Site-Specific Internal Motions in GB1 Protein Microcrystals Revealed by 3D 2H–13C–13C Solid-State NMR Spectroscopy
Abstract

2H quadrupolar line shapes deliver rich information about protein dynamics. A newly designed 3D 2H–13C–13C solid-state NMR magic angle spinning (MAS) experiment is presented and demonstrated on the microcrystalline β1 immunoglobulin binding domain of protein G (GB1). The implementation of 2H–13C adiabatic rotor-echo-short-pulse-irradiation cross-polarization (RESPIRATION CP) ensures the accuracy of the extracted line shapes and provides enhanced sensitivity relative to conventional CP methods. The 3D 2H–13C–13C spectrum reveals 2H line shapes for 140 resolved aliphatic deuterium sites. Motional-averaged 2H quadrupolar parameters obtained from the line-shape fitting identify side-chain motions. Restricted side-chain dynamics are observed for a number of polar residues including K13, D22, E27, K31, D36, N37, D46, D47, K50, and E56, which we attribute to the effects of salt bridges and hydrogen bonds. In contrast, we observe significantly enhanced side-chain flexibility for Q2, K4, K10, E15, E19, N35, N40, and E42, due to solvent exposure and low packing density. T11, T16, and T17 side chains exhibit motions with larger amplitudes than other Thr residues due to solvent interactions. The side chains of L5, V54, and V29 are highly rigid because they are packed in the core of the protein. High correlations were demonstrated between GB1 side-chain dynamics and its biological function. Large-amplitude side-chain motions are observed for regions contacting and interacting with immunoglobulin G (IgG). In contrast, rigid side chains are primarily found for residues in the structural core of the protein that are absent from protein binding and interactions.
Introduction



Materials and Methods
Preparation of Protein GB1
Solid-State NMR Experiments
Results and Discussion
2H–13C Adiabatic RESPIRATION CP Enabling Accurate Indirect Detection of 2H Line Shapes



3D 2H–13C–13C NMR Correlation Experiment Designed for Studying Protein Backbone and Side-Chain Dynamics

Figure 1

Figure 1. (A) 3D 2H–13C–13C solid-state MAS NMR pulse sequence. (B) 2D 2H–13C planes of the 3D 2H–13C–13C spectrum collected for microcrystalline GB1 and the extracted 2H line shapes for the E56 residue. Experimental line shapes (black), fits (red), and fitting residuals (blue) are displayed in the right column. In the 2D planes, signals with positive and negative intensities are shown in black and green, respectively.
Lys Side-Chain Dynamics in GB1









For deuterium undergoing large-amplitude motion, η̅ was set to zero for line-shape fitting.
The 13Cγ–13Cδ and 13Cδ−13Cξ cross peaks of K4 severely overlap with those of K10. The same 2Hδ and 2Hξ quadrupolar values were assigned to K4 and K10, which were obtained from the overlapped 2H line shapes.
Figure 2

Figure 2. (A) Crystal structure of GB1 (PDB: 2LGI) with Lys aliphatic groups shown in van der Waals spheres and coded with color scaling to values. Two different perspectives are shown for better visualization. (B) K50, (C) K31, (D) K13, (E) K4, (F) K10, and (G) K28 local chemical environment in GB1. The K31 (CH2)β group is color coded with cyan as the
value is not determined. The residues having atoms within 5 Å away for the corresponding Lys residue are shown in sticks. The salt bridges between K50 and D47, E27, and K31 and the hydrogen bond between K13 side chain NH3 and G9 backbone CO are displayed by dash lines.








For deuterium undergoing large-amplitude motion, η̅ was set to zero for line-shape fitting.
Asp and Asn Side-Chain Dynamics in GB1


Figure 3

Figure 3. (A) Crystal structure of GB1 (PDB: 2LGI) with Asp and Asn aliphatic groups shown in van der Waals spheres and coded with color scaling to values. Two different perspectives are shown for better visualization. (B) D47, (C) D22, (D) D46, (E) D36, (F) N37, (G) N35, (H) D40, and (I) N8 local chemical environment in GB1. The residues having atoms within 5 Å away for the corresponding Asn or Asp residue are shown in sticks. The salt bridge between D47 and K50 and the hydrogen bonds between D22 and T25, D46 and A48/T49, and N8 and T55 are presented by dashed lines.
Glu and Gln Side-Chain Dynamics in GB1




For deuterium undergoing large-amplitude motion, η̅ was set to zero for line-shape fitting.





Figure 4

Figure 4. (A) Crystal structure of GB1 (PDB: 2LGI) with Glu and Gln aliphatic groups shown in van der Waals spheres and coded with color scaling to values. Two different perspectives are shown for better visualization. (B) E27, (C) E56, (D) Q32, (E) E42, (F) E15, (G) E19, and (H) Q2 local chemical environment in GB1. Aliphatic groups having deuterium
values undetermined are color coded with cyan. The residues having atoms within 5 Å away for the corresponding Glu or Gln residue are shown in sticks. The salt bridge between E27 and K31 and the hydrogen bond between E56 and D40/K10 are presented by dashed lines.
Thr Side-Chain Dynamics in GB1













For deuterium undergoing large-amplitude motion, η̅ was set to zero for line-shape fitting.
Figure 5

Figure 5. (A) Crystal structure of GB1 (PDB: 2LGI) with Thr aliphatic groups shown in van der Waals spheres and coded with color scaling to values. Two different perspectives are shown for better visualization. (B) T55, (C) T25, (D) T49, (E) T18, (F) T44, (G) T49, (H) T11, (I) T16, (J) T17, and (K) T53 local chemical environment in GB1. Aliphatic groups having deuterium
values undetermined are color coded with cyan. The residues having atoms within 5 Å away for the corresponding Thr residues are shown in sticks. The hydrogen bonds between T55 and N8, T25 and D22, and T51 and T49 are presented by dashed lines.
Side-Chain Dynamics of Nonpolar Residues in GB1













For deuterium undergoing large-amplitude motion, η̅ was set to zero for line-shape fitting. Quadrupolar coupling parameters were not determined for L7 (CH2)δ1 due to signal overlap.
Figure 6

Figure 6. (A) Crystal structure of GB1 (PDB: 2LGI) with Leu and Ile aliphatic groups shown in van der Waals spheres and coded with color scaling to values. (B) L5, (C) L12, (D) L7, and (E) I6 local chemical environment in GB1. Aliphatic groups having deuterium
values undetermined are color coded with cyan. The residues having atoms within 5 Å away for the corresponding Leu/Ile residue are shown in sticks.



For deuterium undergoing large-amplitude motion, η̅ was set to zero for line-shape fitting.
Figure 7

Figure 7. (A) Crystal structure of GB1 (PDB: 2LGI) with Val aliphatic groups shown in van der Waals spheres and coded with color scaling to values. (B) V21, (C) V54, (D) V29, and (E) V39 local chemical environment in GB1. Aliphatic groups having deuterium
values undetermined are color coded with cyan. The residues having atoms within 5 Å away for the corresponding Val residue are shown in sticks.


Backbone Dynamics of Microcrystalline GB1





Figure 8

Figure 8. 2Hα values determined for amino acid residues in microcrystalline GB1. The value for residue F52 was not determined due to signal overlap.



Correlations between the Dynamics of GB1 and Its Structure and Biological Function

Figure 9

Figure 9. Protein side-chain dynamics map for microcrystalline GB1, shown in three different viewpoints. Lys, Asn, Asp, Gln, Glu, Thr, Leu, and Val residues exhibiting large, moderate, and small amplitude side-chain motions are highlighted in red, orange, and blue, respectively. The rest of the residues (gray) are not considered in this illustration.
Conclusions


Supporting Information
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.5b12974.
Uniform 2H magnetization transfer in 2H–13C adiabatic RESPIRATION CP (Figures S1, S2, and S3 and Table S1), Ala 2H one-pulse spectrum fit (Figure S4), GB1 2H line shape fits (Figure S5), Ala dynamics in GB1 (Table S2 and Figure S6), backbone order parameters derived from 2H measurement and CH dipolar measurement (Figure S7), GB1 2Hα η̅ values (Figure S8) (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgment
This research is supported by R01-HL103999, R01-GM073770, R01-GM112845, and R21-GM107905. XS is an American Heart Association Postdoctoral Fellow (15POST25700070). We thank Dr. Deborah Berthold for help with GB1 protein sample preparation. We also thank Dr. Kristin Nuzzio, Mr. Dennis Piehl, and Ms. Lisa Della Ripa for careful reading of the manuscript.
References
This article references 113 other publications.
- 1Teilum, K.; Olsen, J. G.; Kragelund, B. B. Cell. Mol. Life Sci. 2009, 66, 2231– 2247 DOI: 10.1007/s00018-009-0014-6Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXot12lsbc%253D&md5=caa1fa9a979d3ab2fef0e0f82b5452c0Functional aspects of protein flexibilityTeilum, Kaare; Olsen, Johan G.; Kragelund, Birthe B.Cellular and Molecular Life Sciences (2009), 66 (14), 2231-2247CODEN: CMLSFI; ISSN:1420-682X. (Birkhaeuser Verlag)A review. Proteins are dynamic entities, and they possess an inherent flexibility that allows them to function through mol. interactions within the cell, among cells, and even between organisms. An appreciation of the non-static nature of proteins is emerging, but to describe and incorporate this into an intuitive perception of protein function is challenging. Flexibility is of overwhelming importance for protein function, and the changes in protein structure during interactions with binding partners can be dramatic. Here, the authors address protein flexibility, focusing on protein-ligand interactions. The thermodn. involved are reviewed, and examples of structure-function studies involving exptl. detd. flexibility descriptions are presented. While much remains to be understood about protein flexibility, it is clear that it is encoded within the amino acid sequence and should be viewed as an integral part of protein structure.
- 2Henzler-Wildman, K.; Kern, D. Nature 2007, 450, 964– 972 DOI: 10.1038/nature06522Google ScholarThere is no corresponding record for this reference.
- 3Reichert, D.; Zinkevich, T.; Saalwachter, K.; Krushelnitsky, A. J. Biomol. Struct. Dyn. 2012, 30, 617– 627 DOI: 10.1080/07391102.2012.689695Google Scholar3https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhs1CjsbfN&md5=3c33cdd703eebcb170f480ed0d10f27aThe relation of the X-ray B-factor to protein dynamics: insights from recent dynamic solid-state NMR dataReichert, Detlef; Zinkevich, Tatiana; Saalwaechter, Kay; Krushelnitsky, AlexeyJournal of Biomolecular Structure and Dynamics (2012), 30 (6), 617-627CODEN: JBSDD6; ISSN:0739-1102. (Taylor & Francis Ltd.)In addressing the potential use of B-factors derived from X-ray scattering data of proteins for the understanding the (functional) dynamics of proteins, we present a comparison of B-factors of five different proteins (SH3 domain, Crh, GB1, ubiquitin and thioredoxin) with data from recent solid-state NMR expts. reflecting true (rotational) dynamics on well-defined timescales. Apart from trivial correlations involving mobile loop regions and chain termini, we find no significant correlation of B-factors with the dynamic data on any of the investigated timescales, concluding that there is no unique and general correlation of B-factors with the internal reorientational dynamics of proteins.
- 4Kuzmanic, A.; Pannu, N. S.; Zagrovic, B. Nat. Commun. 2014, 5, 3220 DOI: 10.1038/ncomms4220Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC2cvivFKgsw%253D%253D&md5=99e76ed6614f1c57b46fa8917bcbcf99X-ray refinement significantly underestimates the level of microscopic heterogeneity in biomolecular crystalsKuzmanic Antonija; Zagrovic Bojan; Pannu Navraj SNature communications (2014), 5 (), 3220 ISSN:.Biomolecular X-ray structures typically provide a static, time- and ensemble-averaged view of molecular ensembles in crystals. In the absence of rigid-body motions and lattice defects, B-factors are thought to accurately reflect the structural heterogeneity of such ensembles. In order to study the effects of averaging on B-factors, we employ molecular dynamics simulations to controllably manipulate microscopic heterogeneity of a crystal containing 216 copies of villin headpiece. Using average structure factors derived from simulation, we analyse how well this heterogeneity is captured by high-resolution molecular-replacement-based model refinement. We find that both isotropic and anisotropic refined B-factors often significantly deviate from their actual values known from simulation: even at high 1.0 ÅA resolution and Rfree of 5.9%, B-factors of some well-resolved atoms underestimate their actual values even sixfold. Our results suggest that conformational averaging and inadequate treatment of correlated motion considerably influence estimation of microscopic heterogeneity via B-factors, and invite caution in their interpretation.
- 5McDermott, A. Annu. Rev. Biophys. 2009, 38, 385– 403 DOI: 10.1146/annurev.biophys.050708.133719Google ScholarThere is no corresponding record for this reference.
- 6Meirovitch, E.; Shapiro, Y. E.; Polimeno, A.; Freed, J. H. Prog. Nucl. Magn. Reson. Spectrosc. 2010, 56, 360– 405 DOI: 10.1016/j.pnmrs.2010.03.002Google ScholarThere is no corresponding record for this reference.
- 7Krushelnitsky, A.; Reichert, D.; Saalwachter, K. Acc. Chem. Res. 2013, 46, 2028– 2036 DOI: 10.1021/ar300292pGoogle ScholarThere is no corresponding record for this reference.
- 8Krushelnitsky, A.; Reichert, D. Prog. Nucl. Magn. Reson. Spectrosc. 2005, 47, 1– 25 DOI: 10.1016/j.pnmrs.2005.04.001Google ScholarThere is no corresponding record for this reference.
- 9Watt, E. D.; Rienstra, C. M. Anal. Chem. 2014, 86, 58– 64 DOI: 10.1021/ac403956kGoogle ScholarThere is no corresponding record for this reference.
- 10Kay, L. E. J. Magn. Reson. 2005, 173, 193– 207 DOI: 10.1016/j.jmr.2004.11.021Google ScholarThere is no corresponding record for this reference.
- 11Kleckner, I. R.; Foster, M. P. Biochim. Biophys. Acta, Proteins Proteomics 2011, 1814, 942– 968 DOI: 10.1016/j.bbapap.2010.10.012Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXnsVCnsbo%253D&md5=2c047b195fd796a815e800bd162a49deAn introduction to NMR-based approaches for measuring protein dynamicsKleckner, Ian R.; Foster, Mark P.Biochimica et Biophysica Acta, Proteins and Proteomics (2011), 1814 (8), 942-968CODEN: BBAPBW; ISSN:1570-9639. (Elsevier B. V.)A review. Proteins are inherently flexible at ambient temp. At equil., they are characterized by a set of conformations that undergo continuous exchange within a hierarchy of spatial and temporal scales ranging from nanometers to micrometers and femtoseconds to hours. Dynamic properties of proteins are essential for describing the structural bases of their biol. functions including catalysis, binding, regulation and cellular structure. NMR spectroscopy represents a powerful technique for measuring these essential features of proteins. Here the authors provide an introduction to NMR-based approaches for studying protein dynamics, highlighting eight distinct methods with recent examples, contextualized within a common exptl. and anal. framework. The selected methods are (1) Real-time NMR, (2) Exchange spectroscopy, (3) Lineshape anal., (4) CPMG relaxation dispersion, (5) Rotating frame relaxation dispersion, (6) Nuclear spin relaxation, (7) Residual dipolar coupling, (8) Paramagnetic relaxation enhancement. This article is part of a Special Issue entitled: Protein Dynamics: Exptl. and Computational Approaches.
- 12Yang, L. Q.; Sang, P.; Tao, Y.; Fu, Y. X.; Zhang, K. Q.; Xie, Y. H.; Liu, S. Q. J. Biomol. Struct. Dyn. 2014, 32, 372– 393 DOI: 10.1080/07391102.2013.770372Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXksFOms78%253D&md5=028706e72d99b9aa4779974289c72e58Protein dynamics and motions in relation to their functions: several case studies and the underlying mechanismsYang, Li-Quan; Sang, Peng; Tao, Yan; Fu, Yun-Xin; Zhang, Ke-Qin; Xie, Yue-Hui; Liu, Shu-QunJournal of Biomolecular Structure and Dynamics (2014), 32 (3), 372-393CODEN: JBSDD6; ISSN:0739-1102. (Taylor & Francis Ltd.)A review. Proteins are dynamic entities in cellular soln. with functions governed essentially by their dynamic personalities. Here, the authors review several dynamics studies on proteinase K and the HIV-1 virus envelope gp120 glycoprotein to demonstrate the importance of investigating the dynamic behaviors and mol. motions for a complete understanding of their structure-function relations. Using computer simulations and essential dynamic (ED) anal. approaches, the dynamics data obtained revealed that: (1) proteinase K has highly flexible substrate-binding site, thus supporting the induced-fit or conformational selection mechanism of substrate binding; (2) Ca2+ removal from proteinase K increases the global conformational flexibility, decreases the local flexibility of the substrate-binding region, and does not influence the thermal motion of the catalytic triad, thus explaining the exptl. detd. decreased thermostability, reduced substrate affinity, and almost unchanged catalytic activity upon Ca2+ removal; (3) substrate binding affects the large concerted motions of proteinase K, and the resulting dynamic pocket can be connected to substrate binding, orientation, and product release; (4) amino acid mutations 375 S/W and 423 I/P of HIV-1 gp120 have distinct effects on the mol. motions of gp120, facilitating the 375 S/W mutant to assume the CD4-bound conformation, while the 423 I/P mutant to prefer for the CD4-unliganded state. The mechanisms underlying protein dynamics and protein-ligand binding, including the concept of the free energy landscape (FEL) of the protein-solvent system, how the ruggedness and variability of FEL dets. protein's dynamics, and how the 3 ligand-binding models (lock-and-key, induced-fit, and conformational selection) are rationalized based on the FEL theory are discussed in depth.
- 13Karplus, M.; Kuriyan, J. Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 6679– 6685 DOI: 10.1073/pnas.0408930102Google ScholarThere is no corresponding record for this reference.
- 14Lam, A. J.; St-Pierre, F.; Gong, Y.; Marshall, J. D.; Cranfill, P. J.; Baird, M. A.; McKeown, M. R.; Wiedenmann, J.; Davidson, M. W.; Schnitzer, M. J.; Tsien, R. Y.; Lin, M. Z. Nat. Methods 2012, 9, 1005– 1012 DOI: 10.1038/nmeth.2171Google Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhtlajs7zI&md5=4f780d655ff3cdada989b5908cefbc67Improving FRET dynamic range with bright green and red fluorescent proteinsLam, Amy J.; St-Pierre, Francois; Gong, Yiyang; Marshall, Jesse D.; Cranfill, Paula J.; Baird, Michelle A.; McKeown, Michael R.; Wiedenmann, Joerg; Davidson, Michael W.; Schnitzer, Mark J.; Tsien, Roger Y.; Lin, Michael Z.Nature Methods (2012), 9 (10), 1005-1012CODEN: NMAEA3; ISSN:1548-7091. (Nature Publishing Group)A variety of genetically encoded reporters use changes in fluorescence (or Foerster) resonance energy transfer (FRET) to report on biochem. processes in living cells. The std. genetically encoded FRET pair consists of CFPs and YFPs, but many CFP-YFP reporters suffer from low FRET dynamic range, phototoxicity from the CFP excitation light and complex photokinetic events such as reversible photobleaching and photoconversion. We engineered two fluorescent proteins, Clover and mRuby2, which are the brightest green and red fluorescent proteins to date and have the highest Foerster radius of any ratiometric FRET pair yet described. Replacement of CFP and YFP with these two proteins in reporters of kinase activity, small GTPase activity and transmembrane voltage significantly improves photostability, FRET dynamic range and emission ratio changes. These improvements enhance detection of transient biochem. events such as neuronal action-potential firing and RhoA activation in growth cones.
- 15Schuler, B. J. Nanobiotechnol. 2013, 11 (Suppl 1) S2 DOI: 10.1186/1477-3155-11-S1-S2Google ScholarThere is no corresponding record for this reference.
- 16Nesmelov, Y. E.; Thomas, D. D. Biophys. Rev. 2010, 2, 91– 99 DOI: 10.1007/s12551-010-0032-5Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXlsVSksrw%253D&md5=324f9f0b67f707de61088b4fb823fa29Protein structural dynamics revealed by site-directed spin labeling and multifrequency EPRNesmelov, Yuri E.; Thomas, David D.Biophysical Reviews (2010), 2 (2), 91-99CODEN: BRIECG; ISSN:1867-2450. (Springer)A review. Multifrequency ESR (EPR), combined with site-directed spin labeling, is a powerful spectroscopic tool to characterize protein dynamics. The lineshape of an EPR spectrum reflects combined rotational dynamics of the spin probe's local motion within a protein, reorientations of protein domains, and overall protein tumbling. All these motions can be restricted and anisotropic, and sepn. of these motions is important for thorough characterization of protein dynamics. Multifrequency EPR distinguishes between different motions of a spin-labeled protein, due to the frequency dependence of EPR resoln. to fast and slow motion of a spin probe. This gives multifrequency EPR its unique capability to characterize protein dynamics in great detail. In this review, we analyze what makes multifrequency EPR sensitive to different rates of spin probe motion and discuss several examples of its usage to sep. spin probe dynamics and overall protein dynamics, to characterize protein backbone dynamics, and to resolve protein conformational states.
- 17Fleissner, M. R.; Bridges, M. D.; Brooks, E. K.; Cascio, D.; Kalai, T.; Hideg, K.; Hubbell, W. L. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 16241– 16246 DOI: 10.1073/pnas.1111420108Google ScholarThere is no corresponding record for this reference.
- 18Allerhand, A.; Doddrell, D.; Glushko, V.; Cochran, D. W.; Wenkert, E.; Lawson, P. J.; Gurd, F. R. J. Am. Chem. Soc. 1971, 93, 544– 546 DOI: 10.1021/ja00731a053Google ScholarThere is no corresponding record for this reference.
- 19Levy, R. M.; Karplus, M.; Mccammon, J. A. J. Am. Chem. Soc. 1981, 103, 994– 996 DOI: 10.1021/ja00394a072Google ScholarThere is no corresponding record for this reference.
- 20Deverell, C.; Morgan, R. E.; Strange, J. H. Mol. Phys. 1970, 18, 553– 559 DOI: 10.1080/00268977000100611Google ScholarThere is no corresponding record for this reference.
- 21Wittebort, R. J.; Szabo, A. J. Chem. Phys. 1978, 69, 1722 DOI: 10.1063/1.436748Google ScholarThere is no corresponding record for this reference.
- 22Carr, H. Y.; Purcell, E. M. Phys. Rev. 1954, 94, 630– 638 DOI: 10.1103/PhysRev.94.630Google Scholar22https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaG2cXksFKjtA%253D%253D&md5=f729c8b5191575d29fdfd489021f05b7Effects of diffusion on free precession in nuclear magnetic resonance experimentsCarr, H. Y.; Purcell, E. M.Physical Review (1954), 94 (), 630-8CODEN: PHRVAO; ISSN:0031-899X.Nuclear resonance techniques involving free precession are examd., and a convenient variation of Hahn's spin-echo method is described (ibid. 80, 580(1950)). This variation employs a combination of pulses of different intensity or duration ("90°" and "180°" pulses). Measurements of the transverse relaxation time T2 in fluids are often severely compromised by mol. diffusion. Hahn's analysis of the effect of diffusion is reformulated and extended, and a new scheme for measuring T2 is described which, as predicted by the extended theory, largely circumvents the diffusion effect. On the other hand, the free precession technique, applied in a different way, permits a direct measurement of the mol. self-diffusion const. in suitable fluids. A measurement of the self-diffusion const. of H2O at 25° yields D = 2.5 ± 0.3 × 10-5 sq. cm./sec., in good agreement with previous detns. The effect of convection on free precession is also analyzed. A null method for measuring the longitudinal relaxation time T1, based on the unequal-pulse technique, is described.
- 23Luz, Z.; Meiboom, S. J. Chem. Phys. 1963, 39, 366– 370 DOI: 10.1063/1.1734254Google ScholarThere is no corresponding record for this reference.
- 24Loria, J. P.; Rance, M.; Palmer, A. G. J. Am. Chem. Soc. 1999, 121, 2331– 2332 DOI: 10.1021/ja983961aGoogle ScholarThere is no corresponding record for this reference.
- 25Tolman, J. R.; Flanagan, J. M.; Kennedy, M. A.; Prestegard, J. H. Proc. Natl. Acad. Sci. U. S. A. 1995, 92, 9279– 9283 DOI: 10.1073/pnas.92.20.9279Google ScholarThere is no corresponding record for this reference.
- 26Saupe, A.; Englert, G. Phys. Rev. Lett. 1963, 11, 462– 464 DOI: 10.1103/PhysRevLett.11.462Google ScholarThere is no corresponding record for this reference.
- 27Woodward, C. K.; Hilton, B. D. Annu. Rev. Biophys. Bioeng. 1979, 8, 99– 127 DOI: 10.1146/annurev.bb.08.060179.000531Google Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE1MXksFWitLo%253D&md5=9c5ffcdf75aef65a03184d7ea950259aHydrogen exchange kinetics and internal motions in proteins and nucleic acidsWoodward, Clare K.; Hilton, Bruce D.Annual Review of Biophysics and Bioengineering (1979), 8 (), 99-127CODEN: ABPBBK; ISSN:0084-6589.A review with 188 refs.
- 28Muhandiram, D. R.; Yamazaki, T.; Sykes, B. D.; Kay, L. E. J. Am. Chem. Soc. 1995, 117, 11536– 11544 DOI: 10.1021/ja00151a018Google ScholarThere is no corresponding record for this reference.
- 29Gobl, C.; Madl, T.; Simon, B.; Sattler, M. Prog. Nucl. Magn. Reson. Spectrosc. 2014, 80, 26– 63 DOI: 10.1016/j.pnmrs.2014.05.003Google Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXpslOis7g%253D&md5=b8b4e2b127b99a42e9973f960aeb7b7cNMR approaches for structural analysis of multidomain proteins and complexes in solutionGobl, Christoph; Madl, Tobias; Simon, Bernd; Sattler, MichaelProgress in Nuclear Magnetic Resonance Spectroscopy (2014), 80 (), 26-63CODEN: PNMRAT; ISSN:0079-6565. (Elsevier B.V.)A review. NMR spectroscopy is a key method for studying the structure and dynamics of (large) multidomain proteins and complexes in soln. It plays a unique role in integrated structural biol. approaches as esp. information about conformational dynamics can be readily obtained at residue resoln. Here, we review NMR techniques for such studies focusing on state-of-the-art tools and practical aspects. An efficient approach for detg. the quaternary structure of multidomain complexes starts from the structures of individual domains or subunits. The arrangement of the domains/subunits within the complex is then defined based on NMR measurements that provide information about the domain interfaces combined with (long-range) distance and orientational restraints. Aspects discussed include sample prepn., specific isotope labeling and spin labeling; detn. of binding interfaces and domain/subunit arrangements from chem. shift perturbations (CSP), nuclear Overhauser effects (NOEs), isotope editing/filtering, cross-satn., and differential line broadening; and based on paramagnetic relaxation enhancements (PRE) using covalent and sol. spin labels. Finally, the utility of complementary methods such as small-angle X-ray or neutron scattering (SAXS, SANS), ESR (EPR) or fluorescence spectroscopy techniques is discussed. The applications of NMR techniques are illustrated with studies of challenging (high mol. wt.) protein complexes.
- 30Mittermaier, A.; Kay, L. E. Science 2006, 312, 224– 228 DOI: 10.1126/science.1124964Google ScholarThere is no corresponding record for this reference.
- 31Sheppard, D.; Sprangers, R.; Tugarinov, V. Prog. Nucl. Magn. Reson. Spectrosc. 2010, 56, 1– 45 DOI: 10.1016/j.pnmrs.2009.07.004Google ScholarThere is no corresponding record for this reference.
- 32Göbl, C.; Tjandra, N. Entropy 2012, 14, 581– 598 DOI: 10.3390/e14030581Google ScholarThere is no corresponding record for this reference.
- 33Li, F.; Grishaev, A.; Ying, J.; Bax, A. J. Am. Chem. Soc. 2015, 137, 14798– 14811 DOI: 10.1021/jacs.5b10072Google ScholarThere is no corresponding record for this reference.
- 34Im, W.; Jo, S.; Kim, T. Biochim. Biophys. Acta, Biomembr. 2012, 1818, 252– 262 DOI: 10.1016/j.bbamem.2011.07.048Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XkslKrsw%253D%253D&md5=4b07dd2cf3479c393f3903d7f63871d2An ensemble dynamics approach to decipher solid-state NMR observables of membrane proteinsIm, Wonpil; Jo, Sunhwan; Kim, TaehoonBiochimica et Biophysica Acta, Biomembranes (2012), 1818 (2), 252-262CODEN: BBBMBS; ISSN:0005-2736. (Elsevier B.V.)A review. Solid-state NMR (SSNMR) is an invaluable tool for detg. orientations of membrane proteins and peptides in lipid bilayers. Such orientational descriptions provide essential information about membrane protein functions. However, when a semi-static single conformer model is used to interpret various SSNMR observables, important dynamics information can be missing, and, sometimes, even orientational information can be misinterpreted. In addn., over the last decade, mol. dynamics (MD) simulation and semi-static SSNMR interpretation have shown certain levels of discrepancies in terms of transmembrane helix orientation and dynamics. Dynamic fitting models have recently been proposed to resolve these discrepancies by taking into account transmembrane helix whole body motions using addnl. parameters. As an alternative approach, the authors have developed SSNMR ensemble dynamics (SSNMR-ED) using multiple conformer models, which generates an ensemble of structures that satisfies the exptl. observables without any fitting parameters. In this review, various computational methods for detg. transmembrane helix orientations are discussed, and the distributions of VpuTM (from HIV-1) and WALP23 (a synthetic peptide) orientations from SSNMR-ED simulations are compared with those from MD simulations and semi-static/dynamic fitting models. Such comparisons illustrate that SSNMR-ED can be used as a general means to ext. both membrane protein structure and dynamics from the SSNMR measurements. This article is part of a Special Issue entitled: Membrane protein structure and function.
- 35Hu, F.; Luo, W.; Hong, M. Science 2010, 330, 505– 508 DOI: 10.1126/science.1191714Google ScholarThere is no corresponding record for this reference.
- 36Bocian, D. F.; Chan, S. I. Annu. Rev. Phys. Chem. 1978, 29, 307– 335 DOI: 10.1146/annurev.pc.29.100178.001515Google ScholarThere is no corresponding record for this reference.
- 37Lewandowski, J. R.; Sass, H. J.; Grzesiek, S.; Blackledge, M.; Emsley, L. J. Am. Chem. Soc. 2011, 133, 16762– 16765 DOI: 10.1021/ja206815hGoogle ScholarThere is no corresponding record for this reference.
- 38Mollica, L.; Baias, M.; Lewandowski, J. R.; Wylie, B. J.; Sperling, L. J.; Rienstra, C. M.; Emsley, L.; Blackledge, M. J. Phys. Chem. Lett. 2012, 3, 3657– 3662 DOI: 10.1021/jz3016233Google Scholar38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhslaksrrE&md5=4de33d158064b091ad375dd5ef2af54bAtomic-Resolution Structural Dynamics in Crystalline Proteins from NMR and Molecular SimulationMollica, Luca; Baias, Maria; Lewandowski, Jozef R.; Wylie, Benjamin J.; Sperling, Lindsay J.; Rienstra, Chad M.; Emsley, Lyndon; Blackledge, MartinJournal of Physical Chemistry Letters (2012), 3 (23), 3657-3662CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)Solid-state NMR can provide at.-resoln. information about protein motions occurring on a vast range of time scales under similar conditions to those of X-ray diffraction studies and therefore offers a highly complementary approach to characterizing the dynamic fluctuations occurring in the crystal. We compare exptl. detd. dynamic parameters, spin relaxation, chem. shifts, and dipolar couplings, to values calcd. from a 200 ns MD simulation of protein GB1 in its cryst. form, providing insight into the nature of structural dynamics occurring within the cryst. lattice. This simulation allows us to test the accuracy of commonly applied procedures for the interpretation of exptl. solid-state relaxation data in terms of dynamic modes and time scales. We discover that the potential complexity of relaxation-active motion can lead to significant under- or overestimation of dynamic amplitudes if different components are not taken into consideration.
- 39Lewandowski, J. R. Acc. Chem. Res. 2013, 46, 2018– 2027 DOI: 10.1021/ar300334gGoogle ScholarThere is no corresponding record for this reference.
- 40Krushelnitsky, A.; Zinkevich, T.; Reif, B.; Saalwachter, K. J. Magn. Reson. 2014, 248, 8– 12 DOI: 10.1016/j.jmr.2014.09.007Google ScholarThere is no corresponding record for this reference.
- 41Lewandowski, J. R.; Halse, M. E.; Blackledge, M.; Emsley, L. Science 2015, 348, 578– 581 DOI: 10.1126/science.aaa6111Google ScholarThere is no corresponding record for this reference.
- 42Tollinger, M.; Sivertsen, A. C.; Meier, B. H.; Ernst, M.; Schanda, P. J. Am. Chem. Soc. 2012, 134, 14800– 14807 DOI: 10.1021/ja303591yGoogle ScholarThere is no corresponding record for this reference.
- 43Agarwal, V.; Xue, Y.; Reif, B.; Skrynnikov, N. R. J. Am. Chem. Soc. 2008, 130, 16611– 16621 DOI: 10.1021/ja804275pGoogle ScholarThere is no corresponding record for this reference.
- 44Haller, J. D.; Schanda, P. J. Biomol. NMR 2013, 57, 263– 280 DOI: 10.1007/s10858-013-9787-xGoogle ScholarThere is no corresponding record for this reference.
- 45Chevelkov, V.; Xue, Y.; Linser, R.; Skrynnikov, N. R.; Reif, B. J. Am. Chem. Soc. 2010, 132, 5015– 5017 DOI: 10.1021/ja100645kGoogle ScholarThere is no corresponding record for this reference.
- 46Hohwy, M.; Jaroniec, C. P.; Reif, B.; Rienstra, C. M.; Griffin, R. G. J. Am. Chem. Soc. 2000, 122, 3218– 3219 DOI: 10.1021/ja9913737Google ScholarThere is no corresponding record for this reference.
- 47Vinogradov, E.; Madhu, P. K.; Vega, S. J. Chem. Phys. 2001, 115, 8983 DOI: 10.1063/1.1408287Google ScholarThere is no corresponding record for this reference.
- 48Gullion, T.; Schaefer, J. J. Magn. Reson. 1989, 81, 196– 200 DOI: 10.1016/0022-2364(89)90280-1Google Scholar48https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXkvV2hsr8%253D&md5=cd503257b5d0237a5f8bd549517a3a76Rotational-echo double-resonance NMRGullion, Terry; Schaefer, JacobJournal of Magnetic Resonance (1969-1992) (1989), 81 (1), 196-200CODEN: JOMRA4; ISSN:0022-2364.Rotational-echo double-resonance (REDOR), a spin-echo double-resonance (SEDOR) NMR expt. with magic-angle sample spinning, is illustrated by using the results of expts. performed on 13C- and 15N-labeled alanines. The pulse sequence for 13C-15N REDOR is shown.
- 49Gullion, T.; Schaefer, J. Adv. Magn. Opt. Reson. 1989, 13, 57– 83 DOI: 10.1016/B978-0-12-025513-9.50009-4Google ScholarThere is no corresponding record for this reference.
- 50Jaroniec, C. P.; Tounge, B. A.; Rienstra, C. M.; Herzfeld, J.; Griffin, R. G. J. Magn. Reson. 2000, 146, 132– 139 DOI: 10.1006/jmre.2000.2128Google ScholarThere is no corresponding record for this reference.
- 51Schanda, P.; Meier, B. H.; Ernst, M. J. Magn. Reson. 2011, 210, 246– 259 DOI: 10.1016/j.jmr.2011.03.015Google ScholarThere is no corresponding record for this reference.
- 52Dvinskikh, S. V.; Zimmermann, H.; Maliniak, A.; Sandström, D. J. Magn. Reson. 2003, 164, 165– 170 DOI: 10.1016/S1090-7807(03)00180-0Google ScholarThere is no corresponding record for this reference.
- 53Dvinskikh, S. V.; Zimmermann, H.; Maliniak, A.; Sandstrom, D. J. Chem. Phys. 2005, 122, 44512 DOI: 10.1063/1.1834569Google ScholarThere is no corresponding record for this reference.
- 54Levitt, M. H. In Encyclopedia of Nuclear Magnetic Resonance; Grant, D. M.; Harris, R. K., Eds.; John Wiley & Sons: Chichester U.K., 2002; Vol. 9, pp 165– 196.Google ScholarThere is no corresponding record for this reference.
- 55Lorieau, J. L.; Day, L. A.; McDermott, A. E. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, 10366– 10371 DOI: 10.1073/pnas.0800405105Google ScholarThere is no corresponding record for this reference.
- 56Lorieau, J. L.; McDermott, A. E. J. Am. Chem. Soc. 2006, 128, 11505– 11512 DOI: 10.1021/ja062443uGoogle ScholarThere is no corresponding record for this reference.
- 57Zhao, X.; Sudmeier, J. L.; Bachovchin, W. W.; Levitt, M. H. J. Am. Chem. Soc. 2001, 123, 11097– 11098 DOI: 10.1021/ja016328pGoogle ScholarThere is no corresponding record for this reference.
- 58Schanda, P.; Meier, B. H.; Ernst, M. J. Am. Chem. Soc. 2010, 132, 15957– 15967 DOI: 10.1021/ja100726aGoogle ScholarThere is no corresponding record for this reference.
- 59Yang, J.; Tasayco, M. L.; Polenova, T. J. Am. Chem. Soc. 2009, 131, 13690– 13702 DOI: 10.1021/ja9037802Google ScholarThere is no corresponding record for this reference.
- 60Chevelkov, V.; Fink, U.; Reif, B. J. Am. Chem. Soc. 2009, 131, 14018– 14022 DOI: 10.1021/ja902649uGoogle ScholarThere is no corresponding record for this reference.
- 61Meirovitch, E.; Liang, Z.; Freed, J. H. J. Phys. Chem. B 2015, 119, 2857– 2868 DOI: 10.1021/jp511386bGoogle ScholarThere is no corresponding record for this reference.
- 62Torchia, D. A. Annu. Rev. Biophys. Bioeng. 1984, 13, 125– 144 DOI: 10.1146/annurev.bb.13.060184.001013Google Scholar62https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2cXkvVKnu7o%253D&md5=e8be982f7ae0fb3e1d33b924886c4751Solid state NMR studies of protein internal dynamicsTorchia, Dennis A.Annual Review of Biophysics and Bioengineering (1984), 13 (), 125-44CODEN: ABPBBK; ISSN:0084-6589.A review with 65 refs.
- 63Torchia, D. A.; Szabo, A. J. Magn. Reson. 1982, 49, 107– 121 DOI: 10.1016/0022-2364(82)90301-8Google ScholarThere is no corresponding record for this reference.
- 64Olympia, P. L.; Wei, I. Y.; Fung, B. M. J. Chem. Phys. 1969, 51, 1610– 1614 DOI: 10.1063/1.1672220Google ScholarThere is no corresponding record for this reference.
- 65Hologne, M.; Chevelkov, V.; Reif, B. Prog. Nucl. Magn. Reson. Spectrosc. 2006, 48, 211– 232 DOI: 10.1016/j.pnmrs.2006.05.004Google ScholarThere is no corresponding record for this reference.
- 66Jelinski, L. W. Annu. Rev. Mater. Sci. 1985, 15, 359– 377 DOI: 10.1146/annurev.ms.15.080185.002043Google Scholar66https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2MXlvFKmurY%253D&md5=1b3fd9ec31842011ea6db576e7f43956Solid state deuterium NMR studies of polymer chain dynamicsJelinski, Lynn W.Annual Review of Materials Science (1985), 15 (), 359-77CODEN: ARMSCX; ISSN:0084-6600.A review with 49 refs. on solid-state D-NMR expts. used for detn. of local mol. dynamics of polymers. Also, briefly reviewed were D-NMR applications on anal. of polymeric liq. crystals, phase sepn. in polymers, and polyurethane structural anal.
- 67Hologne, M.; Hirschinger, J. Solid State Nucl. Magn. Reson. 2004, 26, 1– 10 DOI: 10.1016/S0926-2040(03)00062-6Google Scholar67https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXkt1Gjtro%253D&md5=488b9d0e6297c31e1f3664e095bc3f2fMolecular dynamics as studied by static-powder and magic-angle spinning 2H NMRHologne, Maggy; Hirschinger, JeromeSolid State Nuclear Magnetic Resonance (2004), 26 (1), 1-10CODEN: SSNRE4; ISSN:0926-2040. (Elsevier Science)The 2H NMR magic-angle spinning (MAS) technique is compared to the static-powder quadrupole echo (QE) and Jeener-Brockaert (JB) pulse sequences for a quant. investigation of mol. dynamics in solids. The linewidth of individual spinning sidebands of the one-dimensional MAS spectra are obsd. to be characteristic of the correlation time from ∼10-2 to ∼10-8 s so that the dynamic range is increased by approx. three orders of magnitude when compared to the QE expt. As a consequence, MAS 2H NMR is found to be more sensitive to the presence of an inhomogeneous distribution of correlation times than the QE and JB expts. which rely upon lineshape distortions due to anisotropic T2 and T1Q relaxation, resp. All these results are demonstrated exptl. and numerically using the two-site flip motion of di-Me sulfone and of the nitrobenzene guest in the α-p-tert-butylcalix[4]arene-nitrobenzene inclusion compd.
- 68Batchelder, L. S.; Niu, C. H.; Torchia, D. A. J. Am. Chem. Soc. 1983, 105, 2228– 2231 DOI: 10.1021/ja00346a021Google ScholarThere is no corresponding record for this reference.
- 69Pines, A.; Gibby, M. G.; Waugh, J. S. J. Chem. Phys. 1973, 59, 569– 590 DOI: 10.1063/1.1680061Google Scholar69https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE3sXkvVGgsLo%253D&md5=43fbff1149b7d4de334d6168521cc10cProton-enhanced NMR of dilute spins in solidsPines, A.; Gibby, M. G.; Waugh, J. S.Journal of Chemical Physics (1973), 59 (2), 569-90CODEN: JCPSA6; ISSN:0021-9606.The NMR signals of isotopically or chem. dil. nuclear spins S in solids can be enhanced by repeatedly transferring polarization from a more abundant species I of high abundance (usually protons) to which they are coupled. The gain in power sensitivity as compared with conventional observation of the rare spins approaches NII(I + 1)γI2/NSS(S + 1)γS2, or ∼103 for S = 13C, I = H in org. solids. The transfer of polarization is accomplished by any of a no. of double resonance methods. High-frequency resoln. of the S-spin signal was obtained by decoupling of the abundant spins. The exptl. requirements of the technique are discussed and a brief comparison of its sensitivity with other procedures is made. Representative applications and exptl. results are mentioned.
- 70Hartmann, S. R.; Hahn, E. L. Phys. Rev. 1962, 128, 2042– 2053 DOI: 10.1103/PhysRev.128.2042Google ScholarThere is no corresponding record for this reference.
- 71Hologne, M.; Chen, Z.; Reif, B. J. Magn. Reson. 2006, 179, 20– 28 DOI: 10.1016/j.jmr.2005.10.014Google ScholarThere is no corresponding record for this reference.
- 72Shi, X.; Yarger, J. L.; Holland, G. P. J. Magn. Reson. 2013, 226, 1– 12 DOI: 10.1016/j.jmr.2012.10.013Google ScholarThere is no corresponding record for this reference.
- 73Hologne, M.; Faelber, K.; Diehl, A.; Reif, B. J. Am. Chem. Soc. 2005, 127, 11208– 11209 DOI: 10.1021/ja051830lGoogle ScholarThere is no corresponding record for this reference.
- 74Shi, X.; Yarger, J. L.; Holland, G. P. Chem. Commun. 2014, 50, 4856– 4859 DOI: 10.1039/c4cc00971aGoogle ScholarThere is no corresponding record for this reference.
- 75Shi, X.; Holland, G. P.; Yarger, J. L. Biomacromolecules 2015, 16, 852– 859 DOI: 10.1021/bm5017578Google ScholarThere is no corresponding record for this reference.
- 76Wei, D.; Akbey, U. m.; Paaske, B.; Oschkinat, H.; Reif, B.; Bjerring, M.; Nielsen, N. C. J. Phys. Chem. Lett. 2011, 2, 1289– 1294 DOI: 10.1021/jz200511bGoogle Scholar76https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXlvFSqtLw%253D&md5=9804df3d0cfc0853c94e0d7e503df8b5Optimal 2H rf Pulses and 2H-13C Cross-Polarization Methods for Solid-State 2H MAS NMR of Perdeuterated ProteinsWei, Daxiu; Akbey, Umit; Paaske, Berit; Oschkinat, Hartmut; Reif, Bernd; Bjerring, Morten; Nielsen, Niels Chr.Journal of Physical Chemistry Letters (2011), 2 (11), 1289-1294CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)We present a novel concept for rf pulses and optimal control designed cross-polarization expts. for quadrupolar nuclei. The methods are demonstrated for 2H CP-MAS and 2H multiple-pulse NMR of perdeuterated proteins, for which sensitivity enhancements up to an order of magnitude are presented relative to commonly used approaches. The so-called RESPIRATION rf pulses combines the concept of short broad-band pulses with generation of pulses with large flip angles through distribution of the rf pulse over several rotor echoes. This leads to close-to-ideal rf pulses, facilitating implementation of expts. relying on the ability to realize high-performance 90 and 180° pulses, as, for example, in refocused INEPT and double-to-single quantum coherence expts., or just pulses that provide a true representation of the quadrupolar powder pattern to ext. information about the structure or dynamics. The optimal control 2H → 13C CP-MAS method demonstrates transfer efficiencies up to around 85% while being extremely robust toward rf inhomogeneity and resonance offsets.
- 77Nielsen, A. B.; Jain, S.; Ernst, M.; Meier, B. H.; Nielsen, N. C. J. Magn. Reson. 2013, 237, 147– 151 DOI: 10.1016/j.jmr.2013.09.002Google ScholarThere is no corresponding record for this reference.
- 78Jain, S.; Bjerring, M.; Nielsen, N. C. J. Phys. Chem. Lett. 2012, 3, 703– 708 DOI: 10.1021/jz3000905Google Scholar78https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XislSqtLk%253D&md5=16f412f704236445de9c532d91ebe4caEfficient and Robust Heteronuclear Cross-Polarization for High-Speed-Spinning Biological Solid-State NMR SpectroscopyJain, Sheetal; Bjerring, Morten; Nielsen, Niels Chr.Journal of Physical Chemistry Letters (2012), 3 (6), 703-708CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)The authors present a new and highly efficient approach for heteronuclear coherence transfer in solid-state NMR spectroscopy under high-speed spinning conditions. The so-called RESPIRATIONCP expt. exploits phase-alternated recoupling on only one of the two rf channels intertwined in a synchronized train of short rf pulses on both channels. The method provides significantly higher efficiencies than state-of-the art techniques including ramped and adiabatic cross-polarization expts. with long durations of intense rf irradn. At the same time, it is easier to setup exptl. and significantly more robust toward imperfections such as rf inhomogeneity, misadjustments, and sample-induced variations in the rf tuning. The method is described anal., numerically, and exptl. for biol. solids. The authors demonstrate sensitivity gains of factors of 1.3 and 1.8 for typical 1H→15N and 15N→13C transfers and a combined gain of a factor of 2-4 for a typical NCA expt. for biol. solid-state NMR.
- 79Jain, S. K.; Nielsen, A. B.; Hiller, M.; Handel, L.; Ernst, M.; Oschkinat, H.; Akbey, U.; Nielsen, N. C. Phys. Chem. Chem. Phys. 2014, 16, 2827– 2830 DOI: 10.1039/c3cp54419bGoogle ScholarThere is no corresponding record for this reference.
- 80Hohwy, M.; Rienstra, C. M.; Griffin, R. G. J. Chem. Phys. 2002, 117, 4973– 4987 DOI: 10.1063/1.1488136Google ScholarThere is no corresponding record for this reference.
- 81Hohwy, M.; Rienstra, C. M.; Jaroniec, C. P.; Griffin, R. G. J. Chem. Phys. 1999, 110, 7983– 7992 DOI: 10.1063/1.478702Google ScholarThere is no corresponding record for this reference.
- 82Mittermaier, A.; Kay, L. E. J. Am. Chem. Soc. 1999, 121, 10608– 10613 DOI: 10.1021/ja9925047Google ScholarThere is no corresponding record for this reference.
- 83Sheppard, D.; Li, D. W.; Bruschweiler, R.; Tugarinov, V. J. Am. Chem. Soc. 2009, 131, 15853– 15865 DOI: 10.1021/ja9063958Google ScholarThere is no corresponding record for this reference.
- 84Franks, W. T.; Zhou, D. H.; Wylie, B. J.; Money, B. G.; Graesser, D. T.; Frericks, H. L.; Sahota, G.; Rienstra, C. M. J. Am. Chem. Soc. 2005, 127, 12291– 12305 DOI: 10.1021/ja044497eGoogle ScholarThere is no corresponding record for this reference.
- 85Van Geet, A. L. Anal. Chem. 1968, 40, 2227– 2229 DOI: 10.1021/ac50158a064Google ScholarThere is no corresponding record for this reference.
- 86Ernst, M.; Samoson, A.; Meier, B. H. J. Magn. Reson. 2003, 163, 332– 339 DOI: 10.1016/S1090-7807(03)00155-1Google ScholarThere is no corresponding record for this reference.
- 87Shaka, A. J.; Keeler, J.; Freeman, R. J. Magn. Reson. 1983, 53, 313– 340 DOI: 10.1016/0022-2364(83)90035-5Google ScholarThere is no corresponding record for this reference.
- 88Delaglio, F.; Grzesiek, S.; Vuister, G. W.; Zhu, G.; Pfeifer, J.; Bax, A. J. Biomol. NMR 1995, 6, 277– 293 DOI: 10.1007/BF00197809Google Scholar88https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXhtVSmurfK&md5=a670fca5b164083e2178fafd2fb951ffNMRPipe: a multidimensional spectral processing system based on UNIX pipesDelaglio, Frank; Grzesiek, Stephan; Vuister, Geerten W.; Zhu, Guang; Pfeifer, John; Bax, AdJournal of Biomolecular NMR (1995), 6 (3), 277-93CODEN: JBNME9; ISSN:0925-2738. (ESCOM)The NMRPipe system is a UNIX software environment of processing, graphics, and anal. tools designed to meet current routine and research-oriented multidimensional processing requirements, and to anticipate and accommodate future demands and developments. The system is based on UNIX pipes, which allow programs running simultaneously to exchange streams of data under user control. In an NMRPipe processing scheme, a stream of spectral data flows through a pipeline of processing programs, each of which performs one component of the overall scheme, such as Fourier transformation or linear prediction. Complete multidimensional processing schemes are constructed as simple UNIX shell scripts. The processing modules themselves maintain and exploit accurate records of data sizes, detection modes, and calibration information in all dimensions, so that schemes can be constructed without the need to explicitly define or anticipate data sizes or storage details of real and imaginary channels during processing. The asynchronous pipeline scheme provides other substantial advantages, including high flexibility, favorable processing speeds, choice of both all-in-memory and disk-bound processing, easy adaptation to different data formats, simpler software development and maintenance, and the ability to distribute processing tasks on multi-CPU computers and computer networks.
- 89Massiot, D.; Fayon, F.; Capron, M.; King, I.; Le Calve, S.; Alonso, B.; Durand, J. O.; Bujoli, B.; Gan, Z. H.; Hoatson, G. Magn. Reson. Chem. 2002, 40, 70– 76 DOI: 10.1002/mrc.984Google Scholar89https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38Xlt1ajuw%253D%253D&md5=d32cb46ca90ac83a8f2d62c95a93a88fModeling one- and two-dimensional solid-state NMR spectraMassiot, Dominique; Fayon, Franck; Capron, Mickael; King, Ian; Le Calve, Stephanie; Alonso, Bruno; Durand, Jean-Olivier; Bujoli, Bruno; Gan, Zhehong; Hoatson, GinaMagnetic Resonance in Chemistry (2002), 40 (1), 70-76CODEN: MRCHEG; ISSN:0749-1581. (John Wiley & Sons Ltd.)A review. With the description of more and more complex 1- and two-dimensional NMR expts. comes the need to develop methods to make a comprehensive interpretation of the various different expts. that can be carried out on the same sample or series of related samples. The authors present some examples of modeling 1- and two-dimensional solid-state NMR spectra of I = 1/2 spin and quadrupolar nuclei, using lab.-developed software that is made available to the NMR community.
- 90Schmidt, H. L. F.; Sperling, L. J.; Gao, Y. G.; Wylie, B. J.; Boettcher, J. M.; Wilson, S. R.; Rienstra, C. A. J. Phys. Chem. B 2007, 111, 14362– 14369 DOI: 10.1021/jp075531pGoogle ScholarThere is no corresponding record for this reference.
- 91Wylie, B. J.; Sperling, L. J.; Nieuwkoop, A. J.; Franks, W. T.; Oldfield, E.; Rienstra, C. M. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 16974– 16979 DOI: 10.1073/pnas.1103728108Google ScholarThere is no corresponding record for this reference.
- 92Schmidt, H. L.; Shah, G. J.; Sperling, L. J.; Rienstra, C. M. J. Phys. Chem. Lett. 2010, 1, 1623– 1628 DOI: 10.1021/jz1004413Google Scholar92https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC2srmtFehtA%253D%253D&md5=574992a689fabc298bf9be2a8fc6da69NMR Determination of Protein pK(a) Values in the Solid StateSchmidt Heather L Frericks; Shah Gautam J; Sperling Lindsay J; Rienstra Chad MThe journal of physical chemistry letters (2010), 1 (10), 1623-1628 ISSN:.Charged residues play an important role in defining key mechanistic features in many biomolecules. Determining the pK(a) values of large, membrane or fibrillar proteins can be challenging with traditional methods. In this study we show how solid-state NMR is used to monitor chemical shift changes during a pH titration for the small soluble β1 immunoglobulin binding domain of protein G. The chemical shifts of all the amino acids with charged side-chains throughout the uniformly-(13)C,(15)N-labeled protein were monitored over several samples varying in pH; pK(a) values were determined from these shifts for E27, D36, and E42, and the bounds for the pK(a) of other acidic side-chain resonances were determined. Additionally, this study shows how the calculated pK(a) values give insights into the crystal packing of the protein.
- 93Salonen, L. M.; Ellermann, M.; Diederich, F. Angew. Chem., Int. Ed. 2011, 50, 4808– 4842 DOI: 10.1002/anie.201007560Google Scholar93https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXlvVekt70%253D&md5=3684d00abc320d1ea0c71565495d5d5cAromatic rings in chemical and biological recognition: energetics and structuresSalonen, Laura M.; Ellermann, Manuel; Diederich, FrancoisAngewandte Chemie, International Edition (2011), 50 (21), 4808-4842CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. This review describes a multidimensional treatment of mol. recognition phenomena involving arom. rings in chem. and biol. systems. It summarizes new results reported since the appearance of an earlier review in 2003 in host-guest chem., biol. affinity assays and biostructural anal., data base mining in the Cambridge Structural Database (CSD) and the Protein Data Bank (PDB), and advanced computational studies. Topics addressed are arene-arene, perfluoroarene-arene, S···arom., cation-π, and anion-π interactions, as well as hydrogen bonding to π systems. The generated knowledge benefits, in particular, structure-based hit-to-lead development and lead optimization both in the pharmaceutical and in the crop protection industry. It equally facilitates the development of new advanced materials and supramol. systems, and should inspire further utilization of interactions with arom. rings to control the stereochem. outcome of synthetic transformations.
- 94Eswar, N.; Ramakrishnan, C. Protein Eng., Des. Sel. 2000, 13, 227– 238 DOI: 10.1093/protein/13.4.227Google ScholarThere is no corresponding record for this reference.
- 95Vijayakumar, M.; Qian, H.; Zhou, H. X. Proteins: Struct., Funct., Genet. 1999, 34, 497– 507 DOI: 10.1002/(SICI)1097-0134(19990301)34:4<497::AID-PROT9>3.0.CO;2-GGoogle ScholarThere is no corresponding record for this reference.
- 96Vugmeyster, L.; Ostrovsky, D.; Lipton, A. S. J. Phys. Chem. B 2013, 117, 6129– 6137 DOI: 10.1021/jp4021596Google ScholarThere is no corresponding record for this reference.
- 97Ottiger, M.; Bax, A. J. Am. Chem. Soc. 1999, 121, 4690– 4695 DOI: 10.1021/ja984484zGoogle ScholarThere is no corresponding record for this reference.
- 98Iijima, T.; Tsuchiya, S. J. Mol. Spectrosc. 1972, 44, 88– 107 DOI: 10.1016/0022-2852(72)90194-4Google ScholarThere is no corresponding record for this reference.
- 99Mittermaier, A.; Kay, L. E.; Forman-Kay, J. D. J. Biomol. NMR 1999, 13, 181– 185 DOI: 10.1023/A:1008387715167Google ScholarThere is no corresponding record for this reference.
- 100Sarkar, S. K.; Young, P. E.; Torchia, D. A. J. Am. Chem. Soc. 1986, 108, 6459– 6464 DOI: 10.1021/ja00281a002Google ScholarThere is no corresponding record for this reference.
- 101Barchi, J. J., Jr.; Grasberger, B.; Gronenborn, A. M.; Clore, G. M. Protein Sci. 1994, 3, 15– 21 DOI: 10.1002/pro.5560030103Google Scholar101https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXlt1yjtLw%253D&md5=1b2ff1da4f2eb56452cca955e9fd453bInvestigation of the backbone dynamics of the IgG-binding domain of streptococcal protein G by heteronuclear two-dimensional 1H-15N nuclear magnetic resonance spectroscopyBarchi, Joseph J., Jr.; Grasberger, Bruce; Gronenborn, Angela M.; Clore, G. MariusProtein Science (1994), 3 (1), 15-21CODEN: PRCIEI; ISSN:0961-8368.The backbone dynamics of the Ig-binding domain (B1) of streptococcal protein G, uniformly labeled with 15N, have been investigated by two-dimensional inverse detected heteronuclear 1H-15N NMR spectroscopy at 500 and 600 MHz. 15N T1, T2, and nuclear Overhauser enhancement data were obtained for all 55 backbone NH vectors of the B1 domain at both field strengths. The overall correlation time obtained from an anal. of the T1/T2 ratios was 3.3 ns at 26°. Overall, the B1 domain is a relatively rigid protein, consistent with the fact that over 95% of the residues participate in secondary structure, comprising a four-stranded sheer arranged in a -1, +3x, -1 topol., on top of which lies a single helix. Residues in the turns and loops connecting the elements of secondary structure tend to exhibit a higher degree of mobility on the picosecond time scale, as manifested by lower values of the overall order parameter. A no. of residues at the ends of the secondary structure elements display two distinct internal motions that are faster than the overall rotational correlation time: one is fast (<20 ps) and lies in the extreme narrowing limit, whereas the other is one to two orders of magnitude slower (1-3 ns) and lies outside the extreme narrowing limit. The slower motion can be explained by large-amplitude (20-40°) lumps in the N-H vectors between states with well-defined orientations that are stabilized by hydrogen bonds. In addn., residues in the helix and in the outer β-strands (particularly β-strand 2) display a small degree of chem. exchange line broadening, possibly due to a minor rotational motion of the helix relative to the sheer that curls around it.
- 102Seewald, M. J.; Pichumani, K.; Stowell, C.; Tibbals, B. V.; Regan, L.; Stone, M. J. Protein Sci. 2000, 9, 1177– 1193 DOI: 10.1110/ps.9.6.1177Google Scholar102https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXks1KqsL0%253D&md5=99d9f1cab2dfc1d2bc07a5f54bd341f6The role of backbone conformational heat capacity in protein stability: temperature dependent dynamics of the B1 domain of Streptococcal protein GSeewald, Michael J.; Pichumani, Kumar; Stowell, Cheri; Tibbals, Benjamin V.; Regan, Lynne; Stone, Martin J.Protein Science (2000), 9 (6), 1177-1193CODEN: PRCIEI; ISSN:0961-8368. (Cambridge University Press)The contributions of backbone NH group dynamics to the conformational heat capacity of the B1 domain of Streptococcal protein G have been estd. from the temp. dependence of 15N NMR-derived order parameters. Longitudinal (R1) and transverse (R2) relaxation rates, transverse cross-relation rates (ηxy), and steady state {1H}-15N nuclear Overhauser effects were measured at temps. of 0, 10, 20, 30, 40, and 50° for 89-100% of the backbone secondary amide nitrogen nuclei in the B1 domain. The ratio R2/ηxy was used to identify nuclei for which conformational exchange makes a significant contribution to R2. Relaxation data were fit to the extended model-free dynamics formalism, incorporating an axially sym. mol. rotational diffusion tensor. The temp. dependence of the order parameter (S2) was used to calc. the contribution of each NH group to conformational heat capacity (Cp) and a characteristic temp. (T*), representing the d. of conformational energy states accessible to each NH group. The heat capacities of the secondary structure regions of the B1 domain are significantly higher than those of comparable regions of other proteins, whereas the heat capacities of less structured regions are similar to those in other proteins. The higher local heat capacities are estd. to contribute up to ∼0.8 kJ/mol K to the total heat capacity of the B1 domain, without which the denaturation temp. would be ∼9° lower (78° rather than 87°). Thus, variation of backbone conformational heat capacity of native proteins may be a novel mechanism that contributes to high temp. stabilization of proteins.
- 103Idiyatullin, D.; Nesmelova, I.; Daragan, V. A.; Mayo, K. H. Protein Sci. 2003, 12, 914– 922 DOI: 10.1110/ps.0228703Google Scholar103https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXjtl2jt7w%253D&md5=6b5ab9f4ada5119f7396d3678d81eb53Comparison of 13CαH and 15NH backbone dynamics in protein GB1Idiyatullin, Djaudat; Nesmelova, Irina; Daragan, Vladimir A.; Mayo, Kevin H.Protein Science (2003), 12 (5), 914-922CODEN: PRCIEI; ISSN:0961-8368. (Cold Spring Harbor Laboratory Press)This study presents a site-resolved exptl. view of backbone CαH and NH internal motions in the 56-residue Ig-binding domain of streptococcal protein G, GB1. Using 13CαH and 15NH NMR relaxation data [T1, T2, and NOE] acquired at three resonance frequencies (1H frequencies of 500, 600, and 800 MHz), spectral d. functions were calcd. as F(ω) = 2ωJ(ω) to provide a model-independent way to visualize and analyze internal motional correlation time distributions for backbone groups in GB1. Line broadening in F(ω) curves indicates the presence of nanosecond time scale internal motions (0.8 to 5 nsec) for all CαH and NH groups. Deconvolution of F(ω) curves effectively separates overall tumbling and internal motional correlation time distributions to yield more accurate order parameters than detd. by using std. model free approaches. Compared to NH groups, CαH internal motions are more broadly distributed on the nanosecond time scale, and larger CαH order parameters are related to correlated bond rotations for CαH fluctuations. Motional parameters for NH groups are more structurally correlated, with NH order parameters for example, being larger for residues in more structured regions of β-sheet and helix and generally smaller for residues in the loop and turns. This is most likely related to the observation that NH order parameters are correlated to hydrogen bonding. This study contributes to the general understanding of protein dynamics and exemplifies an alternative and easier way to analyze NMR relaxation data.
- 104Schanda, P.; Huber, M.; Boisbouvier, J.; Meier, B. H.; Ernst, M. Angew. Chem., Int. Ed. 2011, 50, 11005– 11009 DOI: 10.1002/anie.201103944Google ScholarThere is no corresponding record for this reference.
- 105Derrick, J. P.; Wigley, D. B. Nature 1992, 359, 752– 754 DOI: 10.1038/359752a0Google ScholarThere is no corresponding record for this reference.
- 106Gronenborn, A. M.; Clore, G. M. J. Mol. Biol. 1993, 233, 331– 335 DOI: 10.1006/jmbi.1993.1514Google ScholarThere is no corresponding record for this reference.
- 107Lian, L. Y.; Barsukov, I. L.; Derrick, J. P.; Roberts, G. C. K. Nat. Struct. Biol. 1994, 1, 355– 357 DOI: 10.1038/nsb0694-355Google ScholarThere is no corresponding record for this reference.
- 108Lamley, J. M.; Iuga, D.; Oster, C.; Sass, H. J.; Rogowski, M.; Oss, A.; Past, J.; Reinhold, A.; Grzesiek, S.; Samoson, A.; Lewandowski, J. R. J. Am. Chem. Soc. 2014, 136, 16800– 16806 DOI: 10.1021/ja5069992Google ScholarThere is no corresponding record for this reference.
- 109Kato, K.; Lian, L. Y.; Barsukov, I. L.; Derrick, J. P.; Kim, H. H.; Tanaka, R.; Yoshino, A.; Shiraishi, M.; Shimada, I.; Arata, Y.; Roberts, G. C. K. Structure 1995, 3, 79– 85 DOI: 10.1016/S0969-2126(01)00136-8Google ScholarThere is no corresponding record for this reference.
- 110Wand, A. J. Curr. Opin. Struct. Biol. 2013, 23, 75– 81 DOI: 10.1016/j.sbi.2012.11.005Google Scholar110https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhvVektrzL&md5=afdc384ca87a383b523699bbc8666f38The dark energy of proteins comes to light: conformational entropy and its role in protein function revealed by NMR relaxationWand, A. JoshuaCurrent Opinion in Structural Biology (2013), 23 (1), 75-81CODEN: COSBEF; ISSN:0959-440X. (Elsevier Ltd.)A review. Historically it has been virtually impossible to exptl. det. the contribution of residual protein entropy to fundamental protein activities such as the binding of ligands. Recent progress has illuminated the possibility of employing NMR relaxation methods to quant. det. the role of changes in conformational entropy in mol. recognition by proteins. The method rests on using fast internal protein dynamics as a proxy. Initial results reveal a large and variable role for conformational entropy in the binding of ligands by proteins. Such a role for conformational entropy in mol. recognition has significant implications for enzymol., signal transduction, allosteric regulation and the development of protein-directed pharmaceuticals.
- 111Marlow, M. S.; Dogan, J.; Frederick, K. K.; Valentine, K. G.; Wand, A. J. Nat. Chem. Biol. 2010, 6, 352– 358 DOI: 10.1038/nchembio.347Google Scholar111https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXksFakt78%253D&md5=e9ee34fffa91f612441a77a7385015ffThe role of conformational entropy in molecular recognition by calmodulinMarlow, Michael S.; Dogan, Jakob; Frederick, Kendra K.; Valentine, Kathleen G.; Wand, A. JoshuaNature Chemical Biology (2010), 6 (5), 352-358CODEN: NCBABT; ISSN:1552-4450. (Nature Publishing Group)The phys. basis for high-affinity interactions involving proteins is complex and potentially involves a range of energetic contributions. Among these are changes in protein conformational entropy, which cannot yet be reliably computed from mol. structures. We have recently used changes in conformational dynamics as a proxy for changes in conformational entropy of calmodulin upon assocn. with domains from regulated proteins. The apparent change in conformational entropy was linearly related to the overall binding entropy. This view warrants a more quant. foundation. Here we calibrate an 'entropy meter' using an exptl. dynamical proxy based on NMR relaxation and show that changes in the conformational entropy of calmodulin are a significant component of the energetics of binding. Furthermore, the distribution of motion at the interface between the target domain and calmodulin is surprisingly noncomplementary. These observations promote modification of our understanding of the energetics of protein-ligand interactions.
- 112Frederick, K. K.; Marlow, M. S.; Valentine, K. G.; Wand, A. J. Nature 2007, 448, 325– 329 DOI: 10.1038/nature05959Google ScholarThere is no corresponding record for this reference.
- 113Lee, A. L.; Wand, A. J. Nature 2001, 411, 501– 504 DOI: 10.1038/35078119Google ScholarThere is no corresponding record for this reference.
Cited By
This article is cited by 38 publications.
- Pravin P. Taware, Mukul G. Jain, Sreejith Raran-Kurussi, Vipin Agarwal, P. K. Madhu, Kaustubh R. Mote. Measuring Dipolar Order Parameters in Nondeuterated Proteins Using Solid-State NMR at the Magic-Angle-Spinning Frequency of 100 kHz. The Journal of Physical Chemistry Letters 2023, 14
(15)
, 3627-3635. https://doi.org/10.1021/acs.jpclett.3c00492
- Yusuke Nishiyama, Guangjin Hou, Vipin Agarwal, Yongchao Su, Ayyalusamy Ramamoorthy. Ultrafast Magic Angle Spinning Solid-State NMR Spectroscopy: Advances in Methodology and Applications. Chemical Reviews 2023, 123
(3)
, 918-988. https://doi.org/10.1021/acs.chemrev.2c00197
- Pu Duan, Aurelio J. Dregni, Mei Hong. Solid-State NMR 19F–1H–15N Correlation Experiments for Resonance Assignment and Distance Measurements of Multifluorinated Proteins. The Journal of Physical Chemistry A 2022, 126
(39)
, 7021-7032. https://doi.org/10.1021/acs.jpca.2c05154
- Bernd Reif. Deuteration for High-Resolution Detection of Protons in Protein Magic Angle Spinning (MAS) Solid-State NMR. Chemical Reviews 2022, 122
(10)
, 10019-10035. https://doi.org/10.1021/acs.chemrev.1c00681
- Sahil Ahlawat, Kaustubh R. Mote, Nils-Alexander Lakomek, Vipin Agarwal. Solid-State NMR: Methods for Biological Solids. Chemical Reviews 2022, 122
(10)
, 9643-9737. https://doi.org/10.1021/acs.chemrev.1c00852
- Collin G. Borcik, Isaac R. Eason, Boden Vanderloop, Benjamin J. Wylie. 2H,13C-Cholesterol for Dynamics and Structural Studies of Biological Membranes. ACS Omega 2022, 7
(20)
, 17151-17160. https://doi.org/10.1021/acsomega.2c00796
- Janet S. Anderson, Griselda Hernández, David M. LeMaster. 13C NMR Relaxation Analysis of Protein GB3 for the Assessment of Side Chain Dynamics Predictions by Current AMBER and CHARMM Force Fields. Journal of Chemical Theory and Computation 2020, 16
(5)
, 2896-2913. https://doi.org/10.1021/acs.jctc.0c00050
- Qingqing Chen, Shiping Xu, Xingyu Lu, Michael V. Boeri, Yuliya Pepelyayeva, Elizabeth L. Diaz, Sunil-Datta Soni, Marc Allaire, Martin B. Forstner, Brian J. Bahnson, Sharon Rozovsky. 77Se NMR Probes the Protein Environment of Selenomethionine. The Journal of Physical Chemistry B 2020, 124
(4)
, 601-616. https://doi.org/10.1021/acs.jpcb.9b07466
- Manali Ghosh and Chad M. Rienstra . 1H-Detected REDOR with Fast Magic-Angle Spinning of a Deuterated Protein. The Journal of Physical Chemistry B 2017, 121
(36)
, 8503-8511. https://doi.org/10.1021/acs.jpcb.7b07313
- Xiangyan Shi, Bhuvaneswari Kannaian, Chinmayi Prasanna, Aghil Soman, Lars Nordenskiöld. Structural and dynamical investigation of histone H2B in well-hydrated nucleosome core particles by solid-state NMR. Communications Biology 2023, 6
(1)
https://doi.org/10.1038/s42003-023-05050-3
- Liliya Vugmeyster, Aryana Rodgers, Dmitry Ostrovsky, C. James McKnight, Riqiang Fu. Deuteron off-resonance rotating frame relaxation for the characterization of slow motions in rotating and static solid-state proteins. Journal of Magnetic Resonance 2023, 352 , 107493. https://doi.org/10.1016/j.jmr.2023.107493
- Ümit Akbey. Site-specific protein backbone deuterium 2Hα quadrupolar patterns by proton-detected quadruple-resonance 3D 2HαcαNH MAS NMR spectroscopy. Solid State Nuclear Magnetic Resonance 2023, 125 , 101861. https://doi.org/10.1016/j.ssnmr.2023.101861
- Sahil Ahlawat, Subbarao Mohana Venkata Mopidevi, Pravin P. Taware, Sreejith Raran-Kurussi, Kaustubh R. Mote, Vipin Agarwal. Assignment of aromatic side-chain spins and characterization of their distance restraints at fast MAS. Journal of Structural Biology: X 2023, 7 , 100082. https://doi.org/10.1016/j.yjsbx.2022.100082
- Olof Stenström, Candide Champion, Marc Lehner, Guillaume Bouvignies, Sereina Riniker, Fabien Ferrage. How does it really move? Recent progress in the investigation of protein nanosecond dynamics by NMR and simulation. Current Opinion in Structural Biology 2022, 77 , 102459. https://doi.org/10.1016/j.sbi.2022.102459
- Kevin Singewald, James A. Wilkinson, Zikri Hasanbasri, Sunil Saxena. Beyond structure: Deciphering site‐specific dynamics in proteins from double histidine‐based
EPR
measurements. Protein Science 2022, 31
(7)
https://doi.org/10.1002/pro.4359
- Xiangyan Shi, Ziwei Zhai, Yinglu Chen, Jindi Li, Lars Nordenskiöld. Recent Advances in Investigating Functional Dynamics of Chromatin. Frontiers in Genetics 2022, 13 https://doi.org/10.3389/fgene.2022.870640
- Bernd Reif, Sharon E. Ashbrook, Lyndon Emsley, Mei Hong. Solid-state NMR spectroscopy. Nature Reviews Methods Primers 2021, 1
(1)
https://doi.org/10.1038/s43586-020-00002-1
- Marta Bonaccorsi, Tanguy Le Marchand, Guido Pintacuda. Protein structural dynamics by Magic-Angle Spinning NMR. Current Opinion in Structural Biology 2021, 70 , 34-43. https://doi.org/10.1016/j.sbi.2021.02.008
- Martin D. Gelenter, Kelly J. Chen, Mei Hong. Off-resonance 13C–2H REDOR NMR for site-resolved studies of molecular motion. Journal of Biomolecular NMR 2021, 75
(8-9)
, 335-345. https://doi.org/10.1007/s10858-021-00377-7
- Ümit Akbey. Dynamics of uniformly labelled solid proteins between 100 and 300 K: A 2D 2H-13C MAS NMR approach. Journal of Magnetic Resonance 2021, 327 , 106974. https://doi.org/10.1016/j.jmr.2021.106974
- Xiangyan Shi, Chinmayi Prasanna, Aghil Soman, Konstantin Pervushin, Lars Nordenskiöld. Dynamic networks observed in the nucleosome core particles couple the histone globular domains with DNA. Communications Biology 2020, 3
(1)
https://doi.org/10.1038/s42003-020-01369-3
- Eva Meirovitch, Jack H. Freed. Local ordering and dynamics in anisotropic media by magnetic resonance: from liquid crystals to proteins. Liquid Crystals 2020, 47
(13)
, 1926-1954. https://doi.org/10.1080/02678292.2019.1622158
- Evgeny Nimerovsky, Corinne P. Soutar. A modification of γ-encoded RN symmetry pulses for increasing the scaling factor and more accurate measurements of the strong heteronuclear dipolar couplings. Journal of Magnetic Resonance 2020, 319 , 106827. https://doi.org/10.1016/j.jmr.2020.106827
- Xiangyan Shi, Chinmayi Prasanna, Konstantin Pervushin, Lars Nordenskiöld. Solid-state NMR 13C, 15N assignments of human histone H3 in the nucleosome core particle. Biomolecular NMR Assignments 2020, 14
(1)
, 99-104. https://doi.org/10.1007/s12104-020-09927-w
- Alexander E. Khudozhitkov, Jan Neumann, Thomas Niemann, Dzmitry Zaitsau, Peter Stange, Dietmar Paschek, Alexander G. Stepanov, Daniil I. Kolokolov, Ralf Ludwig. Hydrogen Bonding Between Ions of Like Charge in Ionic Liquids Characterized by NMR Deuteron Quadrupole Coupling Constants—Comparison with Salt Bridges and Molecular Systems. Angewandte Chemie 2019, 131
(49)
, 18027-18035. https://doi.org/10.1002/ange.201912476
- Alexander E. Khudozhitkov, Jan Neumann, Thomas Niemann, Dzmitry Zaitsau, Peter Stange, Dietmar Paschek, Alexander G. Stepanov, Daniil I. Kolokolov, Ralf Ludwig. Hydrogen Bonding Between Ions of Like Charge in Ionic Liquids Characterized by NMR Deuteron Quadrupole Coupling Constants—Comparison with Salt Bridges and Molecular Systems. Angewandte Chemie International Edition 2019, 58
(49)
, 17863-17871. https://doi.org/10.1002/anie.201912476
- Maria Makrinich, Amir Goldbourt. 1
H-Detected quadrupolar spin–lattice relaxation measurements under magic-angle spinning solid-state NMR. Chemical Communications 2019, 55
(39)
, 5643-5646. https://doi.org/10.1039/C9CC01176E
- Rachel W. Martin, John E. Kelly, Jessica I. Kelz. Advances in instrumentation and methodology for solid-state NMR of biological assemblies. Journal of Structural Biology 2019, 206
(1)
, 73-89. https://doi.org/10.1016/j.jsb.2018.09.003
- Tom Aharoni, Amir Goldbourt. Rapid automated determination of chemical shift anisotropy values in the carbonyl and carboxyl groups of fd-y21m bacteriophage using solid state NMR. Journal of Biomolecular NMR 2018, 72
(1-2)
, 55-67. https://doi.org/10.1007/s10858-018-0206-1
- Irina Matlahov, Patrick C.A. van der Wel. Hidden motions and motion-induced invisibility: Dynamics-based spectral editing in solid-state NMR. Methods 2018, 148 , 123-135. https://doi.org/10.1016/j.ymeth.2018.04.015
- Xiangyan Shi, Chinmayi Prasanna, Toshio Nagashima, Toshio Yamazaki, Konstantin Pervushin, Lars Nordenskiöld. Structure and Dynamics in the Nucleosome Revealed by Solid‐State NMR. Angewandte Chemie 2018, 130
(31)
, 9882-9886. https://doi.org/10.1002/ange.201804707
- Xiangyan Shi, Chinmayi Prasanna, Toshio Nagashima, Toshio Yamazaki, Konstantin Pervushin, Lars Nordenskiöld. Structure and Dynamics in the Nucleosome Revealed by Solid‐State NMR. Angewandte Chemie International Edition 2018, 57
(31)
, 9734-9738. https://doi.org/10.1002/anie.201804707
- Eva Meirovitch, Zhichun Liang, Jack H. Freed. Protein dynamics in the solid-state from 2H NMR lineshape analysis. III. MOMD in the presence of Magic Angle Spinning. Solid State Nuclear Magnetic Resonance 2018, 89 , 35-44. https://doi.org/10.1016/j.ssnmr.2017.11.001
- Bernd Reif. Proton-Detection in Biological MAS Solid-State NMR Spectroscopy. 2018, 879-910. https://doi.org/10.1007/978-3-319-28388-3_69
- Kelsey A. Collier, Suvrajit Sengupta, Catalina A. Espinosa, John E. Kelly, Jessica I. Kelz, Rachel W. Martin. Design and construction of a quadruple-resonance MAS NMR probe for investigation of extensively deuterated biomolecules. Journal of Magnetic Resonance 2017, 285 , 8-17. https://doi.org/10.1016/j.jmr.2017.10.002
- Martin D. Gelenter, Tuo Wang, Shu-Yu Liao, Hugh O’Neill, Mei Hong. 2H–13C correlation solid-state NMR for investigating dynamics and water accessibilities of proteins and carbohydrates. Journal of Biomolecular NMR 2017, 68
(4)
, 257-270. https://doi.org/10.1007/s10858-017-0124-7
- Bernd Reif. Proton-Detection in Biological MAS Solid-State NMR Spectroscopy. 2017, 1-33. https://doi.org/10.1007/978-3-319-28275-6_69-1
- Ruth Bärenwald, Anja Achilles, Frank Lange, Tiago Ferreira, Kay Saalwächter. Applications of Solid-State NMR Spectroscopy for the Study of Lipid Membranes with Polyphilic Guest (Macro)Molecules. Polymers 2016, 8
(12)
, 439. https://doi.org/10.3390/polym8120439
Abstract
Figure 1
Figure 1. (A) 3D 2H–13C–13C solid-state MAS NMR pulse sequence. (B) 2D 2H–13C planes of the 3D 2H–13C–13C spectrum collected for microcrystalline GB1 and the extracted 2H line shapes for the E56 residue. Experimental line shapes (black), fits (red), and fitting residuals (blue) are displayed in the right column. In the 2D planes, signals with positive and negative intensities are shown in black and green, respectively.
Figure 2
Figure 2. (A) Crystal structure of GB1 (PDB: 2LGI) with Lys aliphatic groups shown in van der Waals spheres and coded with color scaling to
values. Two different perspectives are shown for better visualization. (B) K50, (C) K31, (D) K13, (E) K4, (F) K10, and (G) K28 local chemical environment in GB1. The K31 (CH2)β group is color coded with cyan as the
value is not determined. The residues having atoms within 5 Å away for the corresponding Lys residue are shown in sticks. The salt bridges between K50 and D47, E27, and K31 and the hydrogen bond between K13 side chain NH3 and G9 backbone CO are displayed by dash lines.
Figure 3
Figure 3. (A) Crystal structure of GB1 (PDB: 2LGI) with Asp and Asn aliphatic groups shown in van der Waals spheres and coded with color scaling to
values. Two different perspectives are shown for better visualization. (B) D47, (C) D22, (D) D46, (E) D36, (F) N37, (G) N35, (H) D40, and (I) N8 local chemical environment in GB1. The residues having atoms within 5 Å away for the corresponding Asn or Asp residue are shown in sticks. The salt bridge between D47 and K50 and the hydrogen bonds between D22 and T25, D46 and A48/T49, and N8 and T55 are presented by dashed lines.
Figure 4
Figure 4. (A) Crystal structure of GB1 (PDB: 2LGI) with Glu and Gln aliphatic groups shown in van der Waals spheres and coded with color scaling to
values. Two different perspectives are shown for better visualization. (B) E27, (C) E56, (D) Q32, (E) E42, (F) E15, (G) E19, and (H) Q2 local chemical environment in GB1. Aliphatic groups having deuterium
values undetermined are color coded with cyan. The residues having atoms within 5 Å away for the corresponding Glu or Gln residue are shown in sticks. The salt bridge between E27 and K31 and the hydrogen bond between E56 and D40/K10 are presented by dashed lines.
Figure 5
Figure 5. (A) Crystal structure of GB1 (PDB: 2LGI) with Thr aliphatic groups shown in van der Waals spheres and coded with color scaling to
values. Two different perspectives are shown for better visualization. (B) T55, (C) T25, (D) T49, (E) T18, (F) T44, (G) T49, (H) T11, (I) T16, (J) T17, and (K) T53 local chemical environment in GB1. Aliphatic groups having deuterium
values undetermined are color coded with cyan. The residues having atoms within 5 Å away for the corresponding Thr residues are shown in sticks. The hydrogen bonds between T55 and N8, T25 and D22, and T51 and T49 are presented by dashed lines.
Figure 6
Figure 6. (A) Crystal structure of GB1 (PDB: 2LGI) with Leu and Ile aliphatic groups shown in van der Waals spheres and coded with color scaling to
values. (B) L5, (C) L12, (D) L7, and (E) I6 local chemical environment in GB1. Aliphatic groups having deuterium
values undetermined are color coded with cyan. The residues having atoms within 5 Å away for the corresponding Leu/Ile residue are shown in sticks.
Figure 7
Figure 7. (A) Crystal structure of GB1 (PDB: 2LGI) with Val aliphatic groups shown in van der Waals spheres and coded with color scaling to
values. (B) V21, (C) V54, (D) V29, and (E) V39 local chemical environment in GB1. Aliphatic groups having deuterium
values undetermined are color coded with cyan. The residues having atoms within 5 Å away for the corresponding Val residue are shown in sticks.
Figure 8
Figure 8. 2Hα
values determined for amino acid residues in microcrystalline GB1. The value for residue F52 was not determined due to signal overlap.
Figure 9
Figure 9. Protein side-chain dynamics map for microcrystalline GB1, shown in three different viewpoints. Lys, Asn, Asp, Gln, Glu, Thr, Leu, and Val residues exhibiting large, moderate, and small amplitude side-chain motions are highlighted in red, orange, and blue, respectively. The rest of the residues (gray) are not considered in this illustration.
References
ARTICLE SECTIONSThis article references 113 other publications.
- 1Teilum, K.; Olsen, J. G.; Kragelund, B. B. Cell. Mol. Life Sci. 2009, 66, 2231– 2247 DOI: 10.1007/s00018-009-0014-6Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXot12lsbc%253D&md5=caa1fa9a979d3ab2fef0e0f82b5452c0Functional aspects of protein flexibilityTeilum, Kaare; Olsen, Johan G.; Kragelund, Birthe B.Cellular and Molecular Life Sciences (2009), 66 (14), 2231-2247CODEN: CMLSFI; ISSN:1420-682X. (Birkhaeuser Verlag)A review. Proteins are dynamic entities, and they possess an inherent flexibility that allows them to function through mol. interactions within the cell, among cells, and even between organisms. An appreciation of the non-static nature of proteins is emerging, but to describe and incorporate this into an intuitive perception of protein function is challenging. Flexibility is of overwhelming importance for protein function, and the changes in protein structure during interactions with binding partners can be dramatic. Here, the authors address protein flexibility, focusing on protein-ligand interactions. The thermodn. involved are reviewed, and examples of structure-function studies involving exptl. detd. flexibility descriptions are presented. While much remains to be understood about protein flexibility, it is clear that it is encoded within the amino acid sequence and should be viewed as an integral part of protein structure.
- 2Henzler-Wildman, K.; Kern, D. Nature 2007, 450, 964– 972 DOI: 10.1038/nature06522Google ScholarThere is no corresponding record for this reference.
- 3Reichert, D.; Zinkevich, T.; Saalwachter, K.; Krushelnitsky, A. J. Biomol. Struct. Dyn. 2012, 30, 617– 627 DOI: 10.1080/07391102.2012.689695Google Scholar3https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhs1CjsbfN&md5=3c33cdd703eebcb170f480ed0d10f27aThe relation of the X-ray B-factor to protein dynamics: insights from recent dynamic solid-state NMR dataReichert, Detlef; Zinkevich, Tatiana; Saalwaechter, Kay; Krushelnitsky, AlexeyJournal of Biomolecular Structure and Dynamics (2012), 30 (6), 617-627CODEN: JBSDD6; ISSN:0739-1102. (Taylor & Francis Ltd.)In addressing the potential use of B-factors derived from X-ray scattering data of proteins for the understanding the (functional) dynamics of proteins, we present a comparison of B-factors of five different proteins (SH3 domain, Crh, GB1, ubiquitin and thioredoxin) with data from recent solid-state NMR expts. reflecting true (rotational) dynamics on well-defined timescales. Apart from trivial correlations involving mobile loop regions and chain termini, we find no significant correlation of B-factors with the dynamic data on any of the investigated timescales, concluding that there is no unique and general correlation of B-factors with the internal reorientational dynamics of proteins.
- 4Kuzmanic, A.; Pannu, N. S.; Zagrovic, B. Nat. Commun. 2014, 5, 3220 DOI: 10.1038/ncomms4220Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC2cvivFKgsw%253D%253D&md5=99e76ed6614f1c57b46fa8917bcbcf99X-ray refinement significantly underestimates the level of microscopic heterogeneity in biomolecular crystalsKuzmanic Antonija; Zagrovic Bojan; Pannu Navraj SNature communications (2014), 5 (), 3220 ISSN:.Biomolecular X-ray structures typically provide a static, time- and ensemble-averaged view of molecular ensembles in crystals. In the absence of rigid-body motions and lattice defects, B-factors are thought to accurately reflect the structural heterogeneity of such ensembles. In order to study the effects of averaging on B-factors, we employ molecular dynamics simulations to controllably manipulate microscopic heterogeneity of a crystal containing 216 copies of villin headpiece. Using average structure factors derived from simulation, we analyse how well this heterogeneity is captured by high-resolution molecular-replacement-based model refinement. We find that both isotropic and anisotropic refined B-factors often significantly deviate from their actual values known from simulation: even at high 1.0 ÅA resolution and Rfree of 5.9%, B-factors of some well-resolved atoms underestimate their actual values even sixfold. Our results suggest that conformational averaging and inadequate treatment of correlated motion considerably influence estimation of microscopic heterogeneity via B-factors, and invite caution in their interpretation.
- 5McDermott, A. Annu. Rev. Biophys. 2009, 38, 385– 403 DOI: 10.1146/annurev.biophys.050708.133719Google ScholarThere is no corresponding record for this reference.
- 6Meirovitch, E.; Shapiro, Y. E.; Polimeno, A.; Freed, J. H. Prog. Nucl. Magn. Reson. Spectrosc. 2010, 56, 360– 405 DOI: 10.1016/j.pnmrs.2010.03.002Google ScholarThere is no corresponding record for this reference.
- 7Krushelnitsky, A.; Reichert, D.; Saalwachter, K. Acc. Chem. Res. 2013, 46, 2028– 2036 DOI: 10.1021/ar300292pGoogle ScholarThere is no corresponding record for this reference.
- 8Krushelnitsky, A.; Reichert, D. Prog. Nucl. Magn. Reson. Spectrosc. 2005, 47, 1– 25 DOI: 10.1016/j.pnmrs.2005.04.001Google ScholarThere is no corresponding record for this reference.
- 9Watt, E. D.; Rienstra, C. M. Anal. Chem. 2014, 86, 58– 64 DOI: 10.1021/ac403956kGoogle ScholarThere is no corresponding record for this reference.
- 10Kay, L. E. J. Magn. Reson. 2005, 173, 193– 207 DOI: 10.1016/j.jmr.2004.11.021Google ScholarThere is no corresponding record for this reference.
- 11Kleckner, I. R.; Foster, M. P. Biochim. Biophys. Acta, Proteins Proteomics 2011, 1814, 942– 968 DOI: 10.1016/j.bbapap.2010.10.012Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXnsVCnsbo%253D&md5=2c047b195fd796a815e800bd162a49deAn introduction to NMR-based approaches for measuring protein dynamicsKleckner, Ian R.; Foster, Mark P.Biochimica et Biophysica Acta, Proteins and Proteomics (2011), 1814 (8), 942-968CODEN: BBAPBW; ISSN:1570-9639. (Elsevier B. V.)A review. Proteins are inherently flexible at ambient temp. At equil., they are characterized by a set of conformations that undergo continuous exchange within a hierarchy of spatial and temporal scales ranging from nanometers to micrometers and femtoseconds to hours. Dynamic properties of proteins are essential for describing the structural bases of their biol. functions including catalysis, binding, regulation and cellular structure. NMR spectroscopy represents a powerful technique for measuring these essential features of proteins. Here the authors provide an introduction to NMR-based approaches for studying protein dynamics, highlighting eight distinct methods with recent examples, contextualized within a common exptl. and anal. framework. The selected methods are (1) Real-time NMR, (2) Exchange spectroscopy, (3) Lineshape anal., (4) CPMG relaxation dispersion, (5) Rotating frame relaxation dispersion, (6) Nuclear spin relaxation, (7) Residual dipolar coupling, (8) Paramagnetic relaxation enhancement. This article is part of a Special Issue entitled: Protein Dynamics: Exptl. and Computational Approaches.
- 12Yang, L. Q.; Sang, P.; Tao, Y.; Fu, Y. X.; Zhang, K. Q.; Xie, Y. H.; Liu, S. Q. J. Biomol. Struct. Dyn. 2014, 32, 372– 393 DOI: 10.1080/07391102.2013.770372Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXksFOms78%253D&md5=028706e72d99b9aa4779974289c72e58Protein dynamics and motions in relation to their functions: several case studies and the underlying mechanismsYang, Li-Quan; Sang, Peng; Tao, Yan; Fu, Yun-Xin; Zhang, Ke-Qin; Xie, Yue-Hui; Liu, Shu-QunJournal of Biomolecular Structure and Dynamics (2014), 32 (3), 372-393CODEN: JBSDD6; ISSN:0739-1102. (Taylor & Francis Ltd.)A review. Proteins are dynamic entities in cellular soln. with functions governed essentially by their dynamic personalities. Here, the authors review several dynamics studies on proteinase K and the HIV-1 virus envelope gp120 glycoprotein to demonstrate the importance of investigating the dynamic behaviors and mol. motions for a complete understanding of their structure-function relations. Using computer simulations and essential dynamic (ED) anal. approaches, the dynamics data obtained revealed that: (1) proteinase K has highly flexible substrate-binding site, thus supporting the induced-fit or conformational selection mechanism of substrate binding; (2) Ca2+ removal from proteinase K increases the global conformational flexibility, decreases the local flexibility of the substrate-binding region, and does not influence the thermal motion of the catalytic triad, thus explaining the exptl. detd. decreased thermostability, reduced substrate affinity, and almost unchanged catalytic activity upon Ca2+ removal; (3) substrate binding affects the large concerted motions of proteinase K, and the resulting dynamic pocket can be connected to substrate binding, orientation, and product release; (4) amino acid mutations 375 S/W and 423 I/P of HIV-1 gp120 have distinct effects on the mol. motions of gp120, facilitating the 375 S/W mutant to assume the CD4-bound conformation, while the 423 I/P mutant to prefer for the CD4-unliganded state. The mechanisms underlying protein dynamics and protein-ligand binding, including the concept of the free energy landscape (FEL) of the protein-solvent system, how the ruggedness and variability of FEL dets. protein's dynamics, and how the 3 ligand-binding models (lock-and-key, induced-fit, and conformational selection) are rationalized based on the FEL theory are discussed in depth.
- 13Karplus, M.; Kuriyan, J. Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 6679– 6685 DOI: 10.1073/pnas.0408930102Google ScholarThere is no corresponding record for this reference.
- 14Lam, A. J.; St-Pierre, F.; Gong, Y.; Marshall, J. D.; Cranfill, P. J.; Baird, M. A.; McKeown, M. R.; Wiedenmann, J.; Davidson, M. W.; Schnitzer, M. J.; Tsien, R. Y.; Lin, M. Z. Nat. Methods 2012, 9, 1005– 1012 DOI: 10.1038/nmeth.2171Google Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhtlajs7zI&md5=4f780d655ff3cdada989b5908cefbc67Improving FRET dynamic range with bright green and red fluorescent proteinsLam, Amy J.; St-Pierre, Francois; Gong, Yiyang; Marshall, Jesse D.; Cranfill, Paula J.; Baird, Michelle A.; McKeown, Michael R.; Wiedenmann, Joerg; Davidson, Michael W.; Schnitzer, Mark J.; Tsien, Roger Y.; Lin, Michael Z.Nature Methods (2012), 9 (10), 1005-1012CODEN: NMAEA3; ISSN:1548-7091. (Nature Publishing Group)A variety of genetically encoded reporters use changes in fluorescence (or Foerster) resonance energy transfer (FRET) to report on biochem. processes in living cells. The std. genetically encoded FRET pair consists of CFPs and YFPs, but many CFP-YFP reporters suffer from low FRET dynamic range, phototoxicity from the CFP excitation light and complex photokinetic events such as reversible photobleaching and photoconversion. We engineered two fluorescent proteins, Clover and mRuby2, which are the brightest green and red fluorescent proteins to date and have the highest Foerster radius of any ratiometric FRET pair yet described. Replacement of CFP and YFP with these two proteins in reporters of kinase activity, small GTPase activity and transmembrane voltage significantly improves photostability, FRET dynamic range and emission ratio changes. These improvements enhance detection of transient biochem. events such as neuronal action-potential firing and RhoA activation in growth cones.
- 15Schuler, B. J. Nanobiotechnol. 2013, 11 (Suppl 1) S2 DOI: 10.1186/1477-3155-11-S1-S2Google ScholarThere is no corresponding record for this reference.
- 16Nesmelov, Y. E.; Thomas, D. D. Biophys. Rev. 2010, 2, 91– 99 DOI: 10.1007/s12551-010-0032-5Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXlsVSksrw%253D&md5=324f9f0b67f707de61088b4fb823fa29Protein structural dynamics revealed by site-directed spin labeling and multifrequency EPRNesmelov, Yuri E.; Thomas, David D.Biophysical Reviews (2010), 2 (2), 91-99CODEN: BRIECG; ISSN:1867-2450. (Springer)A review. Multifrequency ESR (EPR), combined with site-directed spin labeling, is a powerful spectroscopic tool to characterize protein dynamics. The lineshape of an EPR spectrum reflects combined rotational dynamics of the spin probe's local motion within a protein, reorientations of protein domains, and overall protein tumbling. All these motions can be restricted and anisotropic, and sepn. of these motions is important for thorough characterization of protein dynamics. Multifrequency EPR distinguishes between different motions of a spin-labeled protein, due to the frequency dependence of EPR resoln. to fast and slow motion of a spin probe. This gives multifrequency EPR its unique capability to characterize protein dynamics in great detail. In this review, we analyze what makes multifrequency EPR sensitive to different rates of spin probe motion and discuss several examples of its usage to sep. spin probe dynamics and overall protein dynamics, to characterize protein backbone dynamics, and to resolve protein conformational states.
- 17Fleissner, M. R.; Bridges, M. D.; Brooks, E. K.; Cascio, D.; Kalai, T.; Hideg, K.; Hubbell, W. L. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 16241– 16246 DOI: 10.1073/pnas.1111420108Google ScholarThere is no corresponding record for this reference.
- 18Allerhand, A.; Doddrell, D.; Glushko, V.; Cochran, D. W.; Wenkert, E.; Lawson, P. J.; Gurd, F. R. J. Am. Chem. Soc. 1971, 93, 544– 546 DOI: 10.1021/ja00731a053Google ScholarThere is no corresponding record for this reference.
- 19Levy, R. M.; Karplus, M.; Mccammon, J. A. J. Am. Chem. Soc. 1981, 103, 994– 996 DOI: 10.1021/ja00394a072Google ScholarThere is no corresponding record for this reference.
- 20Deverell, C.; Morgan, R. E.; Strange, J. H. Mol. Phys. 1970, 18, 553– 559 DOI: 10.1080/00268977000100611Google ScholarThere is no corresponding record for this reference.
- 21Wittebort, R. J.; Szabo, A. J. Chem. Phys. 1978, 69, 1722 DOI: 10.1063/1.436748Google ScholarThere is no corresponding record for this reference.
- 22Carr, H. Y.; Purcell, E. M. Phys. Rev. 1954, 94, 630– 638 DOI: 10.1103/PhysRev.94.630Google Scholar22https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaG2cXksFKjtA%253D%253D&md5=f729c8b5191575d29fdfd489021f05b7Effects of diffusion on free precession in nuclear magnetic resonance experimentsCarr, H. Y.; Purcell, E. M.Physical Review (1954), 94 (), 630-8CODEN: PHRVAO; ISSN:0031-899X.Nuclear resonance techniques involving free precession are examd., and a convenient variation of Hahn's spin-echo method is described (ibid. 80, 580(1950)). This variation employs a combination of pulses of different intensity or duration ("90°" and "180°" pulses). Measurements of the transverse relaxation time T2 in fluids are often severely compromised by mol. diffusion. Hahn's analysis of the effect of diffusion is reformulated and extended, and a new scheme for measuring T2 is described which, as predicted by the extended theory, largely circumvents the diffusion effect. On the other hand, the free precession technique, applied in a different way, permits a direct measurement of the mol. self-diffusion const. in suitable fluids. A measurement of the self-diffusion const. of H2O at 25° yields D = 2.5 ± 0.3 × 10-5 sq. cm./sec., in good agreement with previous detns. The effect of convection on free precession is also analyzed. A null method for measuring the longitudinal relaxation time T1, based on the unequal-pulse technique, is described.
- 23Luz, Z.; Meiboom, S. J. Chem. Phys. 1963, 39, 366– 370 DOI: 10.1063/1.1734254Google ScholarThere is no corresponding record for this reference.
- 24Loria, J. P.; Rance, M.; Palmer, A. G. J. Am. Chem. Soc. 1999, 121, 2331– 2332 DOI: 10.1021/ja983961aGoogle ScholarThere is no corresponding record for this reference.
- 25Tolman, J. R.; Flanagan, J. M.; Kennedy, M. A.; Prestegard, J. H. Proc. Natl. Acad. Sci. U. S. A. 1995, 92, 9279– 9283 DOI: 10.1073/pnas.92.20.9279Google ScholarThere is no corresponding record for this reference.
- 26Saupe, A.; Englert, G. Phys. Rev. Lett. 1963, 11, 462– 464 DOI: 10.1103/PhysRevLett.11.462Google ScholarThere is no corresponding record for this reference.
- 27Woodward, C. K.; Hilton, B. D. Annu. Rev. Biophys. Bioeng. 1979, 8, 99– 127 DOI: 10.1146/annurev.bb.08.060179.000531Google Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE1MXksFWitLo%253D&md5=9c5ffcdf75aef65a03184d7ea950259aHydrogen exchange kinetics and internal motions in proteins and nucleic acidsWoodward, Clare K.; Hilton, Bruce D.Annual Review of Biophysics and Bioengineering (1979), 8 (), 99-127CODEN: ABPBBK; ISSN:0084-6589.A review with 188 refs.
- 28Muhandiram, D. R.; Yamazaki, T.; Sykes, B. D.; Kay, L. E. J. Am. Chem. Soc. 1995, 117, 11536– 11544 DOI: 10.1021/ja00151a018Google ScholarThere is no corresponding record for this reference.
- 29Gobl, C.; Madl, T.; Simon, B.; Sattler, M. Prog. Nucl. Magn. Reson. Spectrosc. 2014, 80, 26– 63 DOI: 10.1016/j.pnmrs.2014.05.003Google Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXpslOis7g%253D&md5=b8b4e2b127b99a42e9973f960aeb7b7cNMR approaches for structural analysis of multidomain proteins and complexes in solutionGobl, Christoph; Madl, Tobias; Simon, Bernd; Sattler, MichaelProgress in Nuclear Magnetic Resonance Spectroscopy (2014), 80 (), 26-63CODEN: PNMRAT; ISSN:0079-6565. (Elsevier B.V.)A review. NMR spectroscopy is a key method for studying the structure and dynamics of (large) multidomain proteins and complexes in soln. It plays a unique role in integrated structural biol. approaches as esp. information about conformational dynamics can be readily obtained at residue resoln. Here, we review NMR techniques for such studies focusing on state-of-the-art tools and practical aspects. An efficient approach for detg. the quaternary structure of multidomain complexes starts from the structures of individual domains or subunits. The arrangement of the domains/subunits within the complex is then defined based on NMR measurements that provide information about the domain interfaces combined with (long-range) distance and orientational restraints. Aspects discussed include sample prepn., specific isotope labeling and spin labeling; detn. of binding interfaces and domain/subunit arrangements from chem. shift perturbations (CSP), nuclear Overhauser effects (NOEs), isotope editing/filtering, cross-satn., and differential line broadening; and based on paramagnetic relaxation enhancements (PRE) using covalent and sol. spin labels. Finally, the utility of complementary methods such as small-angle X-ray or neutron scattering (SAXS, SANS), ESR (EPR) or fluorescence spectroscopy techniques is discussed. The applications of NMR techniques are illustrated with studies of challenging (high mol. wt.) protein complexes.
- 30Mittermaier, A.; Kay, L. E. Science 2006, 312, 224– 228 DOI: 10.1126/science.1124964Google ScholarThere is no corresponding record for this reference.
- 31Sheppard, D.; Sprangers, R.; Tugarinov, V. Prog. Nucl. Magn. Reson. Spectrosc. 2010, 56, 1– 45 DOI: 10.1016/j.pnmrs.2009.07.004Google ScholarThere is no corresponding record for this reference.
- 32Göbl, C.; Tjandra, N. Entropy 2012, 14, 581– 598 DOI: 10.3390/e14030581Google ScholarThere is no corresponding record for this reference.
- 33Li, F.; Grishaev, A.; Ying, J.; Bax, A. J. Am. Chem. Soc. 2015, 137, 14798– 14811 DOI: 10.1021/jacs.5b10072Google ScholarThere is no corresponding record for this reference.
- 34Im, W.; Jo, S.; Kim, T. Biochim. Biophys. Acta, Biomembr. 2012, 1818, 252– 262 DOI: 10.1016/j.bbamem.2011.07.048Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XkslKrsw%253D%253D&md5=4b07dd2cf3479c393f3903d7f63871d2An ensemble dynamics approach to decipher solid-state NMR observables of membrane proteinsIm, Wonpil; Jo, Sunhwan; Kim, TaehoonBiochimica et Biophysica Acta, Biomembranes (2012), 1818 (2), 252-262CODEN: BBBMBS; ISSN:0005-2736. (Elsevier B.V.)A review. Solid-state NMR (SSNMR) is an invaluable tool for detg. orientations of membrane proteins and peptides in lipid bilayers. Such orientational descriptions provide essential information about membrane protein functions. However, when a semi-static single conformer model is used to interpret various SSNMR observables, important dynamics information can be missing, and, sometimes, even orientational information can be misinterpreted. In addn., over the last decade, mol. dynamics (MD) simulation and semi-static SSNMR interpretation have shown certain levels of discrepancies in terms of transmembrane helix orientation and dynamics. Dynamic fitting models have recently been proposed to resolve these discrepancies by taking into account transmembrane helix whole body motions using addnl. parameters. As an alternative approach, the authors have developed SSNMR ensemble dynamics (SSNMR-ED) using multiple conformer models, which generates an ensemble of structures that satisfies the exptl. observables without any fitting parameters. In this review, various computational methods for detg. transmembrane helix orientations are discussed, and the distributions of VpuTM (from HIV-1) and WALP23 (a synthetic peptide) orientations from SSNMR-ED simulations are compared with those from MD simulations and semi-static/dynamic fitting models. Such comparisons illustrate that SSNMR-ED can be used as a general means to ext. both membrane protein structure and dynamics from the SSNMR measurements. This article is part of a Special Issue entitled: Membrane protein structure and function.
- 35Hu, F.; Luo, W.; Hong, M. Science 2010, 330, 505– 508 DOI: 10.1126/science.1191714Google ScholarThere is no corresponding record for this reference.
- 36Bocian, D. F.; Chan, S. I. Annu. Rev. Phys. Chem. 1978, 29, 307– 335 DOI: 10.1146/annurev.pc.29.100178.001515Google ScholarThere is no corresponding record for this reference.
- 37Lewandowski, J. R.; Sass, H. J.; Grzesiek, S.; Blackledge, M.; Emsley, L. J. Am. Chem. Soc. 2011, 133, 16762– 16765 DOI: 10.1021/ja206815hGoogle ScholarThere is no corresponding record for this reference.
- 38Mollica, L.; Baias, M.; Lewandowski, J. R.; Wylie, B. J.; Sperling, L. J.; Rienstra, C. M.; Emsley, L.; Blackledge, M. J. Phys. Chem. Lett. 2012, 3, 3657– 3662 DOI: 10.1021/jz3016233Google Scholar38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhslaksrrE&md5=4de33d158064b091ad375dd5ef2af54bAtomic-Resolution Structural Dynamics in Crystalline Proteins from NMR and Molecular SimulationMollica, Luca; Baias, Maria; Lewandowski, Jozef R.; Wylie, Benjamin J.; Sperling, Lindsay J.; Rienstra, Chad M.; Emsley, Lyndon; Blackledge, MartinJournal of Physical Chemistry Letters (2012), 3 (23), 3657-3662CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)Solid-state NMR can provide at.-resoln. information about protein motions occurring on a vast range of time scales under similar conditions to those of X-ray diffraction studies and therefore offers a highly complementary approach to characterizing the dynamic fluctuations occurring in the crystal. We compare exptl. detd. dynamic parameters, spin relaxation, chem. shifts, and dipolar couplings, to values calcd. from a 200 ns MD simulation of protein GB1 in its cryst. form, providing insight into the nature of structural dynamics occurring within the cryst. lattice. This simulation allows us to test the accuracy of commonly applied procedures for the interpretation of exptl. solid-state relaxation data in terms of dynamic modes and time scales. We discover that the potential complexity of relaxation-active motion can lead to significant under- or overestimation of dynamic amplitudes if different components are not taken into consideration.
- 39Lewandowski, J. R. Acc. Chem. Res. 2013, 46, 2018– 2027 DOI: 10.1021/ar300334gGoogle ScholarThere is no corresponding record for this reference.
- 40Krushelnitsky, A.; Zinkevich, T.; Reif, B.; Saalwachter, K. J. Magn. Reson. 2014, 248, 8– 12 DOI: 10.1016/j.jmr.2014.09.007Google ScholarThere is no corresponding record for this reference.
- 41Lewandowski, J. R.; Halse, M. E.; Blackledge, M.; Emsley, L. Science 2015, 348, 578– 581 DOI: 10.1126/science.aaa6111Google ScholarThere is no corresponding record for this reference.
- 42Tollinger, M.; Sivertsen, A. C.; Meier, B. H.; Ernst, M.; Schanda, P. J. Am. Chem. Soc. 2012, 134, 14800– 14807 DOI: 10.1021/ja303591yGoogle ScholarThere is no corresponding record for this reference.
- 43Agarwal, V.; Xue, Y.; Reif, B.; Skrynnikov, N. R. J. Am. Chem. Soc. 2008, 130, 16611– 16621 DOI: 10.1021/ja804275pGoogle ScholarThere is no corresponding record for this reference.
- 44Haller, J. D.; Schanda, P. J. Biomol. NMR 2013, 57, 263– 280 DOI: 10.1007/s10858-013-9787-xGoogle ScholarThere is no corresponding record for this reference.
- 45Chevelkov, V.; Xue, Y.; Linser, R.; Skrynnikov, N. R.; Reif, B. J. Am. Chem. Soc. 2010, 132, 5015– 5017 DOI: 10.1021/ja100645kGoogle ScholarThere is no corresponding record for this reference.
- 46Hohwy, M.; Jaroniec, C. P.; Reif, B.; Rienstra, C. M.; Griffin, R. G. J. Am. Chem. Soc. 2000, 122, 3218– 3219 DOI: 10.1021/ja9913737Google ScholarThere is no corresponding record for this reference.
- 47Vinogradov, E.; Madhu, P. K.; Vega, S. J. Chem. Phys. 2001, 115, 8983 DOI: 10.1063/1.1408287Google ScholarThere is no corresponding record for this reference.
- 48Gullion, T.; Schaefer, J. J. Magn. Reson. 1989, 81, 196– 200 DOI: 10.1016/0022-2364(89)90280-1Google Scholar48https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXkvV2hsr8%253D&md5=cd503257b5d0237a5f8bd549517a3a76Rotational-echo double-resonance NMRGullion, Terry; Schaefer, JacobJournal of Magnetic Resonance (1969-1992) (1989), 81 (1), 196-200CODEN: JOMRA4; ISSN:0022-2364.Rotational-echo double-resonance (REDOR), a spin-echo double-resonance (SEDOR) NMR expt. with magic-angle sample spinning, is illustrated by using the results of expts. performed on 13C- and 15N-labeled alanines. The pulse sequence for 13C-15N REDOR is shown.
- 49Gullion, T.; Schaefer, J. Adv. Magn. Opt. Reson. 1989, 13, 57– 83 DOI: 10.1016/B978-0-12-025513-9.50009-4Google ScholarThere is no corresponding record for this reference.
- 50Jaroniec, C. P.; Tounge, B. A.; Rienstra, C. M.; Herzfeld, J.; Griffin, R. G. J. Magn. Reson. 2000, 146, 132– 139 DOI: 10.1006/jmre.2000.2128Google ScholarThere is no corresponding record for this reference.
- 51Schanda, P.; Meier, B. H.; Ernst, M. J. Magn. Reson. 2011, 210, 246– 259 DOI: 10.1016/j.jmr.2011.03.015Google ScholarThere is no corresponding record for this reference.
- 52Dvinskikh, S. V.; Zimmermann, H.; Maliniak, A.; Sandström, D. J. Magn. Reson. 2003, 164, 165– 170 DOI: 10.1016/S1090-7807(03)00180-0Google ScholarThere is no corresponding record for this reference.
- 53Dvinskikh, S. V.; Zimmermann, H.; Maliniak, A.; Sandstrom, D. J. Chem. Phys. 2005, 122, 44512 DOI: 10.1063/1.1834569Google ScholarThere is no corresponding record for this reference.
- 54Levitt, M. H. In Encyclopedia of Nuclear Magnetic Resonance; Grant, D. M.; Harris, R. K., Eds.; John Wiley & Sons: Chichester U.K., 2002; Vol. 9, pp 165– 196.Google ScholarThere is no corresponding record for this reference.
- 55Lorieau, J. L.; Day, L. A.; McDermott, A. E. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, 10366– 10371 DOI: 10.1073/pnas.0800405105Google ScholarThere is no corresponding record for this reference.
- 56Lorieau, J. L.; McDermott, A. E. J. Am. Chem. Soc. 2006, 128, 11505– 11512 DOI: 10.1021/ja062443uGoogle ScholarThere is no corresponding record for this reference.
- 57Zhao, X.; Sudmeier, J. L.; Bachovchin, W. W.; Levitt, M. H. J. Am. Chem. Soc. 2001, 123, 11097– 11098 DOI: 10.1021/ja016328pGoogle ScholarThere is no corresponding record for this reference.
- 58Schanda, P.; Meier, B. H.; Ernst, M. J. Am. Chem. Soc. 2010, 132, 15957– 15967 DOI: 10.1021/ja100726aGoogle ScholarThere is no corresponding record for this reference.
- 59Yang, J.; Tasayco, M. L.; Polenova, T. J. Am. Chem. Soc. 2009, 131, 13690– 13702 DOI: 10.1021/ja9037802Google ScholarThere is no corresponding record for this reference.
- 60Chevelkov, V.; Fink, U.; Reif, B. J. Am. Chem. Soc. 2009, 131, 14018– 14022 DOI: 10.1021/ja902649uGoogle ScholarThere is no corresponding record for this reference.
- 61Meirovitch, E.; Liang, Z.; Freed, J. H. J. Phys. Chem. B 2015, 119, 2857– 2868 DOI: 10.1021/jp511386bGoogle ScholarThere is no corresponding record for this reference.
- 62Torchia, D. A. Annu. Rev. Biophys. Bioeng. 1984, 13, 125– 144 DOI: 10.1146/annurev.bb.13.060184.001013Google Scholar62https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2cXkvVKnu7o%253D&md5=e8be982f7ae0fb3e1d33b924886c4751Solid state NMR studies of protein internal dynamicsTorchia, Dennis A.Annual Review of Biophysics and Bioengineering (1984), 13 (), 125-44CODEN: ABPBBK; ISSN:0084-6589.A review with 65 refs.
- 63Torchia, D. A.; Szabo, A. J. Magn. Reson. 1982, 49, 107– 121 DOI: 10.1016/0022-2364(82)90301-8Google ScholarThere is no corresponding record for this reference.
- 64Olympia, P. L.; Wei, I. Y.; Fung, B. M. J. Chem. Phys. 1969, 51, 1610– 1614 DOI: 10.1063/1.1672220Google ScholarThere is no corresponding record for this reference.
- 65Hologne, M.; Chevelkov, V.; Reif, B. Prog. Nucl. Magn. Reson. Spectrosc. 2006, 48, 211– 232 DOI: 10.1016/j.pnmrs.2006.05.004Google ScholarThere is no corresponding record for this reference.
- 66Jelinski, L. W. Annu. Rev. Mater. Sci. 1985, 15, 359– 377 DOI: 10.1146/annurev.ms.15.080185.002043Google Scholar66https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2MXlvFKmurY%253D&md5=1b3fd9ec31842011ea6db576e7f43956Solid state deuterium NMR studies of polymer chain dynamicsJelinski, Lynn W.Annual Review of Materials Science (1985), 15 (), 359-77CODEN: ARMSCX; ISSN:0084-6600.A review with 49 refs. on solid-state D-NMR expts. used for detn. of local mol. dynamics of polymers. Also, briefly reviewed were D-NMR applications on anal. of polymeric liq. crystals, phase sepn. in polymers, and polyurethane structural anal.
- 67Hologne, M.; Hirschinger, J. Solid State Nucl. Magn. Reson. 2004, 26, 1– 10 DOI: 10.1016/S0926-2040(03)00062-6Google Scholar67https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXkt1Gjtro%253D&md5=488b9d0e6297c31e1f3664e095bc3f2fMolecular dynamics as studied by static-powder and magic-angle spinning 2H NMRHologne, Maggy; Hirschinger, JeromeSolid State Nuclear Magnetic Resonance (2004), 26 (1), 1-10CODEN: SSNRE4; ISSN:0926-2040. (Elsevier Science)The 2H NMR magic-angle spinning (MAS) technique is compared to the static-powder quadrupole echo (QE) and Jeener-Brockaert (JB) pulse sequences for a quant. investigation of mol. dynamics in solids. The linewidth of individual spinning sidebands of the one-dimensional MAS spectra are obsd. to be characteristic of the correlation time from ∼10-2 to ∼10-8 s so that the dynamic range is increased by approx. three orders of magnitude when compared to the QE expt. As a consequence, MAS 2H NMR is found to be more sensitive to the presence of an inhomogeneous distribution of correlation times than the QE and JB expts. which rely upon lineshape distortions due to anisotropic T2 and T1Q relaxation, resp. All these results are demonstrated exptl. and numerically using the two-site flip motion of di-Me sulfone and of the nitrobenzene guest in the α-p-tert-butylcalix[4]arene-nitrobenzene inclusion compd.
- 68Batchelder, L. S.; Niu, C. H.; Torchia, D. A. J. Am. Chem. Soc. 1983, 105, 2228– 2231 DOI: 10.1021/ja00346a021Google ScholarThere is no corresponding record for this reference.
- 69Pines, A.; Gibby, M. G.; Waugh, J. S. J. Chem. Phys. 1973, 59, 569– 590 DOI: 10.1063/1.1680061Google Scholar69https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE3sXkvVGgsLo%253D&md5=43fbff1149b7d4de334d6168521cc10cProton-enhanced NMR of dilute spins in solidsPines, A.; Gibby, M. G.; Waugh, J. S.Journal of Chemical Physics (1973), 59 (2), 569-90CODEN: JCPSA6; ISSN:0021-9606.The NMR signals of isotopically or chem. dil. nuclear spins S in solids can be enhanced by repeatedly transferring polarization from a more abundant species I of high abundance (usually protons) to which they are coupled. The gain in power sensitivity as compared with conventional observation of the rare spins approaches NII(I + 1)γI2/NSS(S + 1)γS2, or ∼103 for S = 13C, I = H in org. solids. The transfer of polarization is accomplished by any of a no. of double resonance methods. High-frequency resoln. of the S-spin signal was obtained by decoupling of the abundant spins. The exptl. requirements of the technique are discussed and a brief comparison of its sensitivity with other procedures is made. Representative applications and exptl. results are mentioned.
- 70Hartmann, S. R.; Hahn, E. L. Phys. Rev. 1962, 128, 2042– 2053 DOI: 10.1103/PhysRev.128.2042Google ScholarThere is no corresponding record for this reference.
- 71Hologne, M.; Chen, Z.; Reif, B. J. Magn. Reson. 2006, 179, 20– 28 DOI: 10.1016/j.jmr.2005.10.014Google ScholarThere is no corresponding record for this reference.
- 72Shi, X.; Yarger, J. L.; Holland, G. P. J. Magn. Reson. 2013, 226, 1– 12 DOI: 10.1016/j.jmr.2012.10.013Google ScholarThere is no corresponding record for this reference.
- 73Hologne, M.; Faelber, K.; Diehl, A.; Reif, B. J. Am. Chem. Soc. 2005, 127, 11208– 11209 DOI: 10.1021/ja051830lGoogle ScholarThere is no corresponding record for this reference.
- 74Shi, X.; Yarger, J. L.; Holland, G. P. Chem. Commun. 2014, 50, 4856– 4859 DOI: 10.1039/c4cc00971aGoogle ScholarThere is no corresponding record for this reference.
- 75Shi, X.; Holland, G. P.; Yarger, J. L. Biomacromolecules 2015, 16, 852– 859 DOI: 10.1021/bm5017578Google ScholarThere is no corresponding record for this reference.
- 76Wei, D.; Akbey, U. m.; Paaske, B.; Oschkinat, H.; Reif, B.; Bjerring, M.; Nielsen, N. C. J. Phys. Chem. Lett. 2011, 2, 1289– 1294 DOI: 10.1021/jz200511bGoogle Scholar76https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXlvFSqtLw%253D&md5=9804df3d0cfc0853c94e0d7e503df8b5Optimal 2H rf Pulses and 2H-13C Cross-Polarization Methods for Solid-State 2H MAS NMR of Perdeuterated ProteinsWei, Daxiu; Akbey, Umit; Paaske, Berit; Oschkinat, Hartmut; Reif, Bernd; Bjerring, Morten; Nielsen, Niels Chr.Journal of Physical Chemistry Letters (2011), 2 (11), 1289-1294CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)We present a novel concept for rf pulses and optimal control designed cross-polarization expts. for quadrupolar nuclei. The methods are demonstrated for 2H CP-MAS and 2H multiple-pulse NMR of perdeuterated proteins, for which sensitivity enhancements up to an order of magnitude are presented relative to commonly used approaches. The so-called RESPIRATION rf pulses combines the concept of short broad-band pulses with generation of pulses with large flip angles through distribution of the rf pulse over several rotor echoes. This leads to close-to-ideal rf pulses, facilitating implementation of expts. relying on the ability to realize high-performance 90 and 180° pulses, as, for example, in refocused INEPT and double-to-single quantum coherence expts., or just pulses that provide a true representation of the quadrupolar powder pattern to ext. information about the structure or dynamics. The optimal control 2H → 13C CP-MAS method demonstrates transfer efficiencies up to around 85% while being extremely robust toward rf inhomogeneity and resonance offsets.
- 77Nielsen, A. B.; Jain, S.; Ernst, M.; Meier, B. H.; Nielsen, N. C. J. Magn. Reson. 2013, 237, 147– 151 DOI: 10.1016/j.jmr.2013.09.002Google ScholarThere is no corresponding record for this reference.
- 78Jain, S.; Bjerring, M.; Nielsen, N. C. J. Phys. Chem. Lett. 2012, 3, 703– 708 DOI: 10.1021/jz3000905Google Scholar78https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XislSqtLk%253D&md5=16f412f704236445de9c532d91ebe4caEfficient and Robust Heteronuclear Cross-Polarization for High-Speed-Spinning Biological Solid-State NMR SpectroscopyJain, Sheetal; Bjerring, Morten; Nielsen, Niels Chr.Journal of Physical Chemistry Letters (2012), 3 (6), 703-708CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)The authors present a new and highly efficient approach for heteronuclear coherence transfer in solid-state NMR spectroscopy under high-speed spinning conditions. The so-called RESPIRATIONCP expt. exploits phase-alternated recoupling on only one of the two rf channels intertwined in a synchronized train of short rf pulses on both channels. The method provides significantly higher efficiencies than state-of-the art techniques including ramped and adiabatic cross-polarization expts. with long durations of intense rf irradn. At the same time, it is easier to setup exptl. and significantly more robust toward imperfections such as rf inhomogeneity, misadjustments, and sample-induced variations in the rf tuning. The method is described anal., numerically, and exptl. for biol. solids. The authors demonstrate sensitivity gains of factors of 1.3 and 1.8 for typical 1H→15N and 15N→13C transfers and a combined gain of a factor of 2-4 for a typical NCA expt. for biol. solid-state NMR.
- 79Jain, S. K.; Nielsen, A. B.; Hiller, M.; Handel, L.; Ernst, M.; Oschkinat, H.; Akbey, U.; Nielsen, N. C. Phys. Chem. Chem. Phys. 2014, 16, 2827– 2830 DOI: 10.1039/c3cp54419bGoogle ScholarThere is no corresponding record for this reference.
- 80Hohwy, M.; Rienstra, C. M.; Griffin, R. G. J. Chem. Phys. 2002, 117, 4973– 4987 DOI: 10.1063/1.1488136Google ScholarThere is no corresponding record for this reference.
- 81Hohwy, M.; Rienstra, C. M.; Jaroniec, C. P.; Griffin, R. G. J. Chem. Phys. 1999, 110, 7983– 7992 DOI: 10.1063/1.478702Google ScholarThere is no corresponding record for this reference.
- 82Mittermaier, A.; Kay, L. E. J. Am. Chem. Soc. 1999, 121, 10608– 10613 DOI: 10.1021/ja9925047Google ScholarThere is no corresponding record for this reference.
- 83Sheppard, D.; Li, D. W.; Bruschweiler, R.; Tugarinov, V. J. Am. Chem. Soc. 2009, 131, 15853– 15865 DOI: 10.1021/ja9063958Google ScholarThere is no corresponding record for this reference.
- 84Franks, W. T.; Zhou, D. H.; Wylie, B. J.; Money, B. G.; Graesser, D. T.; Frericks, H. L.; Sahota, G.; Rienstra, C. M. J. Am. Chem. Soc. 2005, 127, 12291– 12305 DOI: 10.1021/ja044497eGoogle ScholarThere is no corresponding record for this reference.
- 85Van Geet, A. L. Anal. Chem. 1968, 40, 2227– 2229 DOI: 10.1021/ac50158a064Google ScholarThere is no corresponding record for this reference.
- 86Ernst, M.; Samoson, A.; Meier, B. H. J. Magn. Reson. 2003, 163, 332– 339 DOI: 10.1016/S1090-7807(03)00155-1Google ScholarThere is no corresponding record for this reference.
- 87Shaka, A. J.; Keeler, J.; Freeman, R. J. Magn. Reson. 1983, 53, 313– 340 DOI: 10.1016/0022-2364(83)90035-5Google ScholarThere is no corresponding record for this reference.
- 88Delaglio, F.; Grzesiek, S.; Vuister, G. W.; Zhu, G.; Pfeifer, J.; Bax, A. J. Biomol. NMR 1995, 6, 277– 293 DOI: 10.1007/BF00197809Google Scholar88https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXhtVSmurfK&md5=a670fca5b164083e2178fafd2fb951ffNMRPipe: a multidimensional spectral processing system based on UNIX pipesDelaglio, Frank; Grzesiek, Stephan; Vuister, Geerten W.; Zhu, Guang; Pfeifer, John; Bax, AdJournal of Biomolecular NMR (1995), 6 (3), 277-93CODEN: JBNME9; ISSN:0925-2738. (ESCOM)The NMRPipe system is a UNIX software environment of processing, graphics, and anal. tools designed to meet current routine and research-oriented multidimensional processing requirements, and to anticipate and accommodate future demands and developments. The system is based on UNIX pipes, which allow programs running simultaneously to exchange streams of data under user control. In an NMRPipe processing scheme, a stream of spectral data flows through a pipeline of processing programs, each of which performs one component of the overall scheme, such as Fourier transformation or linear prediction. Complete multidimensional processing schemes are constructed as simple UNIX shell scripts. The processing modules themselves maintain and exploit accurate records of data sizes, detection modes, and calibration information in all dimensions, so that schemes can be constructed without the need to explicitly define or anticipate data sizes or storage details of real and imaginary channels during processing. The asynchronous pipeline scheme provides other substantial advantages, including high flexibility, favorable processing speeds, choice of both all-in-memory and disk-bound processing, easy adaptation to different data formats, simpler software development and maintenance, and the ability to distribute processing tasks on multi-CPU computers and computer networks.
- 89Massiot, D.; Fayon, F.; Capron, M.; King, I.; Le Calve, S.; Alonso, B.; Durand, J. O.; Bujoli, B.; Gan, Z. H.; Hoatson, G. Magn. Reson. Chem. 2002, 40, 70– 76 DOI: 10.1002/mrc.984Google Scholar89https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38Xlt1ajuw%253D%253D&md5=d32cb46ca90ac83a8f2d62c95a93a88fModeling one- and two-dimensional solid-state NMR spectraMassiot, Dominique; Fayon, Franck; Capron, Mickael; King, Ian; Le Calve, Stephanie; Alonso, Bruno; Durand, Jean-Olivier; Bujoli, Bruno; Gan, Zhehong; Hoatson, GinaMagnetic Resonance in Chemistry (2002), 40 (1), 70-76CODEN: MRCHEG; ISSN:0749-1581. (John Wiley & Sons Ltd.)A review. With the description of more and more complex 1- and two-dimensional NMR expts. comes the need to develop methods to make a comprehensive interpretation of the various different expts. that can be carried out on the same sample or series of related samples. The authors present some examples of modeling 1- and two-dimensional solid-state NMR spectra of I = 1/2 spin and quadrupolar nuclei, using lab.-developed software that is made available to the NMR community.
- 90Schmidt, H. L. F.; Sperling, L. J.; Gao, Y. G.; Wylie, B. J.; Boettcher, J. M.; Wilson, S. R.; Rienstra, C. A. J. Phys. Chem. B 2007, 111, 14362– 14369 DOI: 10.1021/jp075531pGoogle ScholarThere is no corresponding record for this reference.
- 91Wylie, B. J.; Sperling, L. J.; Nieuwkoop, A. J.; Franks, W. T.; Oldfield, E.; Rienstra, C. M. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 16974– 16979 DOI: 10.1073/pnas.1103728108Google ScholarThere is no corresponding record for this reference.
- 92Schmidt, H. L.; Shah, G. J.; Sperling, L. J.; Rienstra, C. M. J. Phys. Chem. Lett. 2010, 1, 1623– 1628 DOI: 10.1021/jz1004413Google Scholar92https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC2srmtFehtA%253D%253D&md5=574992a689fabc298bf9be2a8fc6da69NMR Determination of Protein pK(a) Values in the Solid StateSchmidt Heather L Frericks; Shah Gautam J; Sperling Lindsay J; Rienstra Chad MThe journal of physical chemistry letters (2010), 1 (10), 1623-1628 ISSN:.Charged residues play an important role in defining key mechanistic features in many biomolecules. Determining the pK(a) values of large, membrane or fibrillar proteins can be challenging with traditional methods. In this study we show how solid-state NMR is used to monitor chemical shift changes during a pH titration for the small soluble β1 immunoglobulin binding domain of protein G. The chemical shifts of all the amino acids with charged side-chains throughout the uniformly-(13)C,(15)N-labeled protein were monitored over several samples varying in pH; pK(a) values were determined from these shifts for E27, D36, and E42, and the bounds for the pK(a) of other acidic side-chain resonances were determined. Additionally, this study shows how the calculated pK(a) values give insights into the crystal packing of the protein.
- 93Salonen, L. M.; Ellermann, M.; Diederich, F. Angew. Chem., Int. Ed. 2011, 50, 4808– 4842 DOI: 10.1002/anie.201007560Google Scholar93https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXlvVekt70%253D&md5=3684d00abc320d1ea0c71565495d5d5cAromatic rings in chemical and biological recognition: energetics and structuresSalonen, Laura M.; Ellermann, Manuel; Diederich, FrancoisAngewandte Chemie, International Edition (2011), 50 (21), 4808-4842CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. This review describes a multidimensional treatment of mol. recognition phenomena involving arom. rings in chem. and biol. systems. It summarizes new results reported since the appearance of an earlier review in 2003 in host-guest chem., biol. affinity assays and biostructural anal., data base mining in the Cambridge Structural Database (CSD) and the Protein Data Bank (PDB), and advanced computational studies. Topics addressed are arene-arene, perfluoroarene-arene, S···arom., cation-π, and anion-π interactions, as well as hydrogen bonding to π systems. The generated knowledge benefits, in particular, structure-based hit-to-lead development and lead optimization both in the pharmaceutical and in the crop protection industry. It equally facilitates the development of new advanced materials and supramol. systems, and should inspire further utilization of interactions with arom. rings to control the stereochem. outcome of synthetic transformations.
- 94Eswar, N.; Ramakrishnan, C. Protein Eng., Des. Sel. 2000, 13, 227– 238 DOI: 10.1093/protein/13.4.227Google ScholarThere is no corresponding record for this reference.
- 95Vijayakumar, M.; Qian, H.; Zhou, H. X. Proteins: Struct., Funct., Genet. 1999, 34, 497– 507 DOI: 10.1002/(SICI)1097-0134(19990301)34:4<497::AID-PROT9>3.0.CO;2-GGoogle ScholarThere is no corresponding record for this reference.
- 96Vugmeyster, L.; Ostrovsky, D.; Lipton, A. S. J. Phys. Chem. B 2013, 117, 6129– 6137 DOI: 10.1021/jp4021596Google ScholarThere is no corresponding record for this reference.
- 97Ottiger, M.; Bax, A. J. Am. Chem. Soc. 1999, 121, 4690– 4695 DOI: 10.1021/ja984484zGoogle ScholarThere is no corresponding record for this reference.
- 98Iijima, T.; Tsuchiya, S. J. Mol. Spectrosc. 1972, 44, 88– 107 DOI: 10.1016/0022-2852(72)90194-4Google ScholarThere is no corresponding record for this reference.
- 99Mittermaier, A.; Kay, L. E.; Forman-Kay, J. D. J. Biomol. NMR 1999, 13, 181– 185 DOI: 10.1023/A:1008387715167Google ScholarThere is no corresponding record for this reference.
- 100Sarkar, S. K.; Young, P. E.; Torchia, D. A. J. Am. Chem. Soc. 1986, 108, 6459– 6464 DOI: 10.1021/ja00281a002Google ScholarThere is no corresponding record for this reference.
- 101Barchi, J. J., Jr.; Grasberger, B.; Gronenborn, A. M.; Clore, G. M. Protein Sci. 1994, 3, 15– 21 DOI: 10.1002/pro.5560030103Google Scholar101https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXlt1yjtLw%253D&md5=1b2ff1da4f2eb56452cca955e9fd453bInvestigation of the backbone dynamics of the IgG-binding domain of streptococcal protein G by heteronuclear two-dimensional 1H-15N nuclear magnetic resonance spectroscopyBarchi, Joseph J., Jr.; Grasberger, Bruce; Gronenborn, Angela M.; Clore, G. MariusProtein Science (1994), 3 (1), 15-21CODEN: PRCIEI; ISSN:0961-8368.The backbone dynamics of the Ig-binding domain (B1) of streptococcal protein G, uniformly labeled with 15N, have been investigated by two-dimensional inverse detected heteronuclear 1H-15N NMR spectroscopy at 500 and 600 MHz. 15N T1, T2, and nuclear Overhauser enhancement data were obtained for all 55 backbone NH vectors of the B1 domain at both field strengths. The overall correlation time obtained from an anal. of the T1/T2 ratios was 3.3 ns at 26°. Overall, the B1 domain is a relatively rigid protein, consistent with the fact that over 95% of the residues participate in secondary structure, comprising a four-stranded sheer arranged in a -1, +3x, -1 topol., on top of which lies a single helix. Residues in the turns and loops connecting the elements of secondary structure tend to exhibit a higher degree of mobility on the picosecond time scale, as manifested by lower values of the overall order parameter. A no. of residues at the ends of the secondary structure elements display two distinct internal motions that are faster than the overall rotational correlation time: one is fast (<20 ps) and lies in the extreme narrowing limit, whereas the other is one to two orders of magnitude slower (1-3 ns) and lies outside the extreme narrowing limit. The slower motion can be explained by large-amplitude (20-40°) lumps in the N-H vectors between states with well-defined orientations that are stabilized by hydrogen bonds. In addn., residues in the helix and in the outer β-strands (particularly β-strand 2) display a small degree of chem. exchange line broadening, possibly due to a minor rotational motion of the helix relative to the sheer that curls around it.
- 102Seewald, M. J.; Pichumani, K.; Stowell, C.; Tibbals, B. V.; Regan, L.; Stone, M. J. Protein Sci. 2000, 9, 1177– 1193 DOI: 10.1110/ps.9.6.1177Google Scholar102https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXks1KqsL0%253D&md5=99d9f1cab2dfc1d2bc07a5f54bd341f6The role of backbone conformational heat capacity in protein stability: temperature dependent dynamics of the B1 domain of Streptococcal protein GSeewald, Michael J.; Pichumani, Kumar; Stowell, Cheri; Tibbals, Benjamin V.; Regan, Lynne; Stone, Martin J.Protein Science (2000), 9 (6), 1177-1193CODEN: PRCIEI; ISSN:0961-8368. (Cambridge University Press)The contributions of backbone NH group dynamics to the conformational heat capacity of the B1 domain of Streptococcal protein G have been estd. from the temp. dependence of 15N NMR-derived order parameters. Longitudinal (R1) and transverse (R2) relaxation rates, transverse cross-relation rates (ηxy), and steady state {1H}-15N nuclear Overhauser effects were measured at temps. of 0, 10, 20, 30, 40, and 50° for 89-100% of the backbone secondary amide nitrogen nuclei in the B1 domain. The ratio R2/ηxy was used to identify nuclei for which conformational exchange makes a significant contribution to R2. Relaxation data were fit to the extended model-free dynamics formalism, incorporating an axially sym. mol. rotational diffusion tensor. The temp. dependence of the order parameter (S2) was used to calc. the contribution of each NH group to conformational heat capacity (Cp) and a characteristic temp. (T*), representing the d. of conformational energy states accessible to each NH group. The heat capacities of the secondary structure regions of the B1 domain are significantly higher than those of comparable regions of other proteins, whereas the heat capacities of less structured regions are similar to those in other proteins. The higher local heat capacities are estd. to contribute up to ∼0.8 kJ/mol K to the total heat capacity of the B1 domain, without which the denaturation temp. would be ∼9° lower (78° rather than 87°). Thus, variation of backbone conformational heat capacity of native proteins may be a novel mechanism that contributes to high temp. stabilization of proteins.
- 103Idiyatullin, D.; Nesmelova, I.; Daragan, V. A.; Mayo, K. H. Protein Sci. 2003, 12, 914– 922 DOI: 10.1110/ps.0228703Google Scholar103https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXjtl2jt7w%253D&md5=6b5ab9f4ada5119f7396d3678d81eb53Comparison of 13CαH and 15NH backbone dynamics in protein GB1Idiyatullin, Djaudat; Nesmelova, Irina; Daragan, Vladimir A.; Mayo, Kevin H.Protein Science (2003), 12 (5), 914-922CODEN: PRCIEI; ISSN:0961-8368. (Cold Spring Harbor Laboratory Press)This study presents a site-resolved exptl. view of backbone CαH and NH internal motions in the 56-residue Ig-binding domain of streptococcal protein G, GB1. Using 13CαH and 15NH NMR relaxation data [T1, T2, and NOE] acquired at three resonance frequencies (1H frequencies of 500, 600, and 800 MHz), spectral d. functions were calcd. as F(ω) = 2ωJ(ω) to provide a model-independent way to visualize and analyze internal motional correlation time distributions for backbone groups in GB1. Line broadening in F(ω) curves indicates the presence of nanosecond time scale internal motions (0.8 to 5 nsec) for all CαH and NH groups. Deconvolution of F(ω) curves effectively separates overall tumbling and internal motional correlation time distributions to yield more accurate order parameters than detd. by using std. model free approaches. Compared to NH groups, CαH internal motions are more broadly distributed on the nanosecond time scale, and larger CαH order parameters are related to correlated bond rotations for CαH fluctuations. Motional parameters for NH groups are more structurally correlated, with NH order parameters for example, being larger for residues in more structured regions of β-sheet and helix and generally smaller for residues in the loop and turns. This is most likely related to the observation that NH order parameters are correlated to hydrogen bonding. This study contributes to the general understanding of protein dynamics and exemplifies an alternative and easier way to analyze NMR relaxation data.
- 104Schanda, P.; Huber, M.; Boisbouvier, J.; Meier, B. H.; Ernst, M. Angew. Chem., Int. Ed. 2011, 50, 11005– 11009 DOI: 10.1002/anie.201103944Google ScholarThere is no corresponding record for this reference.
- 105Derrick, J. P.; Wigley, D. B. Nature 1992, 359, 752– 754 DOI: 10.1038/359752a0Google ScholarThere is no corresponding record for this reference.
- 106Gronenborn, A. M.; Clore, G. M. J. Mol. Biol. 1993, 233, 331– 335 DOI: 10.1006/jmbi.1993.1514Google ScholarThere is no corresponding record for this reference.
- 107Lian, L. Y.; Barsukov, I. L.; Derrick, J. P.; Roberts, G. C. K. Nat. Struct. Biol. 1994, 1, 355– 357 DOI: 10.1038/nsb0694-355Google ScholarThere is no corresponding record for this reference.
- 108Lamley, J. M.; Iuga, D.; Oster, C.; Sass, H. J.; Rogowski, M.; Oss, A.; Past, J.; Reinhold, A.; Grzesiek, S.; Samoson, A.; Lewandowski, J. R. J. Am. Chem. Soc. 2014, 136, 16800– 16806 DOI: 10.1021/ja5069992Google ScholarThere is no corresponding record for this reference.
- 109Kato, K.; Lian, L. Y.; Barsukov, I. L.; Derrick, J. P.; Kim, H. H.; Tanaka, R.; Yoshino, A.; Shiraishi, M.; Shimada, I.; Arata, Y.; Roberts, G. C. K. Structure 1995, 3, 79– 85 DOI: 10.1016/S0969-2126(01)00136-8Google ScholarThere is no corresponding record for this reference.
- 110Wand, A. J. Curr. Opin. Struct. Biol. 2013, 23, 75– 81 DOI: 10.1016/j.sbi.2012.11.005Google Scholar110https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhvVektrzL&md5=afdc384ca87a383b523699bbc8666f38The dark energy of proteins comes to light: conformational entropy and its role in protein function revealed by NMR relaxationWand, A. JoshuaCurrent Opinion in Structural Biology (2013), 23 (1), 75-81CODEN: COSBEF; ISSN:0959-440X. (Elsevier Ltd.)A review. Historically it has been virtually impossible to exptl. det. the contribution of residual protein entropy to fundamental protein activities such as the binding of ligands. Recent progress has illuminated the possibility of employing NMR relaxation methods to quant. det. the role of changes in conformational entropy in mol. recognition by proteins. The method rests on using fast internal protein dynamics as a proxy. Initial results reveal a large and variable role for conformational entropy in the binding of ligands by proteins. Such a role for conformational entropy in mol. recognition has significant implications for enzymol., signal transduction, allosteric regulation and the development of protein-directed pharmaceuticals.
- 111Marlow, M. S.; Dogan, J.; Frederick, K. K.; Valentine, K. G.; Wand, A. J. Nat. Chem. Biol. 2010, 6, 352– 358 DOI: 10.1038/nchembio.347Google Scholar111https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXksFakt78%253D&md5=e9ee34fffa91f612441a77a7385015ffThe role of conformational entropy in molecular recognition by calmodulinMarlow, Michael S.; Dogan, Jakob; Frederick, Kendra K.; Valentine, Kathleen G.; Wand, A. JoshuaNature Chemical Biology (2010), 6 (5), 352-358CODEN: NCBABT; ISSN:1552-4450. (Nature Publishing Group)The phys. basis for high-affinity interactions involving proteins is complex and potentially involves a range of energetic contributions. Among these are changes in protein conformational entropy, which cannot yet be reliably computed from mol. structures. We have recently used changes in conformational dynamics as a proxy for changes in conformational entropy of calmodulin upon assocn. with domains from regulated proteins. The apparent change in conformational entropy was linearly related to the overall binding entropy. This view warrants a more quant. foundation. Here we calibrate an 'entropy meter' using an exptl. dynamical proxy based on NMR relaxation and show that changes in the conformational entropy of calmodulin are a significant component of the energetics of binding. Furthermore, the distribution of motion at the interface between the target domain and calmodulin is surprisingly noncomplementary. These observations promote modification of our understanding of the energetics of protein-ligand interactions.
- 112Frederick, K. K.; Marlow, M. S.; Valentine, K. G.; Wand, A. J. Nature 2007, 448, 325– 329 DOI: 10.1038/nature05959Google ScholarThere is no corresponding record for this reference.
- 113Lee, A. L.; Wand, A. J. Nature 2001, 411, 501– 504 DOI: 10.1038/35078119Google ScholarThere is no corresponding record for this reference.
Supporting Information
Supporting Information
ARTICLE SECTIONSThe Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.5b12974.
Uniform 2H magnetization transfer in 2H–13C adiabatic RESPIRATION CP (Figures S1, S2, and S3 and Table S1), Ala 2H one-pulse spectrum fit (Figure S4), GB1 2H line shape fits (Figure S5), Ala dynamics in GB1 (Table S2 and Figure S6), backbone order parameters derived from 2H measurement and CH dipolar measurement (Figure S7), GB1 2Hα η̅ values (Figure S8) (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.