Next Article in Journal
Using Imidazolium in the Construction of Hybrid 2D and 3D Lead Bromide Pseudoperovskites
Previous Article in Journal
Fe Doping Enhances the Peroxidase-Like Activity of CuO for Ascorbic Acid Sensing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multiple Intramolecular Hydrogen Bonding in Large Biomolecules: DFT Calculations and Deuterium Isotope Effects on 13C Chemical Shifts as a Tool in Structural Studies

by
Poul Erik Hansen
1,* and
Fadhil S. Kamounah
2
1
Department of Science and Environment, Roskilde University, DK-4000 Roskilde, Denmark
2
Department of Chemistry, University of Copenhagen, Universitetsparken 5, DK-2100 Copenhagen Ø, Denmark
*
Author to whom correspondence should be addressed.
Chemistry 2023, 5(2), 1317-1328; https://doi.org/10.3390/chemistry5020089
Submission received: 30 March 2023 / Revised: 15 May 2023 / Accepted: 17 May 2023 / Published: 23 May 2023
(This article belongs to the Section Theoretical and Computational Chemistry)

Abstract

:
Large biomolecules often have multiple intramolecular hydrogen bonds. In the cases where these interact, it requires special tools to disentangle the patterns. Such a tool could be deuterium isotope effects on chemical shifts. The use of theoretical calculations is an indispensable tool in such studies. The present paper illustrates how DFT calculations of chemical shifts and deuterium isotope effects on chemical shifts in combination with measurements of these effects can establish the complex intramolecular hydrogen bond patterns of rifampicin as an example) The structures were calculated using DFT theoretical calculations, performed with the Gaussian 16 software. The geometries were optimized using the B3LYP functional and the Pople basis set 6-31G(d) and the solvent (DMSO) was taken into account in the PCM approach. Besides the 6-31G(d) basis set, the 6-31 G(d,p) and the 6-3111G(d,p) basis sets were also tested. The nuclear shieldings were calculated using the GIAO approach. Deuteriation was simulated by shortening the X-H bond lengths by 0.01 Å.

Graphical Abstract

1. Introduction

Intramolecular hydrogen bonding is a very important structural parameter in many biomolecules, e.g., peptides, proteins, RNA, DNA and smaller molecules with specific biological effects such as rifampicin. [1]. Furthermore, the correct structure is a prerequisite for binding studies of molecules with biological activity [2] such as rifampicin. Rifampicin has important biological actions and is mentioned as a possible remedy against tuberculosis in combination with other drugs [3]. The introduction of deuterium in OH and NH groups is straightforward and is a minimal perturbation but still one that creates easily observable changes, e.g., in 13C chemical shifts. Therefore, deuterium isotope effects on chemical shifts have turned out to be a useful tool in establishing the presence and the characteristics of intramolecular hydrogen bonds [4]. Examples include the use of deuterium isotope effects on 13C chemical shifts as well as on 15N chemical shifts in proteins [5,6]. Such isotope effects also have advantages in studies of DNA [7,8]. Intramolecular hydrogen bonding may also lead to tautomerism. This can also be studied with deuterium isotope effects on chemical shifts [9]. The situation can be complex in the cases in which several donors and acceptors forming hydrogen bonds with each other. Such a situation is found in rifampicin [2,10]. The use of theoretical calculations to calculate structures and NMR parameters have reached a very high level [11]. DFT calculations of structures, chemical shifts and isotope effects on chemical shifts combined with experimental values is particularly useful in the present study to disentangle the hydrogen bonding pattern of such a complex system as rifampicin and to determine if tautomerism is at play. Rifampicin has shown the very interesting feature that it may take up a zwitterionic form in polar solvents (Figure 1) [8]. Density functional theory (DFT) calculations [12] are a very useful tool to analyze the complex situations both with regard to hydrogen bonding [13] and tautomerism [7]. Both nuclear shieldings and isotope effects on nuclear shieldings can be calculated [1]. Numerous examples of multiple hydrogen bonding involving biologically relevant structures can be found. Examples of calculations of structures and hydrogen bond energies including compounds with pyrroles were performed by the Afonin group [14,15,16].
The measurement of deuterium isotope effects on chemical shifts can be performed in two different ways. If exchange of the label is slow, the measurement can be performed in a one-tube NMR experiment containing both isotopomers. However, if the exchange is fast, e.g., with presence of water, one can use a variation of the percentage of deuterium in the solvent, typically H2O/D2O or CH3OH/CH3OD, or use these as co-solvents in DMSO-d6. Having performed experiments using 0 to 100% D2O, the isotope effects can be obtained as the difference between these two situations. For systems showing changes upon the addition of water, the solvent effect of adding water must be determined. It is of course important to realize that the measured isotope effects are the sum of all possible unresolved isotope effects.
In the case of a tautomeric equilibrium, deuteriation will lead to a change in the chemical equilibrium. This, in turn, will lead to a change in the chemical shifts. This type of isotope effect mainly depends on the chemical shift differences of the same nucleus in the two equilibrating species and is hence a very good monitor of the existence of a tautomeric equilibrium.
The use of a combination of DFT calculations, and measurements of chemical shifts and isotope effects on 13C chemical shifts is a general method that can be used in other biological systems, e.g., more complex derivatives of rifampicin such as the aldehyde [2].

2. Experimental

2.1. Calculations

The structures are calculated using DFT theoretical calculations [9] performed with the Gaussian 16 software [17]. The geometries were optimized using the B3LYP functional [18,19] and the Pople basis set 6-31G(d) [20], and the solvent (DMSO) was taken into account in the PCM [21,22] approach. X, Y, Z coordinates of structure A (see Section 3.4) are given in the Supplementary Materials. Besides the 6-31G(d) basis set, the 6-31 G(d,p) and the 6-3111G(d,p) basis sets were also tested. The nuclear shieldings were calculated using the GIAO approach [23,24]. Deuteriation was simulated by shortening the X-H bond lengths by 0.01 Å [3].

2.2. NMR

One-dimensional 1H and 13 C NMR spectra were recorded using a 300 MHz spectrometer (Bruker, Fallaenden, Germany) recorded at 300.08 MHz and 75.46 MHz, respectively in DMSO-d6 using TMS as a reference. Examples of the 1H and 13C NMR spectra are shown in Section 3.1 and Section 3.2.

2.3. Deuteriation

Deuteriation was achieved in a series of experiments by adding 2, 4 or 9 μL, and in some cases 20 or 30 μL, D2O to the DMSO-d6 solutions. The latter additions were used just to confirm that the effects at such high additions were linear. The solvent effects were estimated by the addition of 10 μL H2O (see Section 3.4).

3. Results

3.1. 1H NMR Spectrum

Both the OH-1 (15.6 ppm) and the NH+ (9.5 ppm) resonances were broad, indicating that they were in exchange. Those of OH-4 (12.5 ppm) and the NH amide (8.4 ppm) protons were sharp suggesting that these were not in exchange. An example of a 1H NMR spectrum of a partially deuteriated species is shown in Figure 2. The degree of deuteriation was determined from the integrals of the exchangeable protons.

3.2. 13C NMR Spectrum

The 13C NMR spectrum of rifampicin in CDCl3 is given in [9]. The assignments used in this paper are very similar except for C-2, C-9 and C-10. A number of the resonances showed splittings due to deuteriation (see Table 1). The signs of the deuterium isotope effects on 13C chemical shifts can be determined by knowing the degree of deuteriation. Other resonances showed a shift as a function of addition of D2O (Table 1) (see Figure 3).

3.3. Assignment of Isotope Effects on 13C Chemical Shifts

In a system with several protons, which can be deuteriated, in this case OH-1, OH-4, NH+, OH-21, OH-23 and the NH amide proton, it is important to be able to determine which deuterium is causing which effect. In the 1H spectrum, OH-1 and NH+ were broad, showing that they were easily exchanged and did not giving rise to observable deuterium isotope effects as doublets. This leaves the observable doublets as being due to OH-4, NHC=O, OH-21 and OH-23. The carbons of the latter two were in the aliphatic region and can be assigned. OH-21 and OH-23 were in a different part of the molecule and will not give rise to isotope effects on carbons of the aromatic rings. In an aliphatic system, the isotope effects are only transmitted over few bonds are thus very local. The effects due to NHC=O were local and restricted to C-2, C=O(NH) and C=C-16.

3.4. Isotope Effects

In a system like the present one, deuterium isotope effects on 13C chemical shifts can be of different kinds. If the deuterium is slowly exchanged or not exchanged at all, intrinsic isotope effects are observed as splittings of the 13C resonances. However, if the deuterium label is exchanged quickly, one observes only a change in the chemical shifts (see Table 2). Both these effects are intrinsic. Intrinsic isotope effects in aliphatic systems are only transmitted over a few bonds, whereas they can be transmitted over many bonds in conjugated systems like aromatic systems. [3] Intrinsic isotope effects are defined as:
nΔC(D) = δC(H) − δC(D)
where n is the number of bonds between the label and the carbon in question.
However, if the system is tautomeric between two species, A and B, equilibrium isotope effects are observed. These are defined as:
nΔX(D)int = (1 − x) nΔX(D)A + x nΔX(D)b
nΔX(D)eq = (δXB − δXA) Δx
nΔX(D)OBS = nΔX(D)int + nΔX(D)eq
In this case, X is 13C and x is the mole fraction of B. Δx is the change in the equilibrium upon deuteration. The interesting part is that the equilibrium isotope effect depends on the chemical shift difference between the two sites (δXB − δXA).
The isotope effects on 13C chemical shifts that are observed as doublets or doublets of doublets can be ascribed to deuteriation either at H-4, the NH or at OH-21 or OH-23 (see assignments, Section 3.3). As described earlier, the effects of deuteriation at the NH are restricted to C-2, C=O(NH) and C=C-16. The remaining observed isotope effects seen in Table 1 were due to deuteriation of OH-4. A plot of the experimental values vs. the calculated ones (see Section 3.5.2) showed a very good agreement (the structure is A of Figure 4). In contrast, a plot based on values of the structure C gave a correlation coefficient (R2) as low as 0.38. In that case, see also the discussion on tautomerism.
Deuteriation at OH-1 did not lead to observable doublets. However, as described in Section 2.2, the isotope effects can be observed by measuring the shift as a function of D2O addition. An example is shown for C-1 in Figure 5. From this plot, it can be seen that the addition of water in itself gave rise to a shift to a higher frequency. This has to be taken into account when estimating the isotope effect. The data are given in Table 2. Clearly, the largest observed isotope effects due to deuteriation occurred at OH-1. The other effects are given in Table 1. The smaller effects were somewhat more uncertain but the signs could, in all the mentioned cases, be determined. The observation of isotope effects for C-40, 41 and 43 was a clear confirmation that the structure was a zwitterion.
In Table 2 it is important to notice that the C-2 doublet of doublets were observed due to isotope effects due to deuteriation both at OH-4 and NH.

3.5. Calculations

The structures were calculated using a truncated version of the molecule leaving out most of the long bridge. In one end, the double bond was kept and in the other end, the chain was replaced by a OCH3 group (see Figure 4). Several structures were tested to obtain the optimal truncation. The test involved the fitting of calculated nuclear shieldings vs. observed 13C chemical shifts (see Section 3.5.1). To determine the best basis set, 6-31G(d,p) and 6-311(G,p) were also tested. For structure 4A, the R2 values were 0.9963 and 0.9935, respectively. If need be, the piperazine ring can be replaced by a NH2 (Figure 4E) producing an R2 of 0.9962. The structures were calculated with and without water molecules close to the O atom. The solvent was also taken into account (see Section 2).

3.5.1. 13C Nuclear Shieldings

The calculated 13C nuclear shieldings for structures C and D of Figure 4 are given in Table 3. The very large differences were related to C-4 and C-38.
A plot of the calculated 13C nuclear shieldings vs. experimental 13C chemical for structure A of Figure 4 is given in Figure 6 (R2 = 0.9949). The carbons included are the core carbons C1-C10 and C-11. As carbons one and two bonds away heavily influence the chemical shifts, only the core carbons were involved as we were using the truncated version of the molecule. For the structure with a water molecule added to the O, R2 = 0.9819. In other words, water did not seem to be bonded in the C1-C8 region. For structure C, R2 = 0.9675, and for structure D, R2 = 0.7884.

3.5.2. Isotope Effects on Chemical Shifts

The calculation of deuterium isotope effects on nuclear shieldings is based on the Jameson theory [25,26]. In the present case, the OH bond lengths were reduced by 0.01 Å to mimic the deuteriation [1]. This will lead to “standard” isotope effects that may need to be scaled by plotting those vs. the observed isotope effects. As seen in Table 1 and Figure 7, no scaling was needed for deuteriation of OH-4. A plot of the corresponding situation with water attached to O gave a slightly lower R2 of 0.9723.

4. Discussion

A breakthrough in understanding the structure of rifampicin (Figure 1) was the finding that the structure is a zwitterion in polar solvents with water added [8]. Taking this into account, a number of intramolecularly hydrogen bonded structures can be formulated as seen in Figure 4 (the structures are truncated versions of rifampicin).
These structures clearly demonstrate that changes in the intramolecular hydrogen bonding in one region will influence the hydrogen bonding pattern in another region. More structures with fewer hydrogen bonds can be drawn, but these are clearly less realistic.
Rifampicin is a large molecule. Here, the calculations were based on a truncated version including all essential features as seen in Figure 4. As water plays an important role in the formation of the zwitterionic form [7,8,27], water was added to the structure hydrogen bonding to C-O. This turned out to have very little effect on the calculated deuterium isotope effects of C-1 or C-4 or the 13C nuclear shieldings (see Section 3.5.1). However, as water is important for the formation of the zwitterionic structure, the effect of water is to solvate the N+ ammonium ion. This is very important as the counter ion is far away.
From the very high chemical shift of OH-1 (15.6 ppm), it can be concluded that OH-1 was hydrogen bonded to O. It was also seen that the OH-1 resonance was broad due to exchange.
With respect to OH-4, this can hydrogen bond to C=O in a seven-membered ring arrangement (Figure 4A or Figure 4E) or to the C=N in a six-membered arrangement (Figure 4C). However, in the latter case, it was very similar to o-hydroxy Schiff bases and can in principle be tautomeric [27] (see Figure 4D). However, in this case, one would expect distinctive equilibrium isotope effects at C-3, C-4, C-10 and C-38 proportional to the chemical shift differences between the carbons in the two equilibrating structures (see Equations (3) and (4)) [28]. The differences in calculated nuclear shielding are given in Table 3. Large isotope effects at C-3, C-4, C-10 and C-38 were clearly not seen in Table 1. Based on the finding that structure C was not fitting, that no equilibrium was established and the fact that the deuterium isotope effects on the 13C chemical shifts of Figure 4A fits nicely (See Figure 5) makes 4A the likely structure. As OH-4 is hydrogen bonded to C-11=O, then the NH logically forms an intramolecular hydrogen bond to the nitrogen of the C=N bond. The observed 2ΔC-2(ND) isotope effect of 0.133 ppm clearly showed that the NH was hydrogen bonded. Values of 2ΔC-2(ND) for non-hydrogen bonded cases are typically 0.08–0.1 ppm [29,30]. Based on these arguments, a full structure similar to structures A or E of Figure 4 is the preferred one. Furthermore, this was confirmed by a comparison of experimental 13C chemical shifts vs. calculated 13C nuclear shieldings, as seen from the R2 of 0.9949, as discussed in Section 3.5.1. The isotope effects at C-40, 41 and 43 (Table 1) confirmed that rifampicin was in a zwitterionic form. These finding are similar to the full structure found in the crystal [9].

5. Conclusions

The search for a suitable basis set to use with the B3LYP functional resulted in G(d) giving good NMR nuclear shieldings and being suitable for the study of large biological systems. The DFT calculations were performed using the mentioned functional and basis set. The calculated nuclear shielding and deuterium isotope effects on 13C nuclear shieldings combined with experimental 1H and 13C chemical shifts and deuterium isotope effects on the latter enabled an analysis of the complex hydrogen bonding pattern of the zwitterionic form of rifampicin. The result was an extended hydrogen bonding network determined to a large extent on the hydrogen bond between O and OH-1. The deuterium isotope effects on nuclear shieldings also confirmed the zwitterionic nature of rifampicin in polar solvents.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemistry5020089/s1, Figure S1: Expanded 13C NMR spectra. Table S1: x, y, z coordinates.

Author Contributions

Conceptualization, P.E.H.; methodology, P.E.H.; writing—original draft preparation, P.E.H.; writing—review and editing, F.S.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank Annette Christensen for the expert recording of the NMR spectra and P. Przybylski for his help.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pyta, K.; Janas, A.; Skrzypczak, N.; Schilf, W.; Wicher, B.; Gdaniec, M.; Bartl, F.; Przybylski, P. Specific Interactions between Rifamycin Antibiotics and Water Influencing Ability to Overcome Natural Cell Barriers and the Range of Antibacterial Potency. ACS Infect. Dis. 2019, 5, 1754–1763. [Google Scholar] [CrossRef] [PubMed]
  2. Kozyra, P.; Kaczor, A.; Karczmarzyk, Z.; Wysocki, W.; Pitucha, M. Experimental and computational studies of tautomerism pyridine carbonyl thiosemicarbazide derivatives. Struct. Chem. 2023. [Google Scholar] [CrossRef]
  3. Pyta, K.; Przybylski, P.; Klich, K.; Stefańska, J. A new model of binding of rifampicin and its amino analogues as zwitterionsto bacterial RNA polymerase. Org. Biomol. Chem. 2012, 10, 8283–8297. [Google Scholar] [CrossRef] [PubMed]
  4. Hansen, P.E. Isotope Effects on Chemical shifts of small molecules. Molecules 2022, 27, 2405. [Google Scholar] [CrossRef]
  5. Hansen, P.E. Isotope Effects on Chemical Shifts of Proteins and Peptides. Magn. Reson. Chem. 2000, 38, 1–10. [Google Scholar] [CrossRef]
  6. Hansen, P.E. Isotope Effects on Chemical Shifts in the Study of Hydrogen Bonded Biological Systems. Progress NMR 2020, 120–121, 109–117. [Google Scholar] [CrossRef]
  7. Kim, Y.-I.; Manalo, M.N.; Perés, M.L.; LiWang, A. Computational and empirical trans-hydrogen bond deuterium isotope shifts suggest that N1–N3 A: U hydrogen bonds of RNA are shorter than those of A:T hydrogen bonds of DNA. J. Biomol. NMR 2006, 34, 229–236. [Google Scholar] [CrossRef]
  8. Manalo, M.N.; Perés, L.M.; LiWang, A. Hydrogen-bonding and base-stacking interactions are coupled in DNA, as suggested by Calculated and experimental trans-Hbond deuterium isotope effects. J. Am. Chem. Soc. 2007, 129, 11298–11299. [Google Scholar] [CrossRef]
  9. Hansen, P.E. Methods to distinguish tautomeric cases from static ones. In Tautomerism: Ideas, Compounds, Applications; Antonov, L., Ed.; Wiley-VCH: Weinheim, Germany, 2016. [Google Scholar]
  10. Pyta, K.; Przybylski, P.; Wicker, B.; Gdaniec, M.; Stefańska, J. Intramolecular proton transfer impact on antibacterial properties of ansamycin antibiotic rifampicin and its new amino analogues. Org. Biomol. Chem. 2012, 10, 2385–2388. [Google Scholar] [CrossRef]
  11. Morgante, P.; Peverat, R. The devil in the details: A tutorial review on some undervalued aspects of density functional theory calculations. Quantum Chem. 2020, 120, e26332. [Google Scholar] [CrossRef]
  12. Becke, A.D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  13. Ireta, J.; Neugebauer, J.; Scheffler, M. On the Accuracy of DFT for Describing Hydrogen Bonds: Dependence on the Bond Directionality. J. Phys. Chem. A 2004, 108, 5692–5698. [Google Scholar] [CrossRef]
  14. Afonin, A.V.; Ushakov, I.A.; Simonenko, D.E.; Shmidt, E.Y.; Zorina, N.V.; Mikhaleva, A.I.; Trofimov, B.A. 1H and 13C NMR Study of Bifurcated Intramolecular Hydrogen Bonds in 2,6-Bis(2-pyrrolyl)pyridine and 2,6-Bis(1-vinyl-2-pyrrolyl)pyridine. Russ. J. Org. Chem. 2005, 41, 1516–1521, Translated from Zhurnal Organicheskoi Khimii, Vol. 41, No. 10, 2005. [Google Scholar] [CrossRef]
  15. Afonin, A.V.; Vashchenko, A.V.; Ushakov, I.A.; Zorina, N.V.; Schmidt, E.Y. Comparative analysis of hydrogen bondingwith participation of the nitrogen, oxygen andsulfur atoms in the 2(2-heteroaryl)pyrroles andtheir trifluoroacetyl derivatives based on the 1H,13C,15N spectroscopy and DFT calculations. Magn. Reson. Chem. 2008, 46, 441–447. [Google Scholar] [CrossRef]
  16. Afonin, A.V.; Vashchenko, A.V.; Sigalov, M.V. Estimating the energy of intramolecular hydrogenbonds from 1H NMR and QTAIM calculations. Org. Biomol. Chem. 2016, 14, 11199. [Google Scholar] [CrossRef] [PubMed]
  17. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  18. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef] [PubMed]
  19. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef]
  20. Ditchfield, R.; Hehre, W.J.; Pople, J.A. Self-consistent Molecular-orbital methods. 9. Extended Gaussian-type Basis for Molecular-Orbital Studies of Organic Molecules. J. Chem. Phys. 1971, 54, 724–728. [Google Scholar] [CrossRef]
  21. Miertus, S.; Scrocco, E.; Tomasi, J. Electrostatic interaction of a solute with a continuum. A direct utilization of AB initio molecular potentials for the prevision of solvent effects. Chem. Phys. 1981, 55, 117–129. [Google Scholar] [CrossRef]
  22. Scalmani, G.; Frisch, M.J. Continuous surface charge polarizable continuum models of solvation. I. General formalism. J. Chem. Phys. 2010, 132, 114110. [Google Scholar] [CrossRef]
  23. Ditchfield, R. Self-consistent perturbation theory of diamagnetism. I. A gauge-invariant LCAO method for N.M.R. chemical shifts. Mol. Phys. 1974, 27, 789–807. [Google Scholar] [CrossRef]
  24. Wolinski, K.; Hilton, F.F.; Pulay, P. Efficient implementation of the gauge-independent atomic orbital method for NMR chemical shift calculations. J. Am. Chem. Soc. 1990, 112, 8251–8260. [Google Scholar] [CrossRef]
  25. Jameson, C.J.; Osten, H.-J. The NMR isotope shift in polyatomic molecules. Estimation of the dynamic factors. J. Chem. Phys. 1984, 81, 4300–4305. [Google Scholar] [CrossRef]
  26. Jameson, C.J. Isotopes in the Physical and Biomedical Sciences. Isotopic Applications in NMR Studies; Buncel, E., Jones, J.R., Eds.; Elsevier: Amsterdam, The Netherlands, 1991. [Google Scholar]
  27. Dziembowska, T.; Rozwadowski, Z.; Filarowski, A.; Hansen, P.E. A multinuclear NMR study of proton transfer equilibrium in Schiff bases derived from 2-hydroxy-1-naphtaldehyde. Deuterium isotope effects on 13C and 15N chemical shifts. Magn. Reson. Chem. 2001, 39, S67–S80. [Google Scholar] [CrossRef]
  28. Filarowski, A.; Koll, A.; Rospenk, M.; Krol-Starzomska, I.; Hansen, P.E. Tautomerism of sterically hindered Schiff bases. Deuterium Isotope Effects on 13C Chemical Shifts. J. Phys. Chem. A 2005, 109, 4464–4473. [Google Scholar] [CrossRef] [PubMed]
  29. Jarret, M.; Sin, N.; Dintzner, M. Deuterium Isotope Effects on13C NMR Chemical Shifts of Amides. Microchem. J. 1997, 56, 19–21. [Google Scholar] [CrossRef]
  30. Newmark, R.A.; Hill, J.R. Assignment of Primary and Secondary Amide Carbonyl Resonances in Carbon-13 NMR. J. Magn. Reson. 1976, 21, 1–7. [Google Scholar] [CrossRef]
Figure 1. Rifampicin.
Figure 1. Rifampicin.
Chemistry 05 00089 g001
Figure 2. 1H NMR spectrum of 45 mg rifampicin in 0.6 mL DMSO-d6 + 4 μL D2O.
Figure 2. 1H NMR spectrum of 45 mg rifampicin in 0.6 mL DMSO-d6 + 4 μL D2O.
Chemistry 05 00089 g002
Figure 3. 13C NMR spectrum of 45 mg rifampicin in 0.6 mL DMSO-d6 + 4 μL D2O.
Figure 3. 13C NMR spectrum of 45 mg rifampicin in 0.6 mL DMSO-d6 + 4 μL D2O.
Chemistry 05 00089 g003
Figure 4. Truncated structures with different hydrogen bond motifs. More structures can clearly be constructed, but these are the most likely ones. Motifs: (A) OH-O, NH–N, OH-C=O; (B) OH-C=O, NH-N, OH-C=O; (C) OH-O, N-OH; (D) OH-O, NH-C=O; (E) Like A but with 1,4-piperazine ring replaced by NH2 group.
Figure 4. Truncated structures with different hydrogen bond motifs. More structures can clearly be constructed, but these are the most likely ones. Motifs: (A) OH-O, NH–N, OH-C=O; (B) OH-C=O, NH-N, OH-C=O; (C) OH-O, N-OH; (D) OH-O, NH-C=O; (E) Like A but with 1,4-piperazine ring replaced by NH2 group.
Chemistry 05 00089 g004aChemistry 05 00089 g004b
Figure 5. Plot of 13C chemical shifts of C-1 vs. addition of heavy water (Series 1 and 3). Series 2 is the addition of H2O.
Figure 5. Plot of 13C chemical shifts of C-1 vs. addition of heavy water (Series 1 and 3). Series 2 is the addition of H2O.
Chemistry 05 00089 g005
Figure 6. Plot of calculated vs. observed 13C chemical shifts for structure A of Figure 4.
Figure 6. Plot of calculated vs. observed 13C chemical shifts for structure A of Figure 4.
Chemistry 05 00089 g006
Figure 7. Plot of calculated deuterium isotope effects vs. observed isotope effects for deuteriation at OH-4. The data are based on structure 4A and with DMSO as the solvent.
Figure 7. Plot of calculated deuterium isotope effects vs. observed isotope effects for deuteriation at OH-4. The data are based on structure 4A and with DMSO as the solvent.
Chemistry 05 00089 g007
Table 1. Observed and calculated deuterium isotope effects on 13C chemical shifts in ppm measured in DMSO-d6 as well as observed 13C chemical shifts.
Table 1. Observed and calculated deuterium isotope effects on 13C chemical shifts in ppm measured in DMSO-d6 as well as observed 13C chemical shifts.
Carbon13C CS aIE OH-1 bIE H-4 cIE NHCO cIE NH+ dIE OH-21,23 c
C-1148.520.33 (0.3)
C-2 d115.51 0.05 (0.07) d0.13
C-3114.24
C-4145.40−0.10 (−0.15)0.31 (0.29)
C-598.510.07 (0.08)0.05 (0.09)
C-6171.77
C-7101.02 −0.09 (0.01)
C-8184.21 0.08 (0.09)
C-9 d117.21−0.09 (−0.14)0.0 (0.02)
C-10 d117.050.07 (−0.11)0.07 (0.07)
C-11184.430.11 (0.07)−0.17 (−0.10)
C-12108.55 0.042
C-13
C-14
C-15166.22 0.099
C-16132.17 0.046
C-21 0.12
C-23 0.12
C-38137.590.07 (0.03)
C-40,4151.21 0.07
C-4342.11 0.04 e
a CS means chemical shift. b IE means isotope effect. These effects were determined as a change in the chemical shift upon D2O addition (see Section 2). Only large effects are given. Values in brackets are calculated values. c Observed as splitting due to slow exchange of the label. Values in brackets are calculated values. d C-2, C-9 and C-10 are interchanged compared to [10]. e Slightly larger than 0.04 ppm.
Table 2. 13C chemical shifts as a function of the addition of D2O (H2O) in μL.
Table 2. 13C chemical shifts as a function of the addition of D2O (H2O) in μL.
0 μL2 μL4 μL9 μL20 μL30 μL10 μL a
C11184.43184.37(H) b
184.57(D)
184.29(H)
184.51(D)
184.31(H)
184.47(D)
184.35(H)
184.55(D)
184.40(H)
184.60(D)
184.44(5)
C8184.21184.21(H)
184.13(D)
184.29 (H)
184.20 (D)
184.31 (D)
184.23(H)
184.30(D)184.35(D)184.25
C6171.77171.77171.78171.77171.83171.89171.80
C35169.48169.49169.51169.54169.64169.72169.53
C15166.22166.25125.26166.29166.42166.52166.28
C1148.52148.45148.34148.28148.28148.32148.57
C4145.40145.48(H)
145.17(D)
145.50(H)
145.19(D)
145.54(D)
145.21(H)
145.63(D)
145.32(H)
145.67(D)
145.36(H)
145.40
C29142.79142.79142.81142.82142.89142.94142.80
C19138.13138.16138.18138.19138.30138.37138.17
C38137.59137.60137.54137.49137.54137.55137.55
C16132.17132.15132.13132.17132.18132.20132.18
C17131.99132.02132.05132.06--132.03
C18127.72127.70127.71127.68127.73127.76127.70
C28117.59117.58117.58117.56117.63117.67117.61
C9117.21117.26(D)
117.17(H)
117.28(D)
117.20(H)
117.30(D)
117.21(H)
117.30(D)
w c (H)
117.35
w(H)
117.23
C10117.05117.00117.01116.98117.04117.09117.08
C2 e115.51115.48(HH)
115.44(DH)
115.34(HD)
115.31(DD)
115.48(HH)
115.44(DH)
115.36(DH)
115.31(DD)
115.46(HH)
115.42(DH)
115.33(HD)
115.29(DD)
115.35(DD)115.51 d
115.38(DD)
115.51
C3114.24114.24114.27114.27114.34114.38114.25
C12108.55108.55108.57108.57108.62108.67108.57
C7101.02101.07101.08101.11101.23101.31101.06
C598.5198.4898.4698.4498.4298.4898.54
C3755.6355.6455.6555.6555.7255.7755.65
C4051.2151.1751.1951.1851.2451.2751.25
C3947.6847.6647.6947.6847.7247.7647.71
N-CH342.1142.0742.0942.0942.1442.2142.16
a 10 μL water added to estimate the solvent effect of adding water/heavy water. b Refers to the resonance due to the H species and to the D deuterated species. c w means weak. d Only one weak signal observed. e Two deuterium isotope effects were observed.
Table 3. 13C nuclear shieldings for structures C and D of Figure 4.
Table 3. 13C nuclear shieldings for structures C and D of Figure 4.
CarbonNS a of Structure CNS of Structure DDifference b
C-186.4679.796.67
C-218.0618.62−0.56
C-373.7377.66−3.93
C-473.3564.57−21.22
C-594.8690.294.57
C-624.9126.62−1.71
C-747.2541.615.64
C-842.1112.0630.05
C-991.1574.9616.19
C-1083.3992.13−8.74
C-1411.6410.141.50
C-1936.7937.93−1.14
C-2127.4054.74−27.34
a NS means nuclear shielding. b The differences were used to judge the possibility of finding an equilibrium isotope effect (see Section 4).
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

Hansen, P.E.; Kamounah, F.S. Multiple Intramolecular Hydrogen Bonding in Large Biomolecules: DFT Calculations and Deuterium Isotope Effects on 13C Chemical Shifts as a Tool in Structural Studies. Chemistry 2023, 5, 1317-1328. https://doi.org/10.3390/chemistry5020089

AMA Style

Hansen PE, Kamounah FS. Multiple Intramolecular Hydrogen Bonding in Large Biomolecules: DFT Calculations and Deuterium Isotope Effects on 13C Chemical Shifts as a Tool in Structural Studies. Chemistry. 2023; 5(2):1317-1328. https://doi.org/10.3390/chemistry5020089

Chicago/Turabian Style

Hansen, Poul Erik, and Fadhil S. Kamounah. 2023. "Multiple Intramolecular Hydrogen Bonding in Large Biomolecules: DFT Calculations and Deuterium Isotope Effects on 13C Chemical Shifts as a Tool in Structural Studies" Chemistry 5, no. 2: 1317-1328. https://doi.org/10.3390/chemistry5020089

Article Metrics

Back to TopTop