Next Article in Journal
Fluorescent Magnetic Mesoporous Nanoprobes for Biotechnological Enhancement Procedures in Gene Therapy
Next Article in Special Issue
Artifacts’ Detection for MRI Non-Metallic Needles: Comparative Analysis for Artifact Evaluation Using K-Means and Manual Quantification
Previous Article in Journal
A Strong Magnetic Field Alters the Activity and Selectivity of the CO2RR by Restraining C–C Coupling
Previous Article in Special Issue
Slow Methyl Axes Motions in Perdeuterated Villin Headpiece Subdomain Probed by Cross-Correlated NMR Relaxation Measurements
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

NMR of Paramagnetic Proteins: 13C Derived Paramagnetic Relaxation Enhancements Are an Additional Source of Structural Information in Solution

1
Magnetic Resonance Center and Department of Chemistry, University of Florence, Via L. Sacconi 6, 50019 Sesto Fiorentino, Italy
2
Instituto de Tecnologia Química e Biológica António Xavier (ITQB-NOVA), Universidade Nova de Lisboa, Av. da República (EAN), 2780-157 Oeiras, Portugal
*
Authors to whom correspondence should be addressed.
Magnetochemistry 2023, 9(3), 66; https://doi.org/10.3390/magnetochemistry9030066
Submission received: 30 January 2023 / Revised: 17 February 2023 / Accepted: 22 February 2023 / Published: 26 February 2023
(This article belongs to the Special Issue NMR Spectroscopy and Imaging in Biological Chemistry and Medicine)

Abstract

:
In paramagnetic metalloproteins, longitudinal relaxation rates of 13C′ and 13Cα nuclei can be measured using 13C detected experiments and converted into electron spin-nuclear spin distance restraints, also known as Paramagnetic Relaxation Enhancement (PRE) restraints. 13C are less sensitive to paramagnetism than 1H nuclei, therefore, 13C based PREs constitute an additional, non-redundant, structural information. We will discuss the complementarity of 13C PRE restraints with 1H PRE restraints in the case of the High Potential Iron Sulfur Protein (HiPIP) PioC, for which the NMR structure of PioC has been already solved by a combination of classical and paramagnetism-based restraints. We will show here that 13C R1 values can be measured also at very short distances from the paramagnetic center and that the obtained set of 13C based restraints can be added to 1H PREs and to other classical and paramagnetism based NMR restraints to improve quality and quantity of the NMR information.

1. Introduction

Metalloproteins play crucial roles in catalysis, electron transfer, metal storage/transport. Many of them are available in two different oxidation states and are paramagnetic in at least one of them [1,2]. NMR can provide the structure in solution at atomic resolution of paramagnetic proteins [3]. Tailored NMR experiments, paramagnetism-based structural restraints, and ad-hoc structure calculation programs allowed obtaining solution structures of paramagnetic proteins with a resolution comparable, if not better, with that of diamagnetic analogues of similar size [4,5,6,7,8,9,10,11,12]. In this scenario, relaxation-based NMR restraints are a powerful class of dipole–dipole restraints, providing through-space connectivities between the NMR active nuclei and the metal center(s) [13,14,15,16,17,18,19,20,21,22,23]. We have shown that, when a sufficient number of Paramagnetic Relaxation Enhancements (PRE, hereafter) effects are measurable throughout the entire protein, PREs restrain the conformational space with an efficiency comparable to Nuclear Overhauser Effects (NOE), even though they originate a set of distances all involving a single point, i.e., the paramagnetic center [24]. In the case of the NMR structure of the small paramagnetic protein PioC [25], obtained with a combination of NOE and paramagnetism-based NMR restraints, we showed that the solution structure obtained with only Paramagnetic NMR restraints is essentially the same as the one obtained with the full set of NMR restraints [24]. However, the only PREs used for structure calculations were R1 and R2 relaxation rates of 1H spins obtained via 13C- and 15N- Heteronuclear Single Quantum Coherences (HSQC) type experiments [26,27]. In order to sample the backbone more accurately, the relaxation rates of 13C′ and 13Cα spins are expected to have two advantages: (i) together with HN and Hα, 13C′ and 13Cα would provide a set of relaxation rates capable to define the relative orientation of the backbone of each aminoacid with respect to the metal center (Figure 1); (ii) because paramagnetic relaxation depends on γ2 of the investigated nucleus, 13C are less sensitive to paramagnetism and, therefore, they are expected to provide complementary information with respect to the 1H based PREs: when paramagnetism is strong and 1H signals are not observable, 13C signals are still detectable [28], therefore 13C based PREs constitute an additional, non-redundant, structural information.
We will test this idea on the protein PioC, that we have already used to study the effects of PREs vs. classical restraints. The Iron-sulfur High Potential Iron Protein (HiPIP) PioC from Rhodopseudomonas palustris TIE-1 [25] contains a [Fe4S4]3+/2+ cluster, being stable in the reduced [Fe4S4]2+ form. The protein has only 54 amino acids and it is an excellent case to validate novel NMR approaches for paramagnetic molecules. Most of the protein is affected by paramagnetism, but the intensive study by NMR has provided an almost complete sequence-specific assignment and the blind sphere, i.e., the region of the protein where, in principle, no information can be obtained by NMR due to paramagnetic induced line broadening, has been essentially eliminated, opening the way to new approaches of investigation. The methods validated using PioC could then be applied in more challenging systems, in which the paramagnetic relaxation prevents the identification of signals in a large sphere around the metal center [29]. Here, 13C direct detection and 13C PREs may become a precious tool to refine the structure in the proximity of the metal center [30,31,32,33,34,35,36]. Towards this aim, here we present how 13C R1 relaxation rates can be measured using 13C detected experiments in highly paramagnetic systems, further discussing their complementarity and consistency with 1H PRE restraints. Finally, we will consider how the NMR structure of PioC can be refined when 13C based NMR restraints are added into structure calculations.

2. Materials and Methods

Sample preparation. PioC was expressed and purified, as previously reported [25]. Uniformly 15N, 13C labelled PioC was expressed and purified in the M9 minimal media with the addition of ammonium sulfate (15N2, 99%) and [U-13C6] D-glucose in the M9 minimal media when labelling was required. The detailed procedure has been previously reported [24,37].
R1 longitudinal relaxation rates of 13C′ spins. A Bruker Avance III 700 MHz spectrometer operating at 700.06 MHz 1H frequency, equipped with a 5 mm, cryogenically cooled, inverse detection probe head (TXI), was used to measure 13C′ R1 longitudinal relaxation rates. Two series of HNCO-T1 experiments were recorded with parameters summarized in Table 1. The series with 7 s of recycle delay was recorded using the following relaxation delays: 2 ms, 50 ms, 150 ms, 200 ms, 400 ms, 600 ms, 1 s, 1.5 s, 2 s, 3 s and 5 s. The series recorded with 2 s of recycle delay, the relaxation delays were: 2 ms, 6 ms, 10 ms, 16 ms, 24 ms, 50 ms, 70 ms, 100 ms, 140 ms, 220 ms, 350 ms, 360 ms, 500 ms and 700 ms. R1 of fast relaxing C′ spins were measured using a tailored IR-COCA-AP. Experiments were recorded at 700 MHz Bruker AVANCE NEO spectrometer, equipped with a 5 mm cryogenically cooled probe head optimized for 13C direct detection (CT-TXO), operating at 176.05 MHz 13C frequency. The 13C′-Cα evolution delay was set to 2.7 ms, the IPAP scheme for virtual homonuclear decoupling was abolished, acquisition and recycle delays were shortened (see Table 1) to increase the number of transients per experimental time. The inversion recovery delays were: 1 ms, 10 ms, 50 ms, 80 ms, 120 ms, 200 ms, 300 ms, 500 ms, 800 ms, 1.2 s and 2 s. Longitudinal relaxation rates of carbonyl spins of Proline preceding residues were measured using a CON-T1 experiment (Table 1) with the following delays for the inversion recovery of the magnetization: 2 ms, 50 ms, 100 ms, 150 ms, 200 ms, 400 ms, 600 ms, 1 s, 1.5 s, 2 s, 3.5 s and 5 s.
R1 longitudinal relaxation rates 13Cα spins. Cα nuclei longitudinal relaxation rates measurements were recorded at a 700 MHz NMR spectrometer, equipped with a probe head optimized for 13C direct detection experiments, as mentioned above. A series of experiments was collected using a IR-CACO-IPAP pulse sequence [38] in which the inversion recovery delays were: 5 ms, 100 ms, 220 ms, 340 ms, 460 ms, 700 ms, 1 s, 1.6 s, 2.4 s, 3.6 s, 4.6 s and 5 s. An IR-CACO-AP series was recorded to sample fast relaxing Cα spins. The experiment is optimized as described above (Table 1). The inversion recovery delays used were: 1 ms, 5 ms, 10 ms, 20 ms, 40 ms, 60 ms, 80 ms, 120 ms, 200 ms, 300 ms, 500 ms, 800 ms, 1.2 s and 2 s. All the experiments were recorded using waltz65 and garp4 decoupling scheme for 1H and 15N decoupling. Smoothed square shape for all gradients was used. Q5- and Q3-shaped pulses, with a duration of 300 and 231 μs respectively, [39] were used for 13C band-selective π/2 and π flip angle pulses.
Data analysis and assessment. All the spectra were processed with a squared cosine weighting functions on Topspin 4.0.8 software. Exponential decays of the inversion recovery experiments were evaluated using cross-peak intensities. The intensity of each signal was integrated using Computed Aided Resonance Assignment (CARA) software [40]. All relaxation data were analyzed using the Origin 2022 software (v9.9.5). Peak intensities were fitted with a three-parameter exponential decay model.
Structure calculation. Structure calculations were performed with the software CYANA 2.1 [41,42,43]. NOEs converted into upper distance limits, backbone dihedral angle constraints and 1H PREs restraints were derived from the previously calculated and deposited structure (PDB: 6XYV) [24]. In the final structure calculation, 2000 conformers are calculated using the standard simulated annealing schedule with 65,000 torsion angle dynamics steps per conformer. The 20 conformers with the lowest Target Function are analyzed and compared with the previous deposited structure that was calculated with the same procedure. Details about the FeS cluster design and insertion into the structure calculation were previously reported [24].

3. Results and Discussion

3.1. Measurements of 13C′ Longitudinal Relaxation Rates via 1H Detection Methods

13C′ longitudinal relaxation rates can be collected via an HNCO-type experiment [44]. The experiment gives rise to an HSQC-type spectrum in which the signal intensity of each Hi-Ni peak is modulated by the R1 relaxation rate of the preceding carbonyl (C′i-1). Proton magnetization is decoupled during the inversion recovery of 13C′ spins, while no inversion pulses are given on 13Cα and 15N. This sequence does not remove cross correlation effects between 13C′z with both 13Cα and 15N, which may be operative during the longitudinal recovery delay of 13C′ spins. In order to properly sample the decay of both fast and slow relaxing signals, the experiment was, indeed, repeated twice with two different set of recovery and recycle delays. Obtained results are summarized in Table 2.
In our hands, the HNCO-T1 experiment was hardly susceptible to experimental optimization with respect to fast relaxing signals. The in-phase carbonyl single quantum coherence is created via a refocused N-C′ INEPT step which requires, for the out-and-back pathway, about 100 ms. A refocused INEPT cannot be significantly shortened without losing the efficiency of the transfer and, consequently, the formation C′z signal is in phase. Accordingly, many R1 values cannot be properly measured due to efficient paramagnetic relaxation effective during the HNCO block.

3.2. Measurements of 13C′ Relaxation Rates via 13C Detection Methods

Paramagnetic relaxation depends on γ2 of the relaxing nucleus, therefore the use of “protonless” experiments in which 1H spins are decoupled throughout the entire sequence, offers the opportunity to design experiments that are more robust than 1H detected experiments to the loss of information due to paramagnetic relaxation. A CON-T1 experiment can be easily implemented [45] to obtain 13C R1 measurements by simply adding an inversion recovery filter prior to the conventional CON sequence. We have used this approach to measure R1 values of Proline preceding residues, which were obviously missing in the HNCO-T1 (Table 2). Vis-à-vis the optimization for paramagnetic relaxation, we have already shown that a paramagnetically tailored CON experiment allows one to obtain signals from residues closer to the metal than a standard CON [46,47]; however, when one aims at monitoring signals that do not “survive” the long coherence transfer periods, C′/N transfer should be replaced by a C′/Cα transfer. To obtain reliable R1 measurements in the presence of efficient paramagnetic relaxation effects, we propose the pulse sequence shown in Figure 2a. The experiment is essentially an IR-COCA-AP experiment: the inversion recovery building block for C′ spins will modulate the intensity of a COCA-AP experiment, in which the Cα is detected in antiphase mode to avoid paramagnetic relaxation during the IPAP block [48]. The spectrum is reported in Figure 3a. This is the shortest and simplest experiment, and the only coherence transfer step can be optimized depending on the relaxation properties of the C′yCαz antiphase magnetization, as shown in Figure 2c. With this experiment we obtained a new set of R1 C′ measurements, that also contained signals unobserved (or measured with a very low precision) in HNCO-T1 and CON-T1. In our hands we found that, for R1 rates larger than 4–5 s−1, the IR-COCA-AP experiment was the most reliable experiment for C′ R1 measurements.

3.3. Measurements of 13Cα Relaxation Rates via 13C Detection Methods

By swapping the frequencies of C′ and Cα, the IR-COCA-AP experiment becomes an IR-CACO-AP experiment, as shown in Figure 2b, where the peak intensity is modulated by 13Cα R1 values. The obtained spectrum is reported in Figure 3b. The evolution delay Δ of the experiment is optimized according to the relaxation properties of the CαyC′z coherence. The effect of paramagnetic relaxation on the coherence transfer function C′/Cα is described by the same function shown in Figure 2c, however the relaxation rates of C′yCαz and the CαyC′z are not the same. Many variants of the CACO experiments have been described in the literature [36]. The CACO-AP that we used here has been widely shown to be the most efficient experiment in the presence of strong paramagnetic relaxation effects [48]. As a proof of concept of the validity of the 13C detected approach, we decided to use here a standard IR-CACO-IPAP [38] to measure relaxation rates of “slow” relaxation 13Cα nuclei and the IR-CACO-AP to obtain 13Cα R1 values of fast relaxing signals. The combined approach provided R1 Cα values in the range 1.5–20 s−1, as summarized in Table 2.

3.4. Relaxation Measurements of 13C of Fe-Bound Cysteine Residues

The delocalization of unpaired electron spin density from each iron ion onto the ligands induces significant contact hyperfine shifts on 13C spins of Cys residues. As the main mechanism for this delocalization is the Fe-S- Cα- Cα-C′ sigma bond pathway, the hyperfine shift values of Cys 13C spins are expected to follow this order. For Cysteine residues bound to a [Fe4S4]2+ cluster we observe, on average, 60–80 ppm, for Cα, 30–35 ppm for Cα and negligible contributions for C′ [49,50]. In order to measure Cys 13C R1 rates, we set the Cα carrier at ca 86 ppm and repeated the experiments discussed above. The four Cα/C′ connectivities are clearly observed, as shown in Figure 3c,d. As expected, the R1 rates of Cysteine C′ and Cα signals, also reported in Table 2, experience the largest 13C PRE values throughout the protein backbone.

3.5. Conversion of Relaxation Rates into Distance Restraints

R1 values summarized in Table 2 can be factorized according to (1):
R obs = R dia + R para ,
where Robs is the experimentally measured longitudinal relaxation rate, Rpara is given by the contributions arising from the hyperfine interaction and Rdia accounts from all the other contributions to longitudinal relaxation. The ideal strategy to obtain Rpara contributions would be that of measuring Rdia by repeating the experiment on the same sample in a diamagnetic state. Unfortunately, in PioC both oxidation states of [Fe4S4]3+/2+ are paramagnetic, and the cluster cannot be removed without substantially the structure of the protein. Thus, to factorize the Rpara contributions, for each series of experiments, the average of Robs values of residues 5–7, that are far from the paramagnetic center, provided an average value for Rdia; that was used in Equation (1) to obtain a list of Rpara values. This is an approximation because several factors, such as internal mobility, chemical shift anisotropy relaxation and cross correlations, modulate the Rdia value of each aminoacid. In turn, Rpara arises from the sum of electron–nucleus dipole–dipole coupling and unpaired electron delocalization onto the investigated nucleus, the so-called Fermi Contact relaxation Equation (2):
R para = R DD + R FC ,
The first contribution has a r−6 dependence from the electron spin (S)—nuclear spin (I) distance, and therefore the contact contribution in principle needs to be factorized out [51]. However, The RFC contact contribution is significant only for Fe-bound Cysteine nuclei, which has been treated separately. All other Rpara were then directly converted into upper limit distances according to Equation (3):
U P L ME = K R par 1 6 + 1.4   Å ,
The K R para 1 6 term converts Rpara into a distance from an iron ion of the cluster, assuming the unpaired spin density in the cluster is localized on the iron ions and that, for each Rpara, the effect is fully due to the distance from the closest iron ion. To minimize errors due to these approximations, and in order not to be biased from any structural model, each nuclear spin is not restrained to a specific iron of the cluster but to the center of the mass of the cubane (ME), which is located at about 1.4 Å from the edges of the cubane where iron ions are located. The UPLME used in the structure calculation are also reported in Table 2. Concerning the eight restraints from 13Cα and 13C′ of cluster bound Cys residues, we used the same approach used for the other Rpara values, and then we empirically applied a tolerance to remove the consistent violations on these distances. We obtained structure with acceptable target functions by increasing the calculated distance of Cα and C′ by 15% and 5%, respectively.

3.6. Structure Calculations

13C derived PREs were included in the structure calculation to refine the NMR structure of PioC and to assess the efficiency of these constraints. The 27 C′ and Cα constraints were added to the structure restraints available [52,53,54] for this protein, that are summarized in Table 3. The structure obtained with this extended set of NMR restraints is shown in Figure 4, together with the structure obtained without 13C PREs. Within the uncertainty, the pairwise RMSD to the mean of the structure obtained with the new set of restraints is the same as the one obtained previously. The backbone RMSD between the mean structures of the two families is 0.43 Å. This compares with the backbone RMSD values of 0.46 ± 0.11 Å and 0.43 ± 0.10 Å, observed for the families with and without 13C PREs, thus indicating that the two structures are identical. Figure 5 shows that, with the exception of a small region between Cys25 and Cys 34, the per-residue RMSD between the two average structures is always lower than the sum of the RMSD of the two families [55]. This means that the two structures are distinguishable only for a small rearrangement in the residues 27–28, which correspond to the protein region following Cys25, while no significant variations are observed for the rest of the protein structure. The small increase in target function (Table 3) indicates that the new set of restraints is fully compatible with the previous one and contribute to finding a convergent energy minimum. Figure 6 shows the 13C PRE-based distance restraints obtained here vs. the ensemble of conformers with and without their use in the structure calculations. 13C PRE-based restraints are fully congruent and integrated into the full set of NMR restraints. The addition of 13C restraints results in a lower dispersion of the calculated distances, particularly relevant for residues located in the 5–9 Å sphere from the cluster. The plot also shows that Tyrosine 50 retains a high divergence, due to the fact that very few restraints are available for this residue, which the new set of 13C restraints was not able to reduce.

4. Conclusions

We have shown here that 13C PRE-based restraints provide a consistent set of NMR restraints that can be added to 1H PREs and to other classical and paramagnetism-based NMR restraints, to improve quality and quantity of the NMR information. The range of distances sampled by 13C PREs, within each paramagnetic center, is different from that of 1H PREs, therefore the consistency between the two sets of data it is not straightforward and needs to be verified and assessed. The synergistic effect of 13C and 1H PREs restraints, that we have demonstrated in PioC as a proof of concept, will be even more beneficial for systems in which the paramagnetic effects are stronger. In those cases, the combined use of 13C and 1H PREs restraints will then be mandatory to reduce the blind sphere around the metal and improve calculation precision and accuracy. It is well known that the redundancy of NOEs is important to obtain accurate NMR solution structures [56,57,58]; this study shows that this also holds for PREs. This widens up the interest for measuring an increased number of relaxation rates. Small metalloproteins like PioC represent a paradigmatic case to assess the effect of PREs. This can be extremely useful for the NMR characterization of native metalloproteins where the metal center cannot be substituted, denovo designed metalloproteins [59], metalloproteins in which the metal center can be replaced with a paramagnetic probe [60,61], paramagnetic proteins of larger size and complexity [62,63,64], and in diamagnetic proteins where the use of metal tags may contribute to elucidate protein–protein interactions [16,65]. Finally, it is worth mentioning that 13C detected NMR is also a robust approach to study high molecular weight systems such as large size proteins, protein–protein complexes and antibodies [66].

Author Contributions

Conceptualization: F.C., R.O.L. and M.P.; Investigation: L.Q., I.B.T., J.M.S. and M.I.; Writing—Original Draft preparation: L.Q., J.M.S. and M.P.; Writing—Review and Editing: L.Q., I.B.T., J.M.S., M.I., F.C., R.O.L. and M.P.; Supervision: F.C., R.O.L. and M.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

This project was supported from the PRIN2020-FAITH (Post-doctoral Fellowship to JMS). This work was funded by National funds through FCT-Fundaҫão para a Ciência e a Tecnologia, I. P. (FCT), project MOSTMICRO-ITQB with refs UIDB/04612/2020 and UIDP/04612/2020, and LS4FUTURE Associated Laboratory (LA/P/0087/2020). This work benefited from access to CERM/CIRMMP, the Instruct-ERIC Italy centre and from COST Action CA21115, supported by COST (European Cooperation in Science and Technology).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Banci, L.; Camponeschi, F.; Ciofi-Baffoni, S.; Piccioli, M. The NMR contribution to protein-protein networking in Fe-S protein maturation. J. Biol. Inorg. Chem. 2018, 23, 665–685. [Google Scholar] [CrossRef] [Green Version]
  2. Piccioli, M.; Turano, P. Transient iron coordination sites in proteins: Exploiting the dual nature of paramagnetic NMR. Coord. Chem. Rev. 2015, 284, 313–328. [Google Scholar] [CrossRef]
  3. Piccioli, M. Paramagnetic NMR Spectroscopy Is a Tool to Address Reactivity, Structure, and Protein-Protein Interactions of Metalloproteins: The Case of Iron-Sulfur Proteins. Magnetochemistry 2020, 6, 46. [Google Scholar] [CrossRef]
  4. Trindade, I.B.; Coelho, A.; Cantini, F.; Piccioli, M.; Louro, R.O. NMR of paramagnetic metalloproteins in solution: Ubi venire, quo vadis? J. Inorg. Biochem. 2022, 234, 111871. [Google Scholar] [CrossRef]
  5. Clore, G.M. Practical Aspects of Paramagnetic Relaxation Enhancement in Biological Macromolecules. Methods Enzymol. 2015, 564, 485–497. [Google Scholar]
  6. Hus, J.C.; Marion, D.; Blackledge, M. De novo determination of protein structure by NMR using orientational and long-range order restraints. J. Mol. Biol. 2000, 298, 927–936. [Google Scholar] [CrossRef] [Green Version]
  7. Ramelot, T.A.; Cort, J.R.; Goldsmith-Fischman, S.; Kornhaber, G.J.; Xiao, R.; Shastry, R.; Acton, T.B.; Honig, B.; Montelione, G.T.; Kennedy, M.A. Solution NMR structure of the iron-sulfur cluster assembly protein U (IscU) with zinc bound at the active site. J. Mol. Biol. 2004, 344, 567–583. [Google Scholar] [CrossRef] [PubMed]
  8. Parigi, G.; Ravera, E.; Luchinat, C. Paramagnetic effects in NMR for protein structures and ensembles: Studies of metalloproteins. Curr. Opin. Struct. Biol. 2022, 74, 102386. [Google Scholar] [CrossRef]
  9. Swartjes, A.; White, P.B.; Bruekers, J.P.J.; Elemans, J.A.A.W.; Nolte, R.J.M. Paramagnetic relaxation enhancement NMR as a tool to probe guest binding and exchange in metallohosts. Nat. Commun. 2022, 13, 1846. [Google Scholar] [CrossRef]
  10. Karschin, N.; Becker, S.; Griesinger, C. Interdomain Dynamics via Paramagnetic NMR on the Highly Flexible Complex Calmodulin/Munc13-1. J. Am. Chem. Soc. 2022, 144, 17041–17053. [Google Scholar] [CrossRef]
  11. Pell, A.J.; Pintacuda, G.; Grey, C.P. Paramagnetic NMR in solution and the solid state. Prog. Nucl. Magn. Reson. Spectrosc. 2019, 111, 1–271. [Google Scholar] [CrossRef]
  12. Pan, B.B.; Yang, F.; Ye, Y.; Wu, Q.; Li, C.; Huber, T.; Su, X.C. 3D structure determination of a protein in living cells using paramagnetic NMR spectroscopy. Chem. Commun. 2016, 52, 10237–10240. [Google Scholar] [CrossRef] [Green Version]
  13. Donaldson, L.W.; Skrynnikov, N.R.; Choy, W.Y.; Muhandiram, D.R.; Sarkar, B.; Forman-Kay, J.D.; Kay, L.E. Structural characterization of proteins with an attached ATCUN motif by paramagnetic relaxation enhancement NMR spectroscopy. J. Am. Chem. Soc. 2001, 123, 9843–9847. [Google Scholar] [CrossRef]
  14. Rumpel, S.; Ravera, E.; Sommer, C.; Reijerse, E.J.; Fares, C.; Luchinat, C.; Lubitz, W. 1H NMR spectroscopy of (FeFe) hydrogenase: Insight into the electronic structure of the active site. J. Am. Chem. Soc. 2018, 140, 131–134. [Google Scholar] [CrossRef] [Green Version]
  15. Joss, D.; Haussinger, D. Design and applications of lanthanide chelating tags for pseudocontact shift NMR spectroscopy with biomacromolecules. Prog. Nucl. Magn. Reson. Spectrosc. 2019, 114–115, 284–312. [Google Scholar] [CrossRef]
  16. Miao, Q.; Nitsche, C.; Orton, H.; Overhand, M.; Otting, G.; Ubbink, M. Paramagnetic Chemical Probes for Studying Biological Macromolecules. Chem. Rev. 2022, 122, 9571–9642. [Google Scholar] [CrossRef] [PubMed]
  17. Orton, H.W.; Otting, G. Accurate Electron–Nucleus Distances from Paramagnetic Relaxation Enhancements. J. Am. Chem. Soc. 2018, 140, 7688–7697. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Toyama, Y.; Rangadurai, A.K.; Kay, L.E. Measurement of 1Hα transverse relaxation rates in proteins: Application to solvent PREs. J. Biomol. NMR 2022, 76, 137–152. [Google Scholar] [CrossRef] [PubMed]
  19. Lenard, A.J.; Mulder, F.A.A.; Madl, T. Solvent paramagnetic relaxation enhancement as a versatile method for studying structure and dynamics of biomolecular systems. Prog. Nucl. Magn. Reson. Spectrosc. 2022, 132–133, 113–139. [Google Scholar] [CrossRef]
  20. Ravera, E.; Gigli, L.; Fiorucci, L.; Luchinat, C.; Parigi, G. The evolution of paramagnetic NMR as a tool in structural biology. Phys. Chem. Chem. Phys. 2022, 24, 17397–17416. [Google Scholar] [CrossRef]
  21. Parker, D.; Suturina, E.A.; Kuprov, I.; Chilton, N.F. How the Ligand Field in Lanthanide Coordination Complexes Determines Magnetic Susceptibility Anisotropy, Paramagnetic NMR Shift, and Relaxation Behavior. Acc. Chem. Res. 2020, 53, 1520–1534. [Google Scholar] [CrossRef]
  22. Vogel, R.; Müntener, T.; Häussinger, D. Intrinsic anisotropy parameters of a series of lanthanoid complexes deliver new insights into the structure-magnetism relationship. Chem 2021, 7, 3144–3156. [Google Scholar] [CrossRef]
  23. Abelein, A.; Ciofi-Baffoni, S.; Mörman, C.; Kumar, R.; Giachetti, A.; Piccioli, M.; Biverstål, H. Molecular Structure of Cu(II)-Bound Amyloid-β Monomer Implicated in Inhibition of Peptide Self-Assembly in Alzheimer’s Disease. JACS Au 2022, 2, 2571–2584. [Google Scholar] [CrossRef]
  24. Trindade, I.B.; Invernici, M.; Cantini, F.; Louro, R.O.; Piccioli, M. PRE-driven protein NMR structures: An alternative approach in highly paramagnetic systems. FEBS J. 2021, 288, 3010–3023. [Google Scholar] [CrossRef] [PubMed]
  25. Bird, L.J.; Saraiva, I.H.; Park, S.; Calcada, E.O.; Salgueiro, C.A.; Nitschke, W.; Louro, R.O.; Newman, D.K. Nonredundant roles for cytochrome c2 and two high-potential iron-sulfur proteins in the photoferrotroph Rhodopseudomonas palustris TIE-1. J. Bacteriol. 2014, 196, 850–858. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Invernici, M.; Trindade, I.B.; Cantini, F.; Louro, R.O.; Piccioli, M. Measuring transverse relaxation in highly paramagnetic systems. J. Biomol. NMR 2020, 74, 431–442. [Google Scholar] [CrossRef]
  27. Ciofi-Baffoni, S.; Gallo, A.; Muzzioli, R.; Piccioli, M. The IR-15N-HSQC-AP experiment: A new tool for NMR spectroscopy of paramagnetic molecules. J. Biomol. NMR 2014, 58, 123–128. [Google Scholar] [CrossRef]
  28. Arnesano, F.; Banci, L.; Piccioli, M. NMR structures of paramagnetic metalloproteins. Q. Rev. Biophys. 2005, 38, 167–219. [Google Scholar] [CrossRef] [PubMed]
  29. Trindade, I.B.; Hernandez, G.; Lebegue, E.; Barriere, F.; Cordeiro, T.; Piccioli, M.; Louro, R.O. Conjuring up a ghost: Structural and functional characterization of FhuF, a ferric siderophore reductase from E. coli. J. Biol. Inorg. Chem. 2021, 26, 313–326. [Google Scholar] [CrossRef]
  30. Arnesano, F.; Banci, L.; Bertini, I.; Felli, I.C.; Luchinat, C.; Thompsett, A.R. A strategy for the NMR characterization of type II copper(II) proteins: The case of the copper trafficking protein CopC from Pseudomonas syringae. J. Am. Chem. Soc. 2003, 125, 7200–7208. [Google Scholar] [CrossRef]
  31. Piccioli, M.; Poggi, L. Tailored HCCH-TOCSY experiment for resonance assignment in the proximity of a paramagnetic center. J. Magn. Reson. 2002, 155, 236–243. [Google Scholar] [CrossRef] [PubMed]
  32. Alik, A.; Bouguechtouli, C.; Julien, M.; Bermel, W.; Ghouil, R.; Zinn-Justin, S.; Theillet, F.X. Sensitivity-Enhanced 13 C-NMR Spectroscopy for Monitoring Multisite Phosphorylation at Physiological Temperature and pH. Angew. Chem. Int. Ed. 2020, 59, 10411–10415. [Google Scholar] [CrossRef]
  33. Cerofolini, L.; Staderini, T.; Giuntini, S.; Ravera, E.; Fragai, M.; Parigi, G.; Pierattelli, R.; Luchinat, C. Long-range paramagnetic NMR data can provide a closer look on metal coordination in metalloproteins. J. Biol. Inorg. Chem. 2018, 23, 71–80. [Google Scholar] [CrossRef] [Green Version]
  34. Bermel, W.; Bertini, I.; Felli, I.; Piccioli, M.; Pierattelli, R. 13C-detected protonless NMR spectroscopy of proteins in solution. Prog. Nucl. Magn. Reson. Spectrosc. 2006, 48, 25–45. [Google Scholar] [CrossRef]
  35. Caillet-Saguy, C.; Delepierre, M.; Lecroisey, A.; Bertini, I.; Piccioli, M.; Turano, P. Direct detected 13C NMR to investigate the Iron(III) hemophore HasA. J. Am. Chem. Soc. 2006, 128, 150–158. [Google Scholar] [CrossRef]
  36. Felli, I.C.; Pierattelli, R. 13C Direct Detected NMR for Challenging Systems. Chem. Rev. 2022, 122, 9468–9496. [Google Scholar] [CrossRef]
  37. Trindade, I.B.; Invernici, M.; Cantini, F.; Louro, R.O.; Piccioli, M. 1H, 13C and 15N assignment of the paramagnetic high potential iron–sulfur protein (HiPIP) PioC from Rhodopseudomonas palustris TIE-1. Biomol. NMR Assign. 2020, 14, 211–215. [Google Scholar] [CrossRef]
  38. Pontoriero, L.; Schiavina, M.; Murrali, M.G.; Pierattelli, R.; Felli, I.C. Monitoring the Interaction of α-Synuclein with Calcium Ions through Exclusively Heteronuclear Nuclear Magnetic Resonance Experiments. Angew. Chem. Int. Ed. 2020, 59, 18537–18545. [Google Scholar] [CrossRef]
  39. Emsley, L.; Bodenhausen, G. Optimization of shaped selective pulses for NMR using a quaternion description of their overall propagators. J. Magn. Reason. 1992, 97, 135–148. [Google Scholar] [CrossRef]
  40. Keller, R. The Computer Aided Resonance Assignment Tutorial; CANTINA Verlag: Goldau, Switzerland, 2004; 81p. [Google Scholar]
  41. Güntert, P. Automated NMR structure calculation with CYANA. Methods Mol. Biol. 2004, 278, 353–378. [Google Scholar]
  42. Guntert, P.; Buchner, L. Combined automated NOE assignment and structure calculation with CYANA. J. Biomol. NMR 2015, 62, 453–471. [Google Scholar] [CrossRef]
  43. Dutta, S.K.; Serrano, P.; Proudfoot, A.; Geralt, M.; Pedrini, B.; Herrmann, T.; Wthrich, K. APSY-NMR for protein backbone assignment in high-throughput structural biology. J. Biomol. NMR 2015, 61, 47–53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Zeng, L.; Fischer, M.W.F.; Zuiderweg, E.R.P. Study of protein dynamics in solution by measurement of 13Cα-13CO NOE and 13CO longitudinal relaxation. J. Biomol. NMR 1996, 7, 157–162. [Google Scholar] [CrossRef] [Green Version]
  45. Bermel, W.; Bertini, I.; Felli, I.C.; Peruzzini, R.; Pierattelli, R. Exclusively heteronuclear NMR experiments to obtain structural and dynamic information on proteins. ChemPhysChem 2010, 11, 689–695. [Google Scholar] [CrossRef]
  46. Camponeschi, F.; Gallo, A.; Piccioli, M.; Banci, L. The long-standing relationship between Paramagnetic NMR and Iron-Sulfur proteins: The mitoNEET example. An old method for new stories or the other way around? Magn. Reson. Discuss. 2021, 2, 203–211. [Google Scholar] [CrossRef]
  47. Grifagni, D.; Silva, J.M.; Cantini, F.; Piccioli, M.; Banci, L. Relaxation-based NMR assignment: Spotlights on ligand binding sites in human CISD3. J. Inorg. Biochem. 2023, 239, 112089. [Google Scholar] [CrossRef]
  48. Bertini, I.; Jimenez, B.; Piccioli, M. 13C direct detected experiments: Optimization for paramagnetic signals. J. Magn. Reson. 2005, 174, 125–132. [Google Scholar] [CrossRef]
  49. Trindade, I.B.; Invernici, M.; Cantini, F.; Louro, R.O.; Piccioli, M. Sequence-specific assignments in NMR spectra of paramagnetic systems: A non-systematic approach. Inorg. Chim. Acta 2021, 514, 119984. [Google Scholar] [CrossRef]
  50. Bertini, I.; Capozzi, F.; Luchinat, C.; Piccioli, M.; Vicens Oliver, M. NMR is a unique and necessary step in the investigation of iron- sulfur proteins: The HiPIP from R. gelatinosus as an example. Inorg. Chim. Acta 1992, 198–200, 483–491. [Google Scholar] [CrossRef]
  51. Bertini, I.; Luchinat, C.; Parigi, G.; Ravera, E. NMR of Paramagnetic Molecules; Elsevier: Amsterdam, The Netherlands, 2017; 508p. [Google Scholar]
  52. Banci, L.; Bertini, I.; Luchinat, C.; Piccioli, M.; Scozzafava, A.; Turano, P. 1H NOE studies on dicopper(II) dicobalt(II) superoxide dismutase. Inorg. Chem. 1989, 28, 4650–4656. [Google Scholar] [CrossRef]
  53. Spronk, C.; Zerko, S.; Gorka, M.; Kozminski, W.; Bardiaux, B.; Zambelli, B.; Musiani, F.; Piccioli, M.; Basak, P.; Blum, F.C.; et al. Structure and dynamics of Helicobacter pylori nickel-chaperone HypA: An integrated approach using NMR spectroscopy, functional assays and computational tools. J. Biol. Inorg. Chem. 2018, 23, 1309–1330. [Google Scholar] [CrossRef]
  54. Bertini, I.; Capozzi, F.; Luchinat, C.; Piccioli, M.; Vila, A.J. The Fe4S4 centers in ferredoxins studied through proton and carbon hyperfine coupling. Sequence specific assignments of cysteines in ferredoxins from Clostridium acidi urici and Clostridium pasteurianum. J. Am. Chem. Soc. 1994, 116, 651–660. [Google Scholar] [CrossRef]
  55. Banci, L.; Bertini, I.; Cantini, F.; Della Malva, N.; Rosato, A.; Herrmann, T.; Wüthrich, K. Solution structure and intermolecular interactions of the third metal-binding domain of ATP7A, the Menkes disease protein. J. Biol. Chem. 2006, 281, 29141–29147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Sala, D.; Huang, Y.J.; Cole, C.A.; Snyder, D.A.; Liu, G.; Ishida, Y.; Swapna, G.V.T.; Brock, K.P.; Sander, C.; Fidelis, K.; et al. Protein structure prediction assisted with sparse NMR data in CASP13. Proteins 2019, 87, 1315–1332. [Google Scholar] [CrossRef] [Green Version]
  57. Duclert-Savatier, N.; Martínez, L.; Nilges, M.; Malliavin, T.E. The redundancy of NMR restraints can be used to accelerate the unfolding behavior of an SH3 domain during molecular dynamics simulations. BMC Struct. Biol. 2011, 11, 46. [Google Scholar] [CrossRef] [Green Version]
  58. Nabuurs, S.B.; Spronk, C.A.; Krieger, E.; Maassen, H.; Vriend, G.; Vuister, G.W. Quantitative evaluation of experimental NMR restraints. J. Am. Chem. Soc. 2003, 125, 12026–12034. [Google Scholar] [CrossRef]
  59. Kim, J.D.; Pike, D.H.; Tyryshkin, A.M.; Swapna, G.V.T.; Raanan, H.; Montelione, G.T.; Nanda, V.; Falkowski, P.G. Minimal Heterochiral de Novo Designed 4Fe-4S Binding Peptide Capable of Robust Electron Transfer. J. Am. Chem. Soc. 2018, 140, 11210–11213. [Google Scholar] [CrossRef] [PubMed]
  60. Balayssac, S.; Jiménez, B.; Piccioli, M. Assignment Strategy for Fast Relaxing Signals: Complete Aminoacid Identification in Thulium Substituted Calbindin D9k. J. Biomol. NMR 2006, 34, 63–73. [Google Scholar] [CrossRef]
  61. Invernici, M.; Selvolini, G.; Silva, J.M.; Marrazza, G.; Ciofi-Baffoni, S.; Piccioli, M. Interconversion between [2Fe-2S] and [4Fe-4S] cluster glutathione complexes. Chem. Commun. 2022, 58, 3533–3536. [Google Scholar] [CrossRef] [PubMed]
  62. Okuno, Y.; Schwieters, C.D.; Yang, Z.; Clore, G.M. Theory and Applications of Nitroxide-based Paramagnetic Cosolutes for Probing Intermolecular and Electrostatic Interactions on Protein Surfaces. J. Am. Chem. Soc. 2022, 144, 21371–21388. [Google Scholar] [CrossRef] [PubMed]
  63. Arthanari, H.; Takeuchi, K.; Dubey, A.; Wagner, G. Emerging solution NMR methods to illuminate the structural and dynamic properties of proteins. Curr. Opin. Struct. Biol. 2019, 58, 294–304. [Google Scholar] [CrossRef] [PubMed]
  64. Bax, A.; Clore, G.M. Protein NMR: Boundless opportunities. J. Magn. Reson. 2019, 306, 187–191. [Google Scholar] [CrossRef] [PubMed]
  65. Deshmukh, L.; Louis, J.M.; Ghirlando, R.; Clore, G.M. Transient HIV-1 Gag-protease interactions revealed by paramagnetic NMR suggest origins of compensatory drug resistance mutations. Proc. Natl. Acad. Sci. USA 2016, 113, 12456–12461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Arbogast, L.W.; Brinson, R.G.; Marino, J.P. Mapping Monoclonal Antibody Structure by 2D 13C NMR at Natural Abundance. Anal. Chem. 2015, 87, 3556–3561. [Google Scholar] [CrossRef]
Figure 1. Schematic drawing of protein backbone vs. a metal ion: the availability of a larger number of R1 values contributes to place the backbone with respect to the metal. The C′ and Cα distances from the iron ion are reported.
Figure 1. Schematic drawing of protein backbone vs. a metal ion: the availability of a larger number of R1 values contributes to place the backbone with respect to the metal. The C′ and Cα distances from the iron ion are reported.
Magnetochemistry 09 00066 g001
Figure 2. (a) IR-COCA-AP; (b) IR-CACO-AP Phases as follows: ϕ1: x,-x; ϕ2: x; ϕ3: x,x,x,x, y,y,y,y; ϕ4: y; ϕ5: x,x,-x,-x; ϕrec x,-x,-x,x, -x,x,x,-x. Semi-constant time evolution is given in the indirect dimension. 90° and 180° pulses are given as Q5 and Q3 band selective pulses, respectively. (c): calculated efficiency of the coherence transfer function for a C′/Cα coupling (53 Hz) versus the INEPT transfer delay under the effect of relaxation. (a: no relaxation; b: 10s−1, c: 50 s−1, d: 100 s−1, e: 200 s−1, f: 500 s−1). Black dashed line indicates the delay typically used in the experiment (4.7 ms), red dashed line indicates the delay that we have used throughout the experiments reported in this article (2.7 ms).
Figure 2. (a) IR-COCA-AP; (b) IR-CACO-AP Phases as follows: ϕ1: x,-x; ϕ2: x; ϕ3: x,x,x,x, y,y,y,y; ϕ4: y; ϕ5: x,x,-x,-x; ϕrec x,-x,-x,x, -x,x,x,-x. Semi-constant time evolution is given in the indirect dimension. 90° and 180° pulses are given as Q5 and Q3 band selective pulses, respectively. (c): calculated efficiency of the coherence transfer function for a C′/Cα coupling (53 Hz) versus the INEPT transfer delay under the effect of relaxation. (a: no relaxation; b: 10s−1, c: 50 s−1, d: 100 s−1, e: 200 s−1, f: 500 s−1). Black dashed line indicates the delay typically used in the experiment (4.7 ms), red dashed line indicates the delay that we have used throughout the experiments reported in this article (2.7 ms).
Magnetochemistry 09 00066 g002
Figure 3. (a) IR-COCA-AP spectra; (b) IR-CACO-AP (c) IR-COCA-AP spectra Cys region; (d) IR-CACO-AP Cys region. The reported spectra were recorded using parameters summarized in Table 1. For each R1 series, the experiment with the longest recovery delay is shown.
Figure 3. (a) IR-COCA-AP spectra; (b) IR-CACO-AP (c) IR-COCA-AP spectra Cys region; (d) IR-CACO-AP Cys region. The reported spectra were recorded using parameters summarized in Table 1. For each R1 series, the experiment with the longest recovery delay is shown.
Magnetochemistry 09 00066 g003
Figure 4. NMR structures of PioC obtained with full set of restraints (beige) and 13C PRE restrains less structure (cyan). Residues 5–49 are shown.
Figure 4. NMR structures of PioC obtained with full set of restraints (beige) and 13C PRE restrains less structure (cyan). Residues 5–49 are shown.
Magnetochemistry 09 00066 g004
Figure 5. Per-residue backbone RMSD (5–50). The RMSD between the two mean structures (white triangles) is compared with the sum of the pairwise backbone RMSD of each ensemble (black triangles).
Figure 5. Per-residue backbone RMSD (5–50). The RMSD between the two mean structures (white triangles) is compared with the sum of the pairwise backbone RMSD of each ensemble (black triangles).
Magnetochemistry 09 00066 g005
Figure 6. PRE-derived distances vs. solution structures distances. The structures have been calculated without the 13C PRE restraints (a) and using the full set of restraints (b). In the 5–9 Å distance range, the dispersion of calculated distances from the structures are significantly decreased. When 13C PREs are used to refine the structure (b) the calculated distances are restrained with a very few violations observed in the b. part of the figure (values above the line representing the upl values).
Figure 6. PRE-derived distances vs. solution structures distances. The structures have been calculated without the 13C PRE restraints (a) and using the full set of restraints (b). In the 5–9 Å distance range, the dispersion of calculated distances from the structures are significantly decreased. When 13C PREs are used to refine the structure (b) the calculated distances are restrained with a very few violations observed in the b. part of the figure (values above the line representing the upl values).
Magnetochemistry 09 00066 g006
Table 1. Experiments performed and relevant parameters.
Table 1. Experiments performed and relevant parameters.
ExperimentTime DomainNumber of ScanSpectral Width (ppm)Acquisition Time (ms)Recycle Delay (s)
F2F1F2F1F2F1
HNCO-T11024641613.7365312.57
HNCO-T1 1024966413.7505312.52
CON-T11024963240.65171.713.36.4
IR-COCA-AP10241284864.5254514.52
IR-COCA-AP *5129619240.64035.86.81
IR-CACO-IPAP512641631.6304665
IR-CACO-AP5121283231.6304612.12
IR-CACO-AP *5129622431.650465.40.5
* Series of experiments to measure Cysteines C′/Cα longitudinal relaxation rates.
Table 2. R1 measured for C′ and Cα nuclei and related upper limit restraint used in structure calculation.
Table 2. R1 measured for C′ and Cα nuclei and related upper limit restraint used in structure calculation.
Residues ExpR1 C′
(s−1)
ErrUpl
(C′-ME)
Residues ExpR1 Cα (s−1)ErrUpl
(Cα-ME)
VAL 1*1.190.1 VAL1~2.090.5
THR2 THR2
LYS3 LYS3
LYS4 LYS4
ALA5*1.060.1 ALA5~1.760.1
SER6*1.230.4 SER6~1.840.4
HIS7 HIS7~1.780.1
LYS8*1.140.2 LYS8
ASP9*1.060.1 ASP9~1.790.1
ALA10*1.550.18.02ALA10~2.310.2
GLY11*1.430.18.43GLY11~3.610.3
TYR 12*1.420.18.46TYR12~3.120.36.82
GLN 13*1.260.1 GLN13~2.370.2
GLU14*0.880 GLU14~1.790.1
SER15¬2.111.5 SER15~1.850.5
PRO16*1.050.1 PRO16~1.870.2
ASN 17*1.080.1 ASN17~2.040.2
GLY18*0.930.1 GLY18~3.250.4
ALA19*1.240.8 ALA19~1.570.1
LYS20 LYS20
ARG21 ARG21
CYS22##5.960.96.11CYS22> 18.321.75.72
GLY23*1.260.19.59GLY23~3.370.2
THR24**5.872.2 THR24~~2.190.4
CYS25##4.930.76.32CYS25> 14.11.95.93
ARG26 ARG26
GLN 27#7.6815.56GLN 27~~7.070.6
PHE 28*2.071.47.16PHE 28~~8.821.15.51
ARG 29 ARG 29
PRO 30¬1.471.2 PRO 30~1.650.2
PRO31*1.110.1 PRO31~1.410.1
SER 32*1.190.1 SER 32
SER 33**4.550.9 SER 33~1.950.39.17
CYS34##3.530.56.75CYS34> 13.440.85.97
ILE 35 ILE 35~2.120.58.26
THR 36 THR 36~~4.350.9
VAL 37**4.340.4 VAL 37~~4.930.4
GLU 38*1.280.19.35GLU 38~2.150.1
SER 39¬1.451.1 SER 39~2.320.5
PRO 40*1.180.1 PRO 40~1.830.2
ILE 41*1.630.27.82ILE 41~2.560.67.34
SER 42*1.860.1 SER 42~3.510.5
GLU43*1.550.18.04GLU43~2.280.2
ASN 44*1.560.18ASN 44~1.680.2
GLY45#4.560.66.04GLY45~~4.420.4
TRP 46#8.7915.45TRP 46~~9.581.75.44
CYS 47##8.541.55.78CYS 47> 18.584.37.71
ARG 48 ARG 48
LEU49 LEU49
TYR 50 TYR 50~3.071.46.86
ALA 51*2.030.1 ALA 51~2.940.2
GLY 52 GLY 52~3.220.3
LYS 53*1.310.5 LYS 53
ALA 54 ALA 54
Experiment are associated with each symbol as follows: * HNCO-T1 (Slow), ** HNCO-T1 (Fast), ¬ CON-T1, # IR-COCA-AP, ## IR-COCA-AP*, ~ IR-CACO-IPAP, ~~ IR-CACO-AP, > IR-CACO-AP*.
Table 3. The table summarizes the structural restraints utilized in ensemble calculations and the results derived from PSVS (v1.5) analysis performed on the obtained family structures.
Table 3. The table summarizes the structural restraints utilized in ensemble calculations and the results derived from PSVS (v1.5) analysis performed on the obtained family structures.
Full Set13C PRE-Less Restraints
Total number of meaningful NOE upper distance constraints:344344
Total meaningful dihedral angle restraints:5151
Total number of paramagnetic NMR restraints:216189
Residual CYANA Target Function:1.47 ± 0.061.07 ± 0.08
Pairwise RMSD to the mean:
Backbone RMSD$ (residues 5–50):0.46 ± 0.070.43 ± 0.10
All heavy at.RMSD$ (residues 5–50):1.03 ± * 0.231.14 ± 0.16
Ramachandran Plot Summary from Procheck *:
Most favored regions:63.70%65.50%
Additionally allowed regions:34.60%32.80%
Generously allowed regions:1.70%1.40%
Disallowed regions:0.00%0.30%
Average no. of distance viol/stru:
>0.5 Å:00
RMS of Distances violations per meaningful distance constraint (Å):0.02040.0146
* Selected residue ranges: 5–50. Calculated using PSVS 1.5 (Bhattacharya et al. 2007).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Querci, L.; Trindade, I.B.; Invernici, M.; Silva, J.M.; Cantini, F.; Louro, R.O.; Piccioli, M. NMR of Paramagnetic Proteins: 13C Derived Paramagnetic Relaxation Enhancements Are an Additional Source of Structural Information in Solution. Magnetochemistry 2023, 9, 66. https://doi.org/10.3390/magnetochemistry9030066

AMA Style

Querci L, Trindade IB, Invernici M, Silva JM, Cantini F, Louro RO, Piccioli M. NMR of Paramagnetic Proteins: 13C Derived Paramagnetic Relaxation Enhancements Are an Additional Source of Structural Information in Solution. Magnetochemistry. 2023; 9(3):66. https://doi.org/10.3390/magnetochemistry9030066

Chicago/Turabian Style

Querci, Leonardo, Inês B. Trindade, Michele Invernici, José Malanho Silva, Francesca Cantini, Ricardo O. Louro, and Mario Piccioli. 2023. "NMR of Paramagnetic Proteins: 13C Derived Paramagnetic Relaxation Enhancements Are an Additional Source of Structural Information in Solution" Magnetochemistry 9, no. 3: 66. https://doi.org/10.3390/magnetochemistry9030066

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop