Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Origin of 17O NMR chemical shifts based on molecular orbital theory: paramagnetic terms of the pre-α, α and β effects from orbital-to-orbital transitions, along with the effects from vinyl, carbonyl and carboxyl groups

Keigo Matsuzaki, Satoko Hayashi* and Waro Nakanishi*
Faculty of Systems Engineering, Wakayama University, 930 Sakaedani, Wakayama 640-8510, Japan. E-mail: hayashi3@wakayama-u.ac.jp; nakanisi@wakayama-u.ac.jp

Received 2nd February 2024 , Accepted 24th April 2024

First published on 30th April 2024


Abstract

17O NMR chemical shifts (δ(O)) were analysed based on the molecular orbital (MO) theory, using the diamagnetic, paramagnetic and total absolute magnetic shielding tensors (σd(O), σp(O) and σt(O), respectively). O2− was selected as the standard for the analysis. An excellent relationship was observed between σd(O) and the charges on O for O6+, O4+, O2+, O0 and O2−. The data from H2O, HO+, HO and H3O+ were on the correlation line. However, such relationship was not observed for the oxygen species, other than above. The pre-α, α and β effects were evaluated bases on σt(O), where the pre-α effect arises from the protonation to a lone pair orbital on O2−, for an example. The 30–40 ppm and 20–40 ppm (downfield shifts) were predicted for the pre-α and β effects, respectively, whereas the values for the α effect was very small in magnitude, where the effect from the hydrogen bond formation should be considered. Similarly, the carbonyl effect in H2C[double bond, length as m-dash]O and the carboxyl effects in H(HO)C[double bond, length as m-dash]O were evaluated from MeOH, together with H2C[double bond, length as m-dash]CHOH from CH3CH2OH. Very large downfield shifts of 752, 425 and 207 ppm were predicted for H2C[double bond, length as m-dash]O*, H(HO)C[double bond, length as m-dash]O* and H(HO*)C[double bond, length as m-dash]O, respectively, together with the 81 ppm downfield shift for H2C[double bond, length as m-dash]CHO*H. The origin of the effect were visualized based on the occupied-to-unoccupied orbital transitions. As a result, the origin of the 17O NMR chemical shifts (δ(17O)) can be more easily imaged and understand through the image of the effects. The results would help to understand the role of O in the specific position of a compound in question and the mechanisms to arise the shift values also for the experimental scientists. The aim of this study is to establish the plain rules founded in theory for δ(17O), containing the origin, which has been achieved through the treatments.


Introduction

NMR spectra are commonly measured and analysed on a daily basis to determine the structures and/or follow up the reactions. Indeed, 1H and 13C NMR spectroscopy is the most important tool for the purposes, but NMR spectra other than above are also measured on a daily basis.1–4 NMR spectroscopy of 15N, 17O and 19F atoms in the second period, has also been a very important technology in current chemical science research.5–9 Among the nuclei, oxygen is the most abundant chemical element and it will form compounds with any other element, except for some atoms of the Group 18 element. It seems somewhat difficult to form compounds between them. Oxygen is also involved in the various biologically important species, such as amino acids and nucleoacid bases,10–16 together with the materials of high functionalities.17,18 Measurements of 17O NMR spectra in the natural abundance are now much easier by the advances in the spectrometer, irrespective of the very low natural abundance with the spin number of 5/2. As a result, lots of 17O NMR chemical shifts (δ(O)) of oxygen species have been reported thus far, of which values spread over 2500 ppm.

The importance of the NMR spectroscopy is widely recognized, as mentioned above. Experimental chemists usually analyse NMR spectra with the guidance of empirical rules.1,2,9 The empirical rules are very useful for assigning the spectra, however, it is difficult to understand the origin of chemical shifts based on the rules. Indeed, only the chemical shift of the reference species is usually provided in such NMR analysis, but any concept and/or data, that help us to image the origin of the chemical shifts, are not provided. As a result, it is very difficult to visualize the origin of the NMR chemical shifts, especially for experimental scientists, who are not the specialists in this field, including the authors. (They are originally experimental chemists, who use calculations extensively to confirm the causality in the experimental results.) This must be the extreme contrast to the cases of the electronic spectra and the infrared spectra, for example. It is easily come to mind the image of the origin for the spectra. They correspond to the electronic transitions between the occupied and unoccupied energy levels and the transitions between the energy levels of internal vibrations, respectively, in molecules and/or atoms.

Our research interested, therefore the aim of this study, is to establish the plain rules founded in theory for the origin of the 17O NMR chemical shifts for the better understanding of the phenomena. The origin should be visualized based on the specific concepts, such as molecular orbitals (MOs). The plain rules with the origin should be easily imaged and understood by the experimental scientists who are not the specialists. This purpose is given more importance, in this work, than the usual calculations of the NMR parameters, reproducing the observed values accurately and/or to predict well the shift values of unknown target compounds. The results should help to understand the role of O in the specific position of a compound in question and the mechanisms to arise the shift values.

Scheme 1 shows the axes in ROR, used for the analysis, together with some MOs and/or AOs (atomic orbitals). The direction of the p-type lone pair orbital (np(O)) in the symmetric ROR was set to the z-axis, which was perpendicular to the molecular plane, the bisected ∠CROCR direction is set to the x-axis, and that perpendicular to the two is set to the y-axis. In the case of unsymmetric ROR′ (R > R′), the z-axis is set to the direction of np(O), while the y- and x-axes are set appropriately in the plane of O–CR and O–CR′. The axes for the species other than above are shown in the individual figures.


image file: d4ra00843j-s1.tif
Scheme 1 Axes in ROR and ROR′, analysed in this work, along with some orbitals. The atomic orbitals (AOs) of 1s (O) and 2s (O) are not drawn, since they overlap 2pz (O), if illustrated.

The α, β, γ and δ effects are well known as the experimental rules, which correspond to the methyl substitutions in the processes of –O–H → –O–CH3, –O–CH3 → –O–CH2–CH3, –O–CH2–CH3 → –O–CH2–CH2–CH3 and –O–CH2–CH2–CH3 → –O–CH2–CH2–CH2–CH3, respectively. The α, β and γ effects in the 17O NMR chemical shifts are typically found at −40 ppm (upfield shifts), +30 ppm (downfield shifts), −6 ppm (upfield shifts), respectively, with the δ effect being negligibly small, based on the observed values. The α, β and γ effects are analysed based on the MO theory. We have proposed the “pre-α effect” to establish the plain rules and understand the mechanisms in a unified form.19 The “pre-α effect” is defined to originate from the protonation to a lone pair orbital of O (O2− → OH, for example). The pre-α, α, β and γ effects are discussed for δ(17O) in R–17O–R′, where R and R′ are the saturated hydrocarbons. The values for the effects are calculated per unit group (per Me or H). The effects on δ(17O) in the unsaturated moieties are also be discussed, exemplified by the vinyl, carbonyl and carboxyl groups, in this paper. The plain rules, established based on the theory, need to be as simple and easily understood.

The chemical shifts of the respective structures can be theoretically calculated. The origin will be elucidated based on the MO theory. The total absolute magnetic shielding tensors (σt) are used for the analysis, since σt can be calculated with satisfactory accuracy. As shown in eqn (1), σt is decomposed into the diamagnetic and paramagnetic shielding tensors (σd and σp, respectively).20–22 The magnetic shielding tensors consist of three components: σxxm, σyym and σzzm (m = d, p and t). Eqn (2) shows the relationship. As shown in eqn (3), σd is simply expressed as the sum of the contributions over the occupied orbitals (ψi, so is ψj), where the contribution from each ψi to σd (σdi) is proportional to the average inverse distance of electrons from nuclei in ψi, <ri−1> (eqn (4)).23 σp is evaluated by the Coupled-Hartree-Fock (CPHF) method. σp can be decomposed into the contributions from the occupied orbitals or the orbital-to-orbital transitions,24 under the DFT levels. σp is shown in eqn (5), where the contributions from the occupied-to-occupied orbital transitions are neglected.19,23 The process to evaluate σp is highly complex, therefore, σp will be discussed based on the approximate image derived from eqn (6),24 where (εaεi)−1 is the reciprocal orbital energy gap, ψk is the k-th orbital function, [L with combining circumflex]z,N is orbital angular momentum around the resonance nucleus N, and rN is the distance from N.

 
σt = σd + σp (1)
 
σm = (σxxm + σyym + σzzm)/3 (m = d, p and t) (2)
 
image file: d4ra00843j-t1.tif(3)
 
image file: d4ra00843j-t2.tif(4)
 
image file: d4ra00843j-t3.tif(5)
 
image file: d4ra00843j-t4.tif(6)

The NMR chemical shifts of the atoms in the higher periods are predominantly controlled by the σp term. The origin and the mechanisms have been thoroughly analysed, such as for δ(Se).19 Contrary to the atoms in the higher period, the NMR chemical shifts of the atoms in the second period are controlled by both the σd and σp terms. Therefore, the mechanisms such as for δ(O) will be more complex. Here, we discuss the origin and mechanisms for δ(O) based on the MO theory, employing the pre-α, α and β effects, together with the effects from the vinyl, carbonyl and carboxyl groups. Our explanation is intended to clarify the shift values, mainly based on the orbital-to-orbital (ψiψa) transitions, as aforementioned. The earlier investigations on δ(Se) will help to understand δ(O) easier, we believe, due to the similarities in the basic structures of the species consisted of the atoms.19

Methodological details in calculations

Calculations were performed using the Gaussian 09 program package, including GaussView.25 The structures were optimized for various oxygen species with the 6–311++G(3df,3pd) (6D10F) basis set (BSS-A). The structural optimizations were performed at the DFT26–29 (L1) and/or MP2[thin space (1/6-em)]30–32 (L2) levels (L = L1 + L2), after some pre-optimizations. The gauge-independent atomic orbital (GIAO) method33–37 was applied to calculate the absolute magnetic shielding tensors of O [σ(O)]. To examine the level dependence on the σ(O), the σ(O) values were calculated at the various L1 levels of B3LYP,26–29 CAM-B3LYP,38 PBE,39 PBE0,40 LC-ωPBE41 and ωB97X-D42 with BSS-A (L1/BSS-A) and the L2 level. The basis set of def2TZVP43,44 was also applied at the B3LYP level (B3LYP/def2TZVP). The solvent effect of CHCl3 was evaluated with the polarizable continuum model (PCM),45 if necessary. The 6-311+G(3d,3p) (6D10F) basis set (BSS-B) operates similarly well to BSS-A, but the results are not discussed.

A utility program46 was applied to evaluate the contributions from each ψi and/or ψiψa transition. The procedure is explained in Appendix of the ESI. The charge on O (Q(O)) was obtained with the natural population analysis (NPA).47

Results and discussion

Search for suitable level in the calculations: setting the standard for the calculated σt(O) values versus the observed δ(O) values

We will tentatively use σt(O: S) and δ(O: S) as the calculated and observed values, respectively, in this paper, to avoid confusing the discussion, although this notation might not be completely theoretically appropriate. In this case, σt(O: S) and δ(O: S), respectively, stand for the shift values of oxygen species, S.

Before detailed discussion to determine the suitable calculation level in this work, it is necessary to set up the appropriate standard for σt(O: S). The δ(O: H2O) value is taken as the standard for δ(O: S). Therefore, it seems good idea, at first glance, that the σt(O: H2O) value is also taken as the standard for σt(O: S), when the σt(O: S) values are compared directly with the δ(O: S) values. However, this choice will not give good results, since the observed and calculated conditions are very different especially for H2O. Water forms poly-clusters through hydrogen bonds (HBs) in liquid,48 but a single molecule in the gas phase is assumed in the calculation conditions.

To avoid large differences in the chemical shifts, due to the differences between the observed and calculated conditions in water, we selected the δ(O: Me2O) value of −52.50 ppm for the common standard of δ(O: S) and σt(O: S). Namely, δ(O: Me2O) = σt(O: Me2O) = −52.50 ppm is chosen at the common standard for both, where σt(O: Me2O) should be denoted by Δσt(O: Me2O), so σt(O: S) is by Δσt(O: S). The treatment leads Δσt(O: H2O) = 0.00 ppm, fictionally. However, the sign of Δσt(O: S) is basically just the opposite to that of δ(O: S). Therefore, −Δσt(O: S) should be used, instead of Δσt(O: S), for the direct comparison between the calculated and observed values, where δ(O: Me2O) = −52.50 ppm is used as the common standard of both observed and calculated values.

It is now possible to search for the suitable level in this work, after setting up the initial research conditions. The σt(O: S) values for various oxygen species S (ROR + ROR′) were calculated at the DFT levels of B3LYP,26–29 CAM-B3LYP,38 PBE,39 PBE0,40 LC-ωPBE41 and ωB97X-D42 (L1) with BSS-A (L1/BSS-A//L1/BSS-A), together with σd(O: S) and σp(O: S). The MP2 level (L2) is also applied for the calculations. However, only σt(O: S) were obtained at the MP2 level (MP2/BSS-A//MP2/BSS-A). The results are collected in Tables S1–S8 of the ESI. The calculated values are very close with each other.

The −Δσt(O: S) values calculated at the L (=L1 + L2) levels are plotted versus the corresponding δ(O: S), respectively. Fig. 1 shows the plots for S of (ROR + ROR′: the 31 species) at B3LYP. The plot is analysed assuming the linear relationship (y = ax + b: Rc2 (the square of the correlation coefficient)), where (a, b, Rc2) = (0.936, 2.88, 0.982) for the plot in Fig. 1. Similar calculations were performed at various L. Table 1 collects the correlations. Judging from the (a, b, Rc2) values in Table 1, B3LYP, CAM-B3LYP and PBE levels seem suitable for our purpose together with others, the b value seems somewhat larger at PBE, and the a values are less than 0.90 at PBE0, LC-ωPBE and ωB97X-D. The MP2 level gave similar results but Rc2 = 0.934, the poorest value in Table 1. The a value amounts to 0.960 at B3LYP, if the solvent effect of CHCl3 is considered. The results with B3LYP/def2TZVP are shown in entry 9 of Table 1. The a and b values seem very good, whereas Rc2 = 0.926. The differences between observed and calculated values are around 20 ppm in magnitudes for s-BuOMe and s-BuOEt. The B3LYP/BSS-A method is selected for the calculations based on the results. Our aim of this work can be achieved even without the solvent effect in the calculations. The level is most popularly accepted also by the experimental researchers, which is significant for our purposes. Not so different results will be obtained when other levels in Table 1 are applied to the calculations.


image file: d4ra00843j-f1.tif
Fig. 1 Plots of the calculated −Δσt(O: S) versus the observed δ(O: S) (S: ROR + ROR′) at the B3LYP level, with (●) and without (○) the solvent effect of CHCl3.
Table 1 Correlations in the plots of calculated –Δσt(O: S) versus observed δ(O: S) for the ether type oxygen species, S (ROR + ROR′)a,b
Entry Level (L) a b Rc2 N
a Calculated with the GIAO method under L/BSS-A.b Observed data are used for the corresponding species in the plot.c Under the solvent effect of CHCl3.d Calculated with B3LYP/def2TZVP.
1 B3LYP 0.936 2.88 0.982 31
2 CAM-B3LYP 0.911 2.30 0.979 31
3 PBE 0.976 5.43 0.982 31
4 PBE0 0.894 1.55 0.978 31
5 LC-ωPBE 0.845 −2.09 0.979 31
6 ωB97X-D 0.886 −0.07 0.982 31
7 MP2 0.933 1.24 0.934 31
8c B3LYP 0.960 3.36 0.984 31
9d B3LYP 0.929 1.22 0.926 31


Analysis of 17O NMR chemical shifts and the standard species

To determine the suitable standard for the analysis of 17O NMR chemical shifts based on σd(O), σp(O) and σt(O), the values were calculated for O6+, O4+, O2+, O0 and O2− with B3LYP/BSS-A and MP2/BSS-A. Table 2 summarizes the results. The σt(O) values for O6+, O4+ and O2−, calculated with the two methods, were very close to each other. O2− was selected as the standard among the three, after the case of σp(Se).19 It is very favourable to use σp(O: O2−) = 0.0 ppm as a standard, especially for our purpose, although σp(O: O4+) and σp(O: O6+) are also 0.0 ppm. The electronic 1So state of O2− with eight valence electrons by the octet rule and its spherical electron distribution are also favourable for the purpose.
Table 2 Absolute shielding tensors for 17O* (* = 6+, 4+, 2+, 0 and 2−) in the singlet statea
Nuclear Configuration σdB3LYP(O: 1s) σdB3LYP(O: 2s) σdB3LYP(O: 2p) σdB3LYP(O) σpB3LYP(O) σtB3LYP(O) σtMP2(O)
a Calculated by applying the GIAO method under B3LYP/BSS-A and MP2/BSS-A.
O6+ (2s)0(2p)0 272.70 0.00 0.00 272.70 0.00 272.70 272.82
O4+ (2s)2(2p)0 271.45 55.54 0.00 327.00 0.00 327.00 327.09
O2+ (2s)2(2p)2 270.87 49.87 46.41 (×1) 367.15 8382.15 8749.31 6551.47
O0 (2s)2(2p)4 270.67 45.42 39.18 (×2) 394.45 6794.55 7189.01 6010.58
O2− (2s)2(2p)6 270.66 43.73 31.31 (×3) 408.33 0.00 408.33 407.67


Table 3 collects the σd(O), σp(O), σt(O), Δσd(O), Δσp(O) (=σp(O) (since σp(O): O2− = 0 ppm)) and Δσt(O) values for various oxygen species of 1–36, calculated with B3LYP/BSS-A, together with the Q(O) values with NPA. The Δσ*(O: S) (* = d, p and t) values are calculated from O2−, according to Δσ*(O: S) = σ*(O: S) – σ*(O: O2−). The extended conformers are selected for the calculations, since they are less three-dimensionally crowded than others, although others would contribute in some cases (Table S9 of the ESI).

Table 3 The σd(O), σp(O), σt(O), Δσd(O)e, Δσp(O)e and Δσt(O)e values for various oxygen species, 1–36, along with the pre-α, α, β, γ and δ effect and the effects from the vinyl, carbonyl and carboxyl groups, based on Δσt (O)ea,b
Species (sym) Q(O) σd(O) σd(O)) σp(O)c σt(O) σt(O)) Δσd(O)ed Δσp(O)ed Δσt(O)ed Effect
a Calculated with the GIAO method under B3LYP/BSS-A.b Δσ*(O: S) = σ*(O: S) – σ*(O: O2−) (* = d, p and t).c Δσp(O) = σp(O), since (σp(O: O2−) = 0 ppm).d Δσ*(O: S)e = (1/n)(Δσ*(O: S) – Δσ*(O: Se)), see text for n, S and Se.e From EtOH.f From H2C[double bond, length as m-dash]CHOH.g From MeOH.
O2− (1: Oh) −2.000 408.33 (0.00) 0.00 408.33 (0.00) 0.00 0.00 0.00
OH (2: C∞v) −1.372 396.59 (–11.74) −19.56 377.03 (–31.29) −11.74 −19.56 −31.29 Pre-α
MeO (3: C3v) −0.976 415.05 (6.72) −133.83 281.22 (–127.11) 18.46 −114.27 −95.81 α
EtO (4: Cs) −0.938 419.47 (11.15) −290.56 128.91 (–279.41) 4.42 −156.73 −152.31 β
i-PrO (5: Cs) −0.942 423.43 (15.11) −297.32 126.12 (–282.21) 4.19 −81.75 −77.55 β
t-BuO (6: Cs) −0.970 427.54 (19.22) −227.30 200.24 (–208.08) 4.17 −31.16 −26.99 β
H2O (7: C2v) −0.929 392.85 (–15.47) −66.72 326.13 (–82.19) −7.74 −33.36 −41.10 Pre-α
MeOH (8: Cs) −0.740 395.07 (–13.26) −72.87 322.20 (–86.13) 2.21 −6.15 −3.94 α
EtOH (9: Cs) −0.751 398.40 (–9.93) −108.33 290.07 (–118.25) 3.33 −35.46 −32.13 β
i-PrOH (10: C1) −0.752 402.81 (–5.52) −152.24 250.57 (–157.75) 3.87 −39.68 −35.81 β
t-BuOH (11: Cs) −0.759 406.99 (–1.34) −180.47 226.52 (–181.81) 3.97 −35.87 −31.89 β
n-PrOH (12: Cs) −0.747 401.99 (–6.33) −110.17 291.82 (–116.51) 3.59 −1.85 1.74 γ
n-BuOH (13: Cs) −0.747 405.29 (–3.03) −112.68 292.62 (–115.71) 3.30 −2.50 0.80 δ
Me2O (14: C2v) −0.599 396.12 (–12.21) −73.37 322.75 (–85.58) 1.63 −3.32 −1.69 α
EtOMe (15: Cs) −0.604 397.46 (–10.86) −105.43 292.04 (–116.29) 1.35 −32.06 −30.71 β
i-PrOMe (16: C1) −0.614 401.79 (–6.53) −128.26 273.53 (–134.79) 2.84 −27.45 −24.61 β
t-BuOMe (17: Cs) −0.622 405.31 (–3.01) −141.91 263.41 (–144.92) 3.07 −22.85 −19.78 β
n-PrOMe (18: Cs) −0.603 400.83 (–7.49) −105.99 294.84 (–113.48) 3.37 −0.57 2.80 γ
n-BuOMe (19: Cs) −0.600 405.13 (–3.20) −110.00 295.13 (–113.19) 4.30 −4.00 0.29 δ
Et2O (20: C2v) −0.618 396.85 (–11.47) −136.13 260.72 (–147.60) 0.37 −31.38 −31.01 β
i-Pr2O (21: C2) −0.631 401.23 (–7.10) −177.41 223.82 (–184.50) −1.95 −33.70 −35.65 β
t-Bu2O (22: C2) −0.656 393.82 (–14.50) −196.90 196.92 (–211.41) −3.77 −28.97 −32.74 β
n-Pr2O (23: C2v) −0.610 397.47 (–10.86) −132.00 265.47 (–142.86) 0.31 2.07 2.37 γ
n-Bu2O (24: C2v) −0.609 407.03 (–1.29) −140.01 267.02 (–141.30) 4.78 −4.01 0.78 δ
H3O+ (25: C3v) −0.748 397.19 (–11.13) −93.28 303.92 (–104.41) −3.71 −31.09 −34.80 Pre-α
MeH2O+ (26: Cs) −0.624 400.40 (–7.93) −94.92 305.48 (–102.85) 3.21 −1.64 1.56 α
EtH2O+ (27: C1) −0.646 408.30 (–0.02) −132.51 275.80 (–132.53) 7.90 −37.59 −29.68 β
Me3O+ (28: C3v) −0.407 403.21 (–5.12) −106.15 297.05 (–111.27) 2.01 −4.29 −2.29 α
Et3O+ (29: C3) −0.457 397.04 (–11.29) −158.79 238.24 (–170.08) −2.06 −17.55 −19.60 β
OH+ (30: C∞v) 0.480 386.73 (–21.60) 1138.35 1525.08 (1116.76) −21.60 1138.35 1116.76 Pre-α
H2C[double bond, length as m-dash]CHOH (31: Cs) −0.695 402.75 (–5.58) −193.80 208.95 (–199.38) 4.35e −85.47e −81.12e C[double bond, length as m-dash]C
H2C[double bond, length as m-dash]CHOMe (32: Cs) −0.561 402.34 (–5.99) −173.97 228.36 (–179.96) −0.41f 19.83f 19.42f C[double bond, length as m-dash]C
PhOH (33: Cs) −0.700 391.76 (–16.57) −183.66 208.10 (–200.23) −3.31g −110.79g −114.10g C6H5
H2C[double bond, length as m-dash]O (34: C2v) −0.499 404.50 (–3.82) −833.77 −429.27 (–837.59) 9.44g −760.90g −751.46g C[double bond, length as m-dash]O
H(HO)C[double bond, length as m-dash]O* (35: Cs) −0.582 404.48 (–3.84) −506.77 −102.29 (–510.62) 9.42g −433.90g −424.48g OC[double bond, length as m-dash]O*
H(HO*)C[double bond, length as m-dash]O (36: Cs) −0.687 399.82 (–8.51) −284.37 115.44 (–292.88) 4.75g −211.50g −206.75g *OC[double bond, length as m-dash]O


Scheme 2 explains the method to calculate the effects, exemplified by the pre-α, α and β effects. The effects are calculated as Δσt(O: S)e = (1/n)[σt(O: S) – σt(O: Se)], where Se are the starting species to give the effects and n is the factor to make Δσ*(O: S)e per unit group. In the case of the β effect from Me2O to Et2O, Et2O, Me2O and 2 correspond to S, Se and n, respectively, in the equation. The difference of Δσt(O: S) between S = Et2O (σt(O) = 261 ppm) and Me2O (σt(O) = 323 ppm) is −62 ppm, which correspond to the 2β effect (=Δσt(O: S) = σt(O: S) – σt(O: O2−)). The Δσ*(O: S) values are abbreviated by Δ in Scheme 2. Therefore, the β effect in this process is evaluated to be 31 ppm (=Δ/2), for example. The Δσd(O: S)e and Δσp(O: S)e values for the effect are calculated similarly.


image file: d4ra00843j-s2.tif
Scheme 2 Evaluation of the pre-α, α and β effects. The σt(O: S) values in ppm are given in red bold and the differences between the two are by Δ.

The pre-α, α, β, γ and δ effects are calculated, according to the method, so are the vinyl, carbonyl and carboxyl effects. The pre-α, α, β, γ and δ effects are calculated for R-O-R′ (R, R′: saturated hydrocarbons), while the unsaturated moieties of the vinyl, carbonyl and carboxyl effects are calculated from EtOH, MeOH and MeOH, respectively. Table 3 collects the values. Scheme 3 visualizes the effects with the values.


image file: d4ra00843j-s3.tif
Scheme 3 Pre-α, α, β, γ and δ effects, along with the effects from the vinyl, carbonyl and carboxyl groups, on the 17O NMR chemical shifts, calculated with the GIAO method under B3LYP/BSS-A.

Behaviour of σd(O)

The behaviour of the calculated σd(O) values can be understood by considering the two factors derived from eqn (3) and (4). If the number of occupied AOs on O increases, the σd(O) values become larger, whereas the magnitude of each σdi(O: AO) becomes smaller, especially that for the outer AOs. The average distance of the electrons from the nucleus O (ri) in each AO becomes larger due to the increase in electron–electron repulsion if the number of occupied AOs increases. In this case, each < ri−1> (and therefore σdi(O)) in eqn (4) decrease. The σdi(O) values in Table 3 are well understood as the total effect of the two.

To examine the effect of the charge on O (Q(O)), the σd(O) values are plotted versus Q(O) for O6+, O4+, O2+, O0 and O2− (1), as shown in Fig. 2; an excellent correlation by a quadratic function was obtained (y = −1.673x2 − 10.24x + 394.5: Rc2 = 1.000). The results show that the σd(O) values are excellently correlated to Q(O) if the oxygen species has no ligands. The σd(O) values for H2O (7), HO+ (30), HO (2) and H3O+ (25) are also plotted versus Q(O) (see Table 3 for the data). The data points appear on or slightly below the regression curve. The data for HO+ (30) and H3O+ (25) are basically located on the regression curve, and those for H2O (7) and HO (2) are located slightly below the curve. The results show that the H atom(s) on O affect somewhat on σd(O), in addition to the effect on Q(O), although the Q(O) value may change depending on the calculation method.


image file: d4ra00843j-f2.tif
Fig. 2 Plot of σd(O) versus Q(O) for O6+, O4+, O2+, O0 and O2− (1), together with H2O (7), HO+ (30), HO (2) and H3O+ (25).

Analysis of δ(O) based on the MO theory

The behaviour of σd(O: S), where S has at least one alkyl group, is examined, next. Fig. 3 shows the plot of σd(O: S) versus Q(O) for 1–36, other than those in Fig. 2. The σd(O: S) values are analysed separately by the types of S: RO (3–6), ROH (8–13), ROMe (14–19) and ROR (20–24), RH2O+ (26 and 27), and R3O+ (28 and 29), along with others (31–36). Each plot for a type of S appears almost the y-direction, except for S of 31–36. The ranges of σd(O) amount to 15 ppm, while those of Q(O) are very small in each group. The σd(O: S) values become larger in the order of R = Me < Et < i-Pr < t-Bu for RO, ROH and ROMe. The structural dependence appears to control the σd(O: S) values.
image file: d4ra00843j-f3.tif
Fig. 3 Plots of σd(O) versus Q(O) for various oxygen species 1–36, other than those in Fig. 2.

As mentioned above, the magnitudes of Δσd(O: S) are less than 15 ppm for most species in each group of species (see Table 3). However, the magnitudes of Δσd(O: S) are larger than 15 ppm for i-PrO (5: Δσd(O) = 15.1 ppm), t-BuO (6: 19.2 ppm), H2O (7: −15.5 ppm), OH+ (30: −21.6 ppm) and PhOH (33: −16.6 ppm). The first two are the RO type, and the last three are H2O, OH+ and PhOH. The results for OH+ are effectively understood based on Q(O), where the larger magnitude in Δσd(O: S) for OH+ (30) potentially comes from the larger positive Q(O) value (=0.482). The magnitudes of Δσd(O: S) are much smaller than those of Δσp(O: S). The contributions from Δσd(O: S) to Δσt(O: S) are less than 10%, except for OH (2: 37.5%), H2O (7: 18.8%), MeOH (8: 15.4%), Me2O (14: 14.3%) and H3O+ (25: 10.7%). Specifically, Δσp(O: S) contributes predominantly to Δσt(O: S), relative to the case of Δσd(O: S). As a result, 17O NMR chemical shifts can be analysed mainly by Δσp(O: S); however, Δσd(O: S) should be considered when necessary.

Effect of hydrogen bonds on 17O NMR chemical shifts

What is the effect from the hydrogen bonds (HBs) on σd(O), σp(O) and σt(O)? The effect is to be clarified before the detailed discussion of the values. The σd(O), σp(O) and σt(O) values for the various ether monomers (ROR + ROR′), calculated with B3LYP/BSS-A, are collected in Table S1 of the ESI (see also Fig. 1). The σt(O) value of the Me2O dimer is calculated to be only 1.4 ppm downfield of that of the monomer, therefore, the effect of the dimer formation in ROR + ROR′ on δ(O) is considered to be negligible. Namely, the data of the monomers can be used for those of ROR + ROR′.

The σt(O) values are calculated for the monomers and the dimers of ROH and RCOOH, together with the differences in σt(O) between the dimers and the monomers Δσt(O)dm [=σt(O: dimer) – σt(O: monomer)]. The solvent effect of CHCl3 on the σt(O) and Δσt(O)dm values are also calculated. The values are collected in Table S10 of the ESI. Fig. 4 illustrates the monomers and dimers, exemplified by H2O (a) and CH3COOH (b) with the σt(O) (in plain) and Δσt(O)dm (in bold) values in ppm, for the better understanding of the discussion. The dimer formation leads to a downfield shift of 7 ppm for H2O (up to 8 ppm for ROH as shown in Table S10 of the ESI) and a upfield shift of 49 ppm for C[double bond, length as m-dash]O* and a downfield shift of 19 ppm for C–O*–H (totally upfield shift by 15 ppm on average) in RCOOH. The analysis for RCOOH would be more complex, since only the averaged data are available due to the interconversion between topological isomers of RCO*OH and RCOO*H. The contribution from HB formation to δ(O) is well demonstrated, although the direction of the effect may depend on the structures (conformers) of the monomers and dimers.


image file: d4ra00843j-f4.tif
Fig. 4 Illustration of monomers and dimers for H2O (a) and CH3COOH (b). The σt(O) (in plain) and Δσt(O)dm (in bold) values are also shown in ppm.

Fig. 5 shows the plot of −Δσt(O: S) versus δ(O: S) for the monomers and the dimers of ROH, with and without considering the solvent effect of CHCl3. Table 4 collects the correlations (entries 1N, 2N, 3Y and 4Y). The correlations seem (very) good. They are very similar with each other, especially for the dimers, with and without considering the solvent effect. The apparent solvent effect on δ(O: S) seems very small, especially for the dimers. The results may show that the monomers and dimers exist (as in equilibrium) in solutions, which controls δ(O: S) and the solvent effect in ROH. Similarly, −Δσt(O: S) are plotted versus δ(O: S) for the RCOOH monomers and the dimers, with and without considering the solvent effect, although not shown n a figure. The correlations are shown in Table 4 (entries 5N, 6N, 7Y and 8Y). The correlations become better in the order of (RCOOH monomer: with the solvent effect) ≈ (RCOOH monomer: without the solvent effect) ≪ (RCOOH dimer: without the solvent effect) ≈ (RCOOH dimer: with the solvent effect). The dimer formation seems very important in RCOOH, relative to the case of ROH, together with the considering the solvent effect.


image file: d4ra00843j-f5.tif
Fig. 5 Plots of calculated −Δσt(O: S) versus observed δ(O: S) for some monomeric (I) and dimeric (II) alcohols with and without the solvent effect of CDCl3: I with (○) and without (●) the solvent effect and II with (■) and without (□) the solvent effect.
Table 4 Correlations in the plots of calculated −Δσt(O: S) versus observed δ(O: S) for the monomers and dimers of ROH and ROOH, with and without considering the solvent effect of CHCl3 under B3LYP/BSS-Aa
Entryb Plot for a b Rc2 N
a Observed data are used for the corresponding species in the plot.b The solvent effect is specified by N (no solvent effect) or Y (solvent effect) after the entry number.
1N ROH monomers 0.921 −19.26 0.988 9
2N ROH dimers 0.920 −16.65 0.991 9
3Y ROH monomers 0.914 −21.44 0.989 9
4Y ROH dimers 0.921 −16.92 0.991 9
5N RCOOH monomers 0.939 27.29 0.929 5
6N RCOOH dimers 1.072 −19.76 0.968 5
7Y RCOOH monomers 0.829 48.25 0.928 5
8Y RCOOH dimers 0.966 4.03 0.960 5


After confirming the basic behaviour of σt(O) for ROR + ROR′, ROH and RCOOH, next extension is to clarify the origin of δ(O) based on the MO theory. The pre-α, α and β effects, along with the vinyl, carbonyl and carboxyl effects, are analysed using an approximated image, derived from eqn (6).24

Origin of the pre-α effect

How are the 17O NMR chemical shifts originated? electrons around a nucleus 17O shield the external magnetic field at the nucleus. The spherical component of the electron distribution arises the diamagnetic terms σd(O), whereas the paramagnetic terms σp(O) are originated from the unsymmetrical component of the electron distribution. In the case of O2−, only σd(O) occurs, since the ten electrons in O2− spherically distribute.

The protonation of O2− yields HO, which introduces the σ(O–H) and σ*(O–H) orbitals, resulting in the unsymmetrical distribution of electrons in HO. The spherical electron distribution of O2− changes to an unsymmetrical distribution in HO, in this process. As a result, the unsymmetrical component produces σp(O), although the spherical component arises σd(O) in HO. The σp(O) terms are caused through the orbital-to-orbital transitions, such as the ψiψa transition, where σ(O–H) and σ*(O–H) operate as the typical ψi and ψa, respectively, in the ψiψa transition.

We focused our attention to the protonation process on O2− in the NMR analysis as the factor to originate σp(O). We proposed to call this process the pre-α effect, when the origin of the 77Se NMR chemical shifts were discussed based on σp(Se).19 The pre-α effect is very important, since it is the starting point to image the origin of all NMR chemical shifts.

As shown in Scheme 3, the pre-α effect is evaluated by the (Δσd(O)e, Δσp(O)e, Δσt(O)e) values, which are (−11.7, −19.6, −31.3 ppm), (−7.7, −33.4, −41.1 ppm) and (−3.7, −31.1, −34.8 ppm) for the processes from O2− to HO, H2O and H3O+, respectively. The values are calculated per unit group (per H in this case). The Δσt(O)e values are all negative, along with Δσd(O)e and Δσp(O)e; therefore, the pre-α effect is theoretically predicted to be the downfield shifts of 31–41 ppm (Δσt(O)e) (see also Table 3). The saturation effect in the pre-α effect on σd(O), σp(O) and σt(O) by the increase of the H atoms seems not so severe in this case. Table 5 lists the σdi(O), σpi(O) and σti(O) (=σdi(O) + σpi(O)) values for O2−, HO, H2O and H3O+, which are separately by ψi. The 1s (O) AO, in the MOs, predominantly contribute to σd(O) for each species, whereas the 2s (O), 2px (O), 2py(O) and 2pz(O) AOs do much smaller to σd(O), as expected. As shown in Table 5, ψ3 greatly contributes to σp(O) (σp3(O) = −85.2 ppm) for HO, along with ψ4 (−21.9 ppm) and ψ5 (−21.9 ppm). For H2O, ψ3 (σp3(O) = −50.2 ppm), ψ4 (−57.2 ppm) and ψ5 (−63.6 ppm) greatly contribute to σp(O). In the case of H3O+, ψ3 (σp3(O) = −43.3 ppm), ψ4 (−43.3 ppm) and ψ5 (−61.9 ppm) greatly contribute to σp(O). The three orbitals must mainly be constructed by the 2px(O), 2py(O) and 2pz(O) AOs.

Table 5 The σd(O), σp(O) and σt(O) values contributed from each MO of O2− (1), HO (2), H2O (7) and H3O+ (25)a
MO (i in ψi) σdi(O) σpi(O) σti(O)
a Calculated with the GIAO method under B3LYP/BSS-A.b The ψ1, ψ2, ψ3, ψ4 and ψ5 MOs of O2− correspond to 1s (O), 2s (O), 2px (O), 2py (O) and 2pz (O) AOs, respectively.c The σd1(O), σd2(O), σd3(O) and σd4(O) values of O0 are evaluated to be 270.67, 45.42, 39.18 and 39.18 ppm, respectively.
O2− (1: Oh)b,c
1 270.67 0.00 270.67
2 43.73 0.00 43.73
3 31.31 0.00 31.31
4 31.31 0.00 31.31
5 31.31 0.00 31.31
Total 408.33 0.00 408.33
[thin space (1/6-em)]
HO (2: C∞v)
1 270.64 0.00 270.64
2 39.36 −4.99 34.37
3 17.78 −85.15 −67.36
4 34.41 −21.92 12.49
5 34.41 −21.92 12.49
ψocc to ψocc   114.41  
Total 392.85 −19.56 377.03
[thin space (1/6-em)]
H2O (7: C2v)
1 270.61 0.00 270.61
2 38.82 −5.24 33.58
3 18.85 −50.23 −31.38
4 27.35 −57.23 −29.65
5 37.22 −63.55 −26.33
ψocc to ψocc   109.30  
Total 392.85 −66.72 326.13
[thin space (1/6-em)]
H3O+ (25: C3v)
1 270.60 0.00 270.60
2 40.44 −1.93 38.50
3 23.69 −43.31 −19.62
4 23.69 −43.30 −19.61
5 38.77 −61.90 −23.13
ψocc to ψocc   57.17  
Total 397.19 −93.28 303.91


Table 6 shows the ψiψa transitions predominantly contributing to σpia:xx(O), σpia:yy(O) and/or σpia:zz(O) for HO and H2O, where the three components yield σpia(O), according to eqn (2). The magnitudes larger than 6 ppm for σpia(O) are provided in Table 6. (The border value for the positive σpia(O) values to list the table is usually not specified, since the positive values contribute to the diamagnetic direction.) The ψ3ψ9 (σp[3 with combining right harpoon above (vector)]9:xx(O) = −83.9 ppm), ψ3ψ10 (σp[3 with combining right harpoon above (vector)]10:zz(O) = −83.9 ppm), ψ4ψ8 (σp[4 with combining right harpoon above (vector)]8:xx(O) = −61.1 ppm) and ψ5ψ8 (σp[5 with combining right harpoon above (vector)]8:zz(O) = −61.1 ppm) transitions greatly contribute to σpia(O) in HO. In the case of H2O, the ψ3ψ8 (σp[3 with combining right harpoon above (vector)]8:zz(O) = −28.1 ppm), ψ3ψ11 (σp[3 with combining right harpoon above (vector)]11:xx(O) = −32.2 ppm), ψ4ψ9 (σp[4 with combining right harpoon above (vector)]9:zz(O) = −40.9 ppm), ψ5ψ8 (σp[5 with combining right harpoon above (vector)]8:yy(O) = −33.4 ppm) and ψ5ψ9 (σp[5 with combining right harpoon above (vector)]9:xx(O) = −58.8 ppm) transitions greatly contribute to σp(O) (see Table 6).

Table 6 Main contributions from the occupied-to-unoccupied orbital transitions on σp(O) for HO (2) and H2O (7)a
iab σpia:xx(O) σpia:yy(O) σpia:zz(O) σpia(O)
a Calculated with the GIAO method under B3LYP/BSS-A. The magnitudes of σpia(O) larger than 6 ppm are shown.b In ψiψa.
HO (2: C∞v)
3 → 9 −83.87 0.00 0.00 −27.96
3 → 10 0.00 0.00 −83.87 −27.96
3 → 22 0.00 0.00 −18.14 −6.05
3 → 23 −18.14 0.00 0.00 −6.05
4 → 6 21.65 0.00 0.00 7.22
4 → 7 40.70 0.00 0.00 13.57
4 → 8 −61.09 0.00 0.00 −20.36
4 → 14 −35.47 0.00 0.00 −11.82
5 → 6 0.00 0.00 21.65 7.22
5 → 7 0.00 0.00 40.70 13.57
5 → 8 0.00 0.00 −61.09 −20.36
5 → 14 0.00 0.00 −35.47 −11.82
[thin space (1/6-em)]
H2O (7: C2v)
3 → 8 0.00 0.00 −28.12 −9.37
3 →11 −32.22 0.00 0.00 −10.74
4 → 9 0.00 0.00 −40.87 −13.62
4 → 11 0.00 −21.09 0.00 −7.03
4 → 13 0.00 0.00 −18.69 −6.23
4 → 17 0.00 0.00 −25.65 −8.55
5 → 6 0.00 −23.27 0.00 −7.76
5 → 8 0.00 −33.37 0.00 −11.12
5 → 9 −58.81 0.00 0.00 −19.60
5 → 17 −32.49 0.00 0.00 −10.83
5 → 18 0.00 −18.91 0.00 −6.30
5 → 21 −18.69 0.00 0.00 −6.23


Fig. 5 and 6 illustrate the selected ψiψa transitions for HO and H2O, respectively, along with the characteristics of ψi and ψa and the orbital energies. Fig. 5 shows the ψ3ψ9 and ψ3ψ10 transitions in HO, which correspond to the transitions from the occupied σ(O–H) orbital to the vacant 3pz and 3px orbitals, respectively, where 3pz and 3px are equivalent in HO. The ψ4ψ8 and ψ5ψ8 transitions correspond to the transitions from the occupied 2pz and 2px orbitals to the vacant orbitals containing the σ*(O–H) character, respectively. The occupied σ(O–H) and vacant σ*(O–H) orbitals operate as the typical donor and acceptor orbitals, respectively, in the transitions to produce the σpia(O) terms.


image file: d4ra00843j-f6.tif
Fig. 6 Main contributions from each ψiψa transition to the components of σp(O) in HO (2).

The σ(O–H) and σ*(O–H) orbitals in H2O similarly act as the typical donor and acceptor orbitals, respectively, according to the C2v symmetry of H2O, as shown in Fig. 6. The ψ3 (B2)→ψ8 (A1) and ψ3 (B2)→ψ11 (B1) transitions correspond to the occupied σ(H–O–H) orbital to the vacant orbitals containing the σ*(H–O–H) and 3pz(O) characters, respectively. While the ψ4 (A1)→ψ9 (B2) transition corresponds to the occupied ns(O) orbital to the vacant orbital containing the σ*(H–O–H) character, ψ5 (B1) in the ψ5 (B1)→ψ8 (A1) and ψ5 (B1)→ψ9 (B2) transitions has the characters of the occupied np(O) (2pz(O)) orbital. As observed, the σ(O–H) orbitals in H2O act as the typical donors in the combined form of C2v, together with 2pz(O), while the σ*(O–H) orbitals operate as the typical acceptors in the transition, although the character seems to fractionalize to some vacant orbitals, containing the higher 3pz(O) orbital (Fig. 7).


image file: d4ra00843j-f7.tif
Fig. 7 Main contributions from each ψiψa transition to the components of σp(O) in H2O (7).

Origin of the α effect

The α effect is evaluated for MeOH, Me2O and MeO+H2, using (Δσd(O)e, Δσp(O)e, Δσt(O)e), of which values are (2.2, −6.2, −3.9 ppm), (1.6, −3.3, −1.7 ppm) and (3.2, −1.6, −1.6 ppm), respectively. The magnitudes of Δσt(O)e are small in magnitudes (less than 4 ppm). The signs of Δσd(O)e and Δσp(O)e are just inverse, where Δσt(O)e = Δσd(O)e + Δσp(O)e; this is likely the reason for the small α effect predicted based on Δσt(O)e. In the case of MeO from OH, the (Δσd(O)e, Δσp(O)e, Δσt(O)e) values are (18.5, −114.3, −95.8 ppm). The large magnitude for Δσt(O)e comes from the large magnitude of Δσp(O)e, where the negative charge on MeO would contribute to the results.

The large upfield shifts observed in ROH as the α effect appear to be difficult to explain based on the calculated Δσt(O)e values, under the calculation conditions employed in this work. The contribution from HB formation and/or the solvent effect under the observed conditions would be responsible for this.

Table 7 lists the σdi(O), σpi(O) and σti(O) (=σdi(O) + σpi(O)) values, separately by ψi, for Me2O. The inner orbital of ψ1 is constructed by the 1s (O) AO; therefore, it greatly contributes to σd(O) but does not contribute to σp(O). Those of ψ2 and ψ3 are constructed by the two 1s (C) AOs; therefore, the contributions to σd(O) and σp(O) are very minimal. ψ5, ψ6, ψ10 and ψ11 are mainly constructed by the 2s (C) and 2p (C) AOs; therefore, the contributions to σd(O) and σp(O) are also minimal. The contributions from ψ7ψ9 and ψ12 to σpi(O) are large (−31 to −69 ppm), where ψ7ψ9 and ψ12 are mainly formed by the 2p (O) AOs. The contributions from ψ4 and ψ13 to σpi(O) are −13.1 and −17.9 ppm, respectively, where ψ4 and ψ13 are mainly constructed by both 2s (O) and 2p (O) AOs. As shown in Table 8, the ψ8ψ34 (σp[8 with combining right harpoon above (vector)]34:zz(O) = −44.9 ppm), ψ9ψ34 (σp[9 with combining right harpoon above (vector)]34:xx(O) = −48.0 ppm), ψ12ψ37 (σp1[2 with combining right harpoon above (vector)]37:zz(O) = −69.4 ppm) and ψ12ψ51 (σp1[2 with combining right harpoon above (vector)]51:zz(O) = −51.6 ppm) transitions greatly contribute to the components of σp(O) in Me2O.

Table 7 The σd(O), σp(O) and σt(O) values contributed from each MO of Me2O (14: C2v)a
MO (i in ψi) σdi(O) σpi(O) σti(O)
a Calculated with the GIAO method under B3LYP/BSS-A.
1 270.62 0.00 270.62
2,3 0.06 0.12 0.18
4 33.03 −13.07 19.96
5 9.48 4.29 13.77
6 9.14 1.85 10.99
7 11.54 −31.13 −19.59
8 9.46 −40.02 −30.56
9 10.83 −47.49 −36.66
10 −1.00 −2.70 −3.71
11 0.33 7.79 8.11
12 13.25 −68.58 −55.33
13 29.39 −17.93 11.46
ψocc to ψocc   133.50  
Total 396.11 −73.37 322.75


Table 8 Main contributions from occupied-to-unoccupied orbital transitions on σp(O) of Me2O (14: C2v)a,b
iab σpia:xx(O) σpia:yy(O) σpia:zz(O) σpia(O)
a Calculated with the GIAO method under B3LYP/BSS-A. The magnitudes of σpia(O) larger than 6 ppm are shown.b In ψiψa.
7 → 30 0.00 0.00 −25.63 −8.54
8 → 34 0.00 0.00 −44.88 −14.96
8 → 37 0.00 0.00 −22.04 −7.35
9 → 34 −47.96 0.00 0.00 −15.99
11 → 28 0.00 0.00 18.47 6.16
11 → 30 0.00 0.00 25.08 8.36
12 → 15 0.00 0.00 −22.81 −7.60
12 → 26 0.00 0.00 −37.56 −12.45
12 → 29 0.00 −21.18 0.00 −7.06
12→37 0.00 0.00 −69.38 −23.13
12 → 38 0.00 0.00 −18.72 −6.24
12 → 51 0.00 0.00 −51.56 −17.19
13 → 14 0.00 −29.14 0.00 −9.71
13 → 15 −45.67 0.00 0.00 −15.22
13 → 18 0.00 24.35 0.00 8.12
13 → 23 34.67 0.00 0.00 11.56
13 → 26 −51.91 0.00 0.00 −17.30
13 → 28 0.00 60.24 0.00 20.08
13 → 30 0.00 −51.05 0.00 −17.02
13 → 34 67.28 0.00 0.00 22.43
13 → 39 0.00 −25.91 0.00 −8.64
13 → 51 −39.27 0.00 0.00 −13.09
13 → 55 38.17 0.00 0.00 12.72


Fig. 8 shows the selected ψiψa transitions in Me2O; these are considered to be the effective transitions. Both occupied and vacant orbitals extend over the entire molecule. Whereas ψ12 (HOMO−1) of the ns(O) type acts as a good donor in Me2O, the vacant orbitals around ψ14 (LUMO) do not operate as the effective acceptors in the transitions. The high electronegativity of O, relative to C, potentially prevents the contribution of 2p (O) in the vacant orbitals around the LUMO. AOs on the higher electronegative atoms are tend to contribute in the occupied MOs but not in the vacant MOs. The large contributions from the vacant orbitals around LUMO to Δσp(O)e are predicted for the formation of MeO from HO, where the high electronegativity of O would be relaxed by the negative charge.


image file: d4ra00843j-f8.tif
Fig. 8 Main contributions from each ψiψa transition to the components of σp(O) in Me2O (14: C2v).

Origin of the β effect

The β effect is discussed first for ROH, ROMe and ROR, where R changes from Me to Et, then i-Pr, and then t-Bu. The calculated (Δσd(O)e, Δσp(O)e, Δσt(O)e) values are (−0.4 ∼ 4.7, −43.4 ∼ −20.6, −38.8 ∼ −19.8 ppm) for the processes (see Table 3 and Scheme 3). The magnitudes of Δσd(O)e are less than 5 ppm (usually positive), while the Δσp(O)e and Δσt(O)e values are approximately −40 ∼ −20 ppm. Specifically, the β effect is recognized as the downfield shift of 40∼20 ppm, based on the calculations; this effectively explains the observed effect. Similar results are predicted for the processes from MeOH2+ to EtOH2+σt(O)e = −29.7 ppm) and from Me3OH+ to Et3OH+σt(O)e = −19.6 ppm). In the case of the processes from MeO to EtO, then i-PrO and then t-BuO, the (Δσd(O)e, Δσp(O)e, Δσt(O)e) values are (4.4, −156.7, −152.3 ppm), (4.2, −81.8, −77.6 ppm) and (4.2, −31.2, −27.0 ppm), respectively. The magnitudes of Δσp(O)e and Δσt(O)e decrease in the order of CH3 > CH2 > CH, of which H is substituted by Me. The large negative values of Δσp(O)e (and Δσt(O)e) lead to the large β effect in EtO and i-PrO, while the effect for t-BuO appears normal.

Table 9 lists the σd(O), σp(O) and σt(O) values, separately by ψi, exemplified by Et2O (C2v). The contributions from ψ11, ψ12, ψ14, ψ20 and ψ21 to σpi(O) are large (−29.1 – −59.3 ppm). Table 10 shows the main ψiψa transitions, contributing to σpia:xx(O), σpia:yy(O), or σpia:zz(O). The main transitions are ψ14ψ57 (σp1[4 with combining right harpoon above (vector)]57:zz(O) = −31.2 ppm), ψ14ψ88 (σp1[4 with combining right harpoon above (vector)]88:zz(O) = −43.8 ppm), ψ20ψ83 (σp2[0 with combining right harpoon above (vector)]83:zz(O) = −37.8 ppm), ψ20ψ85 (σp2[0 with combining right harpoon above (vector)]85:zz(O) = −37.5 ppm), ψ21ψ22 (σp2[1 with combining right harpoon above (vector)]22:yy(O) = −53.9 ppm), ψ21ψ37 (σp2[1 with combining right harpoon above (vector)]37:yy(O) = −36.6 ppm), ψ21ψ44 (σp2[1 with combining right harpoon above (vector)]44:yy(O) = −30.1 ppm) and ψ21ψ57 (σp2[1 with combining right harpoon above (vector)]57:xx(O) = −35.0 ppm), together with ψ21ψ54 (σp2[1 with combining right harpoon above (vector)]54:yy(O) = 52.0 ppm) and ψ21ψ58 (σp2[1 with combining right harpoon above (vector)]58:xx(O) = 49.1 ppm), which contribute to the diamagnetic direction.

Table 9 The σd(O), σp(O) and σt(O) values contributed from each MO of Et2O (20: C2v)a
MO (i in ψi) σdi(O) σpi(O) σti(O)
a Using the GIAO method under B3LYP/BSS-A.
1 270.61 0.00 270.61
2–5 0.11 0.25 0.35
6 33.53 −8.62 24.91
7 8.26 3.14 11.39
8 5.77 −0.96 4.81
9 3.90 −6.62 −2.71
10 8.46 −0.72 7.74
11 6.58 −59.26 −52.68
12 8.50 −34.89 −26.40
13 0.79 −2.06 −1.28
14 4.31 −36.72 −32.41
15 0.08 −12.83 −12.74
16 2.96 −20.13 −17.17
17 4.94 −13.49 −8.56
18 −0.22 −0.06 −0.28
19 −2.57 −4.12 −6.69
20 12.47 −52.83 −40.36
21 28.38 −29.12 −0.74
ψocc to ψocc   142.90  
Total 396.85 −136.13 260.72


Table 10 Main contributions from the ψoccψunocc transitions on σp(O) in Et2O (20: C2v)a
iab σpia:xx(O) σpia:yy(O) σpia:zz(O) σpia(O) iab σpia:xx(O) σpia:yy(O) σpia:zz(O) σpia(O)
a Calculated with the GIAO method under B3LYP/BSS-A. The magnitudes of σpia(O) larger than 6 ppm are shown.b In ψiψa.
11 → 34 0.00 0.00 −19.97 −6.66 20 → 65 0.00 0.00 −19.34 −6.45
11 → 64 0.00 0.00 −18.68 −6.23 20 → 83 0.00 0.00 −37.75 −12.58
12 → 57 −24.38 0.00 0.00 −8.13 20 → 85 0.00 0.00 −37.45 −12.48
12 → 58 −21.43 0.00 0.00 −7.14 21 → 22 0.00 −53.89 0.00 −17.96
14 → 57 0.00 0.00 −31.15 −10.38 21 → 37 0.00 −36.58 0.00 −12.19
14 → 58 0.00 0.00 −29.21 −9.74 21 → 44 0.00 −30.09 0.00 −10.03
14 → 88 0.00 0.00 −43.76 −14.59 21→54 0.00 52.02 0.00 17.34
15 → 58 0.00 0.00 −23.00 −7.67 21 → 55 0.00 −19.12 0.00 −6.37
16 → 58 −30.60 0.00 0.00 −10.20 21 → 57 −34.98 0.00 0.00 −11.66
17 → 51 −17.84 0.00 0.00 −5.95 21 → 58 49.06 0.00 0.00 16.35
17→ 54 0.00 0.00 −24.15 −8.05 21 → 64 0.00 −28.92 0.00 −9.64
20 → 51 0.00 −29.16 0.00 −9.72 21 → 83 −29.31 0.00 0.00 −9.77


Fig. 9 draws the selected ψiψa transitions in Et2O, together with the characters of ψi and ψa and the orbital energies. It is expected to clarify the mechanisms for the β effect. Similar to the case of Me2O, the occupied and vacant orbitals in Et2O extend over the whole molecule. It is also curious that the vacant orbitals around ψ22 (LUMO) do not operate effectively as acceptors in the transitions. However, the ethyl groups in Et2O seem to play an important role in the (large) β effect, contrary to the case of the Me groups in Me2O, which seem not to play an important role in the α effect, for example.


image file: d4ra00843j-f9.tif
Fig. 9 Main contributions from each ψiψa transition to the components of σp(O) in Et2O (20).

In the case of Et2O, ψ11, ψ12, ψ14, ψ20 and ψ21 contribute to σp(O), over ∼30 ppm in magnitude and the σp(O) value is −136.1 ppm, as the total contribution. The contributions in Et2O are compared with those in Me2O and H2O. The ψ7, ψ8, ψ9 and ψ12 orbitals in Me2O contribute to σp(O), over 30 ppm in magnitude and the σp(O) values −73.4 ppm, as the total contribution. In the case of H2O, ψ3, ψ4 and ψ5 contribute to σp(O), over 50 ppm in magnitude, which leads to the total contribution of σp(O) of −66.7 ppm. The σp(O) values of (Me2O from H2O) and (Et2O from Me2O) are calculated to be −3.3 and −31.4 ppm (per Me), respectively. The values correspond to the minimal α effect in Me2O and the large β effect in Et2O, in magnitudes, based on the calculations. The minimal α effect potentially originates from the cancelling of many (complex) transitions to produce σp(O), while this cancelling would be avoided in the β effect.

Origin of the γ and δ effects

The upfield shifts of 1.7, 2.8 and 2.4 ppm by Δσt(O)e were predicted for the γ effect in the formation of n-PrOH from EtOH, n-PrOMe from EtOMe and n-Pr2O from Et2O, respectively. Similarly, the upfield shifts of 0.8, 0.3 and 0.8 ppm by Δσt(O)e are for the δ effect in n-BuOH formed from n-PrOH, n-BuOMe from n-PrOMe and n-Bu2O from n-Pr2O, respectively. The predicted magnitudes of Δσt(O)e are very small. The magnitudes of Δσd(O)e and Δσp(O)e are also very small, and the signs are the inverse to each other. The mechanisms for the γ and δ effects are not analysed further, due to the negligibly small magnitudes.

Effect from the vinyl group

Large downfield shifts in δ(17O) (∼80 ppm) are reported for vinyl ethers.49 The effect from the vinyl group is calculated, exemplified by the process from EtOH to H2C[double bond, length as m-dash]CHOH (Cs), although the process can be diversely described. The (Δσd(O)e, Δσp(O)e, Δσt(O)e) values for the process are calculated to be (4.4, −85.5, −81.1 ppm), which effectively reproduced the observed results.

Table 11 lists the σd(O), σp(O) and σt(O) values of H2C[double bond, length as m-dash]CHOH (Cs), separately by ψi. The contributions from ψ7, ψ10 and ψ11 to σpi(O) are (very) large, of which values are −48.0, −67.3 and −56.9 ppm, respectively. As shown in Table 12, the ψ10ψ30 (σp1[0 with combining right harpoon above (vector)]30:xx(O) = −21.3 ppm and σp1[0 with combining right harpoon above (vector)]30:yy(O) = −28.7 ppm) and ψ11ψ14 (σp1[1 with combining right harpoon above (vector)]14:xx(O) = −175.5 ppm) transitions provide great contributions.

Table 11 The σd(O), σp(O) and σt(O) values of H2Ce001CHOH (31: Cs), given separately by each ψia
MO (i in ψi) σdi(O) σpi(O) σti(O)
a Using the GIAO method under B3LYP/BSS-A.
1 270.61 0.00 270.61
2 0.02 0.05 0.08
3 0.01 0.01 0.02
4 36.10 −6.49 29.61
5 5.35 4.45 9.80
6 10.22 −14.77 −4.55
7 9.55 −48.00 −38.45
8 9.56 −23.28 −13.71
9 6.00 −18.04 −12.04
10 28.45 −67.27 −38.82
11 17.59 −56.91 −39.32
12 9.29 4.93 14.22
ψocc to ψocc   31.50  
Total 402.75 −193.80 208.95


Table 12 Main contributions from occupied-to-unoccupied orbital transitions on σp(O) of H2C[double bond, length as m-dash]CHOH (31: Cs)a
iab σpia:xx(O) σpia:yy(O) σpia:zz(O) σpia(O)
a Calculated with the GIAO method under B3LYP/BSS-A. The magnitudes of σpia(O) larger than 6 ppm are shown.b In ψiψa.
6 → 14 −0.15 −34.97 0.00 −11.70
7 → 29 −0.29 −21.78 0.00 −7.36
7 → 30 0.00 0.00 −18.56 −6.19
8 → 14 19.02 −54.77 0.00 −11.92
10 → 30 −21.27 −28.67 0.00 −16.65
10 → 31 0.25 −23.28 0.00 −7.68
11 → 13 0.00 0.00 −23.14 −7.71
11 → 14 −175.47 1.34 0.00 −58.04
11 → 30 0.00 0.00 44.83 14.94
11 → 46 0.00 0.00 −27.56 −9.19
12 → 13 5.81 −23.50 0.00 −5.90
12 → 30 15.83 32.73 0.00 16.18


Fig. 10 illustrates the ψiψa transitions in H2C[double bond, length as m-dash]CHOH (Cs) with the axes. The main characters of ψ10, ψ11, ψ14 and ψ30 are the occupied π(C[double bond, length as m-dash]C–O), occupied ns(O), vacant π*(C[double bond, length as m-dash]C–O) and vacant σ*(C[double bond, length as m-dash]C–O) orbitals, respectively, and they extend over the entire molecule. For the large σp(O) values in H2C = CHOH (Cs), ψ10 (π(C[double bond, length as m-dash]C–O)) and ψ11(ns(O)) act as excellent donors, while ψ14 (π*(C[double bond, length as m-dash]C–O)) and ψ30 (σ*(C[double bond, length as m-dash]C–O)) operate as good acceptors. In particular, the ψ11 (HOMO−1)→ψ14 (LUMO+1) transition greatly contributes to σp1[1 with combining right harpoon above (vector)]14(O) of −58.0 ppm.


image file: d4ra00843j-f10.tif
Fig. 10 Main contributions from each ψiψa transition to the components of σp(O) in H2C[double bond, length as m-dash]CHOH (31), with the axes.

Effect from the carbonyl group

Very large downfield shifts in δ(17O) (200–400 ppm) are reported for the species containing the carbonyl group.50 The mechanisms are discussed, exemplified by the formation of H2C[double bond, length as m-dash]O from MeOH, first.

Table 13 lists the σd(O), σp(O) and σt(O) values of H2C[double bond, length as m-dash]O, separately by ψi. The contributions from ψ6 and ψ8 on σpi(O) are very large, which amount to −264.8 and −480.4 ppm, respectively. The (Δσd(O)e, Δσp(O)e, Δσt(O)e) values are (9.4, −760.9, −751.5 ppm) for H2C[double bond, length as m-dash]O from MeOH. As shown in Table 14, the ψ6ψ9 (σp[6 with combining right harpoon above (vector)]9:yy(O) = −647.6 ppm) and ψ8ψ9 (σp[8 with combining right harpoon above (vector)]9:xx(O) = −1385.1 ppm) transitions are the predominant contributors.

Table 13 The σd(O), σp(O) and σt(O) values of H2C[double bond, length as m-dash]O (34: C2v), given separately by each ψia
MO (i in ψi) σdi(O) σpi(O) σti(O)
a Using the GIAO method under B3LYP/BSS-A.
1 270.61 0.00 270.61
2 0.03 0.05 0.08
3 32.51 −25.86 6.65
4 8.91 8.95 17.87
5 10.38 9.93 20.31
6 25.20 −264.75 −239.55
7 28.07 −41.88 −13.81
8 28.79 −480.42 −451.63
ψocc to ψocc   −39.79  
Total 404.50 −833.77 −429.27


Table 14 Main contributions from occupied-to-unoccupied orbital transitions on σp(O) of H2C[double bond, length as m-dash]O (34: C2v)a
iab σpia:xx(O) σpia:yy(O) σpia:zz(O) σpia(O)
a Calculated with the GIAO method under B3LYP/BSS-A. The magnitudes of σpia(O) larger than 10 ppm are shown.b In ψiψa.
3 → 9 0.00 −37.45 0.00 −12.48
5 → 9 93.99 0.00 0.00 31.33
6 → 9 0.00 −647.61 0.00 −215.87
6 → 18 0.00 0.00 −38.06 −12.69
7 → 33 0.00 −31.32 0.00 −10.44
8 → 9 −1385.05 0.00 0.00 −461.68
8 → 10 0.00 0.00 −40.47 −13.49
8 → 13 −43.55 0.00 0.00 −14.52
8 → 19 −31.54 0.00 0.00 −10.51
8 → 33 0.00 0.00 38.09 12.70


Fig. 11 illustrates the selected ψiψa transitions of ψ6ψ9 and ψ8ψ9 in H2C[double bond, length as m-dash]O, along with the molecular axes. The ψ6, ψ8 and ψ9 orbitals mainly have the occupied σ(C[double bond, length as m-dash]O), occupied npy(O) and vacant π*(C[double bond, length as m-dash]O) characters, respectively. The ψ6 (σ(C[double bond, length as m-dash]O)) and ψ8 (npy(O)) orbitals act as excellent donors, while ψ9 (π*(C[double bond, length as m-dash]O)) does as an excellent acceptor to produce the very large σp (O) in H2C[double bond, length as m-dash]O. However, ψ7 (π(C[double bond, length as m-dash]O)) seems not a good donor in the transitions, relative to the case of ψ6 and ψ8.


image file: d4ra00843j-f11.tif
Fig. 11 Main contributions from each ψiψa transition to the components of σp(O) in H2C[double bond, length as m-dash]O (34), together with the axes.

The origin for the very large downfield shift for σp(O: H2C[double bond, length as m-dash]O) is effectively analysed, along with the mechanism.

Effect from the carboxyl group

The carboxyl effect is closely related to the carbonyl effect, which is discussed for the formation of H(HO*)C[double bond, length as m-dash]O* from MeOH.

Table 15 lists the σd(O), σp(O) and σt(O) values of H(HO)C[double bond, length as m-dash]O* and H(HO*)C[double bond, length as m-dash]O, separately by ψi. The contributions from ψ10 and ψ12 to σpi(O) are very large for H(HO)C[double bond, length as m-dash]O*, which amounts to −130.4 and −234.1 ppm, respectively, while that from ψ10 to σpi(O) is also very large for H(HO*)C[double bond, length as m-dash]O, which amounts to −120.2 ppm. Table 16 shows the ψiψa transitions, mainly contributing to σpia:xx(O), σpia:yy(O) and/or σpia:zz(O), in H(HO)C[double bond, length as m-dash]O* and H(HO*)C[double bond, length as m-dash]O. In the case of H(HO)C[double bond, length as m-dash]O*, the ψ10ψ13 (σp1[0 with combining right harpoon above (vector)]13:yy(O) = −279.3 ppm; σp1[0 with combining right harpoon above (vector)]13:xx(O) = −45.9 ppm) and ψ12ψ13 (σp1[2 with combining right harpoon above (vector)]13:xx(O) = −626.0 ppm; σp1[2 with combining right harpoon above (vector)]13:yy(O) = −31.5 ppm) transitions predominantly contribute to σp(O); additionally, the ψ10ψ13 (σp1[0 with combining right harpoon above (vector)]13:yy(O) = −198.3 ppm; σp1[0 with combining right harpoon above (vector)]13:xx(O) = −53.0 ppm) transition predominantly contributes to σp(O) of H(HO*)C[double bond, length as m-dash]O.

Table 15 The σd(O), σp(O) and σt(O) values of H(HO)C[double bond, length as m-dash]O (Cs), given separately by each ψia
MO (i in ψi) σdi(O) σpi(O) σti(O)
a Using the GIAO method under B3LYP/BSS-A.
H(HO)C[double bond, length as m-dash]O* (35: Cs)
1,3 0.01 0.00 0.01
2 270.61 0.00 270.61
4 9.56 −6.12 3.44
5 23.30 −21.95 1.35
6 4.29 3.74 8.03
7 7.16 15.69 22.85
8 12.87 −83.26 −70.39
9 12.53 −10.06 2.47
10 15.65 −130.36 −114.71
11 18.37 −18.43 −0.07
12 30.15 −234.06 −203.91
ψocc to ψocc   −21.96  
Total 404.48 −506.77 −102.29
[thin space (1/6-em)]
H(HO*)C[double bond, length as m-dash]O (36: Cs)
1 270.61 0.00 270.61
2,3 0.02 0.02 0.04
4 22.33 −10.63 11.70
5 13.44 −6.15 7.29
6 15.46 −10.24 5.23
7 11.01 −40.96 −29.95
8 10.18 −8.84 1.35
9 17.85 −33.89 −16.04
10 9.22 −120.24 −111.02
11 18.74 −29.95 −11.22
12 10.95 −42.16 −31.21
ψocc to ψocc   18.67  
Total 399.81 −284.37 115.44


Table 16 Main contributions from occupied to unoccupied orbital transitions on σp(O) of H(HO)C[double bond, length as m-dash]O (Cs)a
iab σpia:xx(O) σpia:yy(O) σpia:zz(O) σpia(O)
a Calculated with the GIAO method under B3LYP/BSS-A. The magnitudes of σpia(O) larger than 10 ppm are shown.b In ψiψa.
H(HO)C[double bond, length as m-dash]O* (35)
5 → 13 0.56 −30.94 0.00 −10.13
7 → 13 85.27 −2.96 0.00 27.44
8 → 13 −0.68 −159.30 0.00 –53.33
10 → 13 −45.89 −279.32 0.00 −108.40
12 → 13 −625.99 −31.46 0.00 −219.15
[thin space (1/6-em)]
H(HO*)C[double bond, length as m-dash]O (36)
7 → 13 −64.11 −35.48 0.00 −33.20
8 → 13 2.20 70.95 0.00 24.38
10 → 13 −53.03 −198.27 0.00 −83.77
10 → 20 0.00 0.00 −45.92 −15.31
12 → 13 −57.30 −32.99 0.00 −30.10


Fig. 12a and b illustrate the selected ψiψa transitions in H(HO)C[double bond, length as m-dash]O* and H(HO*)C[double bond, length as m-dash]O, respectively, along with the molecular axes. While ψ10 and ψ12, in the ψ10ψ13 and ψ12ψ13 transitions of H(HO)C[double bond, length as m-dash]O*, have the main character of occupied σ(C[double bond, length as m-dash]O) and np(O), respectively, ψ13 has the vacant π*(C[double bond, length as m-dash]O) character. Therefore, ψ10 (σ(C[double bond, length as m-dash]O)) and ψ12 (np(O)) act as excellent donors, while ψ13 (π*(C[double bond, length as m-dash]O)) operates as an excellent acceptor in H(HO)C[double bond, length as m-dash]O*. However, ψ11 (π(C[double bond, length as m-dash]O)) seems not a good donor in the transitions, again, if compared with ψ10 and ψ12. Similarly, ψ10 and ψ13, in the ψ10ψ13 transition of H(HO*)C[double bond, length as m-dash]O, have the main character of the occupied σ(C[double bond, length as m-dash]O) and the vacant π*(C[double bond, length as m-dash]O), respectively. Thus, ψ10 (σ(C[double bond, length as m-dash]O)) acts as a good donor and ψ13 (π*(C[double bond, length as m-dash]O)) operates as a good acceptor in the transitions to produce (large) σp(O) of H(HO*)C[double bond, length as m-dash]O.


image file: d4ra00843j-f12.tif
Fig. 12 Main contributions from each ψiψa transition to the components of σp(O) in H(HO)C[double bond, length as m-dash]O* (35) (a) and H(HO*)C[double bond, length as m-dash]O (36) (b), together with the axes.

The (Δσd(O)e, Δσp(O)e, Δσt(O)e) values for the process from H2C[double bond, length as m-dash]O to H(HO)C[double bond, length as m-dash]O* are also of interest. The values are (0.0, 327.0, 327.0 ppm), which means that the H(HO)C[double bond, length as m-dash]O* signal will appear at much higher field of 327.0 ppm from that of H2C[double bond, length as m-dash]O*.

The specific π-type O–C[double bond, length as m-dash]O interaction is responsible for the results. The charge on H(HO)C[double bond, length as m-dash]O* is less positive than that on H2C[double bond, length as m-dash]O, due to the donation from HO to C[double bond, length as m-dash]O in H(HO)C[double bond, length as m-dash]O, which leads to the upfield shift. The wider extension of the MOs over the entire molecule in H(HO)C[double bond, length as m-dash]O needs to be considered, again, although it would be complex. The smaller occupancy of an important orbital in H(HO)C[double bond, length as m-dash]O*, relative to that in H2C[double bond, length as m-dash]O*, would not effectively operate to produce a larger σp(O) in magnitude. The energy differences in the transitions also affect on σp(O), along with the charge on O. A more upfield σp(O) shift is predicted if the charge on O becomes less positive, although the energy term would show the inverse direction from the factor of the charge.

The much larger downfield shift for H2C[double bond, length as m-dash]O relative to H(HO)C[double bond, length as m-dash]O* is effectively reproduced in the calculations. The large upfield shifts in RC(=O)NHR′ and ROC(=O)OR′, relative to H2C[double bond, length as m-dash]O, can also be understood based on the structural similarities to H(HO)C[double bond, length as m-dash]O, relative to H2C[double bond, length as m-dash]O. Specifically, the analysis of H2C[double bond, length as m-dash]O and H(HO)C[double bond, length as m-dash]O can aid in the understanding of the 17O NMR chemical shifts of similar structures. However, further investigations are needed to understand the much higher downfield shifts of the nitroso species and ozone.

Visualization of Δσd(O), Δσp(O) and Δσt(O) in some oxygen containing species

The Δσd(O), Δσp(O) (=σp(O)), Δσt(O) values and the components are plotted for Me2O, Et2O, H2C[double bond, length as m-dash]CHOH, H(HO)C[double bond, length as m-dash]O* and H(HO*)C[double bond, length as m-dash]O. Fig. 13 shows the plot. The contributions from the occupied-to-occupied orbital (ψiψj) transitions, shown in green in Fig. 13, are all positive, except for H(HO)C[double bond, length as m-dash]O*. The σp(O) values are all negative, of which magnitude is small for Me2O but very large for H(HO)C[double bond, length as m-dash]O*. The contributions from the ψiψj transitions seem to decrease as the magnitudes of σp(O) increase. MOs, mainly constructed by the 2px(O), 2py(O) and 2pz(O) AOs, should contribute much on σp(O). The contributions to σp(O) are well visualized, which helps us to understand the rule and the mechanism of σ*(O: * = d, p and t).
image file: d4ra00843j-f13.tif
Fig. 13 Plots of Δσd(O), Δσp(O) (=σp(O)), Δσt(O) and the components, for Me2O, Et2O, H2C[double bond, length as m-dash]CHOH, H(HO)C[double bond, length as m-dash]O* and H(HO*)C[double bond, length as m-dash]O. Each MO contributing to σp(O) is shown by -n in HOMO-n.

Contributions from occupied-to-occupied orbital transitions to σp(O)

The occupied-to-occupied orbital (ψiψj) transitions are usually not considered to be important; therefore, they are often neglected in the discussion. However, they contribute to σp(O) more than expected, in some cases. Such transitions should arise through the redistribution of electrons in a species under an applied magnetic field. Table 17 summarizes the values, again, which are shown in some Tables in the text.
Table 17 The contributions from the occupied-to-occupied orbital (ψiψj) transitions to σp(O) in some oxygen containing speciesa
Species σp(O)o–o Species σp(O)o–o
a Using the GIAO method under B3LYP/BSS-A.
HO (2: C∞v) 114.41 H2C[double bond, length as m-dash]CHOH (31: Cs) 31.50
H2O (7: C2v) 109.30 H2C[double bond, length as m-dash]O (34: C2v) −39.79
H3O+ (25: C3v) 57.17 H(HO)C[double bond, length as m-dash]O* (35: Cs) −21.96
Me2O (14: C2v) 133.50 H(HO*)C[double bond, length as m-dash]O (36: Cs) 18.67
Et2O (20: C2v) 142.90    


The paramagnetic contributions from the occupied-to-occupied transitions, σp(O)o-o, are larger than 100 ppm for HO (C∞v), H2O (C2v), Me2O (C2v) and Et2O (C∞v), which form group A (g(A)). The σp(O)o–o values are less than 60 ppm for H2C[double bond, length as m-dash]CHOH (Cs), H2C[double bond, length as m-dash]O (C2v), H(HO)C[double bond, length as m-dash]O* (Cs), H(HO*)C[double bond, length as m-dash]O (Cs) and H3O+ (C3v), which belong to g(B). According to the discussion about the data in Fig. 13, σp(O)o–o are plotted versus σp(O), to examine the relationship between the two. The plot is shown in Fig. S11 of the ESI. While the correlation was poor for g(A) (y = 105.70–0.261x: Rc2 = 0.625), whereas a very good correlation was obtained for g(B), if analysed with a quadric function (y = 82.87 + 0.288x + 0.00017x2: Rc2 = 0.996). Indeed σp(O)o–o will change depending on σp(O), but the behaviour seems complex.

Conclusions

The 1H NMR chemical shifts are controlled predominantly by the σd term; therefore, δ(1H) are explained mainly by Q(H), especially for saturated organic species, although other terms, such as the aromatic ring current effect, become important in some cases. For the atoms of the third and higher periods, the spectra can be analysed based on the σp term, predominantly, neglecting the σd term as the relative values. In the case of the atoms of the second period, the NMR chemical shifts are controlled by both σd and σp terms, where the contribution ratios will change depending on the atoms and the species containing the atoms. Namely, the analysis of the spectra for the atoms of the second period will be essentially more complex relative to other cases. NMR spectra, containing δ(17O), are usually analysed with the guidance of empirical rules. Indeed, the empirical rules are useful for assigning the spectra, but the origins of chemical shifts are difficult to understand based on such rules. Then, our research interested is to establish the plain rules founded in theory with the origin of the 17O NMR chemical shifts for the better understanding of the phenomena, which is the aim of this study. The origin should be visualized based on the specific concepts, such as molecular orbitals. This purpose is given more importance than the usual one in NMR calculations to reproduce the observed values accurately and/or to predict well the shift values of unknown target compounds.

NMR chemical shifts of 17O are analysed employing the calculated σd, σp and σt terms. The contributions from σd(O) to σt(O) are approximately one tenth of those from σp(O), although the ratio changes depending on the oxygen containing species. The plots of σd(O) versus Q(O) for Ox (x = −2, 0, 2, 4 and 6) effectively follow a quadratic regression curve, and those for H2O, HO+, HO, and H3O+ are located (very) near the curve. Therefore, the σd(O) values can be understood based on Q(O) for the species. However, σd(O) values of ROH and ROR′ (R, R′: alkyl group) change depending on R and R′ but not on Q(O). The σp values were analysed based on the occupied-to-unoccupied orbital (ψiψa) transitions, which arose σp(O). The relationship between σp(O) and Q(O) was not examined, which would be hidden in the complex combinations of in the ψiψa transitions, as shown in eqn (6). Specifically, a broad (but not so strong) relationship between δ(O) and Q(O) has been reported, as expected, if the conditions are satisfied; however, an explicit relationship is not observed for most cases. The occupied-to-occupied orbital (ψiψj) transitions are also examined, of which contributions to σp(O) are denoted by σp(O)o–o. The good proportionality between σp(O)o–o and σp(O) was confirmed in some cases, but not widely. The treatments provided useful information for σp(O), where the contributions from the ψiψj transitions are usually neglected. The relationships between Q(O) and between σp(O) and the orbital–orbital transitions (interactions) are widely clarified, in this work.

The origin of the effects is visualized based on the occupied-to-unoccupied orbital (ψiψa) transitions, where σp(O) arises from the transitions. As a result, the plain rules with the origin can be more easily imaged and understood through the contributions of transitions to the effects also by the experimental scientists, including the authors. The results will help to understand the role of O in the specific position of a compound in question and the mechanisms to arise the shift values. This work also has the potential to provide an understanding of the δ(O) values of unknown species and facilitate new concepts for the strategies to create highly functional materials based the observed δ(O) values, along with the calculated σd(O) and σp(O) values.

Author contributions

W. N. and S. H. formulated the project. K. M. contributed to investigation and writing. W. N. and S. H. organized the data, contributed to supervision, data curation, resources and writing – review & editing the paper. All authors have read and agreed to the published version of the manuscript.

Conflicts of interest

The authors declare no conflicts of interest.

Acknowledgements

The authors are very grateful to Prof. Masahiko Hada and Dr Daisuke Yamaki of Tokyo Metropolitan University for the utility programs.

Notes and references

  1. Encyclopedia of Nuclear Magnetic Resonance, ed. D. M. Grant and R. K. Harris, John Wiley & Sons, New York, 1996 Search PubMed.
  2. Encyclopedia of Nuclear Magnetic Resonance, ed. R. K. Harris and R. E. Wasylishen, John Wiley & Sons, New York, 2012 Search PubMed.
  3. Nuclear Magnetic Shieldings and Molecular Structure, ed. J. A. Tossell, Kluwer Academic Publishers, Dordrecht, Boston, London, 1993 Search PubMed.
  4. Calculation of NMR and EPR Parameters; Theory and Applications, ed. M. Kaupp, M. Bühl and V. G. Malkin, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, 2004 Search PubMed.
  5. C. Geletneky and S. Berger, Eur. J. Org Chem., 1998, 1625–1627 CrossRef CAS.
  6. F. Mocci, G. Uccheddu, A. Frongia and G. Cerioni, J. Org. Chem., 2007, 72, 4163–4168 CrossRef CAS PubMed.
  7. I. P. Gerothanassis, Prog. Nucl. Magn. Reson. Spectrosc., 2010, 56, 95–197 CrossRef CAS PubMed.
  8. I. P. Gerothanassis, Prog. Nucl. Magn. Reson. Spectrosc., 2010, 57, 1–110 CrossRef CAS PubMed.
  9. L. B. Krivdin, Magn. Reson. Spectrosc., 2023, 61, 507–529 CAS.
  10. W. Saenger, Principles of Nucleic Acid Structures, Springer, Berlin, 1984 Search PubMed.
  11. X. Wang, A. R. Chandrasekaran, Z. Shen, Y. P. Ohayon, T. Wang, M. E. Kizer, R. Sha, C. Mao, H. Yan, X. Zhang, S. Liao, B. Ding, B. Chakraborty, N. Jonoska, D. Niu, H. Gu, J. Chao, X. Gao, Y. Li, T. Ciengshin and N. C. Seeman, Chem. Rev., 2019, 119, 6273–6289 CrossRef CAS PubMed.
  12. J.-L. Mergny and D. Sen, Chem. Rev., 2019, 119, 6290–6325 CrossRef CAS PubMed.
  13. F. C. Simmel, B. Yurke and H. R. Singh, Chem. Rev., 2019, 119, 6326–6369 CrossRef CAS PubMed.
  14. S. S. Wang and A. D. Ellington, Chem. Rev., 2019, 119, 6370–6383 CrossRef CAS PubMed.
  15. M. Madsen and K. V. Gothelf, Chem. Rev., 2019, 119, 6384–6458 CrossRef CAS PubMed.
  16. S. Muniyappan, Y. Lin, Y.-H. Lee and J. H. Kim, Biology, 2021, 10, 453–461 CrossRef CAS PubMed.
  17. B. Lin, I. Hung, Z. Gan, P.-H. Chien, H. L. Spencer, S. P. Smith and G. Wu, ChemBioChem, 2021, 22, 826–829 CrossRef CAS PubMed.
  18. V. Balevičius, K. Aidas, A. Maršalka, F. Kuliešius, V. Jakubkienė and S. Tumkevičius, Lith. J. Phys., 2022, 62, 114–125 Search PubMed.
  19. W. Nakanishi, S. Hayashi and M. Hada, Chem.–Eur. J., 2007, 13, 5282–5293 CrossRef CAS PubMed.
  20. K. Kanda, H. Nakatsuji and T. Yonezawa, J. Am. Chem. Soc., 1984, 106, 5888–5892 CrossRef CAS.
  21. Molecular Quantum Mechanics, ed. P. W. Atkins and R. S. Friedman, Oxford: New York, 3rd edn, 1997, ch. 13 Search PubMed.
  22. Indeed, this decomposition includes small arbitrariness due to the coordinate origin dependence, but it does not damage our insights into the 17O NMR spectroscopy.
  23. The occupied-to-unoccupied orbital (ψi → ψa) transitions mainly contribute to σp, whereas the occupied-to-occupied orbital (ψi → ψj) transitions sometimes play an important role in σp.
  24. Based on the second-order perturbation theory at the level of the HF and single-excitation CI approximation, σpia on a resonance nucleus N is shown to be proportional to reciprocal orbital energy gap (εaεi)−1 as expressed in eqn (5), where ψk is the k-th orbital function, [L with combining circumflex]z,N is orbital angular momentum around the resonance nucleus, and rN is the distance from the nucleus N.
  25. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, Revision D.01, Gaussian, Inc., Wallingford CT, 2009 Search PubMed.
  26. A. D. Becke, Phys. Rev., 1988, 38, 3098–3100 CAS.
  27. A. D. Becke, J. Chem. Phys., 1993, 98, 5648–5652 CrossRef CAS.
  28. C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785–789 CrossRef CAS PubMed.
  29. B. Miehlich, A. Savin, H. Stall and H. Preuss, Chem. Phys. Lett., 1989, 157, 200–206 CrossRef CAS.
  30. C. Møller and M. S. Plesset, Phys. Rev., 1934, 46, 618–622 CrossRef.
  31. J. Gauss, J. Chem. Phys., 1993, 99, 3629–3643 CrossRef CAS.
  32. J. Gauss and B. Bunsenges, Phys. Chem., 1995, 99, 1001–1008 CAS.
  33. K. Wolinski, J. F. Hinton and P. Pulay, J. Am. Chem. Soc., 1990, 112, 8251–8260 CrossRef CAS.
  34. K. Wolinski and A. Sadlej, Mol. Phys., 1980, 41, 1419–1430 CrossRef CAS.
  35. R. Ditchfield, Mol. Phys., 1974, 27, 789–807 CrossRef CAS.
  36. R. McWeeny, Phys. Rev., 1962, 126, 1028–1034 CrossRef.
  37. F. London, J. Phys. Radium, 1937, 8, 397–409 CrossRef CAS.
  38. T. Yanai, D. P. Tew and N. C. A. Handy, Chem. Phys. Lett., 2004, 393, 51–57 CrossRef CAS.
  39. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
  40. C. Adamo and V. Barone, J. Chem. Phys., 1999, 110, 6158–6169 CrossRef CAS.
  41. T. M. Henderson, A. F. Izmaylov, G. Scalmani and G. E. Scuseria, J. Chem. Phys., 2009, 131, 044108 CrossRef PubMed.
  42. J.-D. Chai and M. Head-Gordon, Phys. Chem. Chem. Phys., 2008, 10, 6615–6620 RSC.
  43. F. Weigend and R. Ahlrichs, Phys. Chem. Chem. Phys., 2005, 7, 3297–3305 RSC.
  44. F. Weigend and R. Ahlrichs, Phys. Chem. Chem. Phys., 2006, 8, 1057–1065 RSC.
  45. J. Tomasi, B. Mennucci and R. Cammi, Chem. Rev., 2005, 105, 2999–3093 CrossRef CAS PubMed.
  46. The utility program was provided by Prof. Hada and Dr Yamaki of Tokyo Metropolitan University. It is derived from the NMR program in the Gaussian program, which made some processes, necessary for our discussion, possible to be printed out, such as the contributions from the occupied orbitals and/or the orbital-to-orbital transitions51.
  47. E. D. Glendening, J. K. Badenhoop, A. E. Reed, J. E. Carpenter, J. A. Bohmann, C. M. Morales, C. R. Landis and F. Weinhold, NBO Version 6.0, 2013 Search PubMed.
  48. R. A. Klein, B. Mennucci and J. Tomasi, J. Phys. Chem. A, 2004, 108, 5851–5863 CrossRef CAS.
  49. E. Taskinen, Magn. Reson. Chem., 1998, 36, 573–578 CrossRef CAS.
  50. J.-C. Zhuo, Molecules, 1999, 4, 320–328 CrossRef CAS.
  51. Essentially the same analysis can be achieved by using the ADF program. The ADF program should be explained here.

Footnote

Electronic supplementary information (ESI) available: Additional tables and figures and the fully optimized structures given by Cartesian coordinates, together with total energies. See DOI: https://doi.org/10.1039/d4ra00843j

This journal is © The Royal Society of Chemistry 2024