Abstract
Viruses, relatively simple pathogens, are able to replicate in many living organisms and to adapt to various environments. Conventional atomic-resolution structural biology techniques, X-ray crystallography and solution NMR spectroscopy provided abundant information on the structures of individual proteins and nucleic acids comprising viruses; however, viral assemblies are not amenable to analysis by these techniques because of their large size, insolubility, and inherent lack of long-range order. In this article, we review the recent advances in magic angle spinning NMR spectroscopy that enabled atomic-resolution analysis of structure and dynamics of large viral systems and give examples of several exciting case studies.
Keywords: viruses, magic angle spinning, HIV-AIDS, CA capsid protein, bacteriophage, influenza
1. Introduction
Viruses are relatively simple pathogens that are intimately linked with all forms of life. They are adept at replicating in bacteria, archea, protists, fungi, plants, and animals while adapting to a variety of environmental conditions, some of which are extremely abrasive [1–5]. Once a virus penetrates a host cell and releases its genetic material, it is wholly reliant on the infected host to carry out its viral life cycle. Viruses seize control of their host cell machinery to replicate and assemble new virus particles for release and propagation. This leads to infections that can have a significantly negative impact on the health of the host cell and eventually lead to cell death [6].
In contrast to the generally negative perception of viruses, they do possess many potentially beneficial applications [7, 8]. Although viruses and living beings share a close relationship, there are still many questions regarding their structure and function. There are a variety of different methods for studying viruses at different levels of spatial and temporal resolution, including: transmission electron microscopy (TEM) [9], cryo-electron microscopy (cryo-EM) [10], cryo-electron tomography (cryo-ET) [11], X-ray crystallography [12], solution-state NMR [13], atomic force microscopy (AFM) [14], mass spectrometry (MS) [15], circular dichroism (CD) [16], optical tweezers [17], molecular dynamics (MD) [18], and solid-state NMR [19–21]. Of these, only X-ray crystallography, solution-state NMR, and solid-state NMR yield atomic level resolution, with recent exciting advances in cryo-EM bringing the resolution close to atomic for that technique as well [22].
X-ray crystallography is limited to samples that can be crystallized. Solution NMR requires that samples be soluble and is limited to relatively small systems. In contrast, solid-state NMR (SSNMR) has no requirement with respect to long-range order, solubility, or the molecular weight of the system under analysis. SSNMR has therefore found an important role in the structural biology of viruses. The rich information content of the experiments, including insights into structure and dynamics as well as interactions with host cell factors and small-molecule inhibitors, coupled with a wide range of sample conditions amenable for characterization, and the high-interest in studying viral systems has led to many exciting studies on viruses.
One of the most powerful and commonly used SSNMR techniques for studying viral systems is magic angle spinning (MAS) [23, 24]. Under MAS conditions, high-resolution NMR spectra of viruses and their constituent macromolecules (proteins, nucleic acids, and their assemblies) can be collected, and site-specific resonance assignments can be attained for many such systems through 2D and 3D correlation experiments. With chemical shift assignments in hand, additional MAS experiments are performed to acquire the distance and orientational information necessary to build atomic-level three-dimensional structures, to gain insight into dynamic processes on a variety of timescales, and reveal interactions with lipids or other molecules that bind to the macromolecules constituting the virus. The recent advances in fast MAS technologies allowing for sample spinning at frequencies of 40 – 110 kHz have enabled atomic-resolution studies of challenging virus systems which would have been considered intractable only several years ago. Microscopic images of some viral systems that have been studied by MAS NMR are shown in Fig. 1.
Fig. 1.
Images of viral systems that contain individual protein domains, peptides, or entire assemblies that have been investigated by solid-state MAS NMR. (A) fd bacteriophage, (B) the T4 bacteriophage, (C) the T7 bacteriophage, (D) the Pf1 bacteriophage, (E) influenza A, (F) parainfluenza PIV5, (G) measles, (H) HIV-1. (A) Reprinted with permission from Wang et al. J. Mol. Biol. 2006, 361 (2), pp 209–215. Copyright 2006 Elsevier. (B) Reprinted with permission from Danev et al. J. Struct. Biol. 2010, 171 (2), pp 174–181. Copyright 2010 Elsevier. (C) Reprinted with permission from Serwer, J. Ultra Mol. Struct. R. 1977, 58 (3), pp 235–243. Copyright 1977 Elsevier. (D) Reprinted with permission from Kneale et al. J. Mol. Biol. 1982, 156 (2) pp 279–292. Copyright 1982 Elsevier. (E) From Centers for Disease Control and Prevention (CDC) Public Health Image Library (PHIL), content provided by CDC and Dr. F. A. Murphy. (F) Reprinted with permission from Terrier et al. Virus Res. 2009, 142 (1) pp 200–203. Copyright Elsevier 2009. (G) From Centers for Disease Control and Prevention (CDC) Public Health Image Library (PHIL), content provided by CDC, Cynthia S. Goldsmith, and William Bellini, Ph. D. (H) From Centers for Disease Control and Prevention (CDC) Public Health Image Library (PHIL), content provided by CDC and Dr. Edwin P. Ewing.
In this article, we review the recent advances in SSNMR methodologies as applied to viral systems and give examples of several exciting case studies. We focus our attention mainly on systems which were studied by MAS NMR. We note that there have been beautiful studies conducted on viral systems using static SSNMR, these are beyond the scope of this review and have been discussed in depth elsewhere [1, 4, 25–53].
2. Recent Methodological Advances for MAS NMR Studies of Viruses
Sensitivity and resolution are significant challenges to the study of large biomolecules by MAS NMR. Advances in hardware, NMR methodology, and sample preparation are continually improving the attainable sensitivity and resolution, permitting analysis of increasingly complex biological systems, including entire viruses and assemblies of their constituent molecules. Key breakthroughs have included the use of extensive and selective labeling including sample deuteration, magnetic fields of up to 1 GHz, development of MAS NMR coils with reduced electric-field penetration into the sample (“low-E”, “EFree, and “scroll coil”) [54–59], fast MAS enabling proton detection, dynamic nuclear polarization (DNP) sensitivity enhancements, as well as nonuniform sampling. In this section we will discuss recent methodological developments in biological MAS NMR and their application to viral systems.
2.1 Sample Preparation
One of the key steps in obtaining useful, interpretable, and relevant information about viruses and their constituent macromolecules and assemblies is the preparation of high-quality samples. In the subsequent discussion we will focus mainly on preparation of samples of proteins and protein assemblies. There are intense developments in the sample preparation of viral DNA and RNA systems [60–62], but these will not be addressed here in any detail.
Protein expression and purification as well as peptide synthesis techniques are crucial to the success of many structural biology methods, including MAS NMR. One key ingredient in the preparation of structurally (morphologically) and functionally relevant viral protein assemblies is to ensure their assembly and/or reconstitution in biologically relevant conditions. Another key consideration is the choice of isotopic labeling scheme to obtain the desired information content with the minimal number of samples. Various procedures have been implemented to prepare viral systems for MAS NMR analysis, and some examples are considered below.
Intact or Assembled Viral Systems
Several of the viral systems studied by MAS NMR are either intact viruses or assemblies of proteins that are prepared for MAS NMR by centrifugation, lyophilization, precipitation [19–21, 63–73], crystallization [74], or sedimentation [75]. Some of these systems are difficult or impossible to crystallize and have molecular weights that are not compatible with solution NMR.
Membrane Reconstitution
For membrane proteins and peptides, the choice of the lipid for reconstitution plays a major role in the physical properties of the system in question. The structure and dynamics of a given protein or peptide can differ dramatically depending on the choice of lipid [43, 44, 76–80].
Uniform 13C/15N Isotopic Labeling
Uniform 13C/15N isotopic labeling is the most common [81]labeling strategy for the MAS NMR study of viruses and viral proteins. 13C and 15N labels are incorporated into the growth media during protein expression as U-13C6 glucose and 15NH4Cl as the sole carbon and nitrogen sources, respectively. If synthetic peptides are investigated, 13C, 15N labeled amino acids could be used during the solid-state synthesis [51]; this approach is limited to relatively short peptides due to high costs. An alternative approach is an expression of U-13C, 15N fusion protein in E. coli, followed by a cleavage of the tag to release the short peptide of interest [81]. Uniform labeling is an attractive option as it provides the richest information content for a given sample. The major drawback is spectral overlap, particularly in large molecules, that may complicate or even preclude resonance assignments for a significant fraction of residues. Spectral overlap can be alleviated with the increased-dimensionality experiments (3D and 4D) albeit the experimental time increases dramatically. A common alternative is selective, extensive, and sparse amino acid labeling considered below.
Selective Amino Acid Labeling
Selective amino acid labeling is used both in recombinant protein expression [70, 82, 83] and in peptide synthesis [43, 44, 77, 84–86]. The choice of amino acid(s) to be incorporated into a system under analysis is dictated by the desired information content as well as by which amino acids can be incorporated recombinantly without scrambling. In a recent study from the authors’ lab, tyrosine residues were used as a probe of internal dynamics in HIV-1 CA capsid protein assemblies [70]. 13C, 15N-Tyrosine was added to the growth media, which contained 14NH4Cl and 12C6-glucose, before induction by IPTG. In this particular example, CA protein contains only four tyrosine residues, which are located in α-helical regions of the N- and C-terminal domains (NTD and CTD, respectively) and in a hinge region connecting these domains [70]. Tyrosine therefore was an ideal probe of the hinge region dynamics occurring on micro- to millisecond timescales, with the three helical tyrosine residues serving as internal controls. In another example, Yu and Schaefer prepared a selectively L-[ε-15N] lysine sample by adding L-[ε-15N] lysine to modified M9 medium during growth for use in REDOR studies of bacteriophage T4 [83]. Morag, Abramov, and Goldbourt recently employed a similar concept where uniform 13C, 15N labels were selected for by the use of U-13C6 glucose and 15NH4Cl but unlabeled aromatic amino acids (Trp, Phe, Tyr) were also added to the growth media. This blocked the production of the aromatic amino acids, making them NMR silent, which allowed for the observation of DNA signals displaying chemical shifts in the region where the protein aromatic signals would have been present if labeled [64].
Even more versatility can be achieved by using solid-state peptide synthesis, where a judicious combination of unlabeled and 13C, 15N labeled amino acids can produce a labeling pattern that highlights a region of the sequence with high biological significance [87, 88]. The Hong lab has used this approach to label selectively different regions of the M2 transmembrane peptide from the Influenza A virus. Additionally, they have incorporated specialized labels to measure intermolecular distances, such as 4-19F-Phe probes [89].
Sparse and Extensive Labeling
The benefits of sparse and extensive labeling are two-fold. First, the reduction of labeled sites reduces the complexity of the spectra making interpretation easier. Second, resolution is improved in the remaining signals due to the removal of one-bond J couplings. Sparse/extensive labeling can be achieved in several different ways. Choosing [1,3-13C]- or [2-13C]-glycerol as the sole carbon source gives rise to distinct and characteristic labeling patterns for each type of label resulting from the metabolic pathways of amino acid production from glycerol in E. coli. Different characteristic patterns are obtained when [1-13C] or [2-13C]-glucose is used as the sole carbon source [72, 90]. Goldbourt, Day, and McDermott have explored the Entner-Doudoroff (ED) pathway for sparse labeling in the Pf1 bacteriophage. Using 1-13C glucose as the sole carbon source, they prepared samples extensively labeled in the backbone carbonyl region while aliphatic and aromatic carbons are almost entirely suppressed [69].
Deuteration
Deuteration of proteins to varying degrees has been applied to several viral assemblies including Pf1 [91], AP205, and the M2 proton conduction channel [74, 75]. Perdeuteration improves resolution by diluting the 1H spins, thus eliminating 1H-1H couplings [92, 93]. During the expression of perdeuterated proteins, 2H2O is used in place of 1H2O, and [2H, 13C]-glucose is used as the sole carbon source. Once the protein is prepared, protons can be reintroduced at exchangeable sites by back exchange against buffers containing the desired ratio of H2O/D2O buffers [94–97].
2.2 Resonance Assignments
Resonance assignments are required for subsequent site-specific structural and dynamic analysis by NMR. For small peptides derived from viral proteins, resonance assignments can be achieved by 2D correlation experiments. For large viral proteins, 3D spectra are often required. As discussed above, isotopic dilution through amino-acid selective, sparse, or extensive labeling can be employed for spectral simplification and to facilitate resonance assignments [69, 70, 72]. For 13C-detected data sets, assignments of viral proteins follow a well-established protocol. 2D 13C-13C correlation spectra provide valuable information about amino acid types and side chain correlations, enabling spin-system assignments. Backbone walks can be achieved in viral proteins by combining three types of 3D experiments: NCACX, NCOCX, and CANCO [19–21, 69, 98]. Recently, Pintacuda and colleagues have reported a resonance assignment protocol based on proton-detected solid-state NMR at 60 kHz MAS, demonstrated on the Acinetobacter phage 205 (AP205) and the M2 protein of influenza A reconstituted in a lipid bilayer [74, 75]. The 3D (H)(CO)CA(CO)NH experiment with CO-CA-CO out-and-back scalar transfer provides inter-residue correlations, while the 3D (H)(CA)CB(CA)NH experiment with CA-CB-CA out-and-back scalar transfer provides intra-residue correlations. Additional experiments included (H)CONH, (H)CANH, (H)CO(CA)NH, and (H)(CA)CB(CACO)NH. Once the full sequence of six experiments was executed, the MATCH program [99] was used to perform residue-specific backbone assignments automatically. For AP205, one week of experiment time was necessary to collect the suite of six experiments, which resulted in assignments for 94 of the 130 (72%) total residues. For M2, two weeks of experiment time resulted in 44 assignments using MATCH.
Compared to proteins, nucleic acids have been less studied by MAS NMR. Currently, there are no well-developed general MAS NMR assignment protocols for nucleic acids. Corresponding signals in nucleotides usually have very similar chemical shifts, yielding highly congested spectra and making assignments difficult. Moreover, inter-nucleotide correlations are hard to obtain due to the phosphate linkage and consequently long 13C-13C/13C-15N distances. Despite these challenges, several reports have presented assignment approaches for single-stranded DNA in bacteriophages [64, 65, 67]. Resonance assignments of nucleic acids based on 2D 13C-13C correlation spectra usually start with nuclei whose chemical shifts are well resolved [100]. Subsequently, nucleotide spin systems can be assigned based on the previously assigned peaks and their intra-nucleotide cross-peaks, similar to the approach used for side chain assignments for proteins. Assignments of inter-nucleotide correlations can then be inferred from unique cross-peaks [64].
2.3 High Magnetic Fields
The recent development of magnetic fields of 17.6 – 28.1 T has been critical for analysis of large biomolecular systems by MAS NMR, including viruses and assemblies of their constituent macromolecules. The work from the authors’ group on HIV-1 CA protein assemblies has underscored the importance of high fields (17.6 – 21.1 T) to attain the requisite sensitivity and resolution [20, 21]. Pintacuda and co-workers have used magnetic fields of 23.5 T in conjunction with fast MAS and 1H detection (discussed below) to study the measles virus (MeV) nucleocapsid, M2, and AP205 bacteriophage, and demonstrated that the combination of these three technologies produced outstanding-quality data with a fraction of sample required for conventional experiments [75, 101]. High magnetic fields are envisioned to be “a must” for atomic-resolution analysis of multicomponent assemblies of viral macromolecules and of intact viruses.
2.4 Fast Magic Angle Spinning
With the advent of probes capable of spinning at MAS frequencies of 40 – 110 kHz, developments of fast MAS experiments and their applications are rapidly gaining momentum as fast MAS offers dramatically enhanced sensitivity and resolution. With MAS frequencies of 40 kHz and above, 1H detection is readily attainable due to the efficient suppression of 1H-1H homonuclear dipolar couplings, resulting in sharp proton lines, particularly at MAS frequencies above 60 kHz. This also results in greatly improved sensitivity, especially in challenging systems such as viral assemblies and membrane proteins. 1H detection can be performed both in fully protonated and perdeuterated samples [92, 95, 102, 103]. At MAS frequencies of 40–110 kHz, many of the canonical dipolar and chemical shift anisotropy (CSA) recoupling experiments fail. Therefore, much emphasis in the field has been placed on the development of effective dipolar and CSA recoupling schemes. Some of the contemporary techniques were reviewed by the authors recently [104].
2.4.1 Spin Diffusion Experiments
Work in the authors’ laboratory employs homonuclear 13C-13C R-symmetry based spin diffusion sequences for correlation spectroscopy, particularly at MAS frequencies of 40 kHz and above where conventional PDSD and DARR experiments fail [71]. In such cases, the R2-symmetry [105] and CORD [106] experiments were demonstrated to work efficiently in a broad range of systems, including HIV-1 CA protein assemblies. Most recently, a combined RFDR-CORD experiment was developed that exhibits superior performance to both RFDR and CORD methods at fast MAS (40–60 kHz), results in fully broadbanded 13C-13C correlation spectra with high cross peak intensities, and yields long-range cross peaks with monotonous dependence of polarization transfer rates on distances [107]. This recent experiment is anticipated to be advantageous for structure determination of viral assemblies.
2.4.2 Dipolar Recoupling Experiments
At fast MAS frequencies (40 kHz and above), 1H-13C and 1H-15N heteronuclear recoupling is very efficient by R-symmetry sequences [105]. R1632-symmetry based 1H-13C and 1H-15N DIPSHIFT experiments were demonstrated in fully protonated conical assemblies of U-13C, 15N tyrosine labeled HIV-1 CA protein assemblies. These experiments revealed that the tyrosine residues in CA are rigid on the timescales of nano- to microseconds, including the hinge residue Y145, which undergoes motions on much slower, micro- to millisecond timescales [70]. Most recently, an improved R-symmetry based experiment dubbed PARS was developed, where the accuracy of the measurements of heteronuclear dipolar couplings is greatly improved compared to the conventional DIPSHIFT [108]. This is accomplished by supercycling the R-symmetry elements in the sequence, which eliminates residual second-order interactions, such as 1H CSA, that introduce non-negligible error in the determination of heteronuclear dipolar couplings [109, 110].
It is important to note that these R-symmetry dipolar recoupling sequences are well-suited for moderate MAS frequencies as well, and the present authors employ them extensively for probing nano- to microsecond timescale mobility in a variety of systems, including HIV-1 protein assemblies [104].
2.4.3 Heteronuclear CSA recoupling
R-sequences of appropriate symmetry are useful for efficient recoupling of the heteronuclear chemical shift anisotropy (CSA), as was demonstrated originally by Zhao et al. [111]. R-symmetry sequences were employed under fast and moderate MAS conditions to recouple 15N and 13C CSAs in the context of 2D and 3D experiments in various systems, including HIV-1 CA protein assemblies [70, 104, 105, 109, 112] Measurement of heteronuclear CSA tensors is essential in viral systems as they provide a probe of hydrogen bonding environments and dynamics [109, 113].
2.4.4 Proton Detection
Proton detection is particularly advantageous for mass-limited samples, such as intact viral assemblies, which can be difficult to obtain in large quantities. 1H detection has found broad applications in MAS NMR including resonance assignments [114–117], structure determination [118, 119] including 1H-1H distance restraints [120], studies of protein dynamics [93, 121–124], and protein-RNA interactions [125].
Opella and coworkers recently demonstrated the application of 1H detection for the measurement of 1H-15N dipolar couplings in filamentous Pf1 at 50 kHz MAS [91]. High sensitivity was attained with ~0.5 mg perdeuterated protein, and the resolution was high with both full and partial amide proton back-exchange. Pintacuda and co-workers utilized 1H detection for studies of both fully protonated and amide 1H back-exchanged AP205 bacteriophage capsids, measles virus nucleocapsid, as well as M2, as shown in Fig. 2 [74, 75, 101]. A combination of deuteration, fast MAS, and high magnetic fields was required in all these studies. Importantly, the high-quality spectra enabled the use of automated assignment protocols permitting the process to be streamlined [99], relative to the time-intensive manual assignment that is generally required in MAS NMR.
Fig. 2.
Proton-detected 15N-1H spectra of [U-HN,2H,13C,15N]-labeled proteins acquired at 1 GHz, 60 kHz MAS. (A) microcrystalline SH3, (B) microcrystalline β2m, (C) sedimented nucleocapsids of AP205 bacteriophage, (D) M2 channel, and (E) OmpG. Reprinted with permission from Barbett-Massin et al, J. Am. Chem. Soc., 2014, 136 (35), pp 12489–12497. Copyright 2014 American Chemical Society.
2.5 Nonuniform Sampling
Nonuniform sampling (NUS) has been routinely applied in solution state NMR [126–130]. Typically NUS has been applied to streamline the data collection process in cases where sensitivity is not a limiting factor. However, in the case of sensitivity-limited MAS NMR experiments on complex biological systems, NUS can be exploited to attain bona fide sensitivity gains in the time domain, provided appropriate sampling schedules, as was derived analytically by Rovnyak [131]. More recently, Rovnyak and colleagues demonstrated experimentally and analytically that 2-fold sensitivity enhancements can be achieved in each indirect dimension using random, exponentially-weighted nonuniform sampling schedules [131, 132]. Processing NUS data requires specialized, typically nonlinear, reconstruction algorithms. Examples of these include maximum entropy (MaxEnt) [133–135], and its linear regime implementation dubbed maximum entropy interpolation (MINT) [136], forward and fast forward maximum entropy [137–140], GFT [141, 142], covariance NMR [143, 144], and SIFT [145, 146]. There are multiple acquisition schedules and data reconstruction algorithms, detailed elsewhere [133, 147–150].
While NUS has been used extensively in solution NMR, its applications to MAS NMR are only now gaining momentum, and the advantages of this approach as well as the development of standard protocol for NUS data collection and processing are not yet fully realized. To date, in addition to the authors’ work on thioredoxin and LC8, [132, 136, 151], NUS and variations of the technique have been applied to several other biological systems in the context of MAS NMR, including GB1 [145, 152] ubiquitin [118], SH3 [44], Aβ fibril [153], proteorhodopsin [152], a truncated MerF construct [154], and magnetically-aligned, phospholipid-embedded Pf1 [144, 154]. It is anticipated that the application of nonuniform sampling to complex viral systems will be a widely used approach for gaining sensitivity or reducing the experiment time in multidimensional MAS NMR experiments.
2.6 Dynamic Nuclear Polarization
Dynamic nuclear polarization (DNP) can provide drastic sensitivity enhancement, theoretically of up to 660 fold compared to the conventional NMR spectroscopy. In DNP experiments, polarization of unpaired electrons is transferred to nuclear spins by microwave irradiation at the EPR frequency, leading to large signal enhancements. These require introduction of radicals or biradicals, such as TEMPO or TOTAPOL, into the system [155, 156]. If an endogenous radical is present in the sample, its unpaired electron can also be used as a polarization source, albeit with much lower efficiency [157–159]. Key to the study of viruses and other complex biological systems with DNP in conjunction with MAS NMR have been developments such as high magnetic fields, high frequency microwave sources, DNP probes, and new radical polarizing agents [160]. Detailed review of DNP theory and practical aspects is beyond the scope of this article, and below only relevant applications to viral systems are discussed very briefly.
DNP studies of fd bacteriophage demonstrated that polarization from TEMPO can be distributed to the encapsulated DNA, despite the absence of the direct contact with the radicals in the solvent [161]. DNP sensitivity enhancement enabled the observation of the significantly weaker DNA peaks inside the capsid of Pf1 as compared to the coat protein resonances. This allowed for nucleotide-type assignment of the DNA. From this, McDermott and co-workers concluded that the DNA of Pf1 is highly extended and twisted as compared to other bacteriophages [67]. With cryogenic DNP, Griffin and co-workers were able to observe the external binding site(s) of rimantadine with the influenza A M2 proton transporter and measure the protein-ligand distance with an accuracy of ±0.2 Å. The application of DNP was particularly powerful for the observation of this weak (mM) protein-ligand interaction [162]. Overall, DNP is rapidly becoming an important technique for the study of large viral and other protein assemblies and even whole cells [163, 164].
3. Structure and Dynamics of Viral Systems with MAS NMR
In this section, we discuss applications of MAS NMR to both proteins and nucleic acids of numerous viral systems, including the retrovirus HIV-1, bacteriophages, and lipid membrane enveloped viruses.
3.1 HIV-1
Human Immunodeficiency Virus Type 1 (HIV-1), the causative agent of an acquired immunodeficiency syndrome (AIDS), is a retrovirus carrying its genetic information in the form of single-stranded RNA. Once inside the cytoplasm of the host cell, viral RNA is released from the capsid core, and reverse transcribed into DNA, and integrated into the host cell’s genome. HIV-1 consists of an outer envelope composed of lipids and glycoprotein, a viral capsid, a dimer RNA, and viral enzymes including protease, reverse transcriptase, and integrase [165]. MAS NMR spectroscopy has emerged as a popular method to study HIV-1. In this section, we review the highlights of MAS NMR studies of HIV-1.
3.1.1 HIV-1 Capsid Protein Assemblies
HIV-1 encodes a polyprotein called Gag, which directs the packaging of viral genome and assembles into immature, noninfectious virions [166]. In the maturation process, the Gag polyprotein is cleaved at five positions into its constituent domains, forming a conical capsid core to enclose the viral genome and proteins, as illustrated in Fig. 3A [21, 167–169]. The HIV-1 capsid contains approximately 1200 copies of 25.6 kDa capsid protein (CA), and in an idealized model can be thought as a ‘fullerene cone’, composed of a hexameric lattice, with 12 pentameric CA subunits incorporated to complete the closed structure [170, 171]. Recently, a pseudoatomic structure was solved of a CA capsid, using a hybrid approach including all-atom molecular dynamics simulations, cryo-EM, and solution NMR spectroscopy [172]. The architecture of the HIV-1 capsid is shown in Fig. 3B [21, 170, 172]. In vitro, CA assemblies can exhibit conical, tubular, spherical, or donut-like morphologies under different assembly conditions [170, 173–175].
Fig. 3.

(A) Domain organization of the Gag polyprotein and schematic illustration of the proteolytic cleavages during the HIV-1 maturation. The proteolytic cleavage sites are indicated by arrows. (B) Top: All-atom molecular dynamics derived model of mature HIV-1 capsid constructed on the basis of cryo-ET studies. The capsid is comprised of 216 CA hexamers (depicted in orange) and 12 CA pentamers (depicted in magenta), PDB ID 3J3Y. Bottom: the CA hexamer of hexamers (HOH) building block, viewed from the top (left) and from the side (right), PDB ID 3J34. Reprinted with permission from Han et al, J. Am. Chem. Soc., 2013, 135 (47), pp 17793–17803. Copyright 2013 American Chemical Society.
Work from the authors’ laboratory established sample preparation protocols for generating conical, spherical, and tubular assemblies of CA, which are particularly suitable for high-resolution structural analysis by MAS NMR spectroscopy [20]. Among these three kinds of CA assemblies of different morphologies, tubular assemblies of CA yielded so far the best quality of MAS NMR spectra with remarkably high resolution, as shown in Fig. 4, while conical and spherical assemblies of CA also exhibit reasonably high resolution, permitting the determination of structure and dynamics at atomic level [20, 21]. Comparison between the spectra of conical and spherical assemblies showed no significant differences, indicating that the structure of CA is very similar in different morphologies. Using data sets acquired at 21.1 T, resonance assignments were completed for 66% of CA protein in conical assemblies [20]. The hinge region between the N-terminal domain and C-terminal domain was found to be flexible on the millisecond timescales as probed through 13C-15N and 1H-15N dipolar coupling measurements in R-based DIPSHIFT experiments discussed above [70]. These hinge motions were proposed to allow the protein to access multiple conformers, permitting the formation of different assembly morphologies.
Fig. 4.
(A) 2D spectra of U-13C,15N tubular assemblies of HIV-1 CA-SP1 A92E acquired at 19.9 T and 4 °C: NCACX (top), NCOCX (middle right and bottom), and DARR (middle left and middle center). The 1D trace at 112.4 ppm along the 15N dimension of the NCACX spectrum is shown on top. The DARR mixing time was 50 ms. Backbone walk for the stretch of residues R100-T110 is shown on the spectra. (B) Expansion of the 2D DARR spectrum around the region containing the Cα-Cβ resonances of Ser and Thr residues. Reprinted with permission from Han et al, J. Am. Chem. Soc., 2013, 135 (47), pp 17793–17803. Copyright 2013 American Chemical Society.
The authors also addressed HIV-1 maturation process. They have examined the Gag cleavage intermediate CA-SP1 assemblies. As the final step in the maturation process, the cleavage of CA-SP1 triggers the formation of the conical capsid core. The detailed mechanism of this process is not fully understood, including the conformation of SP1 peptide in CA-SP1, which has been a subject of debate [176–178]. To investigate the conformation and dynamics of CA-SP1 assemblies, dipolar- and scalar-based correlation experiments were employed; the measurements revealed that the SP1 peptide adopts a dynamic random coil conformation in CA-SP1 assemblies [21]. Furthermore, analysis of two variants of the CA protein with slightly different primary sequences indicated that sequence variations at distal sites induce conformational plasticity in the SP1 region, in the Cyclophilin A-binding loop, and in the loop preceding helix 8.
Chen and Tycko prepared tubular CA assemblies, observed by dark-field transmission electron microscopy to be mixtures of single- and multiwall tubes, which yielded congested MAS NMR spectra [73]. Subsequently, Bayro et al. improved the sample preparation and acquired well-resolved homo- and heteronuclear correlation spectra [72]. Using the resonance assignment program MCASSIGN2 [179], along with traditional, manual spectral analysis, Bayro et al. assigned about 69% of the residues. They also suggested that the structures of the N-terminal domain and C-terminal domain of CA protein in tubular assemblies remain largely the same as observed in solution, and consistent with the observations from the authors’ laboratory [20, 21]. From the resonance assignments and additional two-dimensional 1H T2-filtered NCA experiments, Bayro et al. distinguished regions that are well-ordered, regions that are dynamically or statically disordered, and regions that are partially mobile. They also carried out 15N-15N backbone recoupling (15N-BARE) experiments to measure distances between sequential amides, thus probing the backbone conformation, as shown in Fig. 5 [72, 180]. By comparing the experimental decay curves obtained in tubular assemblies with those calculated from solution NMR and crystal structures, local conformational changes were detected, including the 310-helix in the C-terminal domain. Structure calculations using resonance assignments and 15N-15N couplings as restraints via the Xplor-NIH program revealed multiple structural changes upon tube formation [72].
Fig. 5.
15N-BARE signal decay curves for the indicated residues. These curves are measurements of sequential amide 15N–15N distances, which are conformation dependent. Filled circles are experimental MAS NMR data of HIV-1 CA tubes. Broken and dotted lines are simulations based on a CA monomer structure from solution NMR (red dots, PDB code 2LF4), a crystal structure of cross-linked CA hexamers (short green dashes, PDB code 3MGE), and a solution NMR structure of CTD dimers (long blue dashes, PDB code 2KOD). Simulations based on the refined structure are shown in gray continuous lines (“Fit”). Reprinted with permission from Bayro et al, J. Mol. Biol., 2014, 426 (5), pp 1109–1127. Copyright 2014 Elsevier.
3.1.2 HIV-1 Rev Protein Assemblies
The 13 kDa HIV-1 Rev protein binds to the Rev response element (RRE) on incompletely spliced viral mRNAs and enables their transport from the nucleus to the cytoplasm [181, 182]. As Rev is prone to aggregate and fibrillize, no high-resolution structural information on full-length Rev had been reported until Blanco et al. applied MAS NMR spectroscopy to the system [183]. In this work, the Tycko group examined Rev filaments in frozen solution with double-quantum chemical shift anisotropy (DQCSA) and constant-time double-quantum-filtered dipolar recoupling (CTDQFD) experiments, which provided site-specific information on backbone torsion angles, supporting the helix-loop-helix model in the N-terminal half of Rev. In a follow-up study, Havlin et al. reported new MAS NMR data on both free Rev and Rev in complex with a 45-base RNA molecule, using selectively 13C labeled samples at all Ala, Val, and Ile positions [82]. Analysis of two-dimensional 13C-13C correlation spectra suggested that along with the helical structure in the N-terminal half of Rev detected in the previous study, the C-terminal half of Rev also adopts a partially helical structure. Moreover, spectral comparisons between free Rev and Rev/RNA co-assemblies showed no significant differences, indicating that Rev protein remains in the same conformation upon RNA binding [82].
3.1.3 HIV-1 Membrane-Associated Proteins
HIV-1 membrane-associated proteins include glycoprotein gp120 and gp41, the trans-activator of transcription Tat, and viral protein Vpu. While we focus on MAS NMR studies of these proteins in this section, we refer our readers to references [29–41] for studies using oriented sample solid-state NMR.
The complex of HIV-1 gp120/gp41 facilitates the fusion of viral and host cell membranes in the first step of HIV infection [184, 185]. Derived from a ~20-residue N-terminal domain of gp41, HIV-1 fusion peptide (HFP) has the ability to bind to the host cell membrane and catalyze the fusion process. A series of MAS NMR studies from the Weliky group, including rotational-echo double resonance (REDOR) measurements and two-dimensional 13C-13C correlation experiments, revealed that HFP largely adopts an antiparallel β-sheet conformation in cholesterol-containing membrane mimics [80, 85, 186]. In addition, 31P relaxation measurements showed increased membrane curvature in the presence of HFP [187]. Qiang et al. synthesized several HFP constructs and examined their fusion rates. They discovered that the smallest oligomer for efficient fusion catalysis is HFPtr [78, 188]. In order to probe the membrane location of HFP, the authors measured the distance between the 13CO of HFP and 31P of the lipid using REDOR experiments, and revealed a positive correlation between the depth of membrane insertion of HFP and their fusogenicities [78, 189]. A later study enabled the measurement of site-specific membrane locations of HPF by examining the distance between specifically 13C labeled HFP and specifically 2H labeled lipids using 13C-2H REDOR experiments [190]. In order to probe the structure of the fusion peptide in large gp41 constructs, Sackett et al. built the N70 construct that models a pre-hairpin intermediate, and the FP-Hairpin construct that models the final fusion state with six-helix bundle structure [191, 192]. MAS NMR probed the secondary structure of the fusion peptide in these large gp41 constructs, and molecules with either α-helical or predominantly antiparallel β-sheet conformation were observed in both [79, 192]. Another study from the Weliky group introduced an approach to examine lyophilized whole cells containing an expressed 154-residue gp41 fragment using MAS NMR, and detected helical structure at the C-terminal of this protein construct by REDOR experiments [193]. In addition to the studies on fusion peptides, MAS NMR was also applied to investigate the V3 loop in gp120 and its interaction with antibody fragments [194, 195].
The HIV-1 Tat protein contains a transduction domain, which is known as the cell-penetrating peptide [196]. Su et al. applied MAS NMR to study the conformation and dynamics of a Tat peptide (residues 48–60) in anionic lipid bilayers [197]. Chemical shifts obtained from 13C-13C and 1H-13C correlation experiments suggested a random coil structure for the Tat peptide. The authors also conducted DIPSHIFT experiments to measure 13C-1H and 15N-1H dipolar couplings whose corresponding order parameters had low values, indicating large-amplitude motions of Tat peptide in membrane lipids. 31P spectra of oriented membrane lipids measured with increasing amount of Tat peptide revealed that the membrane remains intact upon the insertion of Tat peptide. 1H spin diffusion measurements suggested that Tat peptide is not deeply inserted into the membrane lipids. In addition, site-specific distances between Tat peptide and the membrane surface were measured by 13C-31P REDOR experiments, which provided evidence for a salt bridge interaction between the Arg residue of Tat peptide and the phosphate group in the membrane.
HIV-1 Vpu is an 81-residue accessory protein containing an N-terminal transmembrane (TM) domain [198]. Sharp et al. employed MAS NMR to examine the structure and dynamics of a Vpu peptide, comprising N-terminal residues 1–40, in phospholipid bilayers [86]. 15N and 13C chemical shifts were obtained based on two-dimensional 13C-13C and 15N-13C spectra acquired on selectively labeled samples. Spectral analysis suggested helical secondary structure and rigid-body dynamics for residues 3–27. In addition, site-specific information about solvent exposure was provided by H-D exchange experiments. Later, Do et al. characterized full-length Vpu in POPC bilayers using MAS NMR [199]. Partial resonance assignments for 13C were obtained by examining uniformly labeled protein, supporting the α-helical conformation in the TM domain. Furthermore, two- and three-dimensional proton spin diffusion experiments confirmed the correct insertion of TM domain of Vpu into the lipid bilayers and the close proximity between water and cytoplasmic domain of Vpu.
3.1.4 HIV-1 TAR RNA
The HIV-1 transactivation response (TAR) RNA is known to bind with the viral regulatory protein Tat via a TAR hairpin and facilitates viral replication [200]. In order to probe the interactions between TAR RNA and Tat peptides, the Drobny group utilized REDOR experiments to obtain distance information. A lyophilized sample of a 29-nucleotide TAR RNA construct, comprising the Tat-binding site, was examined using 31P-19F REDOR measurements, with and without the presence of a Tat-derived peptide. A nearly 4 Å decrease of distance between the 2′F label at U23 and the phosphorothioate at A27 revealed a significant conformational change upon peptide binding [201]. In a later study, TAR RNA with 5-fluorouridine incorporated at U23 in complex with a Tat-derived peptide with one crucial arginine labeled with 13C and 15N, was examined using 13C/15N-19F REDOR experiments, providing intermolecular distance information between RNA and the peptide [202]. In addition, Olsen et al. also performed site-specific deuterium solid-state MAS NMR experiments to probe the dynamics of TAR, and discovered motions on the μs-ns timescale, which could not be detected by solution NMR [203]. Motional models of residues on the TAR binding interface were established on the basis of deuterium MAS NMR data, which indicates a conformational capture mechanism for the recognition of Tat [204]. The authors also probed the dynamics of TAR RNA as a function of hydration by measuring deuterium lineshapes and relaxation times, and observed sudden dynamics changes when hydration levels were coincident with the initial bulk hydration [205]. Furthermore, the Drobny group presented a sample preparation protocol using polyethylene glycol precipitation, which improved the 13C line widths of TAR RNA spectra from > 6 ppm to ~1 ppm [60].
3.2 Bacteriophages
Bacteriophages are a class of viruses that target and destroy bacterial cells. Their general structure consists of a single- or double-stranded DNA or RNA genome surrounded by an assembly of coat protein subunits. This diverse class of viruses exists in a variety of shapes and sizes, but can be divided into 3 general structural groups: filamentous, icosahedral with tail, and icosahedral without tail. Filamentous bacteriophages, the class that includes Pf1, fd, and M13, possess a rod-like structure, 800–2000 nm in length, with several thousand copies of an α-helical coat protein in a helical array around the ssDNA [206]. Caudovirales, such as T4 and T7, are tailed bacteriophages with DNA stored inside an icosahedral capsid, with a tail to attach to the host cell [207]. They range from approximately 100 to 500 nm in length.
There is a broad range of scientific interest in bacteriophages [208], including applications in nanotechnology [209–212], phage display and targeted gene therapy [213–215], as well as DNA cloning and sequencing [216–218]. Bacteriophages are also widely used as alignment media for residual dipolar coupling measurements in solution-state NMR [219, 220]. Structural analysis of bacteriophages can lend valuable insight into applications of bacteriophages in these fields. MAS NMR studies of bacteriophages have provided insight into the structure and dynamics of bacteriophage coat proteins [19, 68, 221], as well as nucleic acid conformation [67], and interactions between coat protein sites and nucleic acids [64–66]. These studies have also contributed to a growing knowledge base of the relationship between DNA chemical shifts and conformation.
3.2.1 Filamentous Bacteriophages
Filamentous bacteriophages that have been characterized by MAS NMR include Pf1, fd, and M13. Pf1 has been studied through a wide range of other spectroscopic and computational methods and has become a model system for the study of filamentous bacteriophages. McDermott and co-workers presented site-specific resonance assignments of the coat protein of intact, PEG-precipitated Pf1 bacteriophage with 2- and 3-dimensional 13C and 15N correlation spectra, as illustrated in Fig. 6 [19]. This was the first study to assign the coat protein of an intact virus with MAS NMR. They were able to assign 92% of the 46-residue protein in its high temperature form, despite significant helical content and resulting spectral overlap. A single observed peak for most sites of the capsid protein indicated that the asymmetric unit was a single copy of the coat protein, rather than a trimer of slightly differing conformers as proposed in a prior study [222]. Based upon the well-known correlation between protein secondary structure and 13C chemical shifts [223], they determined that the Pf1 coat protein is unstructured at the N-terminus, and a continuous α helix from residues 6–46, in agreement with structural models derived from static SSNMR and X-ray fiber diffraction [27, 222]. Further studies by Lorieau et al. probed the site-specific dynamics of the intact Pf1 coat protein with 1H-13C dipolar order parameters [221]. They demonstrated that while the protein backbone is very rigid, there are several amino acid sidechains that exhibit conformational mobility. These include solvent-exposed residues, and notably, Arg 44 and Lys 45, which interact with the viral DNA inside the capsid. The observed sub-microsecond dynamics of these 2 residues may be an important feature of capsid-DNA interactions in Pf1.
Fig. 6.
3D heteronuclear correlation spectra of Pf1 acquired at 750 MHz. (A) Strip plots for residues 8–15 at the indicated 15N frequency with 13CO chemical shifts on the horizontal axis, and 13Cα shifts on the vertical axis, from NCACX and NCOCX experiments, illustrating the sequential assignment. (B) The projection down the CO dimension of the NCOCX sequential experiment, showing regions corresponding to crosspeaks between the Gly/Ser 15N shifts and the CO-Cα resonance of the preceding residue. (C) Projection down the 15N dimension of the NCACX experiment, showing crosspeaks from many 2- and 3-bond transfers. The gray box indicates aliased peaks arising from N-Cβ and N-Cγ transfers. Reprinted with permission from Goldbourt et al, J. Am. Chem. Soc., 2007, 129 (8), pp 2338–2344. Copyright 2007 American Chemical Society.
Pf1 undergoes a 2-state structural transition near 10°C. Goldbourt et al. probed the effects of temperature on Pf1 coat protein structure with 13C-13C DARR spectra at high and low temperature [68]. Several reversible chemical shift changes as a function of sample temperature were observed, notably among sidechain sites of hydrophobic amino acids. A total of 14 amino acids were observed to have chemical shift changes. Many of the resonances with significant chemical shift perturbations were located near the C-terminus of the coat protein. At longer mixing times (200 ms), some intersubunit contacts could be observed. Their results suggest that the observed chemical shift changes occur due to changes in intersubunit hydrophobic interactions rather than changes in backbone structure or solvent exposure, in agreement with previous studies by Marvin and co-workers [224].
Sergeyev et al. applied DNP to the study of DNA conformation in Pf1 [67]. DNP signal enhancement of the otherwise weak DNA resonances was key for resonance identification and assignment (Fig. 7). It is to be noted that under DNP conditions, sigificant line broadening was observed, as evident in the 2D contour plots shown in Fig. 7. Nevertheless, Sergeyev et al. were able to obtain nucleotide-type assignment of the Pf1 DNA base region with 13C-13C DARR spectra (Fig. 7). Ribose 13C chemical shifts have been shown to be sensitive reporters of nucleic acid conformation [225, 226]. The downfield-shifted SC3′ and SC5′ resonances indicated that the DNA is in a C2′-endo/gauche conformation. Many observed 13C and 15N chemical shifts in the DNA base region lie outside the range of shifts previously reported in the BioMagResBank (BMRB) [100]. This chemical shift behavior indicates an unusual DNA conformation where Watson-Crick base pairing is absent, but base-stacking is still observed, in agreement with studies concluding that Pf1 has an unusually extended and twisted DNA conformation [227].
Fig. 7.
DNP-enhanced 13C-13C DARR spectra of Pf1, indicating the nucleotide-type assignments of the DNA. (A) dC/dT base resonances, (B) dA/dG base resonances, (C) sugar spin systems. Reprinted with permission from Sergeyev et al, J. Am. Chem. Soc., 2011, 133 (50), pp 20208–20217. Copyright 2011 American Chemical Society.
Purusottam et al. characterized water-protein interactions of Pf1 using 2D 1H-15N dipolar-edited medium- and long-distance (MELODI) HETCOR experiments [228]. 13C and 15N dipolar dephasing was utilized to reduce polarization arising from aliphatic and amino protons, such that cross peaks at 4.78 ppm in the proton dimension predominantly arise from water-protein contacts. This dephasing made it possible to work with 100% protonated samples. 1H-15N cross peaks were observed between hydration waters and 7 protein residues, including the dynamic Arg 44. They propose that the capsid-DNA interaction at this site is water-mediated, and hydration waters may play a role in sequence-specific recognition of DNA.
The McDermott group has recently published studies of Pf1 hydration as well [229]. Using 1H-13C HETCOR measurements, water-protein cross peaks of 25 amino acids were observed, including 6 residues in the virion core. Water-DNA contacts were also observed. These results demonstrate that the capsid interior is water-accessible and support the hypothesis that hydration waters have an important role in mediating protein-DNA interactions.
The Goldbourt lab has examined structural features of both the capsid protein and ssDNA of the fd bacteriophage, as well as interactions between the two [64–66]. Fd and other Ff class bacteriophages including M13 have important applications in phage display and nanotechnology [230]. Previously, structural inhomogeneity had limited high-resolution characterization of the wild-type fd capsid [231, 232]. Abramov et al. obtained high-resolution spectra of intact phage particles composed of the wild type capsid, allowing for almost complete site-specific assignment of the capsid protein, as well as qualitative assignment of DNA signals [65]. Characteristic chemical shifts indicated the capsid is predominantly α-helical with a mobile N-terminus and slight curvature. Interestingly, they observed no evidence of structural inhomogeneity. Downfield-shifted C3′ and C5′ 13C peaks of DNA suggest the nucleic acid is in a C2′-endo or C3′-exo sugar pucker conformation, in contrast to previous studies [233, 234]. Subsequent work probed interactions between the protein capsid and nucleic acid [64]. Utilizing 13C-13C and PDSD-mediated 31P-13C correlations, they demonstrated that the primary protein-DNA interaction in fd is an electrostatic interaction between lysine sidechains and the DNA phosphate backbone. A key feature of this study was the use of natural abundance aromatic amino acids to avoid spectral overlap with DNA 13C resonances. They demonstrated that even at moderate spinning frequencies (12 kHz), efficient polarization transfer within DNA nucleotides can be obtained with 13C-13C COmbined R2nν-Driven (CORD) experiments [106], particularly important given the scarcity of protons within the base region of nucleic acids. Using the unique chemical shifts of some nucleic acid sites (i.e. deoxythymidine C7), and nucleotide chemical shifts reported in the BMRB and B-DNA chemical shifts compiled by Sergeyev et al. [67] for reference, almost complete nucleotide-type assignment of DNA base resonances was possible. Internucleotide correlations indicative of base pairing or stacking were also observed. Capsid-DNA interactions were observed for the sugar region and base region, as well as the phosphate backbone of the DNA, based upon cross peaks in 13C-13C and 13C-31P correlation spectra. 13C-13C correlations of note included DNA sugar and base to lysine sidechain cross peaks near the C terminus. Proton-mediated 31P-13C correlations (Fig. 8) also indicate interactions of the phosphate backbone with Lys residues as well as other amino acid residues near the C terminus, elucidating the capsid-DNA interface.
Fig. 8.

2D PHHC spectrum of YFWunlabeled-[13C,15N] fd bacteriophage. The spectrum shows crosspeaks between the phosphorous backbone linkage of the DNA with both DNA sugar and base sites, as well as protein capsid residues. Reprinted with permission from Morag et al, J. Am. Chem. Soc., 2014, 136 (6), pp 2292–2301. Copyright 2014 American Chemical Society.
In further studies, Morag et al. characterized the structure of the coat protein in infectious M13 particles [66], which differs from fd by only 1 amino acid residue (fd’s negatively charged Asp12 is Asn in M13). In addition to the traditional experiments for resonance assignment, J-based z-filter refocused INADEQUATE was applied to M13, allowing assignment of the mobile N-terminus and confirmation of additional assignments. The general chemical shift agreement between fd and M13 indicates overall structural similarity, as well as similar hydrophobic packing and mobility. Amino acids for which chemical shift differences were observed are in the region of residue Asx12, including Lys8 which is involved in intra- and inter-subunit electrostatic interactions, leading to changes in local conformation, packing, and rigidity between fd and M13 in this region. Very recently the Goldbourt lab published a structural model of M13 using structural restraints derived from MAS NMR data and Rosetta for model building [235]. The structure demonstrates the important role of hydrophobic residues for capsid stabilization. This work reports the first MAS NMR structure of an intact viral capsid.
Opella and co-workers have extensively characterized Pf1 and fd with static SSNMR, including structure determination of the Pf1 and fd coat proteins as both an intact viral particle and in a membrane-bound form [25–28, 236]. They have utilized methods such as PISEMA to characterize magnetically aligned samples [237, 238]. MAS and static NMR studies of the M13 coat protein embedded in a phospholipid bilayer have also been performed [239, 240]. This work is beyond the scope of the current review.
3.2.2 Tailed Bacteriophages
Tailed bacteriophages (Caudovirales) such as T4 and T7 have a highly condensed genome within the viral capsid [241]. The mechanism of genome packaging and structure of the DNA within the tightly packaged capsid are of particular interest. Abramov et al. characterized the structure of double-stranded DNA in bacteriophage T7, including development of a large-scale sample preparation protocol [63]. 13C-13C dipolar-assisted rotational resonance (DARR) and 15N-13C transferred echo double resonance (TEDOR) experiments were used for complete 13C and near-complete 15N nucleotide-type assignment of the 40 kbp DNA. As compared to other phages, DNA resonances in T7 are more readily detected given that the nucleic acid comprises > 50% of the virion’s mass. 13C and 15N chemical shift reporters in both the sugar and base regions indicated B-form DNA and the presence of expected Watson-Crick base pairing. This work is an additional contribution to the somewhat limited database of 13C and 15N chemical shifts for nucleic acids, particularly for native systems.
Schaefer and co-workers characterized DNA packaging in bacteriophage T4 with rotational echo double resonance (REDOR) dephasing experiments [83]. 31P-15N REDOR dephasing of thymine and guanosine N1 coupled to all neighboring backbone phosphorous atoms indicated that DNA packaged inside the phage is B form. The observation of REDOR dephasing for ε-lysyl and polyamines also suggested a role for these protein residues in DNA backbone charge stabilization.
3.3 Lipid Membrane Enveloped Viruses
Enveloped viruses contain at the very least the viral genome, a capsid shell composed of proteins, and a lipid membrane [8]. Additional proteins and molecules can also be present. For instance, glycoproteins can decorate the lipid membrane helping the virus attach to a host cell. The lipid membranes that make up the viral envelope are generally derived from lipids present within the host cell.
Many of the viruses that pose major human health concerns are enveloped. These include: herpes simplex, hepatitis B & C, rubella, influenza A, B, & C, measles, mumps, rabies, and Ebola. Several groups have invested time into understanding the structure, dynamics, and function of enveloped viruses. MAS NMR studies of influenza A, parainfluenza 5, and measles viruses, which have been studied by SSNMR, are described below.
3.3.1 Influenza A
One of the most extensively studied virus related systems by solid-state NMR is the transmembrane peptide of the M2 protein of Influenza A (M2TMP). The M2 protein of influenza acts as an ion channel and is responsible for the conduction of protons to the interior of the virion. This lowering of pH promotes uncoating and is critical for the viability of the virus. Antiviral drugs based on adamantane derivatives such as amantadine and rimantadine, which were once effective at binding M2 and inhibiting viral replication, have since found a reduction in clinical use due to many influenza A strains displaying drug resistance.
Beginning in the late 1990s, the laboratory of Tim Cross was the first to begin intense investigation of the structural features of M2TMP using oriented sample SSNMR. These groundbreaking studies, conducted on synthesized peptides containing selective amino acid labels and reconstituted in lipids, served as the foundation for many of the studies that have followed. The original experiments resulted in the first SSNMR evidence for symmetry, orientation, and helical packing in the M2TMP tetrameric bundle [44, 242, 243]. In subsequent studies, Cross and co-workers developed the first high-resolution structure of the backbone for the M2TMP monomer as determined using orientational restraints from PISEMA spectra [51]. They were also the first to report on an interhelical distances in M2TMP. Using a peptide labeled with 13C, 15N Trp and His amino acids, REDOR measurements were conducted under MAS NMR spectroscopy which yielded a maximum distance of 3.9 Angstroms between the 15Nπ site of His37 and the 13Cγ site of Trp41 [45]. In 2003, they published structural and dynamic results characterizing the full-length M2 protein using PISEMA and hydrogen/deuterium exchange experiments [48]. Recently, Cross and coworkers have continued to contribute to the understanding of the influenza A M2 protein with studies including investigations on the role of histidine side chains in proton conduction by monitoring His37 sites using 15N CPMAS experiments [244], the backbone structure of the amantadine blocked proton channel [245], dynamics of residues in the proton channel [246], the role of W41 in proton conduction [52], structural investigations of full length M2 in synthetic bilayers and E. coli membranes [247], and the quaternary structure of the transmembrane bundle [248]. To explore the quaternary structure, 2D 13C-13C as well as 3D NCOCX and NCACX MAS correlation experiments were conducted on a M2 construct expressed in E. coli. Using these correlation data, isotropic chemical shifts were assigned to residues in regions of the spectra with high resolution. Additional 2D 13C-13C correlation experiments were collected with different mixing times, as shown in Fig. 9, to obtain distance restraints for the quaternary structure [248].
Fig. 9.
13C-13C DARR spectra for the M2 protein (spanning residues 22–62) from influenza A as expressed in E. coli. (A) Aliphatic to carbonyl correlations are shown for experiments with a short mixing time 9 ms (blue) and 50 ms (red). (B) Correlations between aliphatic and aromatic sidechains are shown with short mixing times of 9 ms (blue) and longer mixing times of 100 ms (red). (C) Aliphatic to aliphatic correlations are shown for shorter mixing times of 20 ms (blue) and longer mixing times of 50 ms (red). Reprinted with permission from Can et al., J. Am. Chem. Soc., 2012, 134, pp 9022–9025. Copyright 2012 American Chemical Society.
In the past decade, Hong and co-workers have investigated a variety of structural, functional, and dynamical features for the M2TMP with MAS NMR. An early study examined the quaternary structure of the M2TMP bound to DMPC bilayers [89]. Peptides were synthesized with an A30F mutation allowing for a 4-19F label to be incorporated. 19F-19F interhelical distances between nearest neighbor Phe30 residues were measured to be between 7.9 and 9.5 Å using a CODEX experiment. Follow up studies examined uniaxial rotational diffusion [76, 249], intermolecular distances, and side-chain conformations in M2TMP [249–251]. In 2008, the effects of amantadine binding on the conformation and dynamics of the transmembrane portion of the M2 channel of Influenza A were explored. The M2 peptides used in this study contained 13C, 15N labels at several of the amino acid sites that are intimately involved in the binding interaction with the drug, including residues lining the proton conduction channel, involved in helix-helix interactions, and those facing the lipid. 13C CPMAS, 13C-13C, and 15N-13C correlation experiments were used to monitor chemical shift changes upon amantadine binding. 13C T2 relaxation measurements were performed to monitor microsecond dynamics where T2 was found to increase in amantadine bound samples [252]. Additional studies related to the structural and dynamic changes occurring upon amantadine binding [77, 253–255] and the derivation of the amantadine binding site [84] were conducted thereafter. One such study monitored the chemical shifts of several amino acids including Ser31. From 2D 15N-13C correlation experiments, large chemical shift deviations were observed for this residue in in the presence of amantadine. To ensure these shifts were caused by direct protein-drug interactions and not as a result of changes to the lipid environment, the authors prepared samples where amantadine was added to the buffers during sample preparation as well as added directly to the membrane pellet after reconstitution. The results of these experiments are summarized in Fig. 10 [253]. In the past few years, proton sensing, gating, and conduction [256, 257], membrane curvature [88], and the conformational analysis of full-length M2 protein [258] have been explored by the Hong laboratory.
Fig. 10.
2D 15N-13C correlation spectra for the selectively labeled (V28, S31, L36) M2 transmembrane peptide (spanning residues 22–46) reconstituted in DLPC lipids. The three spectra display chemical shift changes based on the amount of amantadine present (none, 10 mM, and 0.4 mg from top to bottom, respectively). (A) No amantadine is present and peak S31 is located at a chemical shift representing the unbound amino acid. (B) As amantadine is added to the buffer at a concentration of 10 mM nearly all of the intensity for S31 is displayed at a chemical shift representing the bound state. (C) This condition is based on the addition of amantadine directly to the reconstituted pellet at a ratio of 1:2 (pellet to drug) and shows 70% of the signal intensity in the bound form. Reprinted with permission from Cady et al., J. Mol. Biol., 2009, 385, pp 1127–1141. Copyright 2008 Elsevier.
Andreas, Griffin, and colleagues investigated an elongated construct of the M2 proton channel spanning residues 18–60 reconstituted in POPC and DPhPC lipids. 15N-13C z-filtered TEDOR and 13C-13C PDSD experiments were conducted in the presence and absence of rimantadine. Large chemical shift changes observed upon binding of rimantadine were indicative of an allosteric mechanism of inhibition [259]. The authors also note that the choice of lipid had very little effect on stability and structure in the case of the elongated construct. Follow up studies included studies of a S31N drug resistant mutant [260] and investigations of inhibitor binding using dynamic nuclear polarization [162]. Taken together, these studies have provided a wealth of information about the transmembrane peptide.
3.3.2 Parainfluenza 5 (PIV5)
Another interesting system that has recently been investigated with MAS NMR is parainfluenza virus 5 (PIV5) from the Paramyxoviridae family. This enveloped virus enters a host cell through a fusion between the viral membrane and the host cell membrane. Hong and co-workers have investigated structural and dynamics aspects of the fusion peptide from the F protein of PIV5. [261, 262].
Using selectively labeled amino acids, peptides corresponding to the F protein of PIV5 were synthesized and reconstituted in a variety of different lipids. A combination of 1D and 2D 13C experiments were used to monitor the conformations of the peptide in different membranes. Dynamics were monitored based on temperature dependent 13C CPMAS measurements and indicated mobility between 273 and 313 K. Peptide depth and hydration levels were monitored through 13C-1H spin diffusion and 1H-31P correlation experiments, respectively [261]. A follow up study provided additional insights into the peptide conformation through the use of 13C CPMAS, 2D 15N-13C, and 2D 13C-13C correlation experiments of peptides in several different lipids. 31P experiments provided evidence for membrane curvature and dehydration in DOPC/DOPG membranes. The data indicate that several different peptide conformations are present depending on the type of lipid that is used for reconstitution [262].
3.3.3 Measles Virus (MeV)
In a very recent report by Pintacuda and coworkers [101], the measles virus (MeV) from the Paramyxoviridae family was examined by MAS NMR. The MeV nucleocapsid is composed of two domains dubbed NCORE and NTAIL and no atomic level structural information was present for the assembled nucleocapsids previously. The authors addressed both fully intact and cleaved MeV nucleocapsids using proton detection to monitor structural and dynamic features of the system. 1H detected experiments were performed at 60 kHz throughout.
Dipolar based 15N-1H correlation experiments were collected for sedimented intact MeV particles and displayed a significant number of signals from the NCORE domain. Scalar based 15N-1H correlation experiments were then collected on the sedimented intact MeV nucleocapsids as well as for MeV nucleocapsids resuspended in solution. From these data, signals from the NTAIL domain were observed. Signals from the sedimented nucleocapsids and those that were resuspeneded in solution were found to be similar.
When comparing 15N-1H dipolar based correlation experiments for sedimented nucleocapsids that were intact to those that were cleaved by trypsin digestion (thus removing the NTAIL domain), it was observed that most of the signals were present in both cases without large chemical shift deviations, suggesting similar structures between the two constructs. Although similar structures were observed, T1 measurements revealed that the local dynamics were significantly different. Additional experiments were conducted which included a spin diffusion step to transfer magnetization from water protons to the protons of the protein. This experiment was used to monitor solvent levels in the intact and cleaved nucleocapsids. The data indicated that the intact nucleocapsid was surrounded by an increased amount of interstitial water providing evidence for a less ordered structure than that observed for the cleaved MeV nucleocapsids.
4. Conclusions and Future Outlook
The work outlined in this review lays a foundation for the expansion of MAS NMR studies to a wide range of viral systems that cannot be investigated with atomic-level resolution under physiological conditions using other biophysical techniques. As demonstrated in these recent developments, contemporary MAS NMR has the ability to provide exquisite information on the structure and dynamics of such systems. With intense methodological developments in magnet technologies, probe hardware, and pulse sequences we anticipate that in the next several years we will witness MAS NMR evolving as a method for analysis of large complex multicomponent assemblies of macromolecules constituting various viruses and even intact virions, at atomic level detail.
Highlights.
MAS NMR can be used to study complex viral assemblies and intact virus particles
MAS NMR provides structural information with atomic level resolution
Recent advances in methodology yield spectra with high sensitivity and resolution
MAS NMR has probed the structure and dynamics of viral systems such as HIV and Pf1
Acknowledgments
This work was supported by the National Institutes of Health (NIGMS Grant P50GM082251) and is a contribution from the Pittsburgh Center for HIV Protein Interactions. We acknowledge the support of the National Science Foundation (NSF Grant CHE0959496) and of the University of Delaware for the acquisition of the 850 MHz NMR spectrometer at the University of Delaware. We acknowledge the support of the core instrumentation facilities by the NIH-COBRE programs (5P30GM103519 and 1P30GM110758).
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Contributor Information
Caitlin Quinn, Email: cmquinn@udel.edu.
Manman Lu, Email: lumm@udel.edu.
Christopher L. Suiter, Email: csuiter@udel.edu.
Guangjin Hou, Email: hou@udel.edu.
Huilan Zhang, Email: zhang@udel.edu.
Tatyana Polenova, Email: tpolenov@udel.edu.
References
- 1.Brussaard CPD, Noordeloos AAM, Sandaa RA, Heldal M, Bratbak G. Discovery of a dsRNA virus infecting the marine photosynthetic protist Micromonas pusilla. Virology. 2004;319:280–291. doi: 10.1016/j.virol.2003.10.033. [DOI] [PubMed] [Google Scholar]
- 2.Nelson RS, Citovsky V. Plant viruses. Invaders of cells and pirates of cellular pathways. Plant Physiol. 2005;138:1809–1814. doi: 10.1104/pp.104.900167. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Pearson MN, Beever RE, Boine B, Arthur K. Mycoviruses of filamentous fungi and their relevance to plant pathology. Mol Plant Pathol. 2009;10:115–128. doi: 10.1111/j.1364-3703.2008.00503.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Prangishvili D. The wonderful world of archaeal viruses. Annu Rev Microbiol. 2013;67:565–585. doi: 10.1146/annurev-micro-092412-155633. [DOI] [PubMed] [Google Scholar]
- 5.Smith AE, Helenius A. How viruses enter animal cells. Science. 2004;304:237–242. doi: 10.1126/science.1094823. [DOI] [PubMed] [Google Scholar]
- 6.Kaminskyy V, Zhivotovsky B. To kill or be killed: how viruses interact with the cell death machinery. Journal of Internal Medicine. 2010;267:473–482. doi: 10.1111/j.1365-2796.2010.02222.x. [DOI] [PubMed] [Google Scholar]
- 7.Harper D. Viruses Biology, Applications, and Control. Garland Science; 2011. [Google Scholar]
- 8.Mateu MG. Structure and Physics of Viruses. Springer; 2013. [Google Scholar]
- 9.Caston JR. Conventional electron microscopy, cryo-electron microscopy and cryo-electron tomography of viruses. In: Mateu MG, editor. Structure and Physics of Viruses. Springer; 2013. pp. 79–115. [DOI] [PubMed] [Google Scholar]
- 10.Chang J, Liu XA, Rochat RH, Baker ML, Chiu W. Reconstructing virus structures from nanometer to near-atomic resolutions with cryo-electron microscopy and tomography. Adv Exp Med Biol. 2012;726:49–90. doi: 10.1007/978-1-4614-0980-9_4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Subramaniam S, Bartesaghi A, Liu J, Bennett AE, Sougrat R. Electron tomography of viruses. Curr Opin Struct Biol. 2007;17:596–602. doi: 10.1016/j.sbi.2007.09.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Reddy VS, Natchiar SK, Stewart PL, Nemerow GR. Crystal structure of human adenovirus at 3.5 angstrom resolution. Science. 2010;329:1071–1075. doi: 10.1126/science.1187292. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Byeon IJL, Meng X, Jung JW, Zhao GP, Yang RF, Ahn JW, Shi J, Concel J, Aiken C, Zhang PJ, Gronenborn AM. Structural convergence between cryo-EM and NMR reveals intersubunit Interactions critical for HIV-1 capsid function. Cell. 2009;139:780–790. doi: 10.1016/j.cell.2009.10.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Kuznetsov YG, Xiao CA, Sun SY, Raoult D, Rossmann M, McPherson A. Atomic force microscopy investigation of the giant mimivirus. Virology. 2010;404:127–137. doi: 10.1016/j.virol.2010.05.007. [DOI] [PubMed] [Google Scholar]
- 15.Tuma R, Coward LU, Kirk MC, Barnes S, Prevelige PE. Hydrogen-deuterium exchange as a probe of folding and assembly in viral capsids. J Mol Biol. 2001;306:389–396. doi: 10.1006/jmbi.2000.4383. [DOI] [PubMed] [Google Scholar]
- 16.Knejzlik Z, Ulbrich P, Strohalm M, Lastuvkova H, Kodicek M, Sakalian M, Ruml T. Conformational changes of the N-terminal part of Mason-Pfizer monkey virus p12 protein during multimerization. Virology. 2009;393:168–176. doi: 10.1016/j.virol.2009.07.014. [DOI] [PubMed] [Google Scholar]
- 17.Rickgauer JP, Fuller DN, Grimes S, Jardine PJ, Anderson DL, Smith DE. Portal motor velocity and internal force resisting viral DNA packaging in bacteriophage phi 29. Biophys J. 2008;94:159–167. doi: 10.1529/biophysj.107.104612. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Zink M, Grubmuller H. Mechanical properties of the icosahedral shell of southern bean mosaic virus: A molecular dynamics study. Biophys J. 2009;96:1350–1363. doi: 10.1016/j.bpj.2008.11.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Goldbourt A, Gross BJ, Day LA, McDermott AE. Filamentous phage studied by magic-angle spinning NMR: Resonance assignment and secondary structure of the coat protein in Pf1. J Am Chem Soc. 2007;129:2338–2344. doi: 10.1021/ja066928u. [DOI] [PubMed] [Google Scholar]
- 20.Han Y, Ahn J, Concel J, Byeon IJL, Gronenborn AM, Yang J, Polenova T. Solid-state NMR studies of HIV-1 capsid protein assemblies. J Am Chem Soc. 2010;132:1976–1987. doi: 10.1021/ja908687k. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Han Y, Hou GJ, Suiter CL, Ahn J, Byeon IJL, Lipton AS, Burton S, Hung I, Gor’kov PL, Gan ZH, Brey W, Rice D, Gronenborn AM, Polenova T. Magic angle spinning NMR reveals sequence-dependent structural plasticity, dynamics, and the spacer peptide 1 conformation in HIV-1 capsid protein assemblies. J Am Chem Soc. 2013;135:17793–17803. doi: 10.1021/ja406907h. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Li XM, Mooney P, Zheng S, Booth CR, Braunfeld MB, Gubbens S, Agard DA, Cheng YF. Electron counting and beam-induced motion correction enable near-atomic-resolution single-particle cryo-EM. Nat Methods. 2013;10:584–590. doi: 10.1038/nmeth.2472. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Andrew ER, Bradbury A, Eades RG. Removal of dipolar broadening of nuclear magnetic resonance spectra of solids by specimen rotation. Nature. 1959;183:1802–1803. [Google Scholar]
- 24.Lowe IJ. Free induction decays of rotating solids. Phys Rev Lett. 1959;2:285–287. [Google Scholar]
- 25.Park SH, Marassi FM, Black D, Opella SJ. Structure and dynamics of the membrane-bound form of Pf1 coat protein: Implications of structural rearrangement for virus assembly. Biophys J. 2010;99:1465–1474. doi: 10.1016/j.bpj.2010.06.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Shon KJ, Kim YG, Colnago LA, Opella SJ. NMR studies of the structure and dynamics of membrane-bound bacteriophage-Pf1 coat protein. Science. 1991;252:1303–1304. doi: 10.1126/science.1925542. [DOI] [PubMed] [Google Scholar]
- 27.Thiriot DS, Nevzorov AA, Zagyanskiy L, Wu CH, Opella SJ. Structure of the coat protein in Pf1 bacteriophage determined by solid-state NMR Spectroscopy. J Mol Biol. 2004;341:869–879. doi: 10.1016/j.jmb.2004.06.038. [DOI] [PubMed] [Google Scholar]
- 28.Zeri AC, Mesleh MF, Nevzorov AA, Opella SJ. Structure of the coat protein in fd filamentous bacteriophage particles determined by solid-state NMR spectroscopy. Proc Natl Acad Sci USA. 2003;100:6458–6463. doi: 10.1073/pnas.1132059100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Skasko M, Wang Y, Tian Y, Tokarev A, Munguia J, Ruiz A, Stephens EB, Opella SJ, Guatelli J. HIV-1 Vpu protein antagonizes innate restriction factor BST-2 via lipid-embedded helix-helix interactions. J Biol Chem. 2012;287:58–67. doi: 10.1074/jbc.M111.296772. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Park SH, Opella SJ. Conformational changes induced by a single amino acid substitution in the trans-membrane domain of Vpu: Implications for HIV-1 susceptibility to channel blocking drugs. Protein Sci. 2007;16:2205–2215. doi: 10.1110/ps.073041107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Park SH, Mrse AA, Nevzorov AA, Mesleh MF, Oblatt-Montal M, Montal M, Opella SJ. Three-dimensional structure of the channel-forming trans-membrane domain of virus protein “u” (Vpu) from HIV-1. J Mol Biol. 2003;333:409–424. doi: 10.1016/j.jmb.2003.08.048. [DOI] [PubMed] [Google Scholar]
- 32.Park SH, De Angelis AA, Nevzorov AA, Wu CH, Opella SJ. Three-dimensional structure of the transmembrane domain of Vpu from HIV-1 in aligned phospholipid bicelles. Biophys J. 2006;91:3032–3042. doi: 10.1529/biophysj.106.087106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Marassi FM, Ma C, Gratkowski H, Straus SK, Strebel K, Oblatt-Montal M, Montal M, Opella SJ. Correlation of the structural and functional domains in the membrane protein Vpu from HIV-1. Proc Natl Acad Sci USA. 1999;96:14336–14341. doi: 10.1073/pnas.96.25.14336. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Ma C, Marassi FM, Jones DH, Straus SK, Bour S, Strebel K, Schubert U, Oblatt-Montal M, Montal M, Opella SJ. Expression, purification, and activities of full-length and truncated versions of the integral membrane protein Vpu from HIV-1. Protein Sci. 2002;11:546–557. doi: 10.1110/ps.37302. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Kochendoerfer GG, Jones DH, Lee S, Oblatt-Montal M, Opella SJ, Montal M. Functional characterization and NMR Spectroscopy on full-length Vpu from HIV-1 prepared by total chemical synthesis. J Am Chem Soc. 2004;126:2439–2446. doi: 10.1021/ja038985i. [DOI] [PubMed] [Google Scholar]
- 36.Henklein P, Kinder R, Schubert U, Bechinger B. Membrane interactions and alignment of structures within the HIV-1 Vpu cytoplasmic domain: effect of phosphorylation of serines 52 and 56. FEBS Lett. 2000;482:220–224. doi: 10.1016/s0014-5793(00)02060-3. [DOI] [PubMed] [Google Scholar]
- 37.Grasnick D, Sternberg U, Strandberg E, Wadhwani P, Ulrich AS. Irregular structure of the HIV fusion peptide in membranes demonstrated by solid-state NMR and MD simulations. European Biophysics Journal with Biophysics Letters. 2011;40:529–543. doi: 10.1007/s00249-011-0676-5. [DOI] [PubMed] [Google Scholar]
- 38.Cook GA, Zhang H, Park SH, Wang Y, Opella SJ. Comparative NMR studies demonstrate profound differences between two viroporins: p7 of HCV and Vpu of HIV-1. Biochim Biophys Acta. 2011;1808:554–560. doi: 10.1016/j.bbamem.2010.08.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Bechinger B. Solid-state NMR investigations of interaction contributions that determine the alignment of helical polypeptides in biological membranes. FEBS Lett. 2001;504:161–165. doi: 10.1016/s0014-5793(01)02741-7. [DOI] [PubMed] [Google Scholar]
- 40.Aisenbrey C, Bertani P, Henklein P, Bechinger B. Structure, dynamics and topology of membrane polypeptides by oriented H-2 solid-state NMR spectroscopy. Eur Biophys J Biophy. 2007;36:451–460. doi: 10.1007/s00249-006-0122-2. [DOI] [PubMed] [Google Scholar]
- 41.Afonin S, Frey A, Bayerl S, Fischer D, Wadhwani P, Weinkauf S, Ulrich AS. The cell-penetrating peptide TAT(48–60) induces a non-lamellar phase in DMPC membranes. ChemPhysChem. 2006;7:2134–2142. doi: 10.1002/cphc.200600306. [DOI] [PubMed] [Google Scholar]
- 42.Hu J, Qin H, Li C, Sharma M, Cross TA, Gao FP. Structural biology of transmembrane domains: Efficient production and characterization of transmembrane peptides by NMR. Protein Sci. 2007;16:2153–2165. doi: 10.1110/ps.072996707. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Kovacs FA, Cross TA. Transmembrane four-helix bundle of influenza A M2 protein channel: Structural implications from helix tilt and orientation. Biophys J. 1997;73:2511–2517. doi: 10.1016/S0006-3495(97)78279-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Kovacs FA, Denny JK, Song Z, Quine JR, Cross TA. Helix tilt of the M2 transmembrane peptide from influenza A virus: An intrinsic property. J Mol Biol. 2000;295:117–125. doi: 10.1006/jmbi.1999.3322. [DOI] [PubMed] [Google Scholar]
- 45.Nishimura K, Kim SG, Zhang L, Cross TA. The closed state of a H+ channel helical bundle combining precise orientational and distance restraints from solid state NMR. Biochemistry. 2002;41:13170–13177. doi: 10.1021/bi0262799. [DOI] [PubMed] [Google Scholar]
- 46.Page RC, Kim S, Cross TA. Transmembrane helix uniformity examined by spectral mapping of torsion angles. Structure. 2008;16:787–797. doi: 10.1016/j.str.2008.02.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Sharma M, Yi MG, Dong H, Qin HJ, Peterson E, Busath DD, Zhou HX, Cross TA. Insight into the mechanism of the influenza A proton channel from a structure in a lipid bilayer. Science. 2010;330:509–512. doi: 10.1126/science.1191750. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Tian CL, Gao PF, Pinto LH, Lamb RA, Cross TA. Initial structural and dynamic characterization of the M2 protein transmembrane and amphipathic helices in lipid bilayers. Protein Sci. 2003;12:2597–2605. doi: 10.1110/ps.03168503. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Tian CL, Tobler K, Lamb RA, Pinto LH, Cross TA. Expression and initial structural insights from solid-state NMR of the M2 proton channel from influenza a virus. Biochemistry. 2002;41:11294–11300. doi: 10.1021/bi025695q. [DOI] [PubMed] [Google Scholar]
- 50.Wang J, Denny J, Tian C, Kim S, Mo Y, Kovacs F, Song Z, Nishimura K, Gan Z, Fu R, Quine JR, Cross TA. Imaging membrane protein helical wheels. J Magn Reson. 2000;144:162–167. doi: 10.1006/jmre.2000.2037. [DOI] [PubMed] [Google Scholar]
- 51.Wang JF, Kim S, Kovacs F, Cross TA. Structure of the transmembrane region of the M2 protein H+ channel. Protein Sci. 2001;10:2241–2250. doi: 10.1110/ps.17901. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Witter R, Nozirov F, Sternberg U, Cross TA, Ulrich AS, Fu RQ. Solid-state F-19 NMR spectroscopy reveals that Trp(41) participates in the gating mechanism of the M2 proton channel of influenza a virus. J Am Chem Soc. 2008;130:918–924. doi: 10.1021/ja0754305. [DOI] [PubMed] [Google Scholar]
- 53.Zhou HX, Cross TA. Modeling the membrane environment has implications for membrane protein structure and function: Influenza A M2 protein. Protein Sci. 2013;22:381–394. doi: 10.1002/pro.2232. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Gor’kov PL, Chekmenev EY, Li CG, Cotten M, Buffy JJ, Traaseth NJ, Veglia G, Brey WW. Using low-E resonators to reduce RF heating in biological samples for static solid-state NMR up to 900 MHz. J Magn Reson. 2007;185:77–93. doi: 10.1016/j.jmr.2006.11.008. [DOI] [PubMed] [Google Scholar]
- 55.McNeill SA, Gor’kov PL, Shetty K, Brey WW, Long JR. A low-E magic angle spinning probe for biological solid state NMR at 750 MHz. J Magn Reson. 2009;197:135–144. doi: 10.1016/j.jmr.2008.12.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Stringer JA, Bronnimann CE, Mullen CG, Zhou DHH, Stellfox SA, Li Y, Williams EH, Rienstra CM. Reduction of RF-induced sample heating with a scroll coil resonator structure for solid-state NMR probes. J Magn Reson. 2005;173:40–48. doi: 10.1016/j.jmr.2004.11.015. [DOI] [PubMed] [Google Scholar]
- 57.Gor’kov PL, Brey WW, Long JR. Probe development for biosolids NMR spectroscopy. In: Harris RK, Wasylishen RE, editors. Encyclopedia of Magnetic Resonance. John Wiley; Chichester: 2012. [Google Scholar]
- 58.Bruker . Efree probes. 2014. [Google Scholar]
- 59.Agilent. High-Power Tools for Solids Samples. Agilent Technologies; 2012. pp. 1–7. [Google Scholar]
- 60.Huang W, Bardaro MF, Varani G, Drobny GP. Preparation of RNA samples with narrow line widths for solid state NMR investigations. J Magn Reson. 2012;223:51–54. doi: 10.1016/j.jmr.2012.07.018. [DOI] [PubMed] [Google Scholar]
- 61.Cherepanov AV, Glaubitz C, Schwalbe H. High-resolution studies of uniformly C-13, N-15-labeled RNA by solid-state NMR spectroscopy. Angew Chem Int Ed. 2010;49:4747–4750. doi: 10.1002/anie.200906885. [DOI] [PubMed] [Google Scholar]
- 62.Dayie KT. Key labeling technologies to tackle sizeable problems in RNA structural biology. Int J Mol Sci. 2008;9:1214–1240. doi: 10.3390/ijms9071214. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Abramov G, Goldbourt A. Nucleotide-type chemical shift assignment of the encapsulated 40 kbp dsDNA in intact bacteriophage T7 by MAS solid-state NMR. J Biomol NMR. 2014;59:219–230. doi: 10.1007/s10858-014-9840-4. [DOI] [PubMed] [Google Scholar]
- 64.Morag O, Abramov G, Goldbourt A. Complete chemical shift assignment of the ssDNA in the filamentous bacteriophage fd reports on its conformation and on its interface with the capsid shell. J Am Chem Soc. 2014;136:2292–2301. doi: 10.1021/ja412178n. [DOI] [PubMed] [Google Scholar]
- 65.Abramov G, Morag O, Goldbourt A. Magic-angle spinning NMR of a class I filamentous bacteriophage virus. J Phys Chem B. 2011;115:9671–9680. doi: 10.1021/jp2040955. [DOI] [PubMed] [Google Scholar]
- 66.Morag O, Abramov G, Goldbourt A. Similarities and differences within members of the Ff family of filamentous bacteriophage viruses. J Phys Chem B. 2011;115:15370–15379. doi: 10.1021/jp2079742. [DOI] [PubMed] [Google Scholar]
- 67.Sergeyev IV, Day LA, Goldbourt A, McDermott AE. Chemical shifts for the unusual DNA structure in Pf1 bacteriophage from dynamic-nuclear-polarization-enhanced solid-state NMR spectroscopy. J Am Chem Soc. 2011;133:20208–20217. doi: 10.1021/ja2043062. [DOI] [PubMed] [Google Scholar]
- 68.Goldbourt A, Day LA, McDermott AE. Intersubunit hydrophobic interactions in Pf1 filamentous phage. J Biol Chem. 2010;285:37051–37059. doi: 10.1074/jbc.M110.119339. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Goldbourt A, Day LA, McDermott AE. Assignment of congested NMR spectra: Carbonyl backbone enrichment via the Entner-Doudoroff pathway. J Magn Reson. 2007;189:157–165. doi: 10.1016/j.jmr.2007.07.011. [DOI] [PubMed] [Google Scholar]
- 70.Byeon IJL, Hou GJ, Han Y, Suiter CL, Ahn J, Jung J, Byeon CH, Gronenborn AM, Polenova T. Motions on the millisecond time scale and multiple conformations of HIV-1 capsid protein: implications for structural polymorphism of CA assemblies. J Am Chem Soc. 2012;134:6455–6466. doi: 10.1021/ja300937v. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Hou GJ, Yan S, Sun SJ, Han Y, Byeon IJL, Ahn J, Concel J, Samoson A, Gronenborn AM, Polenova T. Spin diffusion driven by R-symmetry sequences: Applications to homonuclear correlation spectroscopy in MAS NMR of biological and organic solids. J Am Chem Soc. 2011;133:3943–3953. doi: 10.1021/ja108650x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Bayro MJ, Chen B, Yau WM, Tycko R. Site-specific structural variations accompanying tubular assembly of the HIV-1 capsid protein. J Mol Biol. 2014;426:1109–1127. doi: 10.1016/j.jmb.2013.12.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Chen B, Tycko R. Structural and dynamical characterization of tubular HIV-1 capsid protein assemblies by solid state nuclear magnetic resonance and electron microscopy. Protein Sci. 2010;19:716–730. doi: 10.1002/pro.348. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Barbet-Massin E, Pell AJ, Jaudzems K, Franks WT, Retel JS, Kotelovica S, Akopjana I, Tars K, Emsley L, Oschkinat H, Lesage A, Pintacuda G. Out-and-back C-13-C-13 scalar transfers in protein resonance assignment by proton-detected solid-state NMR under ultra-fast MAS. J Biomol NMR. 2013;56:379–386. doi: 10.1007/s10858-013-9757-3. [DOI] [PubMed] [Google Scholar]
- 75.Barbet-Massin E, Pell AJ, Retel JS, Andreas LB, Jaudzems K, Franks WT, Nieuwkoop AJ, Hiller M, Higman V, Guerry P, Bertarello A, Knight MJ, Felletti M, Le Marchand T, Kotelovica S, Akopjana I, Tars K, Stoppini M, Bellotti V, Bolognesi M, Ricagno S, Chou JJ, Griffin RG, Oschkinat H, Lesage A, Emsley L, Herrmann T, Pintacuda G. Rapid proton-detected NMR assignment for proteins with fast magic angle spinning. J Am Chem Soc. 2014;136:12489–12497. doi: 10.1021/ja507382j. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Luo WB, Cady SD, Hong M. Immobilization of the influenza A M2 transmembrane peptide in virus envelope-mimetic lipid membranes: A solid-state NMR investigation. Biochemistry. 2009;48:6361–6368. doi: 10.1021/bi900716s. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Cady SD, Hong M. Effects of amantadine on the dynamics of membrane-bound influenza A M2 transmembrane peptide studied by NMR relaxation. J Biomol NMR. 2009;45:185–196. doi: 10.1007/s10858-009-9352-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Qiang W, Sun Y, Weliky DP. A strong correlation between fusogenicity and membrane insertion depth of the HIV fusion peptide. Proc Natl Acad Sci USA. 2009;106:15314–15319. doi: 10.1073/pnas.0907360106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Sackett K, Nethercott MJ, Zheng ZX, Weliky DP. Solid-state NMR spectroscopy of the HIV gp41 membrane fusion protein supports intermolecular antiparallel 13 sheet fusion peptide structure in the final six-helix bundle state. J Mol Biol. 2014;426:1077–1094. doi: 10.1016/j.jmb.2013.11.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Yang J, Weliky DP. Solid-state nuclear magnetic resonance evidence for parallel and antiparallel strand arrangements in the membrane-associated HIV-1 fusion peptide. Biochemistry. 2003;42:11879–11890. doi: 10.1021/bi0348157. [DOI] [PubMed] [Google Scholar]
- 81.Wagstaff JL, Howard MJ, Williamson RA. Production of recombinant isotopically labelled peptide by fusion to an insoluble partner protein: generation of integrin alpha v beta 6 binding peptides for NMR. Mol Biosyst. 2010;6:2380–2385. doi: 10.1039/c0mb00105h. [DOI] [PubMed] [Google Scholar]
- 82.Havlin RH, Blanco FJ, Tycko R. Constraints on protein structure in HIV-1 Rev and Rev-RNA supramolecular assemblies from two-dimensional solid state nuclear magnetic resonance. Biochemistry. 2007;46:3586–3593. doi: 10.1021/bi0622928. [DOI] [PubMed] [Google Scholar]
- 83.Yu TY, Schaefer J. REDOR NMR characterization of DNA packaging in bacteriophage T4. J Mol Biol. 2008;382:1031–1042. doi: 10.1016/j.jmb.2008.07.077. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Cady SD, Schmidt-Rohr K, Wang J, Soto CS, DeGrado WF, Hong M. Structure of the amantadine binding site of influenza M2 proton channels in lipid bilayers. Nature. 2010;463:689–692. doi: 10.1038/nature08722. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Qiang W, Bodner ML, Weliky DP. Solid-state NMR Spectroscopy of human immunodeficiency virus fusion peptides associated with host-cell-like membranes: 2D correlation spectra and distance measurements support a fully extended conformation and models for specific antiparallel strand registries. J Am Chem Soc. 2008;130:5459–5471. doi: 10.1021/ja077302m. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Sharpe S, Yau WM, Tycko R. Structure and dynamics of the HIV-1 Vpu transmembrane domain revealed by solid-state NMR with magic-angle spinning. Biochemistry. 2006;45:918–933. doi: 10.1021/bi051766k. [DOI] [PubMed] [Google Scholar]
- 87.Cady S, Wang T, Hong M. Membrane-dependent effects of a cytoplasmic helix on the structure and drug binding of the influenza virus M2 protein. J Am Chem Soc. 2011;133:11572–11579. doi: 10.1021/ja202051n. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Wang T, Cady SD, Hong M. NMR determination of protein partitioning into membrane domains with different curvatures and application to the Influenza M2 peptide. Biophys J. 2012;102:787–794. doi: 10.1016/j.bpj.2012.01.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Luo W, Hong M. Determination of the oligomeric number and intermolecular distances of membrane protein assemblies by anisotropic (1)H-driven spin diffusion NMR spectroscopy. J Am Chem Soc. 2006;128:7242–7251. doi: 10.1021/ja0603406. [DOI] [PubMed] [Google Scholar]
- 90.Lundstrom P, Teilum K, Carstensen T, Bezsonova I, Wiesner S, Hansen DF, Religa TL, Akke M, Kay LE. Fractional C-13 enrichment of isolated carbons using [1-C-13]- or [2-C-13]-glucose facilitates the accurate measurement of dynamics at backbone C-alpha and side-chain methyl positions in proteins. J Biomol NMR. 2007;38:199–212. doi: 10.1007/s10858-007-9158-6. [DOI] [PubMed] [Google Scholar]
- 91.Park SH, Yang C, Opella SJ, Mueller LJ. Resolution and measurement of heteronuclear dipolar couplings of a noncrystalline protein immobilized in a biological supramolecular assembly by proton-detected MAS solid-state NMR spectroscopy. J Magn Reson. 2013;237:164–168. doi: 10.1016/j.jmr.2013.10.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Reif B, Griffin RG. H-1 detected H-1, N-15 correlation spectroscopy in rotating solids. J Magn Reson. 2003;160:78–83. doi: 10.1016/s1090-7807(02)00035-6. [DOI] [PubMed] [Google Scholar]
- 93.Chevelkov V, Fink U, Reif B. Accurate determination of order parameters from H-1, N-15 dipolar couplings in MAS solid-state NMR experiments. J Am Chem Soc. 2009;131:14018–14022. doi: 10.1021/ja902649u. [DOI] [PubMed] [Google Scholar]
- 94.Lewandowski JR, Dumez JN, Akbey U, Lange S, Emsley L, Oschkinat H. Enhanced resolution and coherence lifetimes in the solid-state NMR spectroscopy of perdeuterated proteins under ultrafast magic-angle spinning. Journal of Physical Chemistry Letters. 2011;2:2205–2211. [Google Scholar]
- 95.Akbey U, Lange S, Franks WT, Linser R, Rehbein K, Diehl A, van Rossum BJ, Reif B, Oschkinat H. Optimum levels of exchangeable protons in perdeuterated proteins for proton detection in MAS solid-state NMR spectroscopy. J Biomol NMR. 2010;46:67–73. doi: 10.1007/s10858-009-9369-0. [DOI] [PubMed] [Google Scholar]
- 96.Chevelkov V, Rehbein K, Diehl A, Reif B. Ultrahigh resolution in proton solid-state NMR spectroscopy at high levels of deuteration. Angew Chem Int Ed. 2006;45:3878–3881. doi: 10.1002/anie.200600328. [DOI] [PubMed] [Google Scholar]
- 97.Reif B. Ultra-high resolution in MAS solid-state NMR of perdeuterated proteins: Implications for structure and dynamics. J Magn Reson. 2012;216:1–12. doi: 10.1016/j.jmr.2011.12.017. [DOI] [PubMed] [Google Scholar]
- 98.Ivanir H, Goldbourt A. Solid state NMR chemical shift assignment and conformational analysis of a cellulose binding protein facilitated by optimized glycerol enrichment. J Biomol NMR. 2014;59:185–197. doi: 10.1007/s10858-014-9838-y. [DOI] [PubMed] [Google Scholar]
- 99.Volk J, Herrmann T, Wuthrich K. Automated sequence-specific protein NMR assignment using the memetic algorithm MATCH. J Biomol NMR. 2008;41:127–138. doi: 10.1007/s10858-008-9243-5. [DOI] [PubMed] [Google Scholar]
- 100.Ulrich EL, Akutsu H, Doreleijers JF, Harano Y, Ioannidis YE, Lin J, Livny M, Mading S, Maziuk D, Miller Z, Nakatani E, Schulte CF, Tolmie DE, Kent Wenger R, Yao H, Markley JL. BioMagResBank. Nucleic Acids Res. 2008;36:D402–408. doi: 10.1093/nar/gkm957. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Barbet-Massin E, Felletti M, Schneider R, Jehle S, Communie G, Martinez N, Jensen MR, Ruigrok RWH, Emsley L, Lesage A, Blackledge M, Pintacuda G. Insights into the structure and dynamics of measles virus nucleocapsids by 1H-detected solid-state NMR. Biophys J. 2014;107:941– 946. doi: 10.1016/j.bpj.2014.05.048. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Holland GP, Cherry BR, Jenkins JE, Yarger JL. Proton-detected heteronuclear single quantum correlation NMR spectroscopy in rigid solids with ultra-fast MAS. J Magn Reson. 2010;202:6471. doi: 10.1016/j.jmr.2009.09.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Ishii Y, Yesinowski JP, Tycko R. Sensitivity enhancement in solid-state C-13 NMR of synthetic polymers and biopolymers by H-1 NMR detection with high-speed magic angle spinning. J Am Chem Soc. 2001;123:29212922. doi: 10.1021/ja015505j. [DOI] [PubMed] [Google Scholar]
- 104.Yan S, Suiter CL, Hou GJ, Zhang HL, Polenova T. Probing structure and dynamics of protein assemblies by magic angle spinning NMR spectroscopy. Acc Chem Res. 2013;46:20472058. doi: 10.1021/ar300309s. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Hou GJ, Byeon IJL, Ahn J, Gronenborn AM, Polenova T. H-1-C-13/H-1-N-15 heteronuclear dipolar recoupling by R-symmetry sequences under fast magic angle spinning for dynamics analysis of biological and organic solids. J Am Chem Soc. 2011;133:1864618655. doi: 10.1021/ja203771a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Hou GJ, Yan S, Trebosc J, Amoureux JP, Polenova T. Broadband homonuclear correlation spectroscopy driven by combined R2(n)(v) sequences under fast magic angle spinning for NMR structural analysis of organic and biological solids. J Magn Reson. 2013;232:18–30. doi: 10.1016/j.jmr.2013.04.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Lu X, Guo C, Hou G, Polenova T. Combined zero-quantum and spin-diffusion mixing for efficient homonuclear correlation spectroscopy under fast MAS: Broadband recoupling and detection of long-range correlations. 2014. Submitted. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Hou G, Lu X, Vega AJ, Polenova T. Accurate measurement of heteronuclear dipolar couplings by phase-alternating R-symmetry (PARS) sequences in magic angle spinning NMR spectroscopy. J Chem Phys. 2014;141:104202. doi: 10.1063/1.4894226. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Hou GJ, Byeon IJL, Ahn J, Gronenborn AM, Polenova T. Recoupling of chemical shift anisotropy by R-symmetry sequences in magic angle spinning NMR spectroscopy. J Chem Phys. 2012;137 doi: 10.1063/1.4754149. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Hou GJ, Paramasivam S, Yan S, Polenova T, Vega AJ. Multidimensional magic angle spinning NMR spectroscopy for site-resolved measurement of proton chemical shift anisotropy in biological solids. J Am Chem Soc. 2013;135:1358–1368. doi: 10.1021/ja3084972. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Zhao X, Eden M, Levitt MH. Recoupling of heteronuclear dipolar interactions in solid-state NMR using symmetry-based pulse sequences. Chem Phys Lett. 2001;342:353–361. [Google Scholar]
- 112.Hou G, Paramasivam S, Byeon IJL, Gronenborn AM, Polenova T. Determination of relative tensor orientations by gamma-encoded chemical shift anisotropy/heteronuclear dipolar coupling 3D NMR spectroscopy in biological solids. PCCP. 2010;12:14873–14883. doi: 10.1039/c0cp00795a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Yang J, Tasayco ML, Polenova T. Dynamics of reassembled thioredoxin studied by magic angle spinning NMR: Snapshots from different time scales. J Am Chem Soc. 2009;131:13690–13702. doi: 10.1021/ja9037802. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Chevelkov V, Habenstein B, Loquet A, Giller K, Becker S, Lange A. Proton-detected MAS NMR experiments based on dipolar transfers for backbone assignment of highly deuterated proteins. J Magn Reson. 2014;242:180–188. doi: 10.1016/j.jmr.2014.02.020. [DOI] [PubMed] [Google Scholar]
- 115.Knight MJ, Webber AL, Pell AJ, Guerry P, Barbet-Massin E, Bertini I, Felli IC, Gonnelli L, Pierattelli R, Emsley L, Lesage A, Herrmann T, Pintacuda G. Fast resonance assignment and fold determination of human superoxide dismutase by high-resolution proton-detected solid-state MAS NMR spectroscopy. Angew Chem Int Ed. 2011;50:11697–11701. doi: 10.1002/anie.201106340. [DOI] [PubMed] [Google Scholar]
- 116.Marchetti A, Jehle S, Felletti M, Knight MJ, Wang Y, Xu ZQ, Park AY, Otting G, Lesage A, Emsley L, Dixon NE, Pintacuda G. Backbone assignment of fully protonated solid proteins by 1H detection and ultrafast magic-angle-spinning NMR spectroscopy. Angew Chem Int Ed. 2012;51:10756–10759. doi: 10.1002/anie.201203124. [DOI] [PubMed] [Google Scholar]
- 117.Zhou DH, Shah G, Cormos M, Mullen C, Sandoz D, Rienstra CM. Proton-detected solid-state NMR Spectroscopy of fully protonated proteins at 40 kHz magic-angle spinning. J Am Chem Soc. 2007;129:11791–11801. doi: 10.1021/ja073462m. [DOI] [PubMed] [Google Scholar]
- 118.Huber M, Hiller S, Schanda P, Ernst M, Bockmann A, Verel R, Meier BH. A proton-detected 4D solid-state NMR experiment for protein structure determination. ChemPhysChem. 2011;12:915–918. doi: 10.1002/cphc.201100062. [DOI] [PubMed] [Google Scholar]
- 119.Zhou DH, Shea JJ, Nieuwkoop AJ, Franks WT, Wylie BJ, Mullen C, Sandoz D, Rienstra CM. Solid-sate protein-structure determination with proton-detected triple-resonance 3D magic-angle-spinning NMR spectroscopy. Angew Chem Int Ed. 2007;46:8380–8383. doi: 10.1002/anie.200702905. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Linser R, Bardiaux B, Higman V, Fink U, Reif B. Structure calculation from unambiguous long-range amide and methyl H-1-H-1 distance restraints for a microcrystalline protein with MAS solid-state NMR spectroscopy. J Am Chem Soc. 2011;133:5905–5912. doi: 10.1021/ja110222h. [DOI] [PubMed] [Google Scholar]
- 121.Chevelkov V, Diehl A, Reif B. Quantitative measurement of differential N-15-H-alpha/beta T-2 relaxation rates in a perdeuterated protein by MAS solid-state NMR spectroscopy. Magn Reson Chem. 2007;45:S156–S160. doi: 10.1002/mrc.2129. [DOI] [PubMed] [Google Scholar]
- 122.Chevelkov V, Fink U, Reif B. Quantitative analysis of backbone motion in proteins using MAS solid-state NMR spectroscopy. J Biomol NMR. 2009;45:197–206. doi: 10.1007/s10858-009-9348-5. [DOI] [PubMed] [Google Scholar]
- 123.Knight MJ, Pell AJ, Bertini I, Felli IC, Gonnelli L, Pierattelli R, Herrmann T, Emsley L, Pintacuda G. Structure and backbone dynamics of a microcrystalline metalloprotein by solid-state NMR. Proc Natl Acad Sci USA. 2012;109:11095–11100. doi: 10.1073/pnas.1204515109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Schanda P, Meier BH, Ernst M. Quantitative analysis of protein backbone dynamics in microcrystalline ubiquitin by solid-state NMR spectroscopy. J Am Chem Soc. 2010;132:15957–15967. doi: 10.1021/ja100726a. [DOI] [PubMed] [Google Scholar]
- 125.Asami S, Rakwalska-Bange M, Carlomagno T, Reif B. Protein-RNA interfaces probed by 1H-detected MAS solid-state NMR spectroscopy. Angew Chem Int Ed. 2013;52:2345–2349. doi: 10.1002/anie.201208024. [DOI] [PubMed] [Google Scholar]
- 126.Sun ZYJ, Hyberts SG, Rovnyak D, Park S, Stern AS, Hoch JC, Wagner G. High-resolution aliphatic side-chain assignments in 3D HCcoNH experiments with joint H-C evolution and non-uniform sampling. J Biomol NMR. 2005;32:55–60. doi: 10.1007/s10858-005-3339-y. [DOI] [PubMed] [Google Scholar]
- 127.Frueh DP, Sun ZYJ, Vosburg DA, Walsh CT, Hoch JC, Wagner G. Non-uniformly sampled double-TROSY hNcaNH experiments for NMR sequential assignments of large proteins. J Am Chem Soc. 2006;128:5757–5763. doi: 10.1021/ja0584222. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Rovnyak D, Frueh DP, Sastry M, Sun ZYJ, Stern AS, Hoch JC, Wagner G. Accelerated acquisition of high resolution triple-resonance spectra using non-uniform sampling and maximum entropy reconstruction. J Magn Reson. 2004;170:15–21. doi: 10.1016/j.jmr.2004.05.016. [DOI] [PubMed] [Google Scholar]
- 129.Rovnyak D, Hoch JC, Stern AS, Wagner G. Resolution and sensitivity of high field nuclear magnetic resonance spectroscopy. J Biomol NMR. 2004;30:1–10. doi: 10.1023/B:JNMR.0000042946.04002.19. [DOI] [PubMed] [Google Scholar]
- 130.Motackova V, Novacek J, Zawadzka-Kazimierczuk A, Kazimierczuk K, Zidek L, Anderova HS, Krasny L, Kozminski W, Sklenar V. Strategy for complete NMR assignment of disordered proteins with highly repetitive sequences based on resolution-enhanced 5D experiments. J Biomol NMR. 2010;48:169–177. doi: 10.1007/s10858-010-9447-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Rovnyak D, Sarcone M, Jiang Z. Sensitivity enhancement for maximally resolved two-dimensional NMR by nonuniform sampling. Magn Reson Chem. 2011;49:483–491. doi: 10.1002/mrc.2775. [DOI] [PubMed] [Google Scholar]
- 132.Suiter CL, Paramasivam S, Hou GJ, Sun SJ, Rice D, Hoch J, Rovnyak D, Polenova T. Sensitivity gains, linearity, and spectral reproducibility in nonuniformly sampled multidimensional MAS NMR spectra of high dynamic range. J Biomol NMR. 2014;59:57–73. doi: 10.1007/s10858-014-9824-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Hoch JC, Maciejewski MW, Mobli M, Schuyler AD, Stern AS. Nonuniform sampling and maximum entropy reconstruction in multidimensional NMR. Acc Chem Res. 2014;47:708–717. doi: 10.1021/ar400244v. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Mobli M, Stern AS, Hoch JC. Spectral reconstruction methods in fast NMR: Reduced dimensionality, random sampling and maximum entropy. J Magn Reson. 2006;182:96–105. doi: 10.1016/j.jmr.2006.06.007. [DOI] [PubMed] [Google Scholar]
- 135.Sibisi S, Skilling J, Brereton RG, Laue ED, Staunton J. Maximum entropy signal processing in practical NMR spectroscopy. Nature. 1984;311:446–447. [Google Scholar]
- 136.Paramasivam S, Suiter CL, Hou GJ, Sun SJ, Palmer M, Hoch JC, Rovnyak D, Polenova T. Enhanced sensitivity by nonuniform sampling enables multidimensional MAS NMR spectroscopy of protein assemblies. J Phys Chem B. 2012;116:7416–7427. doi: 10.1021/jp3032786. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Balsgart NM, Vosegaard T. Fast forward maximum entropy reconstruction of sparsely sampled data. J Magn Reson. 2012;223:164–169. doi: 10.1016/j.jmr.2012.07.002. [DOI] [PubMed] [Google Scholar]
- 138.Hyberts SG, Frueh DP, Arthanari H, Wagner G. FM reconstruction of non-uniformly sampled protein NMR data at higher dimensions and optimization by distillation. J Biomol NMR. 2009;45:283–294. doi: 10.1007/s10858-009-9368-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.Hyberts SG, Heffron GJ, Tarragona NG, Solanky K, Edmonds KA, Luithardt H, Fejzo J, Chorev M, Aktas H, Colson K, Falchuk KH, Halperin JA, Wagner G. Ultrahigh-resolution H-1-C-13 HSQC spectra of metabolite mixtures using nonlinear sampling and forward maximum entropy reconstruction. J Am Chem Soc. 2007;129:5108–5116. doi: 10.1021/ja068541x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Hyberts SG, Takeuchi K, Wagner G. Poisson-Gap sampling and forward maximum entropy reconstruction for enhancing the resolution and sensitivity of protein NMR data. J Am Chem Soc. 2010;132:2145–2147. doi: 10.1021/ja908004w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Franks WT, Atreya HS, Szyperski T, Rienstra CM. GFT projection NMR spectroscopy for proteins in the solid state. J Biomol NMR. 2010;48:213–223. doi: 10.1007/s10858-010-9451-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Kim S, Szyperski T. GFT NMR, a new approach to rapidly obtain precise high-dimensional NMR spectral information. J Am Chem Soc. 2003;125:1385–1393. doi: 10.1021/ja028197d. [DOI] [PubMed] [Google Scholar]
- 143.Bruschweiler R, Zhang FL. Covariance nuclear magnetic resonance spectroscopy. J Chem Phys. 2004;120:5253–5260. doi: 10.1063/1.1647054. [DOI] [PubMed] [Google Scholar]
- 144.Lin EC, Opella SJ. Covariance spectroscopy in high-resolution multi-dimensional solid-state NMR. J Magn Reson. 2014;239:57–60. doi: 10.1016/j.jmr.2013.11.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Matsuki Y, Eddy MT, Griffin RG, Herzfeld J. Rapid three-dimensional MAS NMR spectroscopy at critical sensitivity. Angew Chem Int Ed. 2010;49:9215–9218. doi: 10.1002/anie.201003329. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Matsuki Y, Eddy MT, Herzfeld J. Spectroscopy by Integration of Frequency and Time Domain Information for Fast Acquisition of High-Resolution Dark Spectra. J Am Chem Soc. 2009;131:4648–4656. doi: 10.1021/ja807893k. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Hyberts SG, Arthanari H, Wagner G. Novel Sampling Approaches in Higher Dimensional NMR. 2012. Applications of non-uniform sampling and processing; pp. 125–148. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Hoch J, Maciejewski MW, Mobli M, Schuyler AD, Stern AS. Nonuniform sampling in multidimensional NMR. In: Harris RK, Wasylishen RE, editors. Encyclopedia of NMR. Wiley; Chichester: 2012. pp. 2999–3013. [Google Scholar]
- 149.Hoch J, Stern AS. NMR Data Processing. Wiley; New York: 1996. [Google Scholar]
- 150.Hyberts SG, Arthanari H, Robson SA, Wagner G. Perspectives in magnetic resonance: NMR in the post-FFT era. J Magn Reson. 2014;241:60–73. doi: 10.1016/j.jmr.2013.11.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Sun SJ, Yan S, Guo CM, Li MY, Hoch JC, Williams JC, Polenova T. A Time-Saving Strategy for MAS NMR Spectroscopy by Combining Nonuniform Sampling and Paramagnetic Relaxation Assisted Condensed Data Collection. J Phys Chem B. 2012;116:13585–13596. doi: 10.1021/jp3005794. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Li YX, Wang Q, Zhang ZF, Yang J, Hu BW, Chen Q, Noda I, Deng F. Covariance spectroscopy with a non-uniform and consecutive acquisition scheme for signal enhancement of the NMR experiments. J Magn Reson. 2012;217:106–111. doi: 10.1016/j.jmr.2012.02.016. [DOI] [PubMed] [Google Scholar]
- 153.Qiang W. Signal enhancement for the sensitivity-limited solid state NMR experiments using a continuous, non-uniform acquisition scheme. J Magn Reson. 2011;213:171–175. doi: 10.1016/j.jmr.2011.08.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Lin EC, Opella SJ. Sampling scheme and compressed sensing applied to solid-state NMR spectroscopy. J Magn Reson. 2013;237:40–48. doi: 10.1016/j.jmr.2013.09.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Hall DA, Maus DC, Gerfen GJ, Inati SJ, Becerra LR, Dahlquist FW, Griffin RG. Polarization-enhanced NMR spectroscopy of biomolecules in frozen solution. Science. 1997;276:930–932. doi: 10.1126/science.276.5314.930. [DOI] [PubMed] [Google Scholar]
- 156.Song CS, Hu KN, Joo CG, Swager TM, Griffin RG. TOTAPOL: A biradical polarizing agent for dynamic nuclear polarization experiments in aqueous media. J Am Chem Soc. 2006;128:11385–11390. doi: 10.1021/ja061284b. [DOI] [PubMed] [Google Scholar]
- 157.Maly T, Cui DT, Griffin RG, Miller AF. H-1 dynamic nuclear polarization based on an endogenous radical. J Phys Chem B. 2012;116:7055–7065. doi: 10.1021/jp300539j. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.McDermott A, Zysmilich MG, Polenova T. Solid state NMR studies of photoinduced polarization in photosynthetic reaction centers: mechanism and simulations. Solid State Nucl Magn Reson. 1998;11:21–47. doi: 10.1016/s0926-2040(97)00094-5. [DOI] [PubMed] [Google Scholar]
- 159.Zysmilich MG, McDermott A. Photochemically induced nuclear spin polarization in bacterial photosynthetic reaction centers: Assignments of the N-15 SSNMR spectra. J Am Chem Soc. 1996;118:5867–5873. [Google Scholar]
- 160.Maly T, Debelouchina GT, Bajaj VS, Hu KN, Joo CG, Mak-Jurkauskas ML, Sirigiri JR, van der Wel PCA, Herzfeld J, Temkin RJ, Griffin RG. Dynamic nuclear polarization at high magnetic fields. J Chem Phys. 2008;128:052211. doi: 10.1063/1.2833582. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161.Rosay M, Zeri AC, Astrof NS, Opella SJ, Herzfeld J, Griffin RG. Sensitivity-enhanced NMR of biological solids: Dynamic nuclear polarization of Y21M fd bacteriophage and purple membrane. J Am Chem Soc. 2001;123:1010–1011. doi: 10.1021/ja005659j. [DOI] [PubMed] [Google Scholar]
- 162.Andreas LB, Barnes AB, Corzilius B, Chou JJ, Miller EA, Caporini M, Rosay M, Griffin RG. Dynamic nuclear polarization study of inhibitor binding to the M2(18–60) proton transporter from influenza A. Biochemistry. 2013;52:2774–2782. doi: 10.1021/bi400150x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163.Janssen GJ, Daviso E, van Son M, de Groot HJM, Alia A, Matysik J. Observation of the solid-state photo-CIDNP effect in entire cells of cyanobacteria Synechocystis. Photosynth Res. 2010;104:275–282. doi: 10.1007/s11120-009-9508-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Thamarath SS, Alia A, Daviso E, Mance D, Golbeck JH, Matysik J. Whole cell nuclear magnetic resonance characterization of two photochemically active states of the photosynthetic reaction center in heliobacteria. Biochemistry. 2012;51:5763–5773. doi: 10.1021/bi300468y. [DOI] [PubMed] [Google Scholar]
- 165.Engelman A, Cherepanov P. The structural biology of HIV-1: mechanistic and therapeutic insights. Nat Rev Microbiol. 2012;10:279–290. doi: 10.1038/nrmicro2747. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Briggs JAG, Krausslich HG. The molecular architecture of HIV. J Mol Biol. 2011;410:491–500. doi: 10.1016/j.jmb.2011.04.021. [DOI] [PubMed] [Google Scholar]
- 167.Gottlinger HG. The HIV-1 assembly machine. AIDS. 2001;15:S13–S20. doi: 10.1097/00002030-200100005-00003. [DOI] [PubMed] [Google Scholar]
- 168.Pettit SC, Moody MD, Wehbie RS, Kaplan AH, Nantermet PV, Klein CA, Swanstrom R. The P2 domain of Human Immunodeficiency Virus Type-1 Gag regulates sequential proteolytic processing and is required to produce fully infectious virions. J Virol. 1994;68:8017–8027. doi: 10.1128/jvi.68.12.8017-8027.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 169.Sundquist WI, Krausslich HG. HIV-1 assembly, budding, and maturation. Cold Spring Harbor Perspectives in Medicine. 2012;2 doi: 10.1101/cshperspect.a006924. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170.Ganser-Pornillos BK, Cheng A, Yeager M. Structure of full-length HIV-1CA: a model for the mature capsid lattice. Cell. 2007;131:70–79. doi: 10.1016/j.cell.2007.08.018. [DOI] [PubMed] [Google Scholar]
- 171.Pornillos O, Ganser-Pornillos BK, Yeager M. Atomic-level modelling of the HIV capsid. Nature. 2011;469:424–427. doi: 10.1038/nature09640. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 172.Zhao GP, Perilla JR, Yufenyuy EL, Meng X, Chen B, Ning JY, Ahn J, Gronenborn AM, Schulten K, Aiken C, Zhang PJ. Mature HIV-1 capsid structure by cryo-electron microscopy and all-atom molecular dynamics. Nature. 2013;497:643–646. doi: 10.1038/nature12162. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Gross I, Hohenberg H, Krausslich HG. In vitro assembly properties of purified bacterially expressed capsid proteins of human immunodeficiency virus. Eur J Biochem. 1997;249:592–600. doi: 10.1111/j.1432-1033.1997.t01-1-00592.x. [DOI] [PubMed] [Google Scholar]
- 174.Barklis E, McDermott J, Wilkens S, Fuller S, Thompson D. Organization of HIV-1 capsid proteins on a lipid monolayer. J Biol Chem. 1998;273:7177–7180. doi: 10.1074/jbc.273.13.7177. [DOI] [PubMed] [Google Scholar]
- 175.Ehrlich LS, Liu TB, Scarlata S, Chu B, Carter CA. HIV-1 capsid protein forms spherical (immature-like) and tubular (mature-like) particles in vitro: Structure switching by pH-induced conformational changes. Biophys J. 2001;81:586–594. doi: 10.1016/S0006-3495(01)75725-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176.Newman JL, Butcher EW, Patel DT, Mikhaylenko Y, Summers MF. Flexibility in the P2 domain of the HIV-1 Gag polyprotein. Protein Sci. 2004;13:2101–2107. doi: 10.1110/ps.04614804. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Datta SAK, Temeselew LG, Crist RM, Soheilian F, Kamata A, Mirro J, Harvin D, Nagashima K, Cachau RE, Rein A. On the role of the SP1 domain in HIV-1 particle assembly: a molecular switch? J Virol. 2011;85:4111–4121. doi: 10.1128/JVI.00006-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178.Morellet N, Druillennec S, Lenoir C, Bouaziz S, Roques BP. Helical structure determined by NMR of the HIV-1 (345–392)Gag sequence, surrounding p2: Implications for particle assembly and RNA packaging. Protein Sci. 2005;14:375–386. doi: 10.1110/ps.041087605. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179.Hu KN, Qiang W, Tycko R. A general Monte Carlo/simulated annealing algorithm for resonance assignment in NMR of uniformly labeled biopolymers. J Biomol NMR. 2011;50:267–276. doi: 10.1007/s10858-011-9517-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.Hu KN, Qiang W, Bermejo GA, Schwieters CD, Tycko R. Restraints on backbone conformations in solid state NMR studies of uniformly labeled proteins from quantitative amide N-15-N-15 and carbonyl C-13-C-13 dipolar recoupling data. J Magn Reson. 2012;218:115–127. doi: 10.1016/j.jmr.2012.03.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181.Daly TJ, Cook KS, Gray GS, Maione TE, Rusche JR. Specific binding of HIV-1 recombinant Rev protein to the Rev-Responsive Element invitro. Nature. 1989;342:816–819. doi: 10.1038/342816a0. [DOI] [PubMed] [Google Scholar]
- 182.Brighty DW, Rosenberg M. A cis-acting repressive sequence that overlaps the Rev-responsive element of Human-Immunodeficiency-Virus Type-1 regulates nuclear retention of Env Messenger-RNAs independently of known splice signals. Proc Natl Acad Sci USA. 1994;91:8314–8318. doi: 10.1073/pnas.91.18.8314. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183.Blanco FJ, Hess S, Pannell LK, Rizzo NW, Tycko R. Solid-state NMR data support a helix-loop-helix structural model for the N-terminal half of HIV-1 Rev in fibrillar form. J Mol Biol. 2001;313:845–859. doi: 10.1006/jmbi.2001.5067. [DOI] [PubMed] [Google Scholar]
- 184.White JM, Delos SE, Brecher M, Schornberg K. Structures and mechanisms of viral membrane fusion proteins: Multiple variations on a common theme. Crit Rev Biochem Mol Biol. 2008;43:287–288. doi: 10.1080/10409230802058320. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185.Mao YD, Wang LP, Gu C, Herschhorn A, Xiang SH, Haim H, Yang XZ, Sodroski J. Subunit organization of the membrane-bound HIV-1 envelope glycoprotein trimer. Nat Struct Mol Biol. 2012;19:893–899. doi: 10.1038/nsmb.2351. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186.Schmick SD, Weliky DP. Major antiparallel and minor parallel beta sheet populations detected in the membrane-associated Human Immunodeficiency Virus fusion peptide. Biochemistry. 2010;49:10623–10635. doi: 10.1021/bi101389r. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187.Gabrys CM, Yang R, Wasniewski CM, Yang J, Canlas CG, Qiang W, Sun Y, Weliky DP. Nuclear magnetic resonance evidence for retention of a lamellar membrane phase with curvature in the presence of large quantities of the HIV fusion peptide. Biochim Biophys Acta. 2010;1798:194–201. doi: 10.1016/j.bbamem.2009.07.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Qiang W, Weliky DP. HIV fusion peptide and its cross-linked oligomers: Efficient syntheses, significance of the trimer in fusion activity, correlation of beta strand conformation with membrane cholesterol, and proximity to lipid headgroups. Biochemistry. 2009;48:289–301. doi: 10.1021/bi8015668. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189.Qiang W, Yang J, Weliky DP. Solid-state nuclear magnetic resonance measurements of HIV fusion peptide to lipid distances reveal the intimate contact of beta strand peptide with membranes and the proximity of the Ala-14-Gly-16 region with lipid headgroups. Biochemistry. 2007;46:4997–5008. doi: 10.1021/bi6024808. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190.Xie L, Ghosh U, Schmick SD, Weliky DP. Residue-specific membrane location of peptides and proteins using specifically and extensively deuterated lipids and C-13-H-2 rotational-echo double-resonance solid-state NMR. J Biomol NMR. 2013;55:11–17. doi: 10.1007/s10858-012-9692-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191.Sackett K, Nethercott MJ, Shai Y, Weliky DP. Hairpin folding of HIV gp41 abrogates lipid mixing function at physiologic pH and inhibits lipid mixing by exposed gp41 constructs. Biochemistry. 2009;48:2714–2722. doi: 10.1021/bi8019492. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192.Sackett K, Nethercott MJ, Epand RF, Epand RM, Kindra DR, Shai Y, Weliky DP. Comparative analysis of membrane-associated fusion peptide secondary structure and lipid mixing function of HIV gp41 constructs that model the early pre-hairpin intermediate and final hairpin conformations. J Mol Biol. 2010;397:301–315. doi: 10.1016/j.jmb.2010.01.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193.Vogel EP, Curtis-Fisk J, Young KM, Weliky DP. Solid-state nuclear magnetic resonance (NMR) spectroscopy of Human Immunodeficiency Virus gp41 protein that includes the fusion peptide: NMR detection of recombinant Fgp41 in inclusion bodies in whole bacterial cells and structural characterization of purified and membrane-associated Fgp41. Biochemistry. 2011;50:10013–10026. doi: 10.1021/bi201292e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 194.Balbach JJ, Yang J, Weliky DP, Steinbach PJ, Tugarinov V, Anglister J, Tycko R. Probing hydrogen bonds in the antibody-bound HIV-1 gp120 V3 loop by solid state NMR REDOR measurements. J Biomol NMR. 2000;16:313–327. doi: 10.1023/a:1008343623240. [DOI] [PubMed] [Google Scholar]
- 195.Sharpe S, Kessler N, Anglister JA, Yau WM, Tycko R. Solid-state NMR yields structural constraints on the V3 loop from HIV-1 Gp120 bound to the 447-52D antibody Fv fragment. J Am Chem Soc. 2004;126:4979–4990. doi: 10.1021/ja0392162. [DOI] [PubMed] [Google Scholar]
- 196.Frankel AD, Pabo CO. Cellular uptake of the Tat protein from Human Immunodeficiency Virus. Cell. 1988;55:1189–1193. doi: 10.1016/0092-8674(88)90263-2. [DOI] [PubMed] [Google Scholar]
- 197.Su YC, Waring AJ, Ruchala P, Hong M. Membrane-bound dynamic structure of an Arginine-rich cell-penetrating peptide, the protein transduction domain of HIV TAT, from solid-state NMR. Biochemistry. 2010;49:6009–6020. doi: 10.1021/bi100642n. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 198.Cohen EA, Terwilliger EF, Sodroski JG, Haseltine WA. Identification of a protein encoded by the Vpu gene of HIV-1. Nature. 1988;334:532–534. doi: 10.1038/334532a0. [DOI] [PubMed] [Google Scholar]
- 199.Do HQ, Wittlich M, Gluck JM, Mockel L, Willbold D, Koenig BW, Heise H. Full-length Vpu and human CD4(372–433) in phospholipid bilayers as seen by magic angle spinning NMR. Biol Chem. 2013;394:1453–1463. doi: 10.1515/hsz-2013-0194. [DOI] [PubMed] [Google Scholar]
- 200.Dingwall C, Ernberg I, Gait MJ, Green SM, Heaphy S, Karn J, Lowe AD, Singh M, Skinner MA. HIV-1 Tat protein stimulates transcription by binding to a U-rich bulge in the stem of the Tar RNA structure. EMBO J. 1990;9:4145–4153. doi: 10.1002/j.1460-2075.1990.tb07637.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 201.Olsen GL, Edwards TE, Deka P, Varani G, Sigurdsson ST, Drobny GP. Monitoring tat peptide binding to TAR RNA by solid-state P-31-F-19 REDOR NMR. Nucleic Acids Res. 2005;33:3447–3454. doi: 10.1093/nar/gki626. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202.Huang W, Varani G, Drobny GP. C-13/N-15-F-19 intermolecular REDOR NMR study of the interaction of TAR RNA with Tat peptides. J Am Chem Soc. 2010;132:17643–17645. doi: 10.1021/ja1051439. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203.Olsen GL, Echodu DC, Shajani Z, Bardaro MF, Varani G, Drobny GP. Solid-state deuterium NMR studies reveal ms-ns motions in the HIV-1 Transactivation response RNA recognition site. J Am Chem Soc. 2008;130:2896–2897. doi: 10.1021/ja0778803. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204.Olsen GL, Bardaro MF, Jr, Echodu DC, Drobny GP, Varani G. Intermediate rate atomic trajectories of RNA by solid-state NMR spectroscopy. J Am Chem Soc. 2010;132:303–308. doi: 10.1021/ja907515s. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 205.Olsen GL, Bardaro MF, Jr, Echodu DC, Drobny GP, Varani G. Hydration dependent dynamics in RNA. J Biomol NMR. 2009;45:133–142. doi: 10.1007/s10858-009-9355-6. [DOI] [PubMed] [Google Scholar]
- 206.Marvin DA. Filamentous phage structure, infection and assembly. Curr Opin Struct Biol. 1998;8:150–158. doi: 10.1016/s0959-440x(98)80032-8. [DOI] [PubMed] [Google Scholar]
- 207.Ackermann HW. Tailed bacteriophages: The order Caudovirales. Advances in Virus Research. 1999;51:135–201. doi: 10.1016/S0065-3527(08)60785-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 208.Henry M, Debarbieux L. Tools from viruses: Bacteriophage successes and beyond. Virology. 2012;434:151–161. doi: 10.1016/j.virol.2012.09.017. [DOI] [PubMed] [Google Scholar]
- 209.Lee SW, Mao CB, Flynn CE, Belcher AM. Ordering of quantum dots using genetically engineered viruses. Science. 2002;296:892–895. doi: 10.1126/science.1068054. [DOI] [PubMed] [Google Scholar]
- 210.Nam KT, Kim DW, Yoo PJ, Chiang CY, Meethong N, Hammond PT, Chiang YM, Belcher AM. Virus-enabled synthesis and assembly of nanowires for lithium ion battery electrodes. Science. 2006;312:885–888. doi: 10.1126/science.1122716. [DOI] [PubMed] [Google Scholar]
- 211.Yu SM. Piezoelectric devices squeezed virus produces electricity. Nature Nanotechnology. 2012;7:343–344. doi: 10.1038/nnano.2012.85. [DOI] [PubMed] [Google Scholar]
- 212.Mao CB, Solis DJ, Reiss BD, Kottmann ST, Sweeney RY, Hayhurst A, Georgiou G, Iverson B, Belcher AM. Virus-based toolkit for the directed synthesis of magnetic and semiconducting nanowires. Science. 2004;303:213–217. doi: 10.1126/science.1092740. [DOI] [PubMed] [Google Scholar]
- 213.Nicastro J, Sheldon K, Slavcev RA. Bacteriophage lambda display systems: developments and applications. Appl Microbiol Biotechnol. 2014;98:2853–2866. doi: 10.1007/s00253-014-5521-1. [DOI] [PubMed] [Google Scholar]
- 214.Barry MA, Dower WJ, Johnston SA. Toward cell-targeting gene therapy vectors: Selection of cell-binding peptides from random peptide-presenting phage libraries. Nat Med. 1996;2:299–305. doi: 10.1038/nm0396-299. [DOI] [PubMed] [Google Scholar]
- 215.McCafferty J, Griffiths AD, Winter G, Chiswell DJ. Phage antibodies - filamentous phage displaying antibody variable domains. Nature. 1990;348:552–554. doi: 10.1038/348552a0. [DOI] [PubMed] [Google Scholar]
- 216.Pierce JC, Sternberg NL. Using bacteriophage-P1 system to clone high-molecular-weight genomic DNA. Methods Enzymol. 1992;216:549–574. doi: 10.1016/0076-6879(92)16049-p. [DOI] [PubMed] [Google Scholar]
- 217.Sanger F, Coulson AR, Barrell BG, Smith AJH, Roe BA. Cloning in single-stranded bacteriophage as an aid to rapid DNA sequencing. J Mol Biol. 1980;143:161–178. doi: 10.1016/0022-2836(80)90196-5. [DOI] [PubMed] [Google Scholar]
- 218.Salas M. Protein priming of DNA replication. Annu Rev Biochem. 1991;60:39–71. doi: 10.1146/annurev.bi.60.070191.000351. [DOI] [PubMed] [Google Scholar]
- 219.Hansen MR, Mueller L, Pardi A. Tunable alignment of macromolecules by filamentous phage yields dipolar coupling interactions. Nat Struct Biol. 1998;5:1065–1074. doi: 10.1038/4176. [DOI] [PubMed] [Google Scholar]
- 220.Clore GM, Starich MR, Gronenborn AM. Measurement of residual dipolar couplings of macromolecules aligned in the nematic phase of a colloidal suspension of rod-shaped viruses. J Am Chem Soc. 1998;120:10571–10572. [Google Scholar]
- 221.Lorieau JL, Day LA, McDermott AE. Conformational dynamics of an intact virus: Order parameters for the coat protein of Pf1 bacteriophage. Proc Natl Acad Sci USA. 2008;105:10366–10371. doi: 10.1073/pnas.0800405105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 222.Welsh LC, Symmons MF, Marvin DA. The molecular structure and structural transition of the alpha-helical capsid in filamentous bacteriophage Pf1. Acta Crystallographica Section D-Biological Crystallography. 2000;56:137–150. doi: 10.1107/s0907444999015334. [DOI] [PubMed] [Google Scholar]
- 223.Wishart DS, Case DA. Use of chemical shifts in macromolecular structure determination. In: James T, Dotsch V, Schmitz U, editors. Nuclear Magnetic Resonance of Biological Macromolecules, Pt A. 2001. pp. 3–34. [DOI] [PubMed] [Google Scholar]
- 224.Hinz HJ, Greulich KO, Ludwig H, Marvin DA. Calorimetric, density and circular-dichroism studies of the reversible structural transition in Pf1 filamentous bacterial-virus. J Mol Biol. 1980;144:281–289. doi: 10.1016/0022-2836(80)90091-1. [DOI] [PubMed] [Google Scholar]
- 225.Lankhorst PP, Erkelens C, Haasnoot CAG, Altona C. C-13 NMR in conformational analysis of nucleic acid fragments - heteronuclear chemical shift correlation spectroscopy of RNA constitutents. Nucleic Acids Res. 1983;11:7215–7230. doi: 10.1093/nar/11.20.7215. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 226.Santos RA, Tang P, Harbison GS. Determination of the DNA sugar pucker using C-13 NMR spectroscopy. Biochemistry. 1989;28:9372–9378. doi: 10.1021/bi00450a018. [DOI] [PubMed] [Google Scholar]
- 227.Liu DJ, Day LA. Pf1 virus structure - helical coat protein and DNA with paraxial phosphates. Science. 1994;265:671–674. doi: 10.1126/science.8036516. [DOI] [PubMed] [Google Scholar]
- 228.Purusottam RN, Rai RK, Sinha N. Mechanistic insights into water-protein interactions of filamentous bacteriophage. J Phys Chem B. 2013;117:2837–2840. doi: 10.1021/jp310921n. [DOI] [PubMed] [Google Scholar]
- 229.Sergeyev IV, Bahri S, Day LA, McDermott AE. Pf1 bacteriophage hydration by magic angle spinning solid-state NMR. J Chem Phys. 2014;141 doi: 10.1063/1.4903230. [DOI] [PubMed] [Google Scholar]
- 230.Messing J, Gronenborn B, Mullerhill B, Hofschneider PH. Filamentous coliphage M13 as a cloning vehicle - Insertion of a HindII fragment of Lac regulatory region in M13 replicative form invitro. Proc Natl Acad Sci USA. 1977;74:3642–3646. doi: 10.1073/pnas.74.9.3642. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 231.Wang YA, Yu X, Overman S, Tsuboi M, Thomas GJ, Egelman EH. The structure of a filamentous bacteriophage. J Mol Biol. 2006;361:209–215. doi: 10.1016/j.jmb.2006.06.027. [DOI] [PubMed] [Google Scholar]
- 232.Welsh LC, Symmons MF, Nave C, Perham RN, Marseglia EA, Marvin DA. Evidence for tilted smectic liquid crystalline packing of fd Inovirus from X-ray fiber diffraction. Macromolecules. 1996;29:7075–7083. [Google Scholar]
- 233.Thomas GJ, Prescott B, Opella SJ, Day LA. Sugar pucker and phosphodiester conformations in viral genomes of filamentous bacteriophages: fd, If1, Ike, Pf1, Xf, and Pf3. Biochemistry. 1988;27:4350–4357. doi: 10.1021/bi00412a023. [DOI] [PubMed] [Google Scholar]
- 234.Wen ZQ, Overman SA, Thomas GJ. Structure and interactions of the single-stranded DNA genome of filamentous virus fd: Investigation by ultraviolet resonance Raman spectroscopy. Biochemistry. 1997;36:7810–7820. doi: 10.1021/bi970342q. [DOI] [PubMed] [Google Scholar]
- 235.Morag O, Sgourakis N, Baker D, Goldbourt A. The NMR-Rosetta capsid model of M13 bacteriophage reveals a quadrupled hydrophobic packing epitope. Proc Natl Acad Sci USA. 2015 doi: 10.1073/pnas.1415393112. Published ahead of print. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 236.Cross TA, Tsang P, Opella SJ. Comparison of protein and deoxyribonucleic-acid backbone structures in bacteriophage-Fd and bacteriophage-Pf1. Biochemistry. 1983;22:721–726. doi: 10.1021/bi00273a002. [DOI] [PubMed] [Google Scholar]
- 237.Tan WM, Jelinek R, Opella SJ, Malik P, Terry TD, Perham RN. Effects of temperature and Y21M mutation on conformational heterogeneity of the major coat protein (pVIII) of filamentous bacteriophage fd. J Mol Biol. 1999;286:787–796. doi: 10.1006/jmbi.1998.2517. [DOI] [PubMed] [Google Scholar]
- 238.Thiriot DS, Nevzorov AA, Opella SJ. Structural basis of the temperature transition of Pf1 bacteriophage. Protein Sci. 2005;14:1064–1070. doi: 10.1110/ps.041220305. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 239.Glaubitz C, Grobner G, Watts A. Structural and orientational information of the membrane embedded M13 coat protein by C-13-MAS NMR spectroscopy. Biochim Biophys Acta. 2000;1463:151–161. doi: 10.1016/s0005-2736(99)00195-9. [DOI] [PubMed] [Google Scholar]
- 240.Bechinger B. Structure and dynamics of the M13 coat signal sequence in membranes by multidimensional high-resolution and solid-state NMR spectroscopy. Proteins: Struct, Funct, Genet. 1997;27:481–492. doi: 10.1002/(sici)1097-0134(199704)27:4<481::aid-prot2>3.0.co;2-e. [DOI] [PubMed] [Google Scholar]
- 241.Black LW, Thomas JA. Condensed genome structure. Adv Exp Med Biol. 2012;726:469–487. doi: 10.1007/978-1-4614-0980-9_21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 242.Kovacs FA, Brenneman MT, Quine J, Cross TA. Structural characterization of the M2 transmembrane peptide backbone using solid state NMR. Biophys J. 1997;72:298. [Google Scholar]
- 243.Song ZY, Kovacs FA, Wang J, Denny JK, Shekar SC, Quine JR, Cross TA. Transmembrane domain of M2 protein from influenza A virus studied by solid-state N-15 polarization inversion spin exchange at magic angle NMR. Biophys J. 2000;79:767–775. doi: 10.1016/S0006-3495(00)76334-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 244.Hu J, Fu R, Nishimura K, Zhang L, Zhou HX, Busath DD, Vijayvergiya V, Cross TA. Histidines, heart of the hydrogen ion channel from influenza A virus: Toward an understanding of conductance and proton selectivity. Proc Natl Acad Sci USA. 2006;103:6865–6870. doi: 10.1073/pnas.0601944103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 245.Hu J, Asbury T, Achuthan S, Li CG, Bertram R, Quine JR, Fu RQ, Cross TA. Backbone structure of the amantadine-blocked trans-membrane domain M2 proton channel from influenza A virus. Biophys J. 2007;92:4335–4343. doi: 10.1529/biophysj.106.090183. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 246.Hu J, Fu RQ, Cross TA. The chemical and dynamical influence of the anti-viral drug amantadine on the M-2 proton channel transmembrane domain. Biophys J. 2007;93:276–283. doi: 10.1529/biophysj.106.102103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 247.Miao YM, Qin HJ, Fu RQ, Sharma M, Can TV, Hung I, Luca S, Gor’kov PL, Brey WW, Cross TA. M2 proton channel Sstructural validation from full-length protein samples in synthetic bilayers and E. coli membranes. Angew Chem Int Ed. 2012;51:8383–8386. doi: 10.1002/anie.201204666. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 248.Can TV, Sharma M, Hung I, Gor’kov PL, Brey WW, Cross TA. Magic angle spinning and oriented sample solid-state NMR structural restraints combine for influenza A M2 protein functional insights. J Am Chem Soc. 2012;134:9022–9025. doi: 10.1021/ja3004039. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 249.Luo WB, Mani R, Hong M. Side-chain conformation of the M2 transmembrane peptide proton channel of influenza a virus from (19)F solid-state NMR. J Phys Chem B. 2007;111:10825–10832. doi: 10.1021/jp073823k. [DOI] [PubMed] [Google Scholar]
- 250.Hong M, Mishanina TV, Cady SD. Accurate measurement of methyl C-13 chemical shifts by solid-state NMR for the determination of protein side chain conformation: The influenza A M2 transmembrane peptide as an example. J Am Chem Soc. 2009;131:7806–7816. doi: 10.1021/ja901550q. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 251.Williams JK, Zhang Y, Schmidt-Rohr K, Hong M. pH-Dependent conformation, dynamics, and aromatic interaction of the gating Tryptophan residue of the influenza M2 proton channel from solid-state NMR. Biophys J. 2013;104:1698–1708. doi: 10.1016/j.bpj.2013.02.054. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 252.Cady SD, Hong M. Amantadine-induced conformational and dynamical changes of the influenza M2 transmembrane proton channel. Proc Natl Acad Sci USA. 2008;105:1483–1488. doi: 10.1073/pnas.0711500105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 253.Cady SD, Mishanina TV, Hong M. Structure of amantadine-bound M2 transmembrane peptide of influenza A in lipid bilayers from magic-angle-spinning solid-state NMR: The role of Ser31 in amantadine binding. J Mol Biol. 2009;385:1127–1141. doi: 10.1016/j.jmb.2008.11.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 254.Cady SD, Wang J, Wu YB, DeGrado WF, Hong M. Specific binding of adamantane drugs and direction of their polar amines in the pore of the influenza M2 transmembrane domain in lipid bilayers and dodecylphosphocholine micelles determined by NMR spectroscopy. J Am Chem Soc. 2011;133:4274–4284. doi: 10.1021/ja102581n. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 255.Williams JK, Tietze D, Wang J, Wu YB, DeGrado WF, Hong M. Drug-induced conformational and dynamical changes of the S31N mutant of the influenza M2 proton channel investigated by solid-state NMR. J Am Chem Soc. 2013;135:9885–9897. doi: 10.1021/ja4041412. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 256.Hu FH, Luo WB, Hong M. Mechanisms of proton conduction and gating in influenza M2 proton channels from solid-state NMR. Science. 2010;330:505–508. doi: 10.1126/science.1191714. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 257.Hu FH, Schmidt-Rohr K, Hong M. NMR detection of pH-dependent Histidine-water proton exchange reveals the conduction mechanism of a transmembrane proton channel. J Am Chem Soc. 2012;134:3703–3713. doi: 10.1021/ja2081185. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 258.Liao SY, Fritzsching KJ, Hong M. Conformational analysis of the full-length M2 protein of the influenza A virus using solid-state NMR. Protein Sci. 2013;22:1623–1638. doi: 10.1002/pro.2368. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 259.Andreas LB, Eddy MT, Pielak RM, Chou J, Griffin RG. Magic angle spinning NMR investigation of influenza A M2(18–60): Support for an allosteric mechanism of inhibition. J Am Chem Soc. 2010;132:10958–10960. doi: 10.1021/ja101537p. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 260.Andreas LB, Eddy MT, Chou JJ, Griffin RG. Magic-angle-spinning NMR of the drug resistant S31N M2 proton transporter from influenza A. J Am Chem Soc. 2012;134:7215–7218. doi: 10.1021/ja3003606. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 261.Yao HW, Hong M. Membrane-dependent conformation, dynamics, and lipid interactions of the fusion peptide of the Paramyxovirus PIV5 from solid-state NMR. J Mol Biol. 2013;425:563–576. doi: 10.1016/j.jmb.2012.11.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 262.Yao HW, Hong M. Conformation and lipid interaction of the fusion peptide of the Paramyxovirus PIV5 in anionic and negative-curvature membranes from solid-state NMR. J Am Chem Soc. 2014;136:2611–2624. doi: 10.1021/ja4121956. [DOI] [PMC free article] [PubMed] [Google Scholar]








