, 34, 52–59. 10.1016/j.sbi.2015.07.001. [DOI] [PubMed] [Google Scholar]
- Gelman H.; Wirth A. J.; Gruebele M. ReAsH as a Quantitative Probe of In-Cell Protein Dynamics. Biochemistry 2016, 55, 1968–1976. 10.1021/acs.biochem.5b01336. [DOI] [PubMed] [Google Scholar]
- Feig M.; Sugita Y. Reaching New Levels of Realism in Modeling Biological Macromolecules in Cellular Environments. J. Mol. Graphics Modell. 2013, 45, 144–156. 10.1016/j.jmgm.2013.08.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Perilla J. R.; Goh B. C.; Cassidy C. K.; Liu B.; Bernardi R. C.; Rudack T.; Yu H.; Wu Z.; Schulten K. Molecular Dynamics Simulations of Large Macromolecular Complexes. Curr. Opin. Struct. Biol. 2015, 31, 64–74. 10.1016/j.sbi.2015.03.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Feig M.; Chocholousova J.; Tanizaki S. Extending the Horizon: Towards the Efficient Modeling of Large Biomolecular Complexes in Atomic Detail. Theor. Chem. Acc. 2006, 116, 194–205. 10.1007/s00214-005-0062-4. [DOI] [Google Scholar]
- Hospital A.; Goñi J. R.; Orozco M.; Gelpi J. L. Molecular Dynamics Simulations: Advances and Applications. Adv. Appl. Bioinform. Chem. 2015, 8, 37–47. 10.2147/AABC.S70333. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gavin A. C.; Aloy P.; Grandi P.; Krause R.; Boesche M.; Marzioch M.; Rau C.; Jensen L. J.; Bastuck S.; Dumpelfeld B.; et al. Proteome Survey Reveals Modularity of the Yeast Cell Machinery. Nature 2006, 440, 631–636. 10.1038/nature04532. [DOI] [PubMed] [Google Scholar]
- Guell M.; van Noort V.; Yus E.; Chen W. H.; Leigh-Bell J.; Michalodimitrakis K.; Yamada T.; Arumugam M.; Doerks T.; Kuhner S.; et al. Transcriptome Complexity in a Genome-Reduced Bacterium. Science 2009, 326, 1268–1271. 10.1126/science.1176951. [DOI] [PubMed] [Google Scholar]
- Yus E.; Maier T.; Michalodimitrakis K.; van Noort V.; Yamada T.; Chen W. H.; Wodke J. A. H.; Guell M.; Martinez S.; Bourgeois R.; et al. Impact of Genome Reduction on Bacterial Metabolism and Its Regulation. Science 2009, 326, 1263–1268. 10.1126/science.1177263. [DOI] [PubMed] [Google Scholar]
- Ye Y.; Liu X.; Zhang Z.; Wu Q.; Jiang B.; Jiang L.; Zhang X.; Liu M.; Pielak G. J.; Li C. 19F NMR Spectroscopy as a Probe of Cytoplasmic Viscosity and Weak Protein Interactions in Living Cells. Chem. - Eur. J. 2013, 19, 12705–12710. 10.1002/chem.201301657. [DOI] [PubMed] [Google Scholar]
- Sarkar M.; Smith A. E.; Pielak G. J. Impact of Reconstituted Cytosol on Protein Stability. Proc. Natl. Acad. Sci. U. S. A. 2013, 110, 19342–19347. 10.1073/pnas.1312678110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guseman A. J.; Pielak G. J. Cosolute and Crowding Effects on a Side-By-Side Protein Dimer. Biochemistry 2017, 56, 971–976. 10.1021/acs.biochem.6b01251. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bai J.; Liu M.; Pielak G. J.; Li C. Macromolecular and Small Molecular Crowding Have Similar Effects on α-Synuclein Structure. ChemPhysChem 2017, 18, 55–58. 10.1002/cphc.201601097. [DOI] [PubMed] [Google Scholar]
- Cohen R. D.; Guseman A. J.; Pielak G. J. Intracellular pH Modulates Quinary Structure. Protein Sci. 2015, 24, 1748–1755. 10.1002/pro.2765. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Monteith W. B.; Pielak G. J. Residue Level Quantification of Protein Stability in Living Cells. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, 11335–11340. 10.1073/pnas.1406845111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sarkar M.; Li C.; Pielak G. J. Soft Interactions and Crowding. Biophys. Rev. 2013, 5, 187–194. 10.1007/s12551-013-0104-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schlesinger A. P.; Wang Y.; Tadeo X.; Millet O.; Pielak G. J. Macromolecular Crowding Fails To Fold a Globular Protein in Cells. J. Am. Chem. Soc. 2011, 133, 8082–8085. 10.1021/ja201206t. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stadmiller S. S.; Gorensek-Benitez A. H.; Guseman A. J.; Pielak G. J. Osmotic Shock Induced Protein Destabilization in Living Cells and Its Reversal by Glycine Betaine. J. Mol. Biol. 2017, 429, 1155–1161. 10.1016/j.jmb.2017.03.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Theillet F.-X.; Binolfi A.; Bekei B.; Martorana A.; Rose H. M.; Stuiver M.; Verzini S.; Lorenz D.; van Rossum M.; Goldfarb D.; et al. Structural Disorder of Monomeric α-Synuclein Persists in Mammalian Cells. Nature 2016, 530, 45–50. 10.1038/nature16531. [DOI] [PubMed] [Google Scholar]
- Soranno A.; Koenig I.; Borgia M. B.; Hofmann H.; Zosel F.; Nettels D.; Schuler B. 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]
- Anunciado D. B.; Nyugen V. P.; Hurst G. B.; Doktycz M. J.; Urban V.; Langan P.; Mamontov E.; O’Neill H. In vivo Protein Dynamics on the Nanometer Length Scale and Nanosecond Time Scale. J. Phys. Chem. Lett. 2017, 8, 1899–1904. 10.1021/acs.jpclett.7b00399. [DOI] [PubMed] [Google Scholar]
- Latham M. P.; Kay L. E. Is Buffer a Good Proxy for a Crowded Cell-Like Environment? A Comparative NMR Study of Calmodulin Side-Chain Dynamics in Buffer and E. coli Lysate. PLoS One 2012, 7, e48226. 10.1371/journal.pone.0048226. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Latham M. P.; Kay L. E. Probing Non-Specific Interactions of Ca2+-Calmodulin in E. coli Lysate. J. Biomol. NMR 2013, 55, 239–247. 10.1007/s10858-013-9705-2. [DOI] [PubMed] [Google Scholar]
- Latham M. P.; Kay L. E. A Similar In Vitro and In Cell Lysate Folding Intermediate for the FF Domain. J. Mol. Biol. 2014, 426, 3214–3220. 10.1016/j.jmb.2014.07.019. [DOI] [PubMed] [Google Scholar]
- Roos M.; Ott M.; Hofmann M.; Link S.; Rössler E.; Balbach J.; Kruschelnitsky A.; Saalwächter K. Coupling and Decoupling of Rotational and Translational Diffusion of Proteins under Crowding Conditions. J. Am. Chem. Soc. 2016, 138, 10365–10372. 10.1021/jacs.6b06615. [DOI] [PubMed] [Google Scholar]
- Tai J.; Dave K.; Hahn V.; Guzman I.; Gruebele M. Subcellular Modulation of Protein VlsE Stability and Folding Kinetics. FEBS Lett. 2016, 590, 1409–1416. 10.1002/1873-3468.12193. [DOI] [PubMed] [Google Scholar]
- Cino E. A.; Karttunen M.; Choy W.-Y. Effects of Molecular Crowding on the Dynamics of Intrinsically Disordered Proteins. PLoS One 2012, 7, e49876. 10.1371/journal.pone.0049876. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Munari F.; Bortot A.; Zanzoni S.; D’Onofrio M.; Fushman D.; Assfalg M. Identification of Primary and Secondary UBA Footprints on the Surface of Ubiquitin in Cell-Mimicking Crowded Solution. FEBS Lett. 2017, 591, 979–990. 10.1002/1873-3468.12615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ito Y.; Selenko P. Cellular Structural Biology. Curr. Opin. Struct. Biol. 2010, 20, 640–648. 10.1016/j.sbi.2010.07.006. [DOI] [PubMed] [Google Scholar]
- Elcock A. H. Models of Macromolecular Crowding Effects and the Need for Quantitative Comparisons with Experiment. Curr. Opin. Struct. Biol. 2010, 20, 196–206. 10.1016/j.sbi.2010.01.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Danielsson J.; Oliveberg M. Comparing Protein Behaviour in vitro and in vivo, What Does the Data Really Tell Us?. Curr. Opin. Struct. Biol. 2017, 42, 129–135. 10.1016/j.sbi.2017.01.002. [DOI] [PubMed] [Google Scholar]
- Roberts E. Cellular and Molecular Structure as a Unifying Framework for Whole-Cell Modeling. Curr. Opin. Struct. Biol. 2014, 25, 86–91. 10.1016/j.sbi.2014.01.005. [DOI] [PubMed] [Google Scholar]
- Im W.; Liang J.; Olson A.; Zhou H.-X.; Vajda S.; Vakser I. A. Challenges in Structural Approaches to Cell Modeling. J. Mol. Biol. 2016, 428, 2943–2964. 10.1016/j.jmb.2016.05.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kyne C.; Crowley P. B. Grasping the Nature of the Cell Interior: From Physiological Chemistry to Chemical Biology. FEBS J. 2016, 283, 3016–3028. 10.1111/febs.13744. [DOI] [PubMed] [Google Scholar]
- Zhou H.-X. Polymer Crowders and Protein Crowders Act Similarly on Protein Folding Stability. FEBS Lett. 2013, 587, 394–397. 10.1016/j.febslet.2013.01.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Miklos A. C.; Li C.; Sharaf N. G.; Pielak G. J. Volume Exclusion and Soft Interaction Effects on Protein Stability under Crowded Conditions. Biochemistry 2010, 49, 6984–6991. 10.1021/bi100727y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ignatova Z.; Krishnan B.; Bombardier J. P.; Marcelino A. M. C.; Hong J.; Gierasch L. M. From the Test Tube to the Cell: Exploring the Folding and Aggregation of a β-Clam Protein. Biopolymers 2007, 88, 157–163. 10.1002/bip.20665. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hong J.; Gierasch L. M. Macromolecular Crowding Remodels the Energy Landscape of a Protein by Favoring a More Compact Unfolded State. J. Am. Chem. Soc. 2010, 132, 10445–10452. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Macdonald B.; McCarley S.; Noeen S.; van Giessen A. E. Protein–Protein Interactions Affect α-Helix Stability in Crowded Environments. J. Phys. Chem. B 2015, 119, 2956–2967. 10.1021/jp512630s. [DOI] [PubMed] [Google Scholar]
- Macdonald B.; McCarley S.; Noeen S.; van Giessen A. E. β-Hairpin Crowding Agents Affect α-Helix Stability in Crowded Environments. J. Phys. Chem. B 2016, 120, 650–659. 10.1021/acs.jpcb.5b10575. [DOI] [PubMed] [Google Scholar]
- Bille A.; Mohanty S.; Irbäck A. Peptide Folding in the Presence of Interacting Protein Crowders. J. Chem. Phys. 2016, 144, 175105. 10.1063/1.4948462. [DOI] [PubMed] [Google Scholar]
- Bille A.; Linse B.; Mohanty S.; Irbäck A. Equilibrium Simulation of Trp-Cage in the Presence of Protein Crowders. J. Chem. Phys. 2015, 143, 175102. 10.1063/1.4934997. [DOI] [PubMed] [Google Scholar]
- Kadumuri R. V.; Gullipalli J.; Subramanian S.; Jaipuria G.; Atreya H. S.; Vadrevu R. Crowding Interactions Perturb Structure and Stability by Destabilizing the Stable Core of the α-Subunit of Tryptophan Synthase. FEBS Lett. 2016, 590, 2096–2105. 10.1002/1873-3468.12259. [DOI] [PubMed] [Google Scholar]
- Qin S.; Zhou H.-X. Effects of Macromolecular Crowding on the Conformational Ensembles of Disordered Proteins. J. Phys. Chem. Lett. 2013, 4, 3429–3434. 10.1021/jz401817x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Goldenberg D. P.; Argyle B. Minimal Effects of Macromolecular Crowding on an Intrinsically Disordered Protein: A Small-Angle Neutron Scattering Study. Biophys. J. 2014, 106, 905–914. 10.1016/j.bpj.2013.12.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nakano S.-i.; Sugimoto N. Model Studies of the Effects of Intracellular Crowding on Nucleic Acid Interactions. Mol. BioSyst. 2017, 13, 32–41. 10.1039/C6MB00654J. [DOI] [PubMed] [Google Scholar]
- Yildirim A.; Sharma M.; Varner B. M.; Fang L.; Feig M. Conformational Preferences of DNA in Reduced Dielectric Environments. J. Phys. Chem. B 2014, 118, 10874–10881. 10.1021/jp505727w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nakano S.-i.; Miyoshi D.; Sugimoto N. Effects of Molecular Crowding on the Structures, Interactions, and Functions of Nucleic Acids. Chem. Rev. 2014, 114, 2733–2758. 10.1021/cr400113m. [DOI] [PubMed] [Google Scholar]
- Gao M.; Gnutt D.; Orban A.; Appel B.; Righetti F.; Winter R.; Narberhaus F.; Müller S.; Ebbinghaus S. RNA Hairpin Folding in the Crowded Cell. Angew. Chem., Int. Ed. 2016, 55, 3224–3228. 10.1002/anie.201510847. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Baltierra-Jasso L. E.; Morten M. J.; Laflör L.; Quinn S. D.; Magennis S. W. Crowding-Induced Hybridization of Single DNA Hairpins. J. Am. Chem. Soc. 2015, 137, 16020–16023. 10.1021/jacs.5b11829. [DOI] [PubMed] [Google Scholar]
- Kilburn D.; Roh J. H.; Guo L.; Briber R. M.; Woodson S. A. Molecular Crowding Stabilizes Folded RNA Structure by the Excluded Volume Effect. J. Am. Chem. Soc. 2010, 132, 8690–8696. 10.1021/ja101500g. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tyrrell J.; McGinnis J. L.; Weeks K. M.; Pielak G. J. The Cellular Environment Stabilizes Adenine Riboswitch RNA Structure. Biochemistry 2013, 52, 8777–8785. 10.1021/bi401207q. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tyrrell J.; Weeks K. M.; Pielak G. J. Challenge of Mimicking the Influences of the Cellular Environment on RNA Structure by PEG-Induced Macromolecular Crowding. Biochemistry 2015, 54, 6447–6453. 10.1021/acs.biochem.5b00767. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bryant J. E.; Lecomte J. T. J.; Lee A. L.; Young G. B.; Pielak G. J. Protein Dynamics in Living Cells. Biochemistry 2005, 44, 9275–9279. 10.1021/bi050786j. [DOI] [PubMed] [Google Scholar]
- Bryant J. E.; Lecomte J. T. J.; Lee A. L.; Young G. B.; Pielak G. J. Cytosol Has a Small Effect on Protein Backbone Dynamics. Biochemistry 2006, 45, 10085–10091. 10.1021/bi060547b. [DOI] [PubMed] [Google Scholar]
- Abriata L.; Spiga E.; Dal Peraro M. All-Atom Simulations of Crowding Effects on Ubiquitin Dynamics. Phys. Biol. 2013, 10, 045006. 10.1088/1478-3975/10/4/045006. [DOI] [PubMed] [Google Scholar]
- Minh D. D. L.; Chang C.-e.; Trylska J.; Tozzini V.; McCammon J. A. The Influence of Macromolecular Crowding on HIV-1 Protease Internal Dynamics. J. Am. Chem. Soc. 2006, 128, 6006–6007. 10.1021/ja060483s. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Makowski L.; Rodi D. J.; Mandava S.; Minh D. D. L.; Gore D. B.; Fischetti R. F. Molecular Crowding Inhibits Intramolecular Breathing Motions in Proteins. J. Mol. Biol. 2008, 375, 529–546. 10.1016/j.jmb.2007.07.075. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li C.; Charlton L. M.; Lakkavaram A.; Seagle C.; Wang G.; Young G. B.; Macdonald J. M.; Pielak G. J. Differential Dynamical Effects of Macromolecular Crowding on an Intrinsically Disordered Protein and a Globular Protein: Implications for In-Cell NMR Spectroscopy. J. Am. Chem. Soc. 2008, 130, 6310–6311. 10.1021/ja801020z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schreiber G. Kinetic Studies of Protein–Protein Interactions. Curr. Opin. Struct. Biol. 2002, 12, 41–47. 10.1016/S0959-440X(02)00287-7. [DOI] [PubMed] [Google Scholar]
- Northrup S. H.; Erickson H. P. Kinetics of Protein-Protein Association Explained by Brownian Dynamics Computer Simulation. Proc. Natl. Acad. Sci. U. S. A. 1992, 89, 3338–3342. 10.1073/pnas.89.8.3338. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhou H. X. Brownian Dynamics Study of the Influences of Electrostatic Interaction and Diffusion on Protein-Protein Association Kinetics. Biophys. J. 1993, 64, 1711–1726. 10.1016/S0006-3495(93)81543-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Slutsky M.; Mirny L. A. Kinetics of Protein-DNA Interaction: Facilitated Target Location in Sequence-Dependent Potential. Biophys. J. 2004, 87, 4021–4035. 10.1529/biophysj.104.050765. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Berg O. G.; Winter R. B.; Von Hippel P. H. Diffusion-Driven Mechanisms of Protein Translocation on Nucleic Acids. 1. Models and Theory. Biochemistry 1981, 20, 6929–6948. 10.1021/bi00527a028. [DOI] [PubMed] [Google Scholar]
- Li C.; Wang Y.; Pielak G. J. Translational and Rotational Diffusion of a Small Globular Protein under Crowded Conditions. J. Phys. Chem. B 2009, 113, 13390–13392. 10.1021/jp907744m. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Y.; Benton L. A.; Singh V.; Pielak G. J. Disordered Protein Diffusion under Crowded Conditions. J. Phys. Chem. Lett. 2012, 3, 2703–2706. 10.1021/jz3010915. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Y.; Li C.; Pielak G. J. Effects of Proteins on Protein Diffusion. J. Am. Chem. Soc. 2010, 132, 9392–9397. 10.1021/ja102296k. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li C.; Pielak G. J. Using NMR to Distinguish Viscosity Effects from Nonspecific Protein Binding under Crowded Conditions. J. Am. Chem. Soc. 2009, 131, 1368–1369. 10.1021/ja808428d. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Geyer T. Mixing Normal and Anomalous Diffusion. J. Chem. Phys. 2012, 137, 115101. 10.1063/1.4753804. [DOI] [PubMed] [Google Scholar]
- Banks D. S.; Fradin C. Anomalous Diffusion of Proteins due to Molecular Crowding. Biophys. J. 2005, 89, 2960–2971. 10.1529/biophysj.104.051078. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hasnain S.; McClendon C. L.; Hsu M. T.; Jacobson M. P.; Bandyopadhyay P. A New Coarse-Grained Model for E. coli Cytoplasm: Accurate Calculation of the Diffusion Coefficient of Proteins and Observation of Anomalous Diffusion. PLoS One 2014, 9, e106466. 10.1371/journal.pone.0106466. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Galanti M.; Fanelli D.; Maritan A.; Piazza F. Diffusion of Tagged Particles in a Crowded Medium. Europhys. Lett. 2014, 107, 20006. 10.1209/0295-5075/107/20006. [DOI] [Google Scholar]
- Wilcox A. E.; LoConte M. A.; Slade K. M. Effects of Macromolecular Crowding on Alcohol Dehydrogenase Activity Are Substrate-Dependent. Biochemistry 2016, 55, 3550–3558. 10.1021/acs.biochem.6b00257. [DOI] [PubMed] [Google Scholar]
- Johansson H.; Jensen M. R.; Gesmar H.; Meier S.; Vinther J. M.; Keeler C.; Hodsdon M. E.; Led J. J. Specific and Nonspecific Interactions in Ultraweak Protein–Protein Associations Revealed by Solvent Paramagnetic Relaxation Enhancements. J. Am. Chem. Soc. 2014, 136, 10277–10286. 10.1021/ja503546j. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cohen R. D.; Pielak G. J. A Cell is More than the Sum of Its (Dilute) Parts: A Brief History of Quinary Structure. Protein Sci. 2017, 26, 403–413. 10.1002/pro.3092. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pérez Santero S.; Favretto F.; Zanzoni S.; Chignola R.; Assfalg M.; D’Onofrio M. Effects of Macromolecular Crowding on a Small Lipid Binding Protein Probed at the Single-Amino Acid Level. Arch. Biochem. Biophys. 2016, 606, 99–110. 10.1016/j.abb.2016.07.017. [DOI] [PubMed] [Google Scholar]
- Kim Y. C.; Bhattacharya A.; Mittal J. Macromolecular Crowding Effects on Coupled Folding and Binding. J. Phys. Chem. B 2014, 118, 12621–12629. 10.1021/jp508046y. [DOI] [PubMed] [Google Scholar]
- Kim Y. C.; Best R. B.; Mittal J. Macromolecular Crowding Effects on Protein-Protein Binding Affinity and Specificity. J. Chem. Phys. 2010, 133, 205101. 10.1063/1.3516589. [DOI] [PubMed] [Google Scholar]
- Qin S. B.; Cai L.; Zhou H. X. A Method for Computing Association Rate Constants of Atomistically Represented Proteins under Macromolecular Crowding. Phys. Biol. 2012, 9, 066008. 10.1088/1478-3975/9/6/066008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Musiani F.; Giorgetti A. Protein Aggregation and Molecular Crowding: Perspectives from Multiscale Simulations. Int. Rev. Cell Mol. Biol. 2017, 329, 49–77. 10.1016/bs.ircmb.2016.08.009. [DOI] [PubMed] [Google Scholar]
- White D. A.; Buell A. K.; Knowles T. P. J.; Welland M. E.; Dobson C. M. Protein Aggregation in Crowded Environments. J. Am. Chem. Soc. 2010, 132, 5170–5175. 10.1021/ja909997e. [DOI] [PubMed] [Google Scholar]
- Ellis R. J.; Minton A. P. Protein Aggregation in Crowded Environments. Biol. Chem. 2006, 387, 485–497. 10.1515/BC.2006.064. [DOI] [PubMed] [Google Scholar]
- Qin S.; Zhou H.-X. Protein Folding, Binding, and Droplet Formation in Cell-Like Conditions. Curr. Opin. Struct. Biol. 2017, 43, 28–37. 10.1016/j.sbi.2016.10.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Spitzer J.; Poolman B. How Crowded is the Prokaryotic Cytoplasm?. FEBS Lett. 2013, 587, 2094–2098. 10.1016/j.febslet.2013.05.051. [DOI] [PubMed] [Google Scholar]
- Duff M. R. Jr.; Grubbs J.; Serpersu E.; Howell E. E. Weak Interactions between Folate and Osmolytes in Solution. Biochemistry 2012, 51, 2309–2318. 10.1021/bi3000947. [DOI] [PubMed] [Google Scholar]
- Miklos A. C.; Sumpter M.; Zhou H.-X. Competitive Interactions of Ligands and Macromolecular Crowders with Maltose Binding Protein. PLoS One 2013, 8, e74969. 10.1371/journal.pone.0074969. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Abriata L. A.; Spiga E.; Dal Peraro M. Molecular Effects of Concentrated Solutes on Protein Hydration, Dynamics, and Electrostatics. Biophys. J. 2016, 111, 743–755. 10.1016/j.bpj.2016.07.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heinen M.; Zanini F.; Roosen-Runge F.; Fedunova D.; Zhang F. J.; Hennig M.; Seydel T.; Schweins R.; Sztucki M.; Antalik M.; et al. Viscosity and Diffusion: Crowding and Salt Effects in Protein Solutions. Soft Matter 2012, 8, 1404–1419. 10.1039/C1SM06242E. [DOI] [Google Scholar]
- Szymanski J.; Patkowski A.; Wilk A.; Garstecki P.; Holyst R. Diffusion and Viscosity in a Crowded Environment: From Nano- to Macroscale. J. Phys. Chem. B 2006, 110, 25593–25597. 10.1021/jp0666784. [DOI] [PubMed] [Google Scholar]
- Despa F.; Fernandez A.; Berry R. S. Dielectric Modulation of Biological Water. Phys. Rev. Lett. 2004, 93, 228104. 10.1103/PhysRevLett.93.228104. [DOI] [PubMed] [Google Scholar]
- Huang K.-Y.; Kingsley C. N.; Sheil R.; Cheng C.-Y.; Bierma J. C.; Roskamp K. W.; Khago D.; Martin R. W.; Han S. Stability of Protein-Specific Hydration Shell on Crowding. J. Am. Chem. Soc. 2016, 138, 5392–5402. 10.1021/jacs.6b01989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guigas G.; Weiss M. Effects of Protein Crowding on Membrane Systems. Biochim. Biophys. Acta, Biomembr. 2016, 1858, 2441–2450. 10.1016/j.bbamem.2015.12.021. [DOI] [PubMed] [Google Scholar]
- Wang Q.; Cheung; Margaret S. A Physics-Based Approach of Coarse-Graining the Cytoplasm of Escherichia coli (CGCYTO). Biophys. J. 2012, 102, 2353–2361. 10.1016/j.bpj.2012.04.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sterpone F.; Derreumaux P.; Melchionna S. Protein Simulations in Fluids: Coupling the OPEP Coarse-Grained Force Field with Hydrodynamics. J. Chem. Theory Comput. 2015, 11, 1843–1853. 10.1021/ct501015h. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Blanco P.; Via M.; Garcés J.; Madurga S.; Mas F. Brownian Dynamics Computational Model of Protein Diffusion in Crowded Media with Dextran Macromolecules as Obstacles. Entropy 2017, 19, 105. 10.3390/e19030105. [DOI] [Google Scholar]
- Feig M.; Tanizaki S.; Sayadi M.. Implicit Solvent Simulations of Biomolecules in Cellular Environments. Annual Reviews in Computational Chemistry; Elsevier: Oxford, U.K., 2008; Vol. 4, pp 107–121. [Google Scholar]
- Starzyk A.; Wojciechowski M.; Cieplak M. Structural Fluctuations and Thermal Stability of Proteins in Crowded Environments: Effects of the Excluded Volume. Phys. Biol. 2016, 13, 066002. 10.1088/1478-3975/13/6/066002. [DOI] [PubMed] [Google Scholar]
- McGuffee S. R.; Elcock A. H. Diffusion, Crowding & Protein Stability in a Dynamic Molecular Model of the Bacterial Cytoplasm. PLoS Comput. Biol. 2010, 6, e1000694. 10.1371/journal.pcbi.1000694. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ando T.; Skolnick J. Crowding and Hydrodynamic Interactions Likely Dominate in vivo Macromolecular Motion. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 18457–18462. 10.1073/pnas.1011354107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mereghetti P.; Wade R. C. Atomic Detail Brownian Dynamics Simulations of Concentrated Protein Solutions with a Mean Field Treatment of Hydrodynamic Interactions. J. Phys. Chem. B 2012, 116, 8523–8533. 10.1021/jp212532h. [DOI] [PubMed] [Google Scholar]
- Mereghetti P.; Gabdoulline R. R.; Wade R. C. Brownian Dynamics Simulation of Protein Solutions Structural and Dynamical Properties. Biophys. J. 2010, 99, 3782–3791. 10.1016/j.bpj.2010.10.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Trovato F.; Tozzini V. Diffusion within the Cytoplasm: A Mesoscale Model of Interacting Macromolecules. Biophys. J. 2014, 107, 2579–2591. 10.1016/j.bpj.2014.09.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qin S. B.; Minh D. D. L.; McCammon J. A.; Zhou H. X. Method to Predict Crowding Effects by Postprocessing Molecular Dynamics Trajectories: Application to the Flap Dynamics of HIV-1 Protease. J. Phys. Chem. Lett. 2010, 1, 107–110. 10.1021/jz900023w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cheung M. S.; Thirumalai D. Effects of crowding and confinement on the structures of the transition state ensemble in proteins. J. Phys. Chem. B 2007, 111, 8250–8257. 10.1021/jp068201y. [DOI] [PubMed] [Google Scholar]
- Homouz D.; Perham M.; Samiotakis A.; Cheung M. S.; Wittung-Stafshede P. Crowded, Cell-Like Environment Induces Shape Changes in Aspherical Protein. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, 11754–11759. 10.1073/pnas.0803672105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Homouz D.; Stagg L.; Wittung-Stafshede P.; Cheung M. S. Macromolecular Crowding Modulates Folding Mechanism of alpha/beta Protein Apoflavodoxin. Biophys. J. 2009, 96, 671–680. 10.1016/j.bpj.2008.10.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McGuffee S. R.; Elcock A. H. Atomically Detailed Simulations of Concentrated Protein Solutions: The Effects of Salt, pH, Point Mutations, and Protein Concentration in Simulations of 1000-Molecule Systems. J. Am. Chem. Soc. 2006, 128, 12098–12110. 10.1021/ja0614058. [DOI] [PubMed] [Google Scholar]
- Gabdoulline R. R.; Wade R. C. Protein-Protein Association: Investigation of Factors Influencing Association Rates by Brownian Dynamics Simulations. J. Mol. Biol. 2001, 306, 1139–1155. 10.1006/jmbi.2000.4404. [DOI] [PubMed] [Google Scholar]
- Kar P.; Feig M. Recent Advances in Transferable Coarse-Grained Modeling of Proteins. Adv. Protein Chem. Struct. Biol. 2014, 96, 143–180. 10.1016/bs.apcsb.2014.06.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kar P.; Gopal S. M.; Cheng Y. M.; Predeus A. V.; Feig M. PRIMO: A Transferable Coarse-Grained Force Field for Proteins. J. Chem. Theory Comput. 2013, 9, 3769–3788. 10.1021/ct400230y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Andrews C. T.; Elcock A. H. COFFDROP: A Coarse-Grained Nonbonded Force Field for Proteins Derived from All-Atom Explicit-Solvent Molecular Dynamics Simulations of Amino Acids. J. Chem. Theory Comput. 2014, 10, 5178–5194. 10.1021/ct5006328. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chiricotto M.; Sterpone F.; Derreumaux P.; Melchionna S. Multiscale Simulation of Molecular Processes in Cellular Environments. Philos. Trans. R. Soc., A 2016, 374, 20160225. 10.1098/rsta.2016.0225. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Predeus A. V.; Gul S.; Gopal S. M.; Feig M. Conformational Sampling of Peptides in the Presence of Protein Crowders from AA/CG-Multiscale Simulations. J. Phys. Chem. B 2012, 116, 8610–8620. 10.1021/jp300129u. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Feig M.; Sugita Y. Variable Interactions between Protein Crowders and Biomolecular Solutes are Important in Understanding Cellular Crowding. J. Phys. Chem. B 2012, 116, 599–605. 10.1021/jp209302e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Harada R.; Tochio N.; Kigawa T.; Sugita Y.; Feig M. Reduced Native State Stability in Crowded Cellular Environment Due to Protein-Protein Interactions. J. Am. Chem. Soc. 2013, 135, 3696–3701. 10.1021/ja3126992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang P.-h.; Yu I.; Feig M.; Sugita Y. Influence of Protein Crowder Size on Hydration Structure and Dynamics in Macromolecular Crowding. Chem. Phys. Lett. 2017, 671, 63–70. 10.1016/j.cplett.2017.01.012. [DOI] [Google Scholar]
- Yu I.; Mori T.; Ando T.; Harada R.; Jung J.; Sugita Y.; Feig M. Biomolecular Interactions Modulate Macromolecular Structure and Dynamics in Atomistic Model of a Bacterial Cytoplasm. eLife 2016, 5, e19274. 10.7554/eLife.19274. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Candotti M.; Orozco M. The Differential Response of Proteins to Macromolecular Crowding. PLoS Comput. Biol. 2016, 12, e1005040. 10.1371/journal.pcbi.1005040. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Virk A. S.; Stait-Gardner T.; Willis S. A.; Torres A. M.; Price W. S. Macromolecular Crowding Studies of Amino Acids Using NMR Diffusion Measurements and Molecular Dynamics Simulations. Front. Phys. 2015, 3, 1. 10.3389/fphy.2015.00001. [DOI] [Google Scholar]
- Karandur D.; Wong K.-Y.; Pettitt B. M. Solubility and Aggregation of Gly5 in Water. J. Phys. Chem. B 2014, 118, 9565–9572. 10.1021/jp503358n. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Petrov D.; Zagrovic B. Are Current Atomistic Force Fields Accurate Enough to Study Proteins in Crowded Environments?. PLoS Comput. Biol. 2014, 10, e1003638. 10.1371/journal.pcbi.1003638. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Andrews C. T.; Elcock A. H. Molecular Dynamics Simulations of Highly Crowded Amino Acid Solutions: Comparisons of Eight Different Force Field Combinations with Experiment and with Each Other. J. Chem. Theory Comput. 2013, 9, 4585–4602. 10.1021/ct400371h. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Abriata L. A.; Dal Peraro M. Assessing the Potential of Atomistic Molecular Dynamics Simulations to Probe Reversible Protein-Protein Recognition and Binding. Sci. Rep. 2015, 5, 10549. 10.1038/srep10549. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Piana S.; Donchev A. G.; Robustelli P.; Shaw D. E. Water Dispersion Interactions Strongly Influence Simulated Structural Properties of Disordered Protein States. J. Phys. Chem. B 2015, 119, 5113–5123. 10.1021/jp508971m. [DOI] [PubMed] [Google Scholar]
- Best R. B.; Mittal J. Protein Simulations with an Optimized Water Model: Cooperative Helix Formation and Temperature-Induced Unfolded State Collapse. J. Phys. Chem. B 2010, 114, 14916–14923. 10.1021/jp108618d. [DOI] [PubMed] [Google Scholar]
- Hermans J.; Berendsen H. J. C.; van Gunsteren W. F.; Postma J. P. M. A Consistent Empirical Potential for Water-Protein Interactions. Biopolymers 1984, 23, 1513–1518. 10.1002/bip.360230807. [DOI] [Google Scholar]
- Jorgensen W. L.; Chandrasekhar J.; Madura J. D.; Impey R. W.; Klein M. L. Comparison of Simple Potential Functions for Simulating Liquid Water. J. Chem. Phys. 1983, 79, 926–935. 10.1063/1.445869. [DOI] [Google Scholar]
- Yeh I. C.; Hummer G. System-Size Dependence of Diffusion Coefficients and Viscosities from Molecular Dynamics Simulations with Periodic Boundary Conditions. J. Phys. Chem. B 2004, 108, 15873–15879. 10.1021/jp0477147. [DOI] [Google Scholar]
- Venable R. M.; Hatcher E.; Guvench O.; MacKerell A. D.; Pastor R. W. Comparing Simulated and Experimental Translation and Rotation Constants: Range of Validity for Viscosity Scaling. J. Phys. Chem. B 2010, 114, 12501–12507. 10.1021/jp105549s. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang J.; Rauscher S.; Nawrocki G.; Ran T.; Feig M.; de Groot B. L.; Grubmüller H.; MacKerell A. D. Jr. CHARMM36m: An Improved Force Field for Folded and Intrinsically Disordered Proteins. Nat. Methods 2017, 14, 71–73. 10.1038/nmeth.4067. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Best R. B.; Zhu X.; Shim J.; Lopes P.; Mittal J.; Feig M.; MacKerell A. D. Jr Optimization of the Additive CHARMM All-Atom Protein Force Field Targeting Improved Sampling of the Backbone ϕ, ψ and Side-Chain χ1 and χ2 Dihedral Angles. J. Chem. Theory Comput. 2012, 8, 3257–3273. 10.1021/ct300400x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Best R. B.; Mittal J.; Feig M.; MacKerell A. D. Inclusion of Many-Body Effects in the Additive CHARMM Protein CMAP Potential Results in Enhanced Cooperativity of α-Helix and β-Hairpin Formation. Biophys. J. 2012, 103, 1045–1051. 10.1016/j.bpj.2012.07.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Harder E.; Damm W.; Maple J.; Wu C.; Reboul M.; Xiang J. Y.; Wang L.; Lupyan D.; Dahlgren M. K.; Knight J. L.; et al. OPLS3: A Force Field Providing Broad Coverage of Drug-like Small Molecules and Proteins. J. Chem. Theory Comput. 2016, 12, 281–296. 10.1021/acs.jctc.5b00864. [DOI] [PubMed] [Google Scholar]
- Margreitter C.; Reif M. M.; Oostenbrink C. Update on Phosphate and Charged Post-Translationally Modified Amino Acid Parameters in the GROMOS Force Field. J. Comput. Chem. 2017, 38, 714–720. 10.1002/jcc.24733. [DOI] [PubMed] [Google Scholar]
- Best R. B.; Zheng W.; Mittal J. Balanced Protein-Water Interactions Improve Properties of Disorderd Proteins and Non-Specific Protein Association. J. Chem. Theory Comput. 2014, 10, 5113–5124. 10.1021/ct500569b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maier J. A.; Martinez C.; Kasavajhala K.; Wickstrom L.; Hauser K. E.; Simmerling C. ff14SB: Improving the Accuracy of Protein Side Chain and Backbone Parameters from ff99SB. J. Chem. Theory Comput. 2015, 11, 3696–3713. 10.1021/acs.jctc.5b00255. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jung J.; Mori T.; Kobayashi C.; Matsunaga Y.; Yoda T.; Feig M.; Sugita Y. GENESIS: A Hybrid-Parallel and Multi-Scale Molecular Dynamics Simulator with Enhanced Sampling Algorithms for Biomolecular and Cellular Simulations. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2015, 5, 310–323. 10.1002/wcms.1220. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jung J.; Naurse A.; Kobayashi C.; Sugita Y. Graphics Processing Unit Acceleration and Parallelization of GENESIS for Large-Scale Molecular Dynamics Simulations. J. Chem. Theory Comput. 2016, 12, 4947–4958. 10.1021/acs.jctc.6b00241. [DOI] [PubMed] [Google Scholar]
- Friedrichs M. S.; Eastman P.; Vaidyanathan V.; Houston M.; Legrand S.; Beberg A. L.; Ensign D. L.; Bruns C. M.; Pande V. S. Accelerating Molecular Dynamic Simulation on Graphics Processing Units. J. Comput. Chem. 2009, 30, 864–872. 10.1002/jcc.21209. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Krieger E.; Vriend G. New Ways to Boost Molecular Dynamics Simulations. J. Comput. Chem. 2015, 36, 996–1007. 10.1002/jcc.23899. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Abraham M. J.; Murtola T.; Schulz R.; Páll S.; Smith J. C.; Hess B.; Lindahl E. GROMACS: High Performance Molecular Simulations through Multi-Level Parallelism from Laptops to Supercomputers. SoftwareX 2015, 1–2, 19–25. 10.1016/j.softx.2015.06.001. [DOI] [Google Scholar]
- Salomon-Ferrer R.; Götz A. W.; Poole D.; Le Grand S.; Walker R. C. Routine Microsecond Molecular Dynamics Simulations with AMBER on GPUs. 2. Explicit Solvent Particle Mesh Ewald. J. Chem. Theory Comput. 2013, 9, 3878–3888. 10.1021/ct400314y. [DOI] [PubMed] [Google Scholar]
- Shaw D. E.; Maragakis P.; Lindorff-Larsen K.; Piana S.; Dror R. O.; Eastwood M. P.; Bank J. A.; Jumper J. M.; Salmon J. K.; Shan Y. B.; et al. Atomic-Level Characterization of the Structural Dynamics of Proteins. Science 2010, 330, 341–346. 10.1126/science.1187409. [DOI] [PubMed] [Google Scholar]
- Abrams C.; Bussi G. Enhanced Sampling in Molecular Dynamics Using Metadynamics, Replica-Exchange, and Temperature-Acceleration. Entropy 2014, 16, 163–199. 10.3390/e16010163. [DOI] [Google Scholar]
- Torrie G. M.; Valleau J. P. Monte-Carlo Free-Energy Estimates using Non-Boltzmann Sampling - Application to Subcritical Lennard-Jones Fluid. Chem. Phys. Lett. 1974, 28, 578–581. 10.1016/0009-2614(74)80109-0. [DOI] [Google Scholar]
- Sugita Y.; Okamoto Y. Replica-Exchange Molecular Dynamics Method for Protein Folding. Chem. Phys. Lett. 1999, 314, 141–151. 10.1016/S0009-2614(99)01123-9. [DOI] [Google Scholar]
- Okamoto Y. Generalized-Ensemble Algorithms: Enhanced Sampling Techniques for Monte Carlo and Molecular Dynamics Simulations. J. Mol. Graphics Modell. 2004, 22, 425–439. 10.1016/j.jmgm.2003.12.009. [DOI] [PubMed] [Google Scholar]
- Micheletti C.; Laio A.; Parrinello M. Reconstructing the Density of States by History-Dependent Metadynamics. Phys. Rev. Lett. 2004, 92, 92. 10.1103/PhysRevLett.92.170601. [DOI] [PubMed] [Google Scholar]
- Guo Z.; Brooks C. L. III Rapid Screening of Binding Affinities: Application of the λ-Dynamics Method to a Trypsin-Inhibitor System. J. Am. Chem. Soc. 1998, 120, 1920–1921. 10.1021/ja973418e. [DOI] [Google Scholar]
- Hamelberg D.; Mongan J.; McCammon J. A. Accelerated Molecular Dynamics: A Promising and Efficient Simulation Method for Biomolecules. J. Chem. Phys. 2004, 120, 11919–11929. 10.1063/1.1755656. [DOI] [PubMed] [Google Scholar]
- Cossins B. P.; Jacobson M. P.; Guallar V. A New View of the Bacterial Cytosol Environment. PLoS Comput. Biol. 2011, 7, e1002066. 10.1371/journal.pcbi.1002066. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Spiga E.; Abriata L. A.; Piazza F.; Dal Peraro M. Dissecting the Effects of Concentrated Carbohydrate Solutions on Protein Diffusion, Hydration, and Internal Dynamics. J. Phys. Chem. B 2014, 118, 5310–5321. 10.1021/jp4126705. [DOI] [PubMed] [Google Scholar]
- Zhou H.-X. Influences of Crowded Cellular Environments on Protein Folding, Binding, and Oligomerization: Biological Consequences and Potentials of Atomistic Modeling. FEBS Lett. 2013, 587, 1053–1061. 10.1016/j.febslet.2013.01.064. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qin S.; Zhou H.-X. Further Development of the FFT-based Method for Atomistic Modeling of Protein Folding and Binding under Crowding: Optimization of Accuracy and Speed. J. Chem. Theory Comput. 2014, 10, 2824–2835. 10.1021/ct5001878. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qin S.; Zhou H.-X. Atomistic Modeling of Macromolecular Crowding Predicts Modest Increases in Protein Folding and Binding Stability. Biophys. J. 2009, 97, 12–19. 10.1016/j.bpj.2009.03.066. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Drake J. A.; Pettitt B. M. Force Field-Dependent Solution Properties of Glycine Oligomers. J. Comput. Chem. 2015, 36, 1275–1285. 10.1002/jcc.23934. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kuroda Y.; Suenaga A.; Sato Y.; Kosuda S.; Taiji M. All-Atom Molecular Dynamics Analysis of Multi-Peptide Systems Reproduces Peptide Solubility In Line with Experimental Observations. Sci. Rep. 2016, 6, 19479. 10.1038/srep19479. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Feig M.; Harada R.; Mori T.; Yu I.; Takahashi K.; Sugita Y. Complete Atomistic Model of a Bacterial Cytoplasm Integrates Physics, Biochemistry, and Systems Biology. J. Mol. Graphics Modell. 2015, 58, 1–9. 10.1016/j.jmgm.2015.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kurniawan N. A.; Enemark S.; Rajagopalan R. Crowding Alters the Folding Kinetics of a β-Hairpin by Modulating the Stability of Intermediates. J. Am. Chem. Soc. 2012, 134, 10200–10208. 10.1021/ja302943m. [DOI] [PubMed] [Google Scholar]
- Carballo-Pacheco M.; Strodel B. Advances in the Simulation of Protein Aggregation at the Atomistic Scale. J. Phys. Chem. B 2016, 120, 2991–2999. 10.1021/acs.jpcb.6b00059. [DOI] [PubMed] [Google Scholar]
- Maffeo C.; Luan B.; Aksimentiev A. End-To-End Attraction of Duplex DNA. Nucleic Acids Res. 2012, 40, 3812–3821. 10.1093/nar/gkr1220. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hajnic M.; Osorio J. I.; Zagrovic B. Interaction Preferences between Nucleobase Mimetics and Amino Acids in Aqueous Solutions. Phys. Chem. Chem. Phys. 2015, 17, 21414–21422. 10.1039/C5CP01486G. [DOI] [PubMed] [Google Scholar]
- Hajnic M.; Osorio J. I.; Zagrovic B. Computational Analysis of Amino Acids and their Sidechain Analogs in Crowded Solutions of RNA Nucleobases with Implications for the mRNA–Protein Complementarity Hypothesis. Nucleic Acids Res. 2014, 42, 12984–12994. 10.1093/nar/gku1035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Palazzesi F.; Prakash M. K.; Bonomi M.; Barducci A. Accuracy of Current All-Atom Force-Fields in Modeling Protein Disordered States. J. Chem. Theory Comput. 2015, 11, 2–7. 10.1021/ct500718s. [DOI] [PubMed] [Google Scholar]
- Henriques J.; Cragnell C.; Skepö M. Molecular Dynamics Simulations of Intrinsically Disordered Proteins: Force Field Evaluation and Comparison with Experiment. J. Chem. Theory Comput. 2015, 11, 3420–3431. 10.1021/ct501178z. [DOI] [PubMed] [Google Scholar]
- Rauscher S.; Gapsys V.; Gajda M. J.; Zweckstetter M.; de Groot B. L.; Grubmüller H. Structural Ensembles of Intrinsically Disordered Proteins Depend Strongly on Force Field: A Comparison to Experiment. J. Chem. Theory Comput. 2015, 11, 5513–5524. 10.1021/acs.jctc.5b00736. [DOI] [PubMed] [Google Scholar]
- Cho S. S.; Reddy G.; Straub J. E.; Thirumalai D. Entropic Stabilization of Proteins by TMAO. J. Phys. Chem. B 2011, 115, 13401–13407. 10.1021/jp207289b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Auton M.; Bolen D. W.; Rosgen J. Structural Thermodynamics of Protein Preferential Solvation: Osmolyte Solvation of Proteins, Aminoacids, and Peptides. Proteins: Struct., Funct., Genet. 2008, 73, 802–813. 10.1002/prot.22103. [DOI] [PubMed] [Google Scholar]
- Record M. T.; Anderson C. F. Interpretation of Preferential Interaction Coefficients of Nonelectrolytes and of Electrolyte Ions in Terms of a Two-Domain Model. Biophys. J. 1995, 68, 786–794. 10.1016/S0006-3495(95)80254-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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]
- Arakawa T.; Timasheff S. N. The Stabilization of Proteins by Osmolytes. Biophys. J. 1985, 47, 411–414. 10.1016/S0006-3495(85)83932-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu I.; Jindo Y.; Nagaoka M. Microscopic Understanding of Preferential Exclusion of Compatible Solute Ectoine: Direct Interaction and Hydration Alteration. J. Phys. Chem. B 2007, 111, 10231–10238. 10.1021/jp068367z. [DOI] [PubMed] [Google Scholar]
- Kokubo H.; Pettitt B. M. Preferential Solvation in Urea Solutions at Different Concentrations: Properties from Simulation Studies. J. Phys. Chem. B 2007, 111, 5233–5242. 10.1021/jp067659x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aburi M.; Smith P. E. A Combined Simulation and Kirkwood-Buff Approach to Quantify Cosolvent Effects on the Conformational Preferences of Peptides in Solution. J. Phys. Chem. B 2004, 108, 7382–7388. 10.1021/jp036582z. [DOI] [Google Scholar]
- Shimizu S.; Matubayasi N. Preferential Hydration of Proteins: A Kirkwood-Buff Approach. Chem. Phys. Lett. 2006, 420, 518–522. 10.1016/j.cplett.2006.01.034. [DOI] [Google Scholar]
- Yu I.; Nakada K.; Nagaoka M. Spatio-Temporal Characteristics of Transfer Free Energy of Apomyoglobin into the Crowding Condition with Trimethylamine N-oxide: A Study with Three Types of the Kirkwood-Buff Integral. J. Phys. Chem. B 2012, 116, 4080–4088. 10.1021/jp300380p. [DOI] [PubMed] [Google Scholar]
- Minton A. P. Excluded Volume as a Determinant of Macromolecular Structure and Reactivity. Biopolymers 1981, 20, 2093–2120. 10.1002/bip.1981.360201006. [DOI] [Google Scholar]
- Koshland D. E. The Key–Lock Theory and the Induced Fit Theory. Angew. Chem., Int. Ed. Engl. 1995, 33, 2375–2378. 10.1002/anie.199423751. [DOI] [Google Scholar]
- Ebbinghaus S.; Gruebele M. Protein Folding Landscapes in the Living Cell. J. Phys. Chem. Lett. 2011, 2, 314–319. 10.1021/jz101729z. [DOI] [Google Scholar]
- Miklos A. C.; Li C.; Sorrell C. D.; Lyon L. A.; Pielak G. J. An Upper Limit for Macromolecular Crowding Effects. BMC Biophys. 2011, 4, 13. 10.1186/2046-1682-4-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Roosen-Runge F.; Hennig M.; Zhang F. J.; Jacobs R. M. J.; Sztucki M.; Schober H.; Seydel T.; Schreiber F. Protein Self-Diffusion in Crowded Solutions. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 11815–11820. 10.1073/pnas.1107287108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dix J. A.; Verkman A. S. Crowding Effects on Diffusion in Solutions and Cells. Annu. Rev. Biophys. 2008, 37, 247–263. 10.1146/annurev.biophys.37.032807.125824. [DOI] [PubMed] [Google Scholar]
- Dauty E.; Verkman A. S. Molecular Crowding Reduces to a Similar Extent the Diffusion of Small Solutes and Macromolecules: Measurement by Fluorescence Correlation Spectroscopy. J. Mol. Recognit. 2004, 17, 441–447. 10.1002/jmr.709. [DOI] [PubMed] [Google Scholar]
- Höfling F.; Franosch T. Anomalous Transport in the Crowded World of Biological Cells. Rep. Prog. Phys. 2013, 76, 046602. 10.1088/0034-4885/76/4/046602. [DOI] [PubMed] [Google Scholar]
- Ridgway D.; Broderick G.; Lopez-Campistrous A.; Ru’aini M.; Winter P.; Hamilton M.; Boulanger P.; Kovalenko A.; Ellison M. J. Coarse-Grained Molecular Simulation of Diffusion and Reaction Kinetics in a Crowded Virtual Cytoplasm. Biophys. J. 2008, 94, 3748–3759. 10.1529/biophysj.107.116053. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shaw D. E.; Grossman J. P.; Bank J. A.; Batson B.; Butts J. A.; Chao J. C.; Deneroff M. M.; Dror R. O.; Even A.; Fenton C. H.; et al. Anton 2: Raising the Bar for Performance and Programmability in a Special-Purpose Molecular Dynamics Supercomputer. Proceedings of the International Conference for High Performance Computing, Networking, Storage and Analysis; IEEE Press: New Orleans, LA, 2014; pp 41–53. [Google Scholar]
- Schutz C. N.; Warshel A. What are the Dielectric ″Constants″ of Proteins and How to Validate Electrostatic Models?. Proteins: Struct., Funct., Genet. 2001, 44, 400–417. 10.1002/prot.1106. [DOI] [PubMed] [Google Scholar]
- Gilson M. K.; Honig B. H. The Dielectric-Constant of a Folded Protein. Biopolymers 1986, 25, 2097–2119. 10.1002/bip.360251106. [DOI] [PubMed] [Google Scholar]
- Horn H. W.; Swope W. C.; Pitera J. W.; Madura J. D.; Dick T. J.; Hura G. L.; Head-Gordon T. Development of an improved four-site water model for biomolecular simulations: TIP4P-Ew. J. Chem. Phys. 2004, 120, 9665–9678. 10.1063/1.1683075. [DOI] [PubMed] [Google Scholar]
- Paschek D.; Day R.; Garcia A. E. Influence of Water-Protein Hydrogen Bonding on the Stability of Trp-Cage Miniprotein. A Comparison between the TIP3P and TIP4P-Ew Water Models. Phys. Chem. Chem. Phys. 2011, 13, 19840–19847. 10.1039/c1cp22110h. [DOI] [PubMed] [Google Scholar]
- Freedberg D. I.; Selenko P. Live Cell NMR. Annu. Rev. Biophys. 2014, 43, 171–192. 10.1146/annurev-biophys-051013-023136. [DOI] [PubMed] [Google Scholar]
- 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]
- Shtengel G.; Galbraith J. A.; Galbraith C. G.; Lippincott-Schwartz J.; Gillette J. M.; Manley S.; Sougrat R.; Waterman C. M.; Kanchanawong P.; Davidson M. W.; et al. Interferometric Fluorescent Super-Resolution Microscopy Resolves 3D Cellular Ultrastructure. Proc. Natl. Acad. Sci. U. S. A. 2009, 106, 3125–3130. 10.1073/pnas.0813131106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cheng Y.; Grigorieff N.; Penczek; Pawel A.; Walz T. A Primer to Single-Particle Cryo-Electron Microscopy. Cell 2015, 161, 438–449. 10.1016/j.cell.2015.03.050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee S. F.; Thompson; Michael A.; Schwartz M. A.; Shapiro L.; Moerner W. E. Super-Resolution Imaging of the Nucleoid-Associated Protein HU in Caulobacter crescentus. Biophys. J. 2011, 100, L31–L33. 10.1016/j.bpj.2011.02.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bakshi S.; Choi H.; Weisshaar J. C. The Spatial Biology of Transcription and Translation in Rapidly Growing Escherichia coli. Front. Microbiol. 2015, 6, 636. 10.3389/fmicb.2015.00636. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang X.; Llopis P. M.; Rudner D. Z. Organization and Segregation of Bacterial Chromosomes. Nat. Rev. Genet. 2013, 14, 191–203. 10.1038/nrg3375. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhou H. X. Crowding Effects of Membrane Proteins. J. Phys. Chem. B 2009, 113, 7995–8005. 10.1021/jp8107446. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim J. S.; Yethiraj A. Crowding Effects on Association Reactions at Membranes. Biophys. J. 2010, 98, 951–958. 10.1016/j.bpj.2009.11.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gane P. J.; Dean P. M. Recent Advances in Structure-Based Rational Drug Design. Curr. Opin. Struct. Biol. 2000, 10, 401–404. 10.1016/S0959-440X(00)00105-6. [DOI] [PubMed] [Google Scholar]
- Taft C. A.; Da Silva V. B.; Da Silva C. Current Topics in Computer-Aided Drug Design. J. Pharm. Sci. 2008, 97, 1089–1098. 10.1002/jps.21293. [DOI] [PubMed] [Google Scholar]