Milena
Barp
a,
Florian
Kreuter
a,
Qian-Rui
Huang
b,
Jiaye
Jin
a,
Franka. E.
Ninov
a,
Jer-Lai
Kuo
*b,
Ralf
Tonner-Zech
*a and
Knut R.
Asmis
*a
aWilhelm-Ostwald-Institut für Physikalische und Theoretische Chemie, Universität Leipzig, Linnéstraße 2, 04103 Leipzig, Germany. E-mail: ralf.tonner@uni-leipzig.de; knut.asmis@uni-leipzig.de
bInstitute of Atomic and Molecular Sciences, Academia Sinica No. 1 Roosevelt Rd, Sec 4, Taipei, 106319, Taiwan. E-mail: jlkuo@gate.sinica.edu.tw
First published on 29th January 2025
We report on the gas phase vibrational spectroscopy (3500–950 cm−1) of halide anion complexes with 1,1,1,3,3,3-hexafluoroisopropanol (HFIP) and its partially deuterated analogue (HFIP-d1). Infrared photodissociation spectra of messenger-tagged X−(HFIP/HFIP-d1), with X− = Cl−, Br−, and I−, together with electronic structure calculations reveal O–H(D) stretching fundamentals that are red-shifted twice as much as those for the corresponding complexes with isopropanol and water, directly reflecting HFIP's enhanced hydrogen-bond donor ability. The harmonic analysis of the bands in the fingerprint region reveals that HFIP assumes a synperiplanar conformation in the complexes. The consideration of anharmonic effects is necessary to recover the efficient coupling between stretching and bending modes in the OH stretching region. An energy decomposition analysis shows that the roughly twice as large binding energy in the HFIP complexes vs. i-PrOH and water is determined mainly by differences in the electrostatic attraction. The observed red-shifts, which reflect the extent of charge transfer along the coordinate of the proton transfer reaction X− + HM → XH + M−, correlate qualitatively with the difference in the proton affinities ΔPA = PA(X−) − PA(M−). A more quantitative agreement requires also considering differences in the hydrogen bond angle.
Infrared (IR) studies show that the isolated HFIP molecule exists in two conformations, an antiperiplanar (AP) and a synclinal (SC) conformer (see Scheme 1).11 The AP conformer is found to be more stable, in contrast to HFIP's non-fluorinated analog isopropanol (i-PrOH), which adopts an SC conformation.12 Shahi and Arunan reported the microwave spectrum of HFIP cooled in a supersonic expansion and assigned it to the AP conformer.13 Their ab initio calculations confirm this assignment and find the SC conformer about 5 kJ mol−1 higher in energy, while the synperiplanar (SP) conformer represents a saddle point on the potential energy surface, only 1 kJ mol−1 above the SC minima.13 HFIP's dipole moment increases along the series AP → SC → SP.14,15 As a result, the higher energy structures are stabilized upon aggregation as a consequence of more favorable electrostatic interactions, leading to the stronger HB donor ability of aggregated HFIP.3,14
Wang and coworkers recently studied orientation-specific charge–dipole interactions in the anion complexes X−(HFIP), X− = F−, Cl−, Br−, I−, and O2−, using anion photoelectron spectroscopy.16,17 They showed that the combination of the charge–dipole interaction with the formation of an ionic hydrogen bond (IHB) leads to a preference for the SP/SC isomer, since these HFIP conformers exhibit a larger dipole moment. The interaction energy decreases with increasing anion size, i.e., with decreasing anion proton affinity (PA). The interaction with the fluoride anion is considerably stronger than with the chloride anion, leading to proton transfer and formation of a complex formally containing HF and deprotonated HFIP.16 This directly raises the question regarding the role of charge transfer (vs. electrostatic interactions) in these model systems containing IHBs.18,19
Here, we apply cryogenic ion trap vibrational spectroscopy20 to study the halide anion complexes X−(HM) with X− = Cl−, Br−, I− and HM = HFIP, HFIP-d1, i-PrOH, i-PrOD, H2O and D2O. The present study builds on the landmark studies of Johnson and coworkers, who studied the IHB interaction in halide anion complexes with water in the gas phase.21,22 They demonstrated that vibrational action spectroscopy allows to systematically characterize IHBs by measuring the red-shift of the corresponding OH stretching frequency (ΔνOH), which is defined as
ΔνOH = νfreeOH − νHBOH, | (1) |
![]() | ||
Fig. 1 IRPD spectra of D2-tagged X−(HFIP), X−(i-PrOH) and X−(H2O) (upper panels, from left to right) and X−(HFIP-d1), X−(i-PrOD) and X−(D2O) complexes (lower panels, from left to right) for X− = Cl−, Br− and I− (top to bottom). Bands assigned to the OH(OD) stretching fundamental are shown in red (see Table 1 for band positions). Bands marked with an asterisk indicate excitation of the D2 stretching mode of the messenger-tag. |
Complex | ν OH/νOD | ΔνOH/ΔνOD | |||
---|---|---|---|---|---|
IRPDa | Harmonicb | VPT2c | DVR-FBRd | IRPDe | |
a Values obtained from IRPD spectra of D2-tagged complexes. The value reported is for the most intense IRPD band assigned to the OH/OD stretch fundamental (see text for details). For I−(H2O) and Cl−(HFIP-d1) we report the center of the doublet, see ref. 22 and 24 for detailed band assignment of I−(H2O). b MP2/aug-cc-pVTZ (X− = Cl−, Br−) or MP2/aug-cc-pVTZ-PP (X− = I−) harmonic frequencies. c VPT2/MP2/aug-cc-pVDZ (X− = Cl−, Br−) or VPT2/MP2/aug-cc-pVDZ-PP (X− = I−) anharmonic frequencies. d DVR-FBR/RI-MP2 + DLPNO-CCSD(T)/aug-cc-pVTZ anharmonic frequencies. The value reported is for the transition with highest OH stretch contribution (see text for details). e Red-shifts are determined with respect to the vibrational frequency of the corresponding free, uncoupled OH or OD oscillator: HFIP (SC, 3668 cm−1),11,25 H2O (3707 cm−1),26,27i-PrOH (3658 cm−1),12 HFIP-d1 (SC 2703 cm−1),25 D2O (2730 cm−1),28 and i-PrOD (2676 cm−1).29 f See Methods section and ESI for computational details and references. | |||||
Cl−(HFIP/HFIP-d1) | 2535/1973 | 2925/2134 | 2441/1880 | 2588/1993 | 1133/730 |
Br−(HFIP/HFIP-d1) | 2748/2092 | 3091/2251 | 2719/2079 | 2731/2080 | 920/611 |
I−(HFIP/HFIP-d1) | 3028/2271 | 3219/2344 | 2928/2204 | 640/432 | |
Cl−(i-PrOH/i-PrOD) | 3087/2303 | 3270/2383 | 2994/2246 | 571/373 | |
Br−(i-PrOH/i-PrOD) | 3208/2393 | 3365/2450 | 3136/2340 | 450/283 | |
I−(i-PrOH/i-PrOD) | 3315/2463 | 3460/2519 | 3230/2403 | 343/213 | |
Cl−(H2O/D2O) | 3158/2350 | 3338/2425 | 3081/2294 | 549/380 | |
Br−(H2O/D2O) | 3308/2449 | 3426/2487 | 3223/2391 | 399/281 | |
I−(H2O/D2O) | 3417/2520 | 3525/2557 | 3328/2461 | 290/210 |
Above 1800 cm−1, excitation of the OH stretching mode (νOH) of the hydroxyl group involved in the IHB is the most prominent and also the most diagnostic feature. Note, multiple bands associated with this excitation are observed in several spectra. Similar observations have been previously reported for halide–water complexes and attributed to the presence of a strong IHB combined with the anharmonic nature of the corresponding O–H oscillator, leading to the excitation of combination bands.21,24 Moreover, Fermi resonances with nearby overtone excitations of the CH and OH bending modes can further complicate the spectral pattern.24 The most intense band (above 1800 cm−1) typically corresponds to the fundamental excitation of the OH stretching mode, except for the spectra of Br−(HFIP) and Br−(HFIP-d1), where strong anharmonic coupling leads to two bands of similar intensity (vide infra). For Cl−(HFIP), Br−(HFIP) and I−(HFIP), the bands assigned to excitation of the OH stretching fundamental are centered at 2535 cm−1, 2748 cm−1 and 3028 cm−1, respectively, corresponding to decreasing red-shifts ΔνOH of 1133 cm−1 (Cl−), 920 cm−1 (Br−) and 640 cm−1 (I−) with increasing halide anion size, as expected. ΔνOH is determined with respect to the OH stretching frequency of the free SC conformer of HFIP (3668 cm−1)25 and is also listed in Table 1. There are two substantially weaker features that we also observed in this spectral region, which correspond to excitation of the CH stretching mode of HFIP (2935–2958 cm−1) as well as the nominally IR-forbidden stretching mode of D2 (2862–2905 cm−1), which gains IR intensity through charge-induced-dipole interactions.30
Upon deuteration of HFIP's hydroxyl group the corresponding IRPD feature, now associated with the OD stretch excitation (ΔνOD), is shifted to lower wavenumbers by a factor of 1.29–1.33, close to the expected ratio of 1.36 for a free OH vs. a free OD oscillator. For chloride, bromide and iodide, ΔνOD is observed at 1973 cm−1 (ΔνOD = 730 cm−1), 2092 cm−1 (611 cm−1) and 2271 cm−1 (432 cm−1), respectively. Moreover, the associated absorption features are simpler, indicating that anharmonic couplings are reduced upon deuteration, as expected.
Similar to the previously discussed IRPD spectra, the most intense transition in this spectral region is due to excitation of the hydrogen-bonded hydroxyl group and observed at 3087 cm−1 (Cl−), 3208 cm−1 (Br−) and 3315 cm−1 (I−) for the three halide complexes. These correspond to red-shifts that are roughly half as large as for the corresponding HFIP complexes. Due to the weaker HBs, the OH stretching features are simpler than those observed for HFIP. Nonetheless, there is some unresolved structure observed in the most intense IRPD feature, indicating efficient coupling to a low frequency mode, presumably to the HB stretching mode. There are also one or two bands at higher wavenumbers, which we attribute to the excitation of combination bands. At lower energies, around 3000 cm−1 and below, excitation of the seven CH stretching modes contributes to a partially resolved feature consisting of multiple vibrational transitions. The D2 stretch of the tagging molecule is also expected in this region (see Fig. 1).
Upon deuteration of i-PrOH's hydroxyl group, an OD stretching band is observed below 2500 cm−1 at 2303 cm−1 (Cl−), 2393 cm−1 (Br−) and 2463 cm−1 (I−), corresponding to νOH/νOD ratios of 1.34–1.35, respectively. Note, the νOD feature in the X−(i-PrOD)·D2 spectra are simpler than those observed for X−(HFIP-d1)·D2 and now consist mainly of a single band, which only remains markedly asymmetric in the Cl−(HFIP-d1)·D2 spectrum (see Fig. 1). In contrast, the CH (and D2) stretching bands remain nearly unchanged in position and intensity.
For the D2-tagged X−(H2O) complexes, we observe νOH at 3158 cm−1 (Cl−), 3308 cm−1 (Br−) and 3417 cm−1 (I−), slightly (<25 cm−1) less red-shifted compared to the previously reported values of 3146 cm−1 (Cl−), 3296 cm−1 (Br−) and 3393 cm−1 (I−) using Ar-tagging, suggesting that the Ar tag is slightly more perturbing than D2. The determined red-shifts νOH are up to 15% smaller to those observed for i-PrOH (see Table 1). Upon deuteration, this difference is reduced to below 2 cm−1.
Summarizing, the red-shifts ΔνOH(D) obtained from IRPD spectroscopy show that HFIP is a strong HB donor, roughly twice as strong compared to i-PrOH and H2O. Hence, the question arises, what is the exact nature of the IHB interaction in X−(HFIP) complexes? Is the IHB strength solely due to differences in electrostatic interactions, what role does charge transfer play, and are there other, not so obvious, contributions to the binding energy?
![]() | ||
Fig. 2 Minimum-energy structures of HFIP, i-PrOH, X−(HFIP) and X−(i-PrOH). The lowest energy conformer for the neutral molecules and the two lowest energy isomers for the anion complexes are shown. See Table 2 for relative energies and geometric parameters. |
System | ΔE | d OH | d HX | θ OHX | φ HCOH | |
---|---|---|---|---|---|---|
Cl−(HFIP) | SP | 0 | 101 | 192 | 162 | 0 |
AP | 28 | 103 | 187 | 175 | 170 | |
Br−(HFIP) | SP | 0 | 100 | 211 | 159 | 0 |
AP | 29 | 101 | 205 | 174 | 170 | |
I−(HFIP) | SP | 0 | 99 | 235 | 158 | 0 |
AP | 30 | 100 | 231 | 172 | 171 | |
Cl−(i-PrOH) | SC | 0 | 99 | 207 | 167 | 51 |
AP | 1.4 | 99 | 211 | 170 | 180 | |
Br−(i-PrOH) | SC | 0 | 99 | 224 | 165 | 50 |
AP | 1.2 | 99 | 228 | 169 | 180 | |
I−(i-PrOH) | SC | 0 | 98 | 250 | 163 | 48 |
AP | 0.6 | 98 | 255 | 170 | 180 | |
Cl−(H2O) | 99 | 212 | 169 | |||
Br−(H2O) | 99 | 229 | 168 | |||
I−(H2O) | 98 | 256 | 165 |
The preferred conformation predicted for all X−(HFIP) complexes is the SP isomer (see Fig. 2), which is found at least 28 kJ mol−1 lower in energy than the AP isomer (see Table 2), in agreement with the results from the previous anion photoelectron spectroscopy (APES) study using DFT calculations.16 Interestingly, the lower energy isomer exhibits a slightly longer and hence weaker IHB than the higher energy one, independent of the nature of the halide anion. This IHB is also less linear (see Table 2), probably due to an additional, albeit, much weaker interaction between halide anion and the CH group, which is not present in the AP isomer. The driving force for formation of the SP isomer in the halide anion complexes is thus not the formation of a stronger IHB, but rather the larger dipole moment of bare HFIP's SP conformer and hence substantially larger charge–dipole interaction in the anion complex (vide infra).
In contrast to X−(HFIP), the SC isomer, the energetically favoured conformation for bare i-PrOH, is predicted as the global minimum-energy structure for all of the X−(i-PrOH) complexes considered here. However, the AP isomer is found only slightly higher in energy (<2 kJ mol−1) and therefore possibly both isomers may be populated in the experiment (see ESI Fig. S10–S12†). Like for neutral i-PrOH, the SP isomer represents a first-order transition state and lies up to 4 kJ mol−1 higher in energy than the two symmetry-equivalent SC isomers. The lower energy SC isomer of X−(i-PrOH) exhibits a shorter and hence stronger IHB than the AP isomer. In general, the IHB in X−(i-PrOH) is roughly 5–10% longer compared to that in X−(HFIP).
![]() | ||
Fig. 3 Unscaled harmonic MP2/aug-cc-pVTZ IR spectra of the AP (top panel) and the SP isomer (see Fig. 2 for geometries) of Cl−(HFIP) (left) and Cl−(HFIP-d1) (right) compared to the IRPD spectrum of the corresponding D2-tagged complex. See Table 3 for band positions, harmonic vibrational frequencies and assignments. The harmonic spectra were convoluted using a Gaussian line-shape function with a full width at half maximum (FWHM) of 8 cm−1. |
Cl− | Br− | I− | Assignmenta | ||||||
---|---|---|---|---|---|---|---|---|---|
Label | IRPD | MP2 | Label | IRPD | MP2 | Label | IRPD | MP2 | |
a Assignment to local stretching (ν) and bending (δ) vibrational modes. See ref. 31 for a detailed description of modes in neutral HFIP molecule. | |||||||||
a8 (b16) | 1453 (1064) | 1513 (1090) | (d15) | - (1063) | 1498 (1079) | (f18) | - (1041) | 1482 (1062) | δ COH (δCOD) |
a9 (b7) | 1395 (1394) | 1429 (1428) | c12 (d6) | 1392 (1394) | 1430 (1430) | e10 (f8) | 1395 (1394) | 1429 (1429) | δ CCH |
a10 (b8) | 1324 (1360) | 1350 (1388) | c13 (d7) | 1313 (1364) | 1346 (1390) | e11 (f9) | 1319 (1369) | 1338 (1395) | δ OCH |
a11 (b9) | 1286 (1287) | 1311 (1313) | c14 (d8) | 1285 (1289) | 1313 (1315) | e12 (f10) | 1290 (1291) | 1314 (1317) | ν CC |
a12 (b10) | 1264 (1265) | 1291 (1291) | c15 (d9) | 1263 (1267) | 1291 (1291) | e13 (f11) | 1267 (1265) | 1291 (1291) | ν CC |
a13 (b11) | 1233 (1233) | 1246 (1247) | c16 (d10) | 1234 (1235) | 1247 (1248) | e14 (f12) | 1238 (1237) | 1249 (1250) | ν CF |
a14 (b12) | 1173 (1186) | 1188 (1198) | c17 (d11) | 1175 (1187) | 1190 (1200) | e16 (f14) | 1177 (1190) | 1191 (1203) | ν CF |
a15 (b13) | 1162 (1162) | 1179 (1181) | c18 (d12) | 1154 (1155) | 1174 (1175) | e17 (f15) | 1152 (1152) | 1169 (1169) | ν CO |
a16 (b14) | 1144 (1143) | 1159 (1158) | c19 (d13) | 1141 (1145) | 1161 (1161) | e18 (f16) | 1137 (1134) | 1164 (1164) | ν CF |
a17 (b15) | 1102 (1104) | 1116 (1116) | c20 (d14) | 1103 (1108) | 1118 (1118) | e19 (f17) | 1107 (1107) | 1120 (1120) | ν CF |
In order to visualize, how the band positions are affected by (i) deuteration and (ii) the nature of the halide anion, the unscaled harmonic IR spectra for all six HFIP-containing complexes are compared to the experimental IRPD spectra in Fig. 4.
The most obvious change predicted upon deuteration is the red-shift of the bending mode δCOH (1513 cm−1) by 423 cm−1, which corresponds reasonably well with the experimental value of 389 cm−1 for the difference in band positions of a8 and b16 (blue bands in Fig. 4). In addition, two other modes, namely δOCH (red band) and the most IR active of the νCF modes (green band), which correspond to the IRPD band pairs a10 (b8) and a14 (b12), respectively, are blue-shifted upon deuteration, indicating that these modes are more delocalized than expected from a local mode picture and also sensitive to deuteration of the O–H moiety. A direct consequence of the latter shift is that the excitation of the CO stretching mode, νCO, (orange bands in Fig. 4), which appears as a shoulder at 1162 cm−1 (a15) in the IRPD spectrum of Cl−(HFIP), is clearly visible as an isolated band (b13) in the IRPD spectrum of Cl−(HFIP-d1).
The spectra for the different halide anions look very similar, the observed effects are small and the agreement between the predicted and experimental spectra remains satisfactory. Small spectral red-shifts (with increasing halide anion size) are predicted and observed for the excitation of the CO stretching mode, νCO, (Cl−: 1162 cm−1, Br−: 1155 cm−1, I−: 1153 cm−1), and the in-plane COD bending mode, δCOD, (Cl−: 1090 cm−1, Br−: 1063 cm−1, I−: 1041 cm−1). Both can be rationalized on the basis of the decreasing HB strength with increasing halide anion size, which results in a stronger O–H bond and hence slightly weaker C–O bond in the first case, and a longer heavy atom distance and hence a weaker cage effect, in the second case.
The predicted IR spectra including anharmonic effects are compared to the IRPD spectra in the OH(D) stretching region for Cl−(HFIP/HFIP-d1) and Br−(HFIP/HFIP-d1) in Fig. 5. While the (unscaled) harmonic frequencies substantially overestimate the experimental values for νOH(D) (see Table 1), VPT2 systematically underestimates these, except for Br−(HFIP-d1). Apart from excitation of the νOH(D) fundamental, the most intense features predicted by the VPT2 method are combinations of νOH(D) with low-frequency, large amplitude ion-molecule modes. In contrast, the first overtone transitions of the δCOH,δCCH, and δOCH bending modes are predicted weak in intensity. Overall, the agreement of the VPT2 spectra with the IRPD spectra in the OH(D) stretching is improved, compared to the harmonic analysis, but is not as good as with DVR-FBR method, discussed below.
To gain a more quantitative understanding of the nature and extent of anharmonic effect, in particular, to elucidate the role of Fermi resonances in the OH(D) stretching region more reliably, we performed DVR-FBR calculations. For the X−(HFIP) complexes, we included five stretching modes (νCH, νOH(D), νCO, νCC, and νOH(D)⋯X) and four bending modes (in-plane δCOH(D), δCCH, δOCH, and out-of-plane ); for the X−(HFIP-d1) complexes, we choose the nine modes above plus one additional CC stretching mode
. It should be noted that although we use same naming convention as for the harmonic calculations, the δCCH, δOCH, δCOD and
bending modes are well-separated in the case of X−(HFIP-d1); in contrast, for X−(HFIP), their counterpart mix strongly with each other. In our experience, the HB stretching mode νOH⋯X usually plays an important role in 1-to-1 complexes, but the harmonic analysis usually separates its contribution into many low-frequency normal modes. Therefore, we adopt the “intermolecular translation mode”39 to represent the contribution from the HB stretching motion.
The satisfactory agreement between the DVR-FBR and corresponding IRPD spectra in Fig. 5 shows that excitation of the νOH(D) fundamental as well as many two and three-quanta states involving excitation of νOH(D) in combination with the above-mentioned modes is well described (see Table 4 for band assignments). For Br−(HFIP), the two most intense peaks in the DVR-FBR spectrum are predicted at 2731 cm−1 and 2772 cm−1. These transitions contain the highest contribution of νOH. However, the relative weight of νOH is only 0.20 and 0.16, respectively, indicating the extend of anharmonic coupling in this particular system, which is also evidenced by pronounced intensity borrowing of several of the two/three-quanta states located between 2600 to 3000 cm−1. In the case of Cl−(HFIP), the “νOH fundamental” is shifted to be below 2600 cm−1, so two-quanta states above 2700 cm−1 do not gain much intensity due to detuning. The most visible feature in this case is the doublet at 2588 cm−1 and 2665 cm−1 assigned to νOH and δ2OCH, respectively. For Cl−(HFIP-d1) and Br−(HFIP-d1), the latter has more complex vibrational feature than the former also due to better resonance condition between νOD and the two/three-quanta states. Since these two/three-quanta states are heavily mixed, we only list the leading components in Table 4, to aid the band assignments.
Since the EDA-NOCV has been developed for DFT-based approaches, we conduct the analysis with B3LYP-D3(BJ)/TZ2P,45–48 which has been found to accurately reproduce the structures from the MP2 approach outlined above (see ESI, Table S4† for comparison of energies and geometrical parameters). In Table 5, we summarise the main findings for the Cl− complex. Similar findings for X = Br− and I− can be found in the ESI (Tables S5 and S6†), the only notable trend being the decreasing bond strength as the halide anion increases in size (Cl− > Br− > I−).
Cl−(HFIP)SP | Cl−(HFIP)AP | Cl−(i-PrOH) | Cl−(H2O) | |||||
---|---|---|---|---|---|---|---|---|
a Energies in kJ mol−1 and bond length in pm. b Percentage values give the relative contributions of dispersion and electronic effects to ΔEint. c Percentage values give the relative contributions to the attractive EDA terms ΔEelstat and ΔEorb. d Percentage values give the relative contributions of the NOCV to ΔEorb. | ||||||||
ΔEint | −164 | −134 | −84 | −73 | ||||
ΔEint(disp)b | −9 | (5%) | −9 | (7%) | −10 | (12%) | −4 | (5%) |
ΔEint(elec)b | −155 | (95%) | −125 | (93%) | −74 | (88%) | −69 | (95%) |
ΔEPauli | +111 | +120 | +75 | +55 | ||||
ΔEelstatc | −171 | (64%) | −140 | (57%) | −87 | (58%) | −82 | (66%) |
ΔEorbc | −95 | (36%) | −106 | (43%) | −63 | (42%) | −43 | (34%) |
ΔE1(Cl− → H–O)d | −61 | (64%) | −72 | (68%) | −33 | (53%) | −32 | (74%) |
ΔE2(Cl− → H–C)d | −9 | (9%) | ||||||
ΔE3(Cl− → H–CMe)d | −8 | (12%) | ||||||
−5 | (8%) | |||||||
ΔEprep | +22 | +21 | +4 | +3 | ||||
E bond | −142 | −113 | −80 | −70 | ||||
d(Cl−–H) | 193 | 187 | 213 | 212 |
First, we find that the bond strength ordering as a function of the ligand is H2O < i-PrOH ≪ HFIP with HFIP showing more than twice the bond energy (Ebond) compared to H2O. This is also reflected in the interaction energy (ΔEint), although HFIP shows a higher deformation upon HB formation reflected in a sizeable preparation energy (ΔEprep = 22 kJ mol−1) – a term which is close to zero for H2O and i-PrOH. This is a result of the configurational change from the AP to the SP isomer (see Fig. 2). The SP isomer shows additional stabilization due to a second (weaker) HB involving the CH group. A structural indicator of this second HB is the deviation of the HB angle (θOHX = 163°) from the ideal linear configuration. However, the small increase of 10 pm in the C–H bond length suggests that the second HB is much weaker.
The first notable observation upon decomposing the bond energy is that the dispersion energy contribution, ΔEint(disp), is nearly negligible for all complexes listed in Table 5. Although taking the DFT-D3 term as indicator of the attractive London forces is an approximation, the low value compared to the covalent bonding contribution, ΔEint(elec), is a strong indicator that the ion-molecule interaction is not governed by dispersion attraction. The EDA procedure clearly shows the cause for the stronger interaction in the most stable complex Cl−(HFIP)SP compared to the non-fluorinated alcohol i-PrOH complexes: the electrostatic attraction term (ΔEelstat) is considerably larger in its absolute terms (+84 kJ mol−1) as well as in its relative contribution (+6%). Due to the shorter HB, the other two EDA terms (ΔEPauli, ΔEorb) are also larger in HFIP compared to i-PrOH complex but the term dominating the trend is decisive here. This makes the HB in the HFIP complex more similar to that in the H2O complex, where the electrostatic term is also the most important attractive interaction.
For the higher energy Cl−(HFIP)AP isomer several characteristic differences are found (Table 5). Even though the HB is shorter, the total interaction energy is smaller. This can be mainly traced back to a significant decrease in electrostatic attraction due to a less favorable dipole–ion interaction in the Cl−(HFIP)AP isomer. The increased orbital term points towards a larger charge-transfer and hence also more pronounced red-shift ΔνOH (see ESI Fig. S3–S6†). However, this increase is not sufficient to compensate for the smaller electrostatic interaction. The SP isomer is further stabilized by a second, albeit considerably weaker, X−⋯H–C HB, which is not present in the AP isomer (ΔE2 in Table 5).
The deformation densities from the NOCV analysis (Fig. 6) show the most important orbital interactions contributing to ΔEorb. In all three cases, the major contribution is the donation from a non-bonding Cl− lone pair orbital into the antibonding σ*(O–H) orbital (ΔE1). Notably, this interaction is of similar magnitude in the H2O (Fig. 6a) and i-PrOH complexes (Fig. 6c), but nearly twice as strong in the HFIP SP-complex (Fig. 6b) with ΔE1 = −61 kJ mol−1. The NOCV analysis also gives a hint towards the strength of the secondary HB interaction, which only appears in the alcohol interacting with Cl−. This interaction is much weaker but still accounts for 9 kJ mol−1 in the HFIP complex (Fig. 6b, ΔE2). For the i-PrOH complex, two weak CMe–H⋯Cl HBs are found with 8 and 5 kJ mol−1 orbital interaction contribution, respectively (Fig. 6c). We thus conclude from the present bonding analysis that, similar to the X−(H2O) complexes, the interactions in the X−(HFIP) complexes are dominated by electrostatic attraction, which overrules the trends from charge transfer effects. Dispersion attraction only plays a minor and non-decisive role. The electrostatic attraction is largest in the SP isomer and hence this represents the most stable complex, even though the X−⋯H–O HB interaction is weaker than in the higher energy AP isomer.
![]() | ||
Fig. 6 Selected deformation densities (Δρi) from EDA-NOCVs with energy contribution (ΔEi, further explained in Table 5) to ΔEorb in kJ mol−1 and eigenvalues (νi). Charge depletion (red) and charge accumulation (blue) for (a) Cl−(H2O), (b) Cl−(HFIP) and (c) Cl−(i-PrOH). Iso values are chosen for clarity. |
The seminal work on the vibrational spectroscopy of X−(H2O) complexes by Johnson and coworkers21 revealed that the vibrational red-shift ΔνOH is indeed correlated with the halide anion proton affinity (PA). This confirmed the predictions by Thompson and Hynes based on a two valence-bond (VB) state model, in which the first VB state has the charge character X−⋯H2O and the second is a charge-transfer VB state with electronic structure XH⋯OH−.19 ΔνOH is governed by the relative energy of the XH⋯OH− diabatic state, which correlates with the, PA of the anion.51
The PA is defined as the negative enthalpy of the gas phase reaction
X + H+ → XH+, |
In order to extend this model to different solvent molecule, the above reaction can be rewritten as
X− + HM → XH + M−. |
ΔPA = PA(X−) − PA(M−), | (2) |
The ΔPA values for the systems studied here are listed in Table 6 and ΔνOH is plotted against ΔPA in Fig. 7. Several interesting observations can be made. First, the set of red-shifts observed for a particular neutral molecule, and also for each halide anion, are consistent in that an increase in ΔPA leads to an increase ΔνOH. However, the overall agreement is less satisfactory. In more detail, the red-shifts observed for the H2O and i-PrOH complexes are similar, even though H2O exhibits a considerably larger PA. Hence, the extent of charge transfer does not only depend on the relative energy of the VB states (as we are assuming here), but also on other parameters, like the HB angle θOHX (see Table 2). The IHB in the water complexes are nearly linear, while they deviate substantially from linearity in the isopropanol (∼165°) and the HFIP (∼160°) complexes, reducing the orbital overlap with the orbitals and consequently the amount of charge transfer.
X−\HM | H2O | i-PrOH | HFIP |
---|---|---|---|
Cl− | −227 | −174 | −48 |
Br− | −269 | −216 | −90 |
I− | −307 | −254 | −128 |
![]() | ||
Fig. 7 Red-shift of the OH stretching frequency (ΔνOH) associated with the IHB as a function of the difference in proton affinities ΔPA (see eqn (2)) associated with the X−(HM) complexes with X− = Cl−, Br−, I− and HM = HFIP (circles), i-PrOH (squares), H2O (triangles). |
The transitions assigned to the OH stretching fundamentals are characterized by substantial mode mixing (see Table 4). To assess, in how far this affects the trends observed in Fig. 7, we compare this data to the corresponding data for the deuterated species in Fig. S15 (see ESI).† The extent of mode mixing is substantially reduced upon deuteration, as evidenced, for example, by the considerably simpler IRPD spectra of the deuterated species. Except for the expected reduction in the absolute red-shift upon deuteration, there is no qualitative difference between the two data sets, indicating that the effect of mode mixing only plays a minor role in the observed trends.
While the vibrational transitions in the fingerprint region are well reproduced within the harmonic approximation, the reliable prediction of the features in the OH(D) stretching region require an anharmonic treatment. DVR calculations allow evaluating the contribution of overtones and combination bands involving various bending modes to Fermi resonances in the O–H stretching region. These insights emphasize the relevance of anharmonic methods that go beyond standard approaches like VPT2 for a reliable prediction of the signal carrier, whenever ions are present, for example, in electrochemical applications.
Finally, we propose a generalized model for qualitatively predicting the vibrational frequency red-shift ΔνOH based on the difference in the proton affinities of the two conjugated base anions of a proton transfer reaction. This model qualitatively reproduces the observed trends, in particular, when the differences in the HB geometries are small.
The beam of anions is skimmed, collimated in a gas-filled radio frequency (RF) quadrupole ion guide, mass-selected using a quadrupole mass-filter and focused in a RF ring-electrode ion trap, held at a temperature of 12–14 K and continuously filled with D2 gas. Many collisions of the trapped ions with the buffer gas provide gentle cooling of the internal degrees of freedom close to the ambient temperature. At sufficiently low ion-trap temperatures, ion–messenger complexes are formed via three-body collisions.55 Every 100 ms, all ions are extracted from the ion trap and focused, both temporally and spatially, into the centre of the extraction region of the orthogonally-mounted double-focussing reflectron time-of-flight (TOF) tandem photofragmentation mass spectrometer and detected using the background-free IR1MS2 detection scheme.56 To this end, the ion packet is accelerated into the reflectron stage. Ions spread out in space according to their mass-to-charge ratio (m/z) and are refocused at the initial extraction region. Prior to be reaccelerated towards the MCP detector, ion–messenger complexes with a particular m/z value are irradiated by a properly timed and widely wavelength tunable IR laser pulse (bandwidth: 3.5 cm−1). The IR pulse is supplied by an optical parametric oscillator/amplifier (LaserVision: OPO/OPA/AgGaSe2) laser system pumped by an unseeded Nd:YAG laser (Continuum Surelite EX).57 IRPD spectra are recorded by monitoring the intensity of the irradiated ions and their photofragments while the laser wavelength is monitored online using a HighFinesse WS6-600 wavelength meter. The wavelength scanned continuously with a scan speed such that an averaged TOF mass spectrum (over 60 laser shots) is obtained every 2 cm−1. Typically, three to five scans are measured and averaged and the photodissociation cross section σIRPD is determined as described previously.20,58
Second, to account for higher order terms in the PES and dipole moment surface (DMS), we applied ab initio anharmonic algorithms developed in previous works37 by some of us in which the PES and DMS along the selected modes were scanned on the discrete variable representation (DVR) quadrature, and the Hamiltonian matrix can be diagonalized to obtain eigenstates. The IR absorption intensities were obtained from the eigenvectors and the DMS accordingly. To construct the PES (and DMS), single-point energy (and dipole) calculations at grid points generated by the Gauss–Hermite quadrature were performed along the selected vibrational modes; the PES is scanned with at the level of RI-MP2/aug-cc-pVTZ with corrections for CH and OH(OD) stretching modes at the level of DLPNO-CCSD(T)/aug-cc-pVTZ.37 The single point calculations for PES and DMS were performed with the ORCA program package.62 Since DLPNO-CCSD(T)/aug-cc-pVTZ is not applicable to I−, we only simulated complexes with X− = Cl−, Br−. The total number of grid points is quite large to diagonalize the Hamiltonian directly; therefore, we solve the Hamiltonian by transforming it into Finite-Basis-Representation (FBR).38 We refer to this approach as DVR-FBR/RI-MP2+DLPNO-CCSD(T)/aug-cc-pVTZ (or just DVR-FBR). The detail of the methodology, and the comparison between these two methods are discussed in the ESI.†
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4sc08456j |
This journal is © The Royal Society of Chemistry 2025 |