Next Article in Journal
A Novel Modified ZIF-8 Nanoparticle with Enhanced Interfacial Compatibility and Pervaporation Performance in a Mixed Matrix Membrane for De-Alcoholization in Low-Concentration Solutions
Next Article in Special Issue
Longitudinal and Transverse 1H Nuclear Magnetic Resonance Relaxivities of Lanthanide Ions in Aqueous Solution up to 1.4 GHz/33 T
Previous Article in Journal
Adsorptive Stripping Voltammetric Quercetin Determination in Pharmaceuticals and Urine Samples Using a Long Service-Life Array of Carbon Composite Microelectrodes
Previous Article in Special Issue
Development of Resorbable Phosphate-Based Glass Microspheres as MRI Contrast Media Agents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Gas-Phase Studies of NMR Shielding and Indirect Spin–Spin Coupling in 13C-Enriched Ethane and Ethylene

Laboratory of NMR Spectroscopy, Faculty of Chemistry, University of Warsaw, Pasteura 1, 02-093 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(18), 4460; https://doi.org/10.3390/molecules29184460
Submission received: 29 June 2024 / Revised: 13 September 2024 / Accepted: 14 September 2024 / Published: 20 September 2024
(This article belongs to the Special Issue Advanced Magnetic Resonance Methods in Materials Chemistry Analysis)

Abstract

:
13C and 1H NMR spectra were observed as the function of density in 1,2-13C-enriched ethane and ethylene for the pure gaseous compounds and their binary mixtures with xenon and carbon dioxide gases as the solvents. All the chemical shifts and indirect spin–spin couplings were linearly dependent on the solvent density. The appropriate NMR parameters (σ and nJ) in isolated 13C2H6 and 13C2H4 molecules and the coefficients responsible for the binary molecular interactions were determined and compared with previous similar measurements and selected calculated shielding data. The newly obtained 13C shielding values in the isolated ethane and ethylene molecules suggest visible secondary isotope effects due to the additional carbon-13 atom. All the investigated shielding parameters depend on intermolecular interactions, and the dependence of 13C shielding is much more marked. In contrast, the indirect spin–spin couplings in 13C2H6 and 13C2H4 molecules are almost independent of solvent molecules. Their nJ values determined in liquids over sixty years ago are generally consistent with the same nJ parameters in isolated 13C2H6 and 13C2H4 molecules.

Graphical Abstract

1. Introduction

Ethane gas (C2H6) is a colorless, odorless, flammable, light-saturated hydrocarbon mostly available on a giant scale from natural gas. It can also be found dissolved in petroleum oil and is separated from the oil by fractional distillation. The carbonization of coal can obtain smaller quantities of ethane. The overwhelming mass of ethane is exploited to produce ethylene (ethene, C2H4) by steam cracking. Ethylene gas has similar physical properties to ethane but is much more reactive, as always occurs with unsaturated hydrocarbons. In the chemical industry, ethylene is used to produce polyethylene [1], the most popular plastic material in our everyday life. Moreover, the catalyzed hydration of ethylene gas over H3PO4 delivers technical ethyl alcohol (C2H5OH) [2] and the oxidation of ethylene gas leads to another important product, ethylene glycol (C2H6O2). Both the hydrocarbons—saturated ethane and unsaturated ethylene—are important industrial chemicals and have jointly almost the same final applications. For a similar reason, ethane and ethylene are the objects of our present experimental study in which the enriched carbon-13 molecules (13C2H6 and 13C2H4) are applied. We expect that the enriched molecules deliver more spectral information and permit us to obtain new insights into the electronic structure of ethane and ethylene.
Nuclear magnetic resonance (NMR) spectroscopy is an important method in analyses of organic molecules, particularly in the gas phase. Raynes, Buckingham, and Bernstein performed the first precise NMR study of gases in 1962 [3] exploring the proton spectra for selected gaseous compounds; ethane and ethylene were included in this research. The authors found that in a gas with low density, the nuclear magnetic shielding σA(X) of X nucleus in A molecules can be written as an expansion in powers of the density ρA:
σA(X) = σ0A(X) + σ1A(X)ρA + σ2A(X)ρA2 + …
where σ0A(X) is the shielding for an isolated A molecule and σ1A(X) measures the intermolecular effect on nuclear shielding due to binary molecular collisions. The higher-order terms, starting from σ2A(X)ρA2, are negligibly small for low-density samples. Then, the density dependence is linear and the two shielding parameters (σ0A and σ1A) are readily available from Equation (1). Let us admit that the index X is added to the original RBB equation [3] because, at present, many types of magnetic nuclei can be observed by NMR methods. An additional index is required for the description of the applied multinuclear experiment.
For a binary mixture with a small amount of gas A, and gas B used as the solvent, the linear part of Equation (1) can be rewritten as follows:
σA(X) = σ0A(X) + σ1AA(X)ρA + σ1AB(X)ρB + …
where ρA and ρB are the densities of the A and B components, respectively, and σ0A(X) is the shielding at the zero-density limit, the same as in Equation (1). For a very low density of the A component, the σ1AA(X)ρA term is negligibly small, and after careful verification, it can be neglected. Finally, we have the following:
σA(X) = σ0A(X) + σ1AB(X)ρB
Both the experiments performed according to Equations (1) and (3) should deliver the same results for the shielding value in an isolated A molecule, σ0A(X). The precise description of σ1AA and σ1AB parameters is a bit more complex due to the binary collisions of A–A and A–B molecules. The parameters contain the change in shielding caused by the molecular collisions (σ1(A-A) and σ1(A-B)) and the macroscopic effect of susceptibility change (σ1b) when the shielding measurements are performed using an external reference standard [3]. The σ1b term can be estimated by knowing the investigated medium’s magnetic susceptibility, the NMR sample’s shape, and its orientation relative to the direction of the external magnetic field [4]. Finally, it should be remembered that all the shielding parameters in Equations (1)–(3) are temperature-dependent and that their experimental values are usually measured at the standard temperature of 300 K [5].

2. Results

2.1. NMR Spectra of 13C-Enriched Ethane and Ethylene

The 1H NMR spectra of ethane and ethylene containing a natural abundance of 13C nuclei (1.07 percent [6]) are easily affordable for direct observation because the proton magnetic moment is the largest among all stable nuclei, the protons in each investigated molecule are magnetically equivalent, and rare possible 1H–13C spin–spin couplings can only insignificantly modify proton signals. This is due to the low natural abundance of 13C isotopes usually present in the samples (1.07%), which is still sufficient for the fast observation of 13C spectra with modern NMR spectrometers. Spectrometers use the Fourier transform (FT) method and broadband decoupling from protons to detect carbon spectra. It is well illustrated by the 1H and 13C NMR spectra of ethane and ethylene as both spectral methods practically give strong singlet signals for C2H6 and C2H4 molecules [7,8]. Everything is changed when the ethane and ethylene molecules are fully enriched in the carbon-13 isotope [9,10]. The 1H and 13C NMR spectra of the 13C2H6 and 13C2H4 samples become more complex because neither two carbon-13 nuclei nor all protons are magnetically equivalent and their spectra have the forms A3A′3XX′ and AA′A″A‴XX′, respectively, where the A symbols stand for protons and X for carbon-13 nuclei. This means that the shielding values of 1H and 13C nuclei practically remain the same as for the samples with a natural abundance of 13C nuclei (assuming that the additional isotope effects from the second 13C nucleus in both the molecules are negligibly small [9]), but many new spin-spin couplings are observed and available for precise determination, as always occurs in such a case [11].
Figure 1 presents the 13C spectrum of gaseous doubly 13C-enriched ethane. The XX′ part of the A3A3′XX′ spin system is observed by 13C NMR, which consists of 67 separate NMR signals. All the spectral lines are well resolved, which helps with their precise analysis. In contrast, the 1H NMR spectrum presented in Figure 2 is less readable because many proton signals are overlapping. This results from the general rule that 13C spectra cover a much larger shielding range (approx. 250 ppm [12]) than the proton spectra with a shielding range of several ppm. It is also seen in the width of the present 13C spectrum of 13C2H6 (≈3.8 ppm in Figure 1) and only approx. 0.40 ppm in Figure 2 for 1H NMR measurement.
The spectra presented in Figure 1 and Figure 2 remain qualitatively unchanged when ethane-13C2 is mixed with some “inert” gases like xenon (Xe) or carbon dioxide (CO2); however, their spectral lines are slightly affected, as predicted by Equation (3). This is due to the binary collisions of the solute and solvent molecules in the gas phase. The 13C and 1H NMR spectra of doubly 13C-enriched ethylene presented in Figure 3 and Figure 4 show the same features as above described for 13C2H6 but are even more complex. It arises from the hindered rotation of the carbon-carbon double bond that makes all the protons in the 13C2H4 molecule magnetically nonequivalent. The increasing number of spin-spin couplings leads to more complicated 13C and 1H NMR spectra (AA′A″A‴XX′), as shown in Figure 3 and Figure 4.

2.2. Density-Dependent NMR Shielding in 13C2H6 and 13C2H4 Molecules

Figure 1 and Figure 2 reveal that the 13C and 1H spectra of ethane-13C2 are quite complex, as usual for the A3A′3XX′ spin system. It makes the above NMR spectra a little more difficult to interpret but delivers full information on all the spin–spin coupling constants with properly performed computer simulation. The same is true for the NMR spectra of ethylene-13C2 presented in Figure 3 and Figure 4, though the case of the AA′A″A‴XX′ spin system is even a little more complicated. However, all the spectra in Figure 1, Figure 2, Figure 3 and Figure 4 have an axis of symmetry and permit one to directly read all the chemical shifts and, consequently, the precise shielding parameters for all investigated samples.
The present 13C and 1H magnetic shielding measurements are established from NMR chemical shifts relating to external liquid TMS. Both the investigated nuclei are “light nuclei”, and the simplified form of Equation (4) is fairly satisfactory:
δ i = σ T M S σ i 1 σ T M S σ T M S σ i
where σTMS and σi are the shielding values of TMS and investigated compounds, respectively. 1H and 13C magnetic shieldings were determined assuming shielding values for liquid TMS at 300 K: 32.815 ppm for protons and 186.37 ppm for carbon-13 nuclei [13]. Then, the obtained shielding parameters were plotted vs. the density of gases according to Equations (1) and (3), as illustrated by Figure 5a,b for ethane-13C2. The plots indicate the same intercept at the zero point density within an experimental error: approx. 180.775 ppm for 13C shielding and 29.889 ppm for 1H shielding. It confirms the equivalence of applications of the two methods in shielding readings: for pure solute compounds (according to Equation (1)) and observation in gaseous solvents (cf. Equation (3)). They have given us the parameters of shielding at the zero-density limit, which are equivalent to the results for the isolated 13C2H6 molecules. Moreover, the slopes recorded on the same figures have delivered the σ1(A-A)(X) and σ1(A-B)(X) coefficients, which show the intermolecular effects on shielding due to the binary molecular collisions. Figure 5a reveals the decrease in 13C shielding in 13C2H6 molecules with an increased gas density. It is a common effect for saturated hydrocarbons and all other chemical compounds without lone pairs of electrons in their molecules. The magnitude of σ1(A-A)(X) or σ1(A-B)(X) coefficient is the averaged effect of all possible bimolecular collisions in the gas phase. Figure 5a shows that the anisotropy of CO2 molecules can slightly diminish the σ1 intermolecular effect compared to the interactions observed in pure ethane molecules. In contrast, the isotropic xenon atoms containing more electrons diminish the 13C shielding in 13C2H6 due to the intermolecular effects.
Figure 5b presents the 1H shielding dependence on gas densities observed for 1,2-13C-ethane. In contrast to the 13C shielding from the previous picture, all the intermolecular effects observed on protons are negligible and remain within experimental errors (±0.01 ppm). Although the proton shielding scale in Figure 5b is five times enlarged relative to the 13C scale in Figure 5a, the effects of the binary molecular collisions are hardly seen for protons. In particular, in our opinion, the small increase in proton shielding with CO2 molecules is rather artificial and is within an experimental error [14,15]. Generally, it can be concluded that the density dependence of proton shielding in 1,2-13C-ethane has no diagnostic information on the intermolecular interactions observed in the gas phase. On the other hand, it proves that the electronic structure of ethane is very stable and fairly resistant to intermolecular interactions.
Below, we present in Figure 6a,b the shielding dependences on density for the 13C and 1H NMR measurements of 1,2-13C-ethylene in gaseous samples.
As seen in Figure 6a,b, the intermolecular interactions diminish the 13C and 1H magnetic shielding parameters in 1,2-13C-ethylene molecules in a very similar way to that for 1,2-13C-ethane: CO2 as the gas solvent provides the smallest decrease in 13C shielding, and Xe gas gives the largest negative effect. The magnitude of all the intermolecular effects in Figure 6a,b is similar to the previously observed changes in Figure 5a,b. The 1H shielding changes in 1,2-13C-ethylene are a bit more distinct (up to −0.03 ppm for CO2 solvent) but still small, as previously observed for 1,2-13C-ethane. Generally, no meaningful effects in the proton spectra can be detected for the intermolecular interactions in gaseous 13C2H4 (and 13C2H6) molecules with the selected gaseous solvents (CO2 or Xe). All the numerical data of our shielding measurements for 1,2-13C-ethane and 1,2-13C-ethylene are given in Table 1.

2.3. Isotropic Spin–Spin Coupling in 13C2H6 and 13C2H4 Molecules

Determining indirect spin–spin coupling constants requires the deconvolution of 13C NMR spectra, as described in the Experimental Section. Then, the observed parameters for all the spin–spin couplings (nJ) can be analyzed similarly as was carried out for shielding using the linear part of density-dependent spin–spin couplings. The method has been previously used to observe 1J(29Si-19F) spin–spin coupling in gaseous SiF4 [16].
nJA(XY) = nJ0A(XY) + nJ1AA(XY)ρA
nJA(XY) = nJ0A(XY) + nJ1AB(XY)ρB
In the above equations, the XY symbols describe the indirect spin–spin coupling between X and Y nuclei across n chemical bonds. In our current presentation, the symbol A stands for the investigated compound (13C2H6 or 13C2H4) and B for a selected gaseous solvent (CO2 or Xe). Generally, the J-couplings in hydrocarbons are fairly stable and frequently remain almost unchanged due to intermolecular interactions in the gas phase [17], as happens for example in the methane molecule [18]. In contrast, heteroatoms in molecules can dramatically change the scale of 1J density dependence; such effects are observed for the 1J(CF) coupling in CD3F [19] or 1J(CH) in CH2F2 [20].
Four different spin–spin couplings are observed for the A3A3′XX′ spin system in 13C2H6 and six others for the AA′A″A‴XX′ spin system in 13C2H4. We have selected only the two most important couplings for the presentation in detail for each of the investigated hydrocarbons: 1J(CH)—because this coupling is available from NMR measurements performed for samples with the natural abundance of carbon-13 nuclei when the 1J(CC) is seen only in the enriched carbon-13 samples. The first case means the possibility of comparison with numerous other measurements, and the second case offers a new value for the parameter that is measured for the first time in the isolated molecules. Below, Figure 7a,b illustrates the linear density dependence of 1J(CH) and 1J(CC) spin–spin coupling in ethane-13C2 on gas density, as shown by Equations (5) and (6).
Figure 7a,b presents the typical value of 1J(CH) coupling (≈125 Hz) for the sp3 carbon hybridization of electrons in saturated hydrocarbons [12]. The 1J(CC) equal to 35 Hz is also typical for the one-bond 13C-13C coupling observed in saturated hydrocarbons [12]. It is more important that both the spin–spin couplings across one chemical bond in ethane are practically independent of intermolecular interactions. It proves ethane’s stable electronic structure, which remains unchanged to external molecular perturbations.
Finally, all the measurements performed on spin–spin coupling for the isolated 13C2H6 and 13C2H4 molecules are presented in Table 2, separately for each chemical compound. As shown in Table 2, all the J-couplings across more than one chemical bond are small and not differentiated by gas solvents as well; their graphical presentation was deliberately omitted in our presentation. In the next section, the new results on J-couplings are discussed and compared with similar earlier measurements and some data from quantum chemical calculations.

3. Discussion

3.1. Nuclear Magnetic Shielding

The current study provides the first measurements of 13C and 1H NMR shielding in 13CH2H6 and 13CH2H4 gases as the function of density according to Equations (1)–(3), presented in Figure 5 and Figure 6, and summarized in Table 1. As seen, the gas phase measurements allow for the precise determination of shielding parameters σ0(C) and σ0(H) free from intermolecular interactions and, therefore, equivalent for isolated molecules. Similar studies were performed earlier but using ethane and ethylene with the natural abundance of carbon-13 isotope. This means that the objects of previous shielding studies were slightly different: 13CH312CH3 and 13CH212CH2 for 13C NMR investigations, and mostly 12C2H6 and 12C2H4 for 1H experiments. The secondary isotope effect from the additional carbon-13 in the studied molecules is expected to be small. Therefore, we can safely compare all the shielding results on proton and carbon-13 in the isolated ethane and ethylene molecules. Table 3 gives the 13C and 1H shielding values from various experimental and theoretical investigations. The experimental data are limited to the σ0(C) and σ0(H) values and calibrated to a new scale of 13C shielding in which σ0(C) for an isolated CO molecule is equal to 0.6 ± 0.9 ppm [21]. The precision of 1H shielding measurements is much better and equal to ±0.005 ppm when liquid external TMS is applied as the reference standard [13]. The experimental σ0(C) and σ0(H) data shown in Table 3 are fairly consistent. It confirms that the secondary isotope effects in shielding are generally small. The calculated data of 13C and 1H shielding in ethane and ethylene molecules are inconsistent and mostly depend on the applied method of calculations.
As mentioned, Table 3 presents consistent σ0(C) and σ0(H) experimental data for ethane and ethylene. The consistency looks even better if the measurements of 13C shielding performed at higher temperatures [25] are excluded from our discussion for a while because it is of a bit lower precision. Such a measurement contains two unknown temperature contributions to shielding, one from the investigated molecule (ethane or ethylene) and the other from the reference molecule (methane in this case). The remaining three other σ0(C) results were obtained practically at the same temperature as for ethane and ethylene. Our present measurements give slightly larger 13C shielding values, 180.775 ppm for 13CH2H6 and 64.367 ppm for 13CH2H4 than the previous results [24,33]. It allows us to estimate the secondary isotope effects in carbon shielding: 1Δ(13/12C) ≈ +0.25 ppm for ethane and 1Δ(13/12C) ≈ +0.20 ppm for ethylene. The molecules with heavier isotopes have generally increased NMR shielding [34], supporting the above conclusion of 1Δ(13/12C) detection in carbon-13 doubly enriched ethane and ethylene.
Table 3 also contains the 13C and 1H shielding values for ethane and ethylene dissolved in liquid chloroform (CHCl3). The deshielding effects are observed in every case typical for aliphatic hydrocarbons [24]. The decrease in 13C shielding is more significant and correlates with the σ1(A-B)(C) parameters given in Table 1. The same is true for proton shielding in the investigated molecules but the decrease in 1H shielding is considerably smaller and limited only to a fraction of 1 ppm. Nevertheless, the intermolecular effects in shielding are significant in every condensed phase and require removal from the experimental results if the exact shielding values in molecules are needed. It is possible only to explore the measurements performed in the gas phase.
The results in Table 3 are given without error bars because the shielding scales of 13C and 1H nuclei have been defined earlier with the error bars as ±0.9 ppm and ±0.005 ppm, respectively. We have used the original suggestions citing other authors’ results and saved our data with the error bars in Table 1. The small error bars in Table 1 make sense because the current measurements are precisely performed relative to liquid TMS as the external reference standard. When the shielding scale of 13C is improved, one can easily obtain better values for ethane-13C2 and ethylene-13C2. Meanwhile, the comparison of various data in Table 3 illustrates only various experimental and theoretical investigations of shielding.
Ethane and ethylene are relatively small molecules, and the advanced ab initio methods based on the GIAO (Gauge Included Atomic Orbitals) approach [35,36], like CCSD (Coupled Cluster Singlets and Doubles) or CCSD(T) (Coupled Cluster Singlets and Doubles with Perturbative Triple Corrections) [37,38], can be applied for their shielding calculations [34]; such an approach to carbon-13 shielding is superior, as seen in Table 3. The modern state-of-the-art shielding calculations are very advanced, including all the intra- and intermolecular contributions, and the relativistic effects in shielding if required [39]. Various other more approximate methods can be used for shielding calculations in larger molecules: from HF (Hartree-Fock) to FCI (Full Configuration Interaction), MCSCF (Multi-Configuration Self-Consistent-Field), CC (Coupled Claster) approximation, and MP (Møller–Plesset) perturbation theory [40].
As shown, the σ0A parameters of Equations (1)–(3) can be precisely determined and compared with previous NMR measurements and advanced quantum chemical calculations. The second-order terms of the same equations are much more complex and deliver many problems connected with the details of experimental studies performed in the gas phase. The σ1 parameters (also known as the second virial coefficient in Equations (1)–(3)) cannot be easily compared from various laboratories because the applied methods, or even different NMR spectrometers, may introduce significant discrepancies in the measurements of intermolecular effects. Precise quantum calculations of interacting large molecules like ethane or ethylene are still unavailable. The above problem certainly requires careful investigation in the future but exceeds the framework of the present work. For this reason, we have transferred the discussion on the σ1 parameters of ethane-13C2 and ethylene-13C2 to the Supplementary Materials.

3.2. Indirect Spin–Spin Coupling

Figure 7 and Figure 8 present the minimal dependence of 1J(CC) and 1J(CH) spin–spin couplings on density in our study. The minimal effect can be better understood if we compare it with the 1H density-dependent shielding in Figure 5b and Figure 6b. Proton shielding is changed in the range of 0.02 ppm for ethane-13C2 and 0.03 ppm for ethylene-13C2. This means a change of approx. 10 and 15 Hz, respectively, in the 1H NMR spectrum (1 ppm ≈ 500 Hz). The 1J(CC) and 1JCH) spin–spin couplings vary only from 0.1 to 0.15 Hz in the same experimental conditions, i.e., two orders of magnitude less than 1H shielding.
Table 4 shows our recent results of spin–spin coupling data based on well-resolved 13C NMR spectra and extrapolated to the zero-density point. They are compared with the results of 13CH2H6 and 13CH2H4 obtained in condensed environments like liquid solutions in CCl4 [9], liquefied hydrocarbons at low temperatures [10], and liquid crystals [25]. Let us note that the change of samples has limited influence on the observed values of spin–spin coupling; quite different experiments show only minimal effects due to intermolecular interactions in this case. One more feature of the discussed data must be especially underlined. The first analyses of spin–spin coupling in 13CH2H6 and 13CH2H4 were performed over 60 years ago [9,10] using only available low-quality 1H NMR spectra. In our opinion, it is something fantastic and unusual in experimental NMR studies.
The comparison of plots in Figure 7b and Figure 8b reveals a more solvent-differentiated 1J(CC) coupling in ethylene-13C2 than the 1J(CC) in ethane-13C2. It is obviously due to a carbon-carbon double bond in the ethylene molecule. This effect is not confirmed in Table 4 if the results from various experiments are compared. However, the 1J(CC) of ethane and ethylene in Figure 7b and Figure 8b are similarly diminished due to intermolecular interactions. Table 4 confirms the decreased 1J(CC) values of 13CH2H6 and 13CH2H4 molecules in the condensed phases [9,10,25].
The results of the spin–spin couplings of 13CH2H6 and 13CH2H4 shown in detail in Table 2 are a bit better for isolated molecules and can be used as benchmarks in quantum chemical calculations. Table 4 also contains some selected calculated data of spin-spin couplings in ethane and ethylene molecules obtained from different computational methods [36,37,38,39,40,41,47]. The discrepancy between the experimental and theoretical results is well-marked. It arises from the difficulty in performing spin-spin coupling calculations utilizing quantum chemical methods [39,48]. In contrast to shielding calculations, more varied methods of calculations are applied and the different basis sets are focused on the inner electrons in the molecules. Approximate DFT (Density-Functional Theory) methods are more frequently used in studies of indirect J-coupling [49]. The above problems of such a comparison were recently discussed in detail when presenting the experimental and calculated nJ values for fluoromethanes [50].

4. Materials and Methods

4.1. Chemical Compounds and Samples Preparations

1,2-13C-ethane and 1,2-13C-ethylene (both with 99% enrichment in carbon-13 nuclei from Sigma-Aldrich, Poznań, Poland) plus carbon dioxide and xenon (99.99% Messer, also delivered by Sigma-Aldrich in lecture bottles) were used to prepare the samples without further purification. The gas samples were prepared via the condensation of gases from the calibrated part of a vacuum line into the 4.0 mm o.d. glass tubes (approx. 5.5 cm long), which were then sealed with a little torch. The volumes of the sample tubes and the vacuum line were measured using mercury. The studied gases (13C2H6 and 13C2H4) were observed as pure compounds in the range of approximately 7 to 50 bar or applied as solutes at low constant pressure (0.7 bar) and mixed with various quantities of the gaseous solvents (CO2 and Xe) up to a similar range of pressure (7–50 bar). The high-quality liquid solvents (Sigma-Aldrich) were carefully dehydrated and degassed in a vacuum line before the preparation of the liquid solutions with the enriched gaseous solutes. The gas samples were fitted into 5 mm o.d. standard NMR tubes (Wilmad 528-PP, Buena, CA, USA) with liquid cyclohexane-d12 in the annular space for the deuterium lock system. It also permitted us to gain additional control of the 13C and 1H chemical shifts because the shielding values of the liquid cyclohexane-d12 are precisely known (σliq(13C) = 160.513 ppm and σliq(1H) = 31.686 ppm) [51].

4.2. NMR Spectra

One-dimensional 13C and 1H NMR spectra were acquired on a Varian UNITYplus-500 FT spectrometer, Santa Clara, CA, USA (11.7 T) at the 125.88 and 500.62 MHz transmitter frequencies for the 13C and 1H nuclei, respectively. Ethane-13C2 was measured using a Varian sw5 (5 mm switchable_nmr_tm) probe, while ethylene-13C2 required the ID_PFG (indirect_nmr_tm) 5 mm probe. All the measurements were performed at the stabilized temperature of 300 K. The FID acquisition time (AQ) and scan numbers (NS) were variable with sample density due to the efficient spin-rotation relaxation and set up as follows: for ethane-13C2 (1H), AQ = 4–7s and NS = 16–256; for ethane-13C2 (13C), AQ = 3–6s and NS = 72–2000; for ethylene-13C2 (1H), AQ = 1.5–3s and NS = 500–2000; and for ethylene-13C2 (13C), AQ = 1.5–3s and NS = 1000–1600. D1 and DS parameters were not applied in this study. The spectral width was used from 400 to 800 Hz. The substitution method was applied to measure the 13C and 1H NMR chemical shifts relative to external liquid TMS [6]. The absolute magnetic shielding of TMS (32.815 ppm for protons and 186.37 ppm for 13C nuclei in a cylindrical tube parallel to the external magnetic field) [13] was used to convert the NMR chemical shifts into the absolute shielding values of 13C2H6 and 13C2H4. All the bulk susceptibility corrections (BSCs) that did not influence shielding in isolated molecules were removed from our experimental results in Figure 5 and Figure 6 and Table 1 and Table 3 using susceptibility data [52] and the appropriate formula for parallel orientation of cylindrical tubes relative to the external magnetic field [4]. More details on the second virial coefficients (σ1) and the BSC estimation are given in the Supplementary Materials.

4.3. Analysis of Spectral Parameters

Three to five spectra were recorded for each sample at various times to avoid accidental errors, and to increase the precision of the obtained results. It has allowed us to estimate the possible maximal errors in the NMR experiments. All the measurements were performed for a random order of solvent density. The total line-shape (TLS) program [53], as part of the PERCH Software (version 1/96) [47,54], was applied for the full deconvolution of the 13C NMR spectra of ethane-13C2 and ethylene-13C2. It permitted a more precise determination of all the spin coupling constants in the investigated compounds. Let us note that the appropriate proton spectra of investigated molecules were less resolved than the carbon spectra, as the 1H shielding scale is much smaller than that for 13C NMR. Finally, we analyzed the results for pure 13C2H6 and 13C2H4 using Equations (1) and (5) in the full density range. The gas mixtures required the application of Equations (3) and (6), and the measurements were performed for shielding and J-couplings in the limited density range, as shown in Figure 7a,b, and Figure 8a,b. This means the molar ratio (xM) 13C2H6 changes from 0.040 to 0.0092 in CO2 and from 0.036 to 0.0094 in Xe. Similar data were obtained for 13C2H4, as follows: 0.009–0.0029 for CO2 as the solvent and 0.011–0.0038 for Xe. Larger xM values were assigned for lower densities of the solvent gases.

5. Conclusions

The present study yields experimental results for the 13C and 1H magnetic shielding and spin–spin coupling parameters of 1,2-13C-enriched ethane and ethylene in the gas phase. For the first time, the 13C and 1H NMR spectra of the investigated compounds were observed for pure compounds and their binary mixtures with xenon and carbon dioxide gases in a wide range of densities. All the shielding and spin–spin couplings were linearly dependent on gas density. The appropriate NMR parameters (σ0 and nJ0) in isolated 13C2H6 and 13C2H4 molecules and the coefficients responsible for the binary molecular interactions were determined and compared with similar previous measurements and selected calculated data based on modern quantum chemical methods. The comparison of intermolecular effects in shielding is less reliable because many experimental details influence the final value of σ1 parameters. We have extended the presentation of intermolecular shielding effects in the Supplementary Materials, adding some previous results obtained for ethane and ethylene with natural abundance of carbon-13 [3,21,47,55,56,57].
The newly-obtained 13C shielding values in the isolated ethane-13C2 and ethylene-13C2 molecules suggest visible secondary isotope effects (1Δ(13/12C)) due to the second carbon-13 atom in the studied molecules. The present study confirms that all the shielding parameters in 13C2H6 and 13C2H4 are dependent on intermolecular interactions and the dependence of 13C shielding is much more marked. Generally, the decrease in shielding is observed in a more dense molecular isotropic environment. In contrast, the indirect spin–spin couplings in 13C2H6 and 13C2H4 molecules are almost independent of solvent molecules. Their nJ values determined in liquids over sixty years ago are consistent with the same nJ parameters in isolated 13C2H6 and 13C2H4 molecules. However, the current measurements reveal that the 1J(CC) coupling in ethylene is more sensitive to molecular interactions than the appropriate 1J(CC) in ethane. The present study confirms that all the parameters of shielding and spin–spin coupling in ethane and ethylene molecules are more or less sensitive to intermolecular interactions, and this fact cannot be neglected if precise NMR measurements are required.
Small molecules like 13C2H6 and 13C2H4 are especially interesting for the development of new quantum-chemical methods of calculations. The theoretical techniques are always more or less approximate tools, and their improvement is crucial for chemical applications. With new experimental data, the various new methods of quantum calculations can be easily verified, especially when the experimental data are available for small isolated molecules. Our present research delivers experimental data on proton and carbon-13 shielding, numerous J-coupling constants, and isotopic effects in shielding for 13C2H6 and 13C2H4 molecules.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29184460/s1, Table S1: The second virial coefficient of shielding for gaseous ethane and ethylene.

Author Contributions

Conceptualization, M.W. and K.J.; methodology, M.W.; software, M.W.; validation, M.W. and K.J.; formal analysis, M.W.; investigation, M.W.; resources, M.W.; data curation, K.J.; writing—original draft preparation, K.J.; writing—review and editing, K.J.; visualization, M.W.; supervision, K.J.; project administration, K.J.; funding acquisition, K.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partially supported by the Polish State Committee for Scientific Research under Grant 3 T09A 008 16.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Salomons, T.W.G.; Fryhle, C.B.; Snyder, S.A. Salomons’ Organic Chemistry, 12th ed.; John Wiley & Sons: Hoboken, NJ, USA, 2017; p. 475. ISBN -13: 978-1119248972. [Google Scholar]
  2. Fougret, C.M.; Atkins, M.P.; Hölderich, W.F. Influence of the Carrier on the Catalytic Performance of Impregnated Phosphoric Acid in the Hydration of Ethylene. Appl. Catal. A Gen. 1999, 181, 145–156. [Google Scholar] [CrossRef]
  3. Raynes, W.T.; Buckingham, A.D.; Bernstein, H.J. Medium Effects in Proton Magnetic Resonance. I. Gases. J. Chem. Phys. 1962, 36, 3481–3488. [Google Scholar] [CrossRef]
  4. Smith, A.; Raynes, W.T. Studies of the Density Dependence of Proton Magnetic Shielding in Gases. J. Crystallogr. Spectrosc. Res. 1983, 13, 77–87. [Google Scholar] [CrossRef]
  5. Jameson, C.J. Fundamental Intramolecular and Intermolecular Information from NMR in the Gas Phase. In Gas Phase NMR—New Developments in NMR; Price, W.S., Jackowski, K., Jaszuński, M., Eds.; Royal Society of Chemistry: Cambridge, UK, 2016; Chapter 1; pp. 1–51. ISBN 978-1-78262-161-4. [Google Scholar] [CrossRef]
  6. Harris, R.K.; Becker, E.D.; Cabral de Menezes, S.M.; Goodfellow, R.; Granger, P. NMR Nomenclature: Nuclear Spin Properties and Conventions for Chemical Shifts. IUPAC Recommendations 2001. Magn. Reson. Chem. 2002, 40, 489–505. [Google Scholar] [CrossRef]
  7. Pople, J.A.; Schneider, W.G.; Bernstein, H.J. High-Resolution Nuclear Magnetic Resonance; McGraw-Hill Book Company Inc.: New York, NY, USA; Toronto, ON, Canada; London, UK, 1959; p. 90. ISBN -13 9780070505162. [Google Scholar]
  8. Kalinowski, H.O.; Berger, S.; Braun, S. Carbon-13 NMR Spectroscopy; John Wiley & Sons: Chichester, UK; New York, NY, USA; Brisbane, Australia; Toronto, ON, Canada; Singapore, 1988; p. 101. ISBN -13 978-0471913061. [Google Scholar]
  9. Lynden-Bell, R.M.; Sheppard, N. High-Resolution Nuclear-Magnetic-Resonance Spectra of Hydrocarbon Grouping. VI. The Hydrogen Spectra of Carbon-13 Substituted Ethane, Ethylene and Acetylene. Proc. R. Soc. A 1962, 269, 385–403. [Google Scholar] [CrossRef]
  10. Graham, D.M.; Holloway, C.E. Nuclear Spin-Spin Coupling in C13-Enriched Acetylene, Ethylene, and Ethane. Can. J. Chem. 1963, 41, 2114–2118. [Google Scholar] [CrossRef]
  11. Kamińska-Trela, K. 13C-13C Coupling Constants in Diacetylene (1,3-Butadiyne) and its Bis(triethylsilyl) Derivative. Org. Magn. Reson. 1980, 14, 398–403. [Google Scholar] [CrossRef]
  12. Wehrli, F.W.; Wirthlin, T. Interpretation of Carbon-13 NMR Spectra; Heyden and Son Ltd.: London, UK; Philadelphia, PA, USA; Rheine, Germany, 1978; Carbon-13 NMR Correlation Chart; pp. 51, 57, 311. ISBN 0-85501-207-2. [Google Scholar]
  13. Makulski, W.; Jackowski, K. 1H, 13C and 29Si Magnetic Shielding in Gaseous and Liquid Tetramethylsilane. J. Magn. Reson. 2020, 313, 106716. [Google Scholar] [CrossRef]
  14. Rummens, F.H.A. Van der Waals Forces in NMR Intermolecular Shielding Effects. In NMR Basic Principles and Progress; Diehl, P., Fluck, E., Kosfeld, R., Eds.; Springer: New York, NY, USA, 1975; Volume 10, pp. 1–118. ISBN -10 3642661785. [Google Scholar]
  15. Raynes, W.T. Electric Field Effects on Shielding Constants, Encyclopedia of NMR, 2nd ed.; Harris, R.K., Wasylishen, R.E., Eds.; John Wiley & Sons: Chichester, UK, 2012; Volume 3, pp. 1378–1388. ISBN 978-0-470-05821-3. [Google Scholar]
  16. Jameson, A.K.; Reger, J.P. Gas-Phase Density-Dependent Directly Bonded Coupling Constant. J. Phys. Chem. 1971, 75, 437–439. [Google Scholar] [CrossRef]
  17. Jackowski, K. Gas-Phase Studies of Spin-Spin Coupling Constants. Int. J. Mol. Sci. 2003, 4, 135–142. [Google Scholar] [CrossRef]
  18. Bennett, B.; Raynes, W.T. Temperature Dependence of the Carbon-13 Proton Spin-Spin Coupling Constant in Gaseous Methane. Mol. Phys. 1987, 61, 1423–1430. [Google Scholar] [CrossRef]
  19. Jackowski, K.; Kubiszewski, M.; Makulski, W. 13C and 19F Nuclear Magnetic Shielding and Spin-Spin Coupling in Gaseous Fluoromethane-d3. J. Mol. Struct. 2002, 614, 267–272. [Google Scholar] [CrossRef]
  20. Kubiszewski, M.; Makulski, W.; Jackowski, K. Intermolecular Effects on Spin-Spin Coupling and Magnetic Shielding Constants in Gaseous Difluoromethane. J. Mol. Struct. 2004, 704, 211–214. [Google Scholar] [CrossRef]
  21. Raynes, W.T.; McVay, R.; Wright, S.J. An Improved Carbon-13 Nuclear Shielding Scale. J. Chem. Soc. Faraday Trans. 1989, 2, 759–763. [Google Scholar] [CrossRef]
  22. Auer, A.A.; Gauss, J.; Stanton, J.F. Quantitative Prediction of Gas-Phase 13C Nuclear Magnetic Shielding Constants. J. Chem. Phys. 2003, 118, 10407–10417. [Google Scholar] [CrossRef]
  23. Chesnut, D.B. On the Calculation of Hydrogen NMR Chemical Shielding. Chem. Phys. 1997, 214, 73–79. [Google Scholar] [CrossRef]
  24. Jackowski, K.; Raynes, W.T. Density-Dependent Magnetic Shielding in Gas Phase 13C N.M.R. Mol. Phys. 1977, 34, 465–475. [Google Scholar] [CrossRef]
  25. Kaski, J.; Lantto, P.; Vaara, J.; Jokisaari, J. Experimental and Theoretical ab Initio Study of the 13C-13C Spin-Spin Coupling and 1H and 13C Shielding Tensors in Ethane, Ethene, and Ethyne. J. Am. Chem. Soc. 1998, 120, 3993–4005. [Google Scholar] [CrossRef]
  26. Bennet, B.; Raynes, W.T. Intermolecular and Intramolecular Effects on the 1H and 13C Shielding in Some Gaseous Hydrocarbons at Various Temperatures- Experimental Results. Magn. Res. Chem. 1991, 29, 946–954. [Google Scholar] [CrossRef]
  27. Garbacz, P.; Jackowski, K.; Makulski, W.; Wasylishen, R.E. Nuclear Magnetic Shielding for Hydrogen in Selected Isolated Molecules. J. Phys. Chem. A 2012, 116, 11896–11904. [Google Scholar] [CrossRef]
  28. Grayson, M.; Raynes, W.T. Electric Field Effects on the Carbon-13 Nuclear Magnetic Shielding in Several Organic Molecules. Mol. Phys. 1994, 81, 533–545. [Google Scholar] [CrossRef]
  29. Spiesecke, H.; Schneider, W.G.J. Effect of Electronegativity and Magnetic Anisotropy of Substituents on C13 and H1 Chemical Shifts in CH3X and CH3CH2X Compounds. J. Chem. Phys. 1961, 35, 722–731. [Google Scholar] [CrossRef]
  30. Grayson, M.; Raynes, W.T. Electric Field Effects on the Shielding of Protons in C- H Bonds. Mag. Reson. Chem. 1995, 33, 138–143. [Google Scholar] [CrossRef]
  31. Makulski, W.; Szyprowska, A.; Jackowski, K. Precise Determination of the 13C Nuclear Magnetic Moment from 13C, 3He and 1H NMR Measurements in the Gas Phase. Chem. Phys. Lett. 2011, 511, 224–228. [Google Scholar] [CrossRef]
  32. Kupka, T.; Stachów, M.; Nieradka, M.; Kaminsky, J.; Pluta, T. Convergence of Nuclear Magnetic Shielding in the Kohn-Sham Limit for Several Small Molecules. J. Chem. Theory Comput. 2010, 6, 1580–1589. [Google Scholar] [CrossRef]
  33. Jameson, A.K.; Jameson, C.J. Gas-Phase 13C Chemical Shifts in the Zero-Pressure Limit: Refinements to the Absolute Shielding Scale for 13C. Chem. Phys. Lett. 1987, 134, 461–466. [Google Scholar] [CrossRef]
  34. Jameson, C.J.; Mason, J. The Chemical Shift. In Multinuclear NMR; Mason, J., Ed.; Plenum Press: New York, NY, USA, 1987; Chapter 3; pp. 80–82. [Google Scholar] [CrossRef]
  35. Woliński, K.; Hinton, J.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]
  36. Gauss, J. Calculation of NMR Chemical Shifts at Second-order Many-body Perturbation Theory Using Gauge-including Atomic Orbitals. Chem. Phys. Lett. 1992, 191, 614–620. [Google Scholar] [CrossRef]
  37. Watts, J.D.; Gauss, J.; Bartlett, R.J. Coupled-cluster Methods with Noniterative Triple Excitations for Restricted Open-shell Hartree–Fock and other General Single Determinant Reference Functions. Energies and Analytical Gradients. J. Chem. Phys. 1993, 98, 8718–8733. [Google Scholar] [CrossRef]
  38. Gauss, J.; Stanton, J.F. Coupled-cluster Calculations of Nuclear Magnetic Resonance Chemical Shifts. J. Chem. Phys. 1995, 103, 3561–3577. [Google Scholar] [CrossRef]
  39. Halgaker, T.; Jaszuński, M.; Ruud, K. Ab Initio Methods for the Calculation of NMR Shielding and Indirect Spin−Spin Coupling Constants. Chem. Rev. 1999, 99, 293–352. [Google Scholar] [CrossRef] [PubMed]
  40. Antušek, A.; Jaszuński, M. Accurate Non-Relativistic Calculations of NMR Shielding Constants. In Gas Phase, N.M.R.—New Developments in NMR; Price, W.S., Jackowski, K., Jaszuński, M., Eds.; Royal Society of Chemistry: Cambridge, UK, 2016; Chapter 6; pp. 186–217. ISBN 978-1-78262-161-4. [Google Scholar] [CrossRef]
  41. Sneskov, K.; Stanton, J.F. Effects of vibrational averaging on coupled cluster calculations of spin-spin coupling constants for hydrocarbons. Mol. Phys. 2012, 110, 2321–2327. [Google Scholar] [CrossRef]
  42. Keal, T.W.; Helgaker, T.; Sałek, P.; Tozer, D.J. Choice of Exchange-Correlation Functional for Computing NMR Indirect Spin-Spin Coupling Constants. Chem. Phys. Lett. 2006, 425, 163–166. [Google Scholar] [CrossRef]
  43. Kupka, T.; Nieradka, M.; Stachów, M.; Pluta, T.; Nowak, P.; Kjær, H.; Kongsted, J.; Kaminsky, J. Basis Set Convergence of Indirect Spin−Spin Coupling Constants in the Kohn−Sham Limit for Several Small Molecules. J. Phys. Chem. 2012, 116, 3728–3738. [Google Scholar] [CrossRef] [PubMed]
  44. Lantto, P.; Vaara, J.; Helgaker, T. Spin-Spin Coupling Tensors by Density-Functional Linear Response Theory. J. Chem. Phys. 2002, 117, 5998–6009. [Google Scholar] [CrossRef]
  45. Helgaker, T.; Watson, M.; Handy, N.C. Analytical Calculation of Nuclear Magnetic Resonance Indirect Spin-Spin Coupling Constants at the Generalized Gradient Approximation and Hybrid Levels of Density-Functional Theory. J. Chem. Phys. 2000, 113, 9402–9409. [Google Scholar] [CrossRef]
  46. Provasi, P.F.; Aucar, G.A.; Sauer, S.P.A. The Effect of Lone Pairs and Electronegativity on the Indirect Nuclear Spin-Spin Coupling Constants in CH2X (X = CH2, NH, O, S): Ab Initio Calculations using Optimized Contracted Basis Sets. J. Chem. Phys. 2001, 115, 1324–1334. [Google Scholar] [CrossRef]
  47. Laatikainen, R.; Niemitz, M.; Sundelin, J.; Hassinen, T. PERCH Software©, 1993–1995, PERCH Project; Department of Chemistry, University of Kuopio: Kuopio, Finland, 1995. [Google Scholar]
  48. Helgaker, T.; Jaszuński, M.; Pecul, M. The Quantum Chemical Calculation of NMR Indirect Spin-Spin Coupling Constants. Prog. Nucl. Res. Magn. Reson. Spectrosc. 2008, 53, 249–268. [Google Scholar] [CrossRef]
  49. Krivdin, L.B. Theoretical Calculations of Carbon-Hydrogen Spin-Spin Coupling Constants. Prog. Nucl. Res. Magn. Reson. Spectrosc. 2018, 108, 17–73. [Google Scholar] [CrossRef]
  50. Jackowski, K.; Słowiński, M.A. Searching for the Best Values of NMR Shielding and Spin-Spin Coupling Parameters: CH4-nFn Series of Molecules as the Example. Molecules 2023, 28, 1499. [Google Scholar] [CrossRef]
  51. Jackowski, K.; Jaszuński, M.; Wilczek, M. Alternative Approach to the Standardization of NMR Spectra. Direct Measurement of Nuclear Magnetic Shielding in Molecules. J. Phys. Chem. A 2010, 114, 2471–2475. [Google Scholar] [CrossRef] [PubMed]
  52. Angus, W.R.; Favede, J.; Hoarau, J.; Pacault, A. Zahlenwerte und Funktionen; Landolt-Börnstein Series; Hellwege, K.H., Hellwege, A.M., Eds.; Springer: Berlin, Germany, 1967; Band II, Teil 10; pp. 12, 55, 68. [Google Scholar] [CrossRef]
  53. Laatikainen, R.; Niemitz, M.; Weber, U.; Sundelin, J.; Hassinen, T.; Vepsäläinen, J. General Strategies for Total-Lineshape-Type Spectral Analysis of NMR Spectra Using Integral-Transform Iterator. J. Magn. Reson. Ser. A 1996, 120, 1–10. [Google Scholar] [CrossRef]
  54. Laatikainen, R.; Niemitz, M.; Malaisse, W.J.; Biesemans, M.; Willem, R. A Computational Strategy for the Deconvolution of NMR Spectra with Multiplet Structures and Constraints: Analysis of Overlapping 13C-2H Multiplets of 13C Enriched Metabolites from Cell Suspensions Incubated in Deuterated Media. Magn. Reson. Med. 1996, 36, 359–365. [Google Scholar] [CrossRef] [PubMed]
  55. Gordon, S.; Dailey, B.P. Density versus Chemical Shift in NMR Spectra. J. Chem. Phys. 1961, 34, 1084–1085. [Google Scholar] [CrossRef]
  56. Rummens, F.H.A. Proton Magnetic Resonance of Ethane in the Liquid and Gaseous States. Mol. Phys. 1971, 21, 535–547. [Google Scholar] [CrossRef]
  57. Trappeniers, N.J. Oldenziel, Physica, High-Resolution Nuclear Magnetic Resonance Spectroscopy in Liquids and Gases at Pressures up to 2500 bar: II. Density Dependence of the Proton Magnetic Shielding Constants in the Nonpolar Gases Methane and Ethylene. Physica A 1976, 82, 581–585. [Google Scholar] [CrossRef]
Figure 1. The 125 MHz 13C NMR spectrum of ethane gas doubly enriched in carbon-13 and without proton decoupling at the spectrometer field 11.7 T. The sample contains pure 13C2H6 at an approximate pressure of 40 bar. Pure liquid TMS is an external reference standard for carbon-13 chemical shifts.
Figure 1. The 125 MHz 13C NMR spectrum of ethane gas doubly enriched in carbon-13 and without proton decoupling at the spectrometer field 11.7 T. The sample contains pure 13C2H6 at an approximate pressure of 40 bar. Pure liquid TMS is an external reference standard for carbon-13 chemical shifts.
Molecules 29 04460 g001
Figure 2. The 500 MHz 1H NMR spectrum (the A3A3′ fragment of the A3A′3XX′ spin system) of the same sample as in Figure 1. The enlarged part shows the complex structure of this spectrum due to all the possible spin-spin couplings between 1H and 13C nuclei. Its 50 NMR lines are partially overlapped. Liquid TMS is applied as an external reference standard of proton chemical shifts.
Figure 2. The 500 MHz 1H NMR spectrum (the A3A3′ fragment of the A3A′3XX′ spin system) of the same sample as in Figure 1. The enlarged part shows the complex structure of this spectrum due to all the possible spin-spin couplings between 1H and 13C nuclei. Its 50 NMR lines are partially overlapped. Liquid TMS is applied as an external reference standard of proton chemical shifts.
Molecules 29 04460 g002
Figure 3. The 125 MHz 13C NMR spectrum (at 11.7 T) of ethylene enriched in carbon-13 and without decoupling from protons (the XX′ part of the AA′A″A‴XX′ spin system, with 144 partially overlapping NMR lines). The gas sample contains pure 13C2H4 gas with a pressure of approx. 45 bar. Liquid TMS is used for the measurements of carbon-13 chemical shifts.
Figure 3. The 125 MHz 13C NMR spectrum (at 11.7 T) of ethylene enriched in carbon-13 and without decoupling from protons (the XX′ part of the AA′A″A‴XX′ spin system, with 144 partially overlapping NMR lines). The gas sample contains pure 13C2H4 gas with a pressure of approx. 45 bar. Liquid TMS is used for the measurements of carbon-13 chemical shifts.
Molecules 29 04460 g003
Figure 4. The 500 MHz 1H NMR spectrum of the same sample is shown in Figure 3. It is the AA′A″A‴ part of the AA′A″A‴XX′ spin system with 132 mostly overlapping signals. The proton chemical shifts are measured relative to external liquid TMS.
Figure 4. The 500 MHz 1H NMR spectrum of the same sample is shown in Figure 3. It is the AA′A″A‴ part of the AA′A″A‴XX′ spin system with 132 mostly overlapping signals. The proton chemical shifts are measured relative to external liquid TMS.
Molecules 29 04460 g004
Figure 5. (a,b) Dependence of the 13C and 1H shielding on density observed in pure 1,2-13C-ethane and its binary mixtures with carbon dioxide (CO2) and xenon (Xe) as the gaseous solvents. The measurements were performed relative to liquid TMS accepting its shielding equal to 186.37 ppm for carbon-13 nuclei (a) and 32.815 ppm for protons (b) [13]. The error bars are marked by the size of experimental points. Let us note that the protons’ shielding scale is 5 times expanded compared to the 13C scale for better visibility.
Figure 5. (a,b) Dependence of the 13C and 1H shielding on density observed in pure 1,2-13C-ethane and its binary mixtures with carbon dioxide (CO2) and xenon (Xe) as the gaseous solvents. The measurements were performed relative to liquid TMS accepting its shielding equal to 186.37 ppm for carbon-13 nuclei (a) and 32.815 ppm for protons (b) [13]. The error bars are marked by the size of experimental points. Let us note that the protons’ shielding scale is 5 times expanded compared to the 13C scale for better visibility.
Molecules 29 04460 g005
Figure 6. (a,b) Density-dependent 13C and 1H shielding of 1,2-13C-ethylene for the pure solute and its gaseous mixtures with CO2 and Xe. Liquid TMS (σC = 186.37 ppm, σH = 32.815 ppm) was applied as the external reference standard. The error bars are marked by the size of experimental points.
Figure 6. (a,b) Density-dependent 13C and 1H shielding of 1,2-13C-ethylene for the pure solute and its gaseous mixtures with CO2 and Xe. Liquid TMS (σC = 186.37 ppm, σH = 32.815 ppm) was applied as the external reference standard. The error bars are marked by the size of experimental points.
Molecules 29 04460 g006
Figure 7. (a,b) The 1J(CH) and 1J(CC) isotropic spin–spin couplings in 13C2H6 as density functions at 300 K. Both plots show the negligible influence of intermolecular interactions on the one-bond couplings (1J(CH) in (a) and 1J(CC) in (b)) in ethane-13C2.
Figure 7. (a,b) The 1J(CH) and 1J(CC) isotropic spin–spin couplings in 13C2H6 as density functions at 300 K. Both plots show the negligible influence of intermolecular interactions on the one-bond couplings (1J(CH) in (a) and 1J(CC) in (b)) in ethane-13C2.
Molecules 29 04460 g007
Figure 8. (a,b) The one-bond spin–spin couplings (1J(CH) in (a) and 1J(CC) in (b)) in ethylene-13C2 molecules. Their density dependence on gas density is still minimal but a little more differentiated by solvents than for ethane-13C2.
Figure 8. (a,b) The one-bond spin–spin couplings (1J(CH) in (a) and 1J(CC) in (b)) in ethylene-13C2 molecules. Their density dependence on gas density is still minimal but a little more differentiated by solvents than for ethane-13C2.
Molecules 29 04460 g008
Table 1. Nuclear magnetic shielding of gaseous 1,2-13C-enriched ethane and ethylene measured at a temperature of 300 K from density dependence of the pure compounds and their binary mixtures with CO2 and Xe.
Table 1. Nuclear magnetic shielding of gaseous 1,2-13C-enriched ethane and ethylene measured at a temperature of 300 K from density dependence of the pure compounds and their binary mixtures with CO2 and Xe.
Measured with Gaseous Solvent (B)
Parameter of A MoleculePure A GasCO2Xe
NMR shielding in 13C2H6
σ0(C) [ppm]180.775(2)180.773(2)180.776(3)
σ1(A-B) (C) [ppm mL mol−1] −193(3) *−95(3)−273(4)
σ0(H) [ppm]
σ1(A-B) (H) [ppm mL mol−1]
29.889(2)
−4(3) *
29.885(2)
−3(3)
29.889(2)
3(4)
NMR shielding in 13C2H4
σ0(C) [ppm]
σ1(A-B) (C) [ppm mL mol−1]
64.367(3)
−176(3) *
64.368(2)
−108(3)
62.364(5)
−213(3)
σ0(H) [ppm]
σ1(A-B) (H) [ppm mL mol−1]
25.462(2)
−8(3) *
25.464(2)
−12(3)
25.462(3)
−1(3)
* σ1(A-A) value for the pure solute gas.
Table 2. Spin–spin couplings of gaseous 1,2-13C-enriched ethane and ethylene are measured at 300 K from their density dependences and pure compounds and in binary mixtures with CO2 and Xe.
Table 2. Spin–spin couplings of gaseous 1,2-13C-enriched ethane and ethylene are measured at 300 K from their density dependences and pure compounds and in binary mixtures with CO2 and Xe.
Measured in Gaseous Solvent (B)
Parameter of A MoleculeFor Pure A GasCO2Xe
Spin–spin coupling in 13C2H6
1J0(CH) [Hz]124.98(3)124.97(3)124.97(3)
1JAB(CH) [Hz mL mol−1]9(4) *4(4)9(4)
1J0(CC) [Hz]35.00(2)34.9935.00
1JAB(CC) [Hz mL mol−1]−24(4) *−16(4)−20(4)
2J0(CH) [Hz]−4.79(2)−4.80(2)−4.80(2)
2JAB(CH) [Hz mL mol−1]7(4) *5(3)9(3)
3J0(HH) [Hz]8.08(2)8.09(2)8.09(2)
3JAB(HH) [Hz mL mol−1]−4(4) *−5(3)−6(3)
Spin–spin coupling in 13C2H4
1J0(CH) [Hz]156.03(2)156.03(2)156.04(3)
1JAB(CH) [Hz mL mol−1]22(9) *32(9)19(12)
1J0(CC) [Hz]67.92(2)67.92(2)67.92(3)
1JAB(CC) [Hz mL mol−1]−17(7) *−43(10)1(14)
2J0(CH) [Hz]−2.55−2.54(2)−2.55(2)
2JAB(CH) [Hz mL mol−1]11(4) *9(11)9(10)
2J0(HH) [Hz]2.53(2)2.53(2)2.54(4)
2JAB(HH) [Hz mL mol−1]−39(8) *−31(10)−36(17)
3J0(HH-cis) [Hz]11.81(2)11.81(2)11.81(3)
3JAB(HH-cis) [Hz mL mol−1]−4(7) *4(11)−6(11)
3J0(HH-trans) [Hz]19.18(2)19.19(2)19.18(3)
3JAB(HH-trans) [Hz mL mol−1]−20(8) *−11(10)−17(12)
* JAA value for the pure solute gas, cf. Equation (4).
Table 3. Experimental a and calculated b shielding parameters of ethane and ethylene molecules.
Table 3. Experimental a and calculated b shielding parameters of ethane and ethylene molecules.
13C Shielding [ppm]1H Shielding [ppm]
exp. σ0(13C) calc. σ0(13C)exp. σ0(1H)calc. σ0(1H)
NMR Shielding in 13C2H6
180.775 c 181.9 d29.889 c 30.84 e
180.55 f184.7 g29.894 h30.92 g
180.5 i185.1 j29.87 k31.77 l
181.04 m182.1 n29.97 m31.05 n
(176.628) o185.5 p(29.526) o30.66 p
NMR Shielding in 13C2H4
64.367 c66.5 d25.462 c 26.06 e
64.24 f73.6 g25.468 h26.18 g
64.1 i 60.8 j25.43 k26.69 l
64.55 m58.0 n25.47 m26.08 n
(60.410) o69.4 p(24.979) o25.93 p
a Gas-phase results extrapolated to zero density and corrected to the absolute scale of 13C shielding (σ0 = 0.6 ± 0.9 ppm for 13CO molecule [21]), otherwise as described; b GIAO calculations as described; c this work; d CCSD(T) including vibrational corrections [22]; e MP2 corrected for rovibrational contributions [23]; f originally measured relative to liquid benzene [24] and converted to shielding as described in ref. [13]; g RAS-II calculations [25]; h measured relative to an isolated methane molecule [26] assuming CH4 proton shielding, σ0 = 30.633 ppm [27]; i measured in low-density gas [28] and corrected according to the actual 13C shielding scale [21]; j ref. [28]; k low-density gas [29]; l SCF [30]; m measured relative to an isolated methane molecule at 348K [26] and accepting carbon-13 σ0 (13CH4) = 195.01 ppm [31]; n SCF [21]; o in liquid CHCl3, present work; p from DFT studies [32].
Table 4. Experimental and ab initio calculated spin–spin coupling parameters of 13C-enriched ethane and ethylene molecules.
Table 4. Experimental and ab initio calculated spin–spin coupling parameters of 13C-enriched ethane and ethylene molecules.
Indirect Spin–Spin Coupling [Hz]
ParameterExperimentalCalculated
Spin–spin coupling in 13C2H6
1J(13C,1H)124.98 a, 124.9 b, 125.3 c, 125.190 d,119.8 e, 119.97 f, 127.5 g, 130.06 h,
1J(13C,13C)35.00 a, 34.6 b, 34.4 c, 34.558 d,38.8 e, 35.20 f, 23.6 g, 38.34 h, 32.6 i
2J(13C,1H)−4.79 a, −4.5 c, −4.655 d,−5.3 e, −5.05 f, −4.3 g, −5.33 h
3J(1H,1H)8.08 a, 8.0 c, 8.002 d,7.2 e, 4.03 and 14.88 f, 5.16 h
Spin–spin coupling in 13C2H4
1J(13C,1H)156.03 a, 156.3 b, 156.4 c, 156.302 d, 147.7 e, 150.21 f, 155.3 g, 154.2 j, 163.4 k
1J(13C,13C)67.92 a, 67.6 b, 67.6 c, 67.54 d,70.2 e, 67.53 f, 70.6 g, 70.5 i, 70.1 j, 71.1 k
2J(13C,1H)−2.55 a, −2.4 b, −2.4 c, −2.408 d,−3.3 e, −3.58 f, −1.0 g, −4.1 h, −1.3 k
2J(1H,1H)2.53 a, 2.5 b, 2.5 c, 2.23 d,0.9 e, 0.98 f, 3.4 g, 3.2 j, 0.2 k
3J(1H,1Hcis)11.81 a, 11.7 b, 11.7 c, 11.62 d,10.4 e, 11.79 f, 11.7 g, 11.0 j, 12.8 k
3J(1H,1Htrans)19.18 a, 19.0 b, 19.1 c, 19.02 d,17.0 e, 18.02 f, 18.3 g, 17.7 j, 19.7 k
a nJ0 from the present 13C NMR experiments; b from 1H NMR spectra of 13C-enriched hydrocarbons in CCl4 solutions [9]; c 1H NMR spectra of liquid 13C-enriched hydrocarbons at −70 °C [10]; d a selected example of 13C-enriched hydrocarbons dissolved in liquid crystals (LC), 1H and 13C NMR experiments [25]; e GIAO MCSCF calculations with more examples in ref. [25]; f CCSD [41], the two results for 3J(1H,1H) in ethane are for the two opposite orientations of H atoms in the stable structure of the 13CH2H6 molecule without averaging; g B97-3 [42]; h BHandH [43]; i B3LYP [44]; j B3LYP [45]; k SOPPA(CCSD) [46].
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

Wilczek, M.; Jackowski, K. Gas-Phase Studies of NMR Shielding and Indirect Spin–Spin Coupling in 13C-Enriched Ethane and Ethylene. Molecules 2024, 29, 4460. https://doi.org/10.3390/molecules29184460

AMA Style

Wilczek M, Jackowski K. Gas-Phase Studies of NMR Shielding and Indirect Spin–Spin Coupling in 13C-Enriched Ethane and Ethylene. Molecules. 2024; 29(18):4460. https://doi.org/10.3390/molecules29184460

Chicago/Turabian Style

Wilczek, Marcin, and Karol Jackowski. 2024. "Gas-Phase Studies of NMR Shielding and Indirect Spin–Spin Coupling in 13C-Enriched Ethane and Ethylene" Molecules 29, no. 18: 4460. https://doi.org/10.3390/molecules29184460

Article Metrics

Back to TopTop