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

Mass-analyzed-threshold-ionization (MATI) spectroscopy of 1,2,3-substituted halogenated benzenes via different intermediate vibrational states in the S1 state

Sascha Krüger , Frank Witte , Jan Helfrich and Jürgen Grotemeyer *
Institut für Physikalische Chemie, Christian-Albrechts-Universität Kiel, Max-Eyth Str. 1, 24118 Kiel, Germany. E-mail: grote@phc.uni-kiel.de

Received 21st October 2014 , Accepted 24th November 2014

First published on 25th November 2014


Abstract

For the first time, two color resonant mass analyzed threshold ionization (MATI) spectroscopy has been applied in order to investigate the ionic properties of 1,3-dichloro-2-fluoro-benzene (1,3,2-DCFB) and 1,3-difluoro-2-chloro-benzene (1,3,2-DFCB) radical cations in their electronic ground state. The ionic ground state of the different samples has been investigated via different S1 intermediate states and compared to 1,2,3-trichlorobenzene measured in previous work. Additionally quantum chemical calculations at DFT (density functional theory) and TDDFT (time-dependent density functional theory) level of theory have been performed to support experimental findings. From the MATI spectra the adiabatic ionization energies of 1,3-dichloro-2-fluorobenzene and 1,3-fluoro-2-chlorobenzene could be determined to be 75.242 ± 6 cm−1 and 75.627 ± 6 cm−1, respectively. Several vibrational modes of both compounds have been assigned by comparison of the experimental and theoretical results.


1. Introduction

Halogenated aromatic molecules became a topic of great interest due to their widespread presence in industrial processes and hence presence in the environment as potentially toxic and cancerogenic pollutants. Moreover halogenated benzenes are important model substances to validate theoretical concepts, such as vibronic coupling and others. As one of the spectroscopically best-investigated molecules of all, benzene can serve as an outstanding reference system to study the influence of tailored perturbations, such as careful choice of substitution pattern, to the (ro)vibronic structure. As an immediate effect a shift in transition energies, ionization energy, molecular geometry and vibrational frequencies can be observed. The characteristics of these effects strongly depend on the number, type and localization of the substituents. As the magnitude of these effects are rather small (from a few wavenumbers up to a hundred wavenumbers), high resolution techniques, such as mass analyzed threshold ionization (MATI) or zero kinetic energy (ZEKE) spectroscopy are ideally suited to investigate such phenomena.1–4 Chloro- and fluoro-benzenes have been subject to various studies investigating the vibronic properties of the first electronic excited state (S1) as well as cationic ground state (D0).1,5–13 In this paper we will focus particularly on out-of-plane b1-symmetric modes. Modes of this symmetry are connected to interesting phenomena observed in a whole variety of halobenzenes. For example, difluorobenzenes exhibit large frequency lowering of b1 (b3u under D2h in case of p-DFB) symmetric modes in going from S0 to S1 or from D0 to S1,14 respectively. The intensity gain of this symmetry forbidden mode was interpreted in terms of the pseudo-Jahn–Teller effect (PJTE). The PJTE is based on the idea that a geometrical distortion blurs the difference in symmetry between two or more electronic states of virtually equal energy. In 1,2,4,5-tetrafluorobenzene (D2h) it was found that the b2g-symmetric mode 11 shows an unusual strong activity in the form of a long progression (v2n11). Recent results show that the planarity of this molecule is distorted along the eigenvector of the 11 mode during excitation to 1B2uS1 state. It is an interesting question to elucidate if these phenomena can be extended to further congeners and identified as a general feature. In particular in the case of the molecular geometry change it is crucial to determine the involved vibrational modes and participating electronic states. For this purpose, the experimental data are compared to theoretical calculations, which enable the possibility of assigning the observed vibrational modes. In this paper we present mass-selected two color two photon Resonance Enhanced Multi Photon Ionization (2C2P-REMPI) spectra of the S1 ← S0 transition and MATI spectra via different S1-intermediate states of 1,3-dichloro-2-fluorobenzene and 1,3-dichloro-2-fluorobenzene. Some assignments of 1,2,3-trichlorobenzene reported previously1 were reconsidered and compared to the newly gained knowledge.

2. Experimental setup

The experimental setup consists of a homemade time-of-flight (TOF) mass spectrometer as described in detail previously.1,15,16 Briefly, the spectrometer consists of a standard second order corrected reflectron time-of-flight mass spectrometer equipped with single stage ion source. The laser system used for excitation and ionization consists of two dye lasers (Laser Analytical Systems LDL 205, Lambda Physics FL 3002). Each dye laser is pumped by a dedicated Nd:YAG laser (Lumonics HY-1200, Continuum Surelite II). A Quantum Composers 9600+ delay pulse generator fed by an external clock operated at 10 Hz controls flash lamp and Q-switch delays. The output of each dye laser is frequency-doubled by a BBO-I crystal yielding tunable ranges from 245 to 285 nm. Wavelength calibration of both dye lasers is performed by recording an opto-galvanic spectrum with a neon hollow cathode lamp yielding accuracy better than 2 cm−1. Initially a supersonic molecular beam of sample molecules and seed gas (argon) with a backing pressure of approximately 2 bar is expanded via a pulsed jet valve (General Valves Series 9) into the ion source. To obtain a sufficient vapor pressure the sample is heated to approximately 80 °C. Excitation or ionization is accomplished by multi photon absorption under field-free conditions. In the MATI modus, promptly generated ions are discriminated against Rydberg neutrals through the application of a weak electrical field of 1–2 V cm−1 by a subsequent delay (∼100 ns) to multi-photon excitation. Finally, a high voltage pulse (890 V cm−1) switched by a Behlke HS56-01 fast thyristor ionizes the Rydberg neutrals and accelerates them into the mass spectrometer. In REMPI modus, the generated ions are accelerated directly into the TOF without the retarding field. The ion signal is detected by a conventional dual micro-channel-plate detector and transferred to a LeCroy 534M digital oscilloscope. A computer, linked to the oscilloscope by a GPIB connection performs the data acquisition and processing. The 123-DCFB was purchased from Aldrich, 126-DFCB from ABCR and were used without further purification.

3. Quantum-chemical calculations

Quantum chemical calculations were performed in order to assign the observed vibrational bands and to support the experimental findings. The software package Turbomole17–20 was used for all quantum chemical calculations. Geometry optimizations and subsequent frequency analyses for molecules in the electronic ground state (S0) and the cationic ground state (D0) were conducted at the density functional theory (DFT) level of theory with both the gradient corrected functional BP86[thin space (1/6-em)]21a and the hybrid functional B3LYP.21b The triple basis set TZVPP21c has been applied to all DFT calculations. Calculations for the S1 were done analogously using the time dependent density functional theory (TDDFT). For comparison reasons, geometry calculations and frequency analyses were also treated at the coupled cluster (CC2) level of theory for the S0- and S1-state using the Ahlrichs basis set cc-pVTZ.21d The nomenclature used for assignment of vibrational bands is according to Varsanyi and Szoke,22 which is derived from Wilson's notation of the Benzene modes.23

We refrained from the use of scaling factors to fit the calculated frequencies.

4. Experimental results

4.1. 1,3-Dichloro-2-fluorobenzene (1,3,2-DCFB)

4.1.1. REMPI spectrum. According to electric dipole selection rules the transition to first excited singlet state 1B2S1(π* ← π) in 1,3,2-DCFB is allowed (y-polarized). For benzene, the corresponding transition A1B2u ← X1A1g is dipole forbidden under D6h symmetry. The reduced symmetry in 1,3,2-DCFB (C2v) leads to an electronically as well as vibronically allowed transition. Starting from the premise that the molecules are sufficiently cooled with regards to their degrees of freedom, only total symmetric (a1) vibrations are allowed according to the Franck–Condon (FC) principle. Due to vibronic coupling to intensive, nearby states, vibrations of a2 and b2 symmetry can gain intensity as well. The REMPI spectrum shown in Fig. 1 in units of internal energy could be recorded in the range up to 1300 cm−1. The excitation energy of the first excited state S1 could be determined for the first time to be 36[thin space (1/6-em)]460 ± 2 cm−1.
image file: c4ra12873g-f1.tif
Fig. 1 (1 + 1′) REMPI spectrum of 1,3-dichloro-2-fluoro-benzene (1,3,2-DCFB).

Normal coordinate analysis for 1,3,2-DCFB yields following distribution of fundamental modes under C2v symmetry: Γvib = 11 × a1, 10 × b2, 3 × a2, 6 × b1. Clearly, the spectrum is dominated by the total symmetric modes 11 (565 cm−1) and 18a (977 cm−1). Moreover the a1-symmetric modes 9a1, 6a1, 7a1, 121 (186 cm−1, 349 cm−1, 792 cm−1, 1104 cm−1) could be identified with lower intensity. With 16a1 (360 cm−1) also an a2-symmetric mode could be assigned. As expected, no modes of b1-symmetry could be assigned to the spectrum in accordance with selection rules. Particular attention should be paid to the low frequency band at 120 cm−1. A band with such a low frequency can with certainty be assigned to an out-of-plain mode. Based on correlation with MATI spectra we assigned the first even overtone of the b1-symmetric mode 17b to this band. The frequency of 52 cm−1 calculated with the coupled cluster method is in reasonable agreement with the experimental value for the half of the overtone for 17b2. The bands at 247 cm−1 (151), 337 cm−1 (6b1), 500 cm−1 (31) and 1236 cm−1 (18b1) were assigned to b2-symmetric modes. The latter two experimental values are in excellent agreement with both calculated values (TDDFT, CC2) for the first excited singlet state of 1,2,3-DCFB. Comparing the calculated values for the 151 and 6b1 mode a major deviation becomes apparent, just like in the case of 17b. All performed calculations show good agreement with the experiment for the modes 9a1, 6a1, 11, 7a1, 18a1, 121, 16a1, whereas the best results were obtained with the CC2 level of theory (see Table 1).

Table 1 Comparison experimental and calculated normal modes observed in the S0, S1 and D0 state of 1,3,2 DCFB
S0(C2v) S1 (C2v) (C2v) (CS) D0 (C2v)
Wilson Sym. exp. B3LYP BP86 CC2 exp. B3LYP BP86 CC2 exp. B3LYP BP86
a IR-band.24 b From combination band. c From overtone.
9a a1 256a 192 186 189 186 241 232 183 190 195 189
6a a1 376a 379 369 380 349 355 361 343 361 381 372
1 a1 597a 601 584 596 565 538 522 558 578 603 586
7a a1 803a 839 813 829 792 831 815 791 798 820 797
18a a1 1058a 1092 1061 1080 977 1013 998 991   1047 1027
12 a1 1112a 1116 1083 1124 1104 1077 1046 1079   1143 1104
13 a1 1205a 1277 1240 1277   1314 1320 1242   1339 1295
19a a1 1462a 1490 1444 1485   1213 1229 1383   1459 1412
8a a1 1572a 1613 1564 1608   1509 1476 1512   1618 1567
20a a1 3084a 3189 3119 3220   3159 3080 3210   3197 3126
2 a1   3212 3142 3243   3220 3147 3246   3218 3147
10a a2 215a,b 208 201 207   120 123 71 174 180 173
16a a2 530a,b 553 534 518 360 458 495 313 512 498 476
17a a2 894a 921 876 890   1045 1060 616   936 897
17b b1 115a,b 107 103 107 60c 31 38 52 95 88 85
10b b1 260a,b 282 271 278 270c 263 259 178 264 274 265
16b b1 513a 533 512 542   502 454 364 452 455 441
4 b1 705a 732 699 680   766 755 460 732 743 715
11 b1 770a 794 758 771   847 827 601 678 826 792
5 b1 965a 983 937 949   1004 966 782   1009 966
15 b2 279a 258 248 252 247 62 61 227 242 233 253
6b b2 401a 404 393 401 337 202 209 319 331 300 331
3 b2 539a 545 528 538 500 500 498 519 471 513 504
7b b2 829a 798 779 810   698 683 759 827 812 796
9b b2 1154a 1180 1145 1170   1177 1144 1105   1083 1104
18b b2 1205a,b 1233 1193 1223 1236 1250 1185 1170   1330 1192
14 b2 1255a 1313 1322 1421   1413 1360 1497   1226 1296
19b b2 1446a 1481 1432 1465   1446 1407 1389   1514 1466
8b b2 1583a 1615 1566 1608   1574 1545 1683   1371 1384
20b b2 3084a 3207 3137 3235   3188 3177 3141   3215 3144


4.1.2. MATI spectra. The MATI spectra via the S1 intermediate states 00, 17b1, 9a1, 11 are shown in Fig. 2. The origin of the D0(2B1) state and with that the adiabatic ionization energy was found to be 75.242 ± 6 cm−1 (9.3288 ± 0.0007 eV). This value is in good accordance with the previously by conventional photoelectron spectroscopy determined value of 9.32 ± 0.02 eV.25
image file: c4ra12873g-f2.tif
Fig. 2 MATI spectra of 1,3,2-DCFB of different vibrational intermediate states of the first excited electronic state: (a) via 11, (b) via 9a1, (c) via 17b2, (d) via the electronic origin (00).

Besides the 00 transition, the MATI spectrum obtained via the electronic origin exhibits several more additional resonances. The most prominent peak in the spectrum is the band assigned to the 17b mode, which is constituting a violation of the Δv = 0 propensity rule. Additionally, the mode 17b1 at 95 cm−1 appears as a combination band (17b19a1) at 288 cm−1. Moreover the spectrum is conspicuously rich in b1-symmetric modes. In addition to 17b1 we assigned the 10b1 (264 cm−1), 16b1 (452 cm−1), 111 (678 cm−1) and 41 (732 cm−1). Another out-of-plane vibration (10a1) with a2-symmetry and low intensity has been assigned to the band at 174 cm−1. The bands at 361 cm−1, 578 cm−1 have been assigned to the total symmetric modes 6a1 and 11, respectively. b2 symmetric modes were identified at 242 cm−1 (151), 331 cm−1 (6b1), 471 cm−1 (31) and 827 cm−1 (7b1). The band at 657 cm−1 fits the overtone 6b1.

The MATI spectrum obtained via the S117b2 shows a short, three-membered regular progression with transitions corresponding to the excitation of one, two and three quanta. In contradiction to the Δv = 0 propensity rule, the band labeled 17b1 is the most prominent peak in the spectrum. It is notable that the 17b mode also appears in the combination bands 17b19a1 (288 cm−1) and 16b16b1 (423 cm−1). Also apparent is the richness in b1-symmetric modes, first and foremost the intense 16b1 (448 cm−1), but also the 10b1 (264 cm−1), 111 (680 cm−1), 41 (728 cm−1). The MATI spectrum obtained via the S19a1 mode also shows a breakdown of the Δv = 0 propensity rule. Not the vertical transition into the D09a1 state is the dominating one, but the band at 242 cm−1 (151). In addition the 17b mode appears again, also as a combination in 17b19a1. A further band at 331 cm−1 could be identified with 6b1. The measured values are in good accordance with the calculated values (195 and 189 cm−1 for the 9a1, 233 and 253 cm−1 for the151 and also 300 and 331 cm−1 for the 6b1).

The MATI spectrum via S111 continues the series of MATI spectra characterized by a breakdown of the Δv = 0 propensity rule in favor for the vibronic transition into the D017b1 state. Further active modes that could be assigned in the recorded range up to 750 cm−1 are the 151 (241 cm−1), 6b1 (331 cm−1) and 11 (578 cm−1). The experimentally determined frequency for the 11 mode is in good accordance with the calculated values of 578 cm−1 or 603 cm−1 respectively. It should be noticed that, owing to the heavy substituents, the displacement pattern of the ‘ring-breathing-mode’ shows a striking deviation from the original benzene pattern. It resembles clearly the pattern of mode 6a found for benzene. The assignment of the band at 578 cm−1 to the mode 6a is excluded since it was already doubtlessly assigned to the band at 361 cm−1.

4.2. 1,3-Difluoro-2-chloro-benzene (1,3,2-DFCB)

4.2.1. REMPI spectrum. 1,3,2-DFCB belongs to C2v point group and has a dipole allowed transition to the 1A1S1(π* ← π) first excited state. According to FC-principle only transitions in total symmetric (a1) modes are allowed, b1 and b2 can gain intensity due to vibronic coupling mechanism, a2 modes are symmetry forbidden. The REMPI spectrum shown in Fig. 3 in units of internal energy could be recorded in the range up to 800 cm−1. The excitation energy of the first excited state S1 could be determined for the first time to be 37[thin space (1/6-em)]449 ± 2 cm−1.
image file: c4ra12873g-f3.tif
Fig. 3 (1 + 1′) REMPI spectrum of 1,3-difluoro-2-chloro-benzene (1,3,2-DFCB).

The electronic origin and the total symmetric mode 9a1 at 228 cm−1 dominates clearly the measured REMPI spectrum. It should be noted that the remaining bands at 381 cm−1 and 607 cm−1 that were both assigned to the total symmetric modes 61 and 11, respectively. Quite unusual for total symmetric modes we found poor consistency between calculated and observed bands: calculation substantially underestimate the 6a1 and 11 with 188 cm−1 to 202 cm−1 and 488 cm−1 to 517 cm−1, respectively. The 9a1 has been overestimated with 364 cm−1 to 374 cm−1. Nevertheless the correlation with MATI spectra backups the assignment of the total symmetric modes. Moreover, the calculations suggest the mode 7a1, that cannot be seen in the spectrum, to appear in the spectrum between 710 cm−1 and 750 cm−1. As expected, no modes of a2 symmetry could be assigned to the spectrum in accordance with selection rules. In contrast to 1,2,3-TCB and 1,3,2-DCFB, transitions to b1-symmetric modes are allowed in 1,3,2-DFCB. The bands at 25, 98 and 404 cm−1 could be identified with the b1-symmetric modes 17b1, 10b1 and 16b1. In this case, the calculated frequencies are in good agreement with the observed ones. The 17b1 has been predicted to appear between 31 cm−1 and 51 cm−1, the 10b1 between 176 cm−1 and 215 cm−1 and the 16b1 between 351 cm−1 and 428 cm−1. With 151 and 6b1 two a2-symmetric modes could be assigned. Noteworthy is the appearance of the 15 mode as a short, three-membered progression of overtones (151, 152, 153). The measured frequencies of 443 cm−1 for 6b1 and 63 cm−1 for 151 are well reproduced by the B3LYP calculation with 439 and 64 cm−1. The BP86 calculation predicts the 151 correctly with 67 cm−1, while it slightly overestimates the 6b1 with 463 cm−1. The CC2 calculation gives good results in predicting the 6b1 with 439 cm−1, while it overestimates the 151 with 154 cm−1.

4.2.2. MATI spectra. The MATI spectra via the S1 intermediate states 00, 17b1, 9a1 are shown in Fig. 4. The origin of the D0(2B1) state and with that the adiabatic ionization energy was found to be 75.627 ± 6 cm−1 (9.3765 ± 0.0007 eV). This value is in good accordance the previously by photoelectron spectroscopy determined value of 9.37 ± 0.02 eV.25
image file: c4ra12873g-f4.tif
Fig. 4 MATI spectra of 1,2,6 DCFB of different vibrational intermediate states of the first excited electronic state: (a) via 9a1, (b) via 17b2, (c) via the electronic origin (00).

In accordance with the Δv = 0 propensity rule, the MATI spectrum obtained via the electronic origin is dominated by the 00-band. Furthermore the spectrum is characterized by strong activity of the two total symmetric modes 9a1 and 6a1. Within the recorded range of 1000 cm−1, the spectrum exhibits a progression in 9a composed of the first three overtones 9a1 (293 cm−1), 9a2 (588 cm−1) and 9a3 (882 cm−1). The 6a appears as ground vibration 6a1 (418 cm−1), first overtone 6a2 (839 cm−1) as well as combination vibration 9a16a1 (706 cm−1).

Additionally we assigned the band at 925 cm−1 to a combination band of modes 9a111. The relative intensive band at 630 cm−1 has been identified with the mode 11. The shoulder at 649 cm−1 in the peak labeled with 11 has been assigned to the mode 41. Exhibiting a similar weak intensity as the 41, another b1-symmetrical mode (7a1) could be identified at 778 cm−1. In the view of previous work,1 the assignment of the bands at 95 cm−1 and 196 cm−1 to the modes 17b1 and 17b2 respectively seems appropriate. Assignments that are more tentative are the 16b1 and 111 to the bands at 439 cm−1 and 839 cm−1. The a2-symmetrical modes 10a1 and 16a1 are assigned to the weak bands at 201 cm−1 and 546 cm−1. With exception of 41 all calculated frequencies are in good accordance with the experimentally determined ones.

Provided the validity of the Δv = 0 propensity rule, the MATI spectrum obtained via the S117b1 mode ought to be characterized by a towering 17b1 band. Contrary to expectations, the band 17b1 could not be observed in the spectrum shown in Fig. 4. Instead, the spectrum shows the 9a1 at 291 cm−1 as the most intensive resonance absorption. As can be seen from the correlation between the MATI spectra in Fig. 4, the first overtone 9a2 appears at 586 cm−1 as well. The signals at 419 cm−1 and 629 cm−1 are identified as the 6a1 and 11 total symmetric modes. Obviously, the signals for the 11-vibration is broadened. Based on the MATI spectrum of the 00 transition, we assigned this signal both to the 11 and 41 vibrations, the latter of which gives rise to the shoulder. The MATI spectrum obtained via the S19a1 shows three prominent peaks: the 6a1, the 9a16a1 and the 18a1. Without folding the spectrum with the laser power, no absolute statement can be made in regards to the peak intensities. The three peaks are of nearly equal intensity (approx. 5% difference), so it is hard to tell which one is the highest in intensity. Nevertheless it can be stated that in contrary to the predictions of the propensity rule, transitions to 6a1 and 9a16a1 are at least as same as intensive while the vertical transition into the D09a1 state is scarcely visible in the spectrum. Nevertheless the mode 9a appears in a series of overtones 9a16a1 (710 cm−1), 9a111 (924 cm−1) and 9a26a1 (1003 cm−1). Remarkably, the spectrum exhibits a strong activity in 6a modes: the bands assigned to the modes 6a1 and 9a16a1 are among the three most intense bands in the spectrum. The 6a1 vibration shows progression activity (6a1 and 6a2 in the recorded range). In addition, the 6a1 mode contributes to overtone bands as already discussed above. The weak features at 97, 193 and 213 cm−1 have been assigned on the basis of our previous studies7,8 to the out-of-plane modes 17b1, 17b2 and in-plane mode 151.

5. Discussion

5.1. 1,3,2-DCFB

Compared to electronic and cationic ground states the 17b exhibits a drastically decreased frequency in the first electronic excited state (see Table 1). Such a frequency lowering suggests a geometrical distortion along the eigenvector of 17b in going from S0 to S1 or from D0 to S1, respectively. As considered earlier by Tsuchiya et al.14 for difluorobenzene it is highly suggested that this phenomenon is the result of strong vibronic coupling between the S1 and nearby states. But the most important indication for a such a distortion shows up in the MATI spectra via the electronic origin and, first and foremost, via the 17b2. The latter exhibits a three-membered progression of the mode 17b. Not just that this progression shows a shift of the Franck–Condon maximum between S1 and D0, the fact that the transition from D017b1 ← S117b2 is favored could be interpreted as a replanarisation of the molecular geometry in going from S1 to D0 along that mode.

Also during ionization via the vibrationless level of the first excited state, the transition D017b1 ← S00 is clearly favored over D000 ← S100 (see MATI via 00Fig. 2). A major role of the 17b mode during ionization is aggravated by the fact the 17b also appears as combination bands (17b19a1 in each of the recorded MATI spectra, 17b16b1 in the MATI via 17b2). The strong activity of the mode 16b1 in the MATI via 17b2 and the general, unusual appearance of multiple other out-of-plane modes throughout the MATI spectra give rise to the assumption that the geometry in the S1 is a product of a distortion along several modes. The CC2 computed geometry optimization seem to support the statements derived from the experimental data: the geometry shown in Fig. 5 is obviously not just the product of a distortion along a single (out-of-plane) normal mode (Table 2).


image file: c4ra12873g-f5.tif
Fig. 5 Excited state geometry of 1,3,2-DCFB obtained by CC2-optimization.
Table 2 Comparison experimental and calculated normal modes observed in the S0, S1 and D0 state of 1,3,2 DFCB
S0(C2v) S1 (C2v) (C2v) (CS) D0 (C2v)
Wilson Sym. exp. B3LYP BP86 CC2 exp. B3LYP BP86 CC2 exp. B3LYP BP86
9a a1   301 290 295 228 374 364 367 293 304 294
6a a1   437 425 436 381 198 202 188 418 435 422
1 a1   617 602 618 607 517 500 488 630 636 620
7a a1   783 760 776   748 729 713 778 784 761
19a a1   1072 1142 1124   1506 1462 1360   1360 1318
18a a1   1121 1042 1065   1113 1083 1059   1045 1021
12 a1   1314 1275 1317   1005 984 949   1150 1114
13 a1   1494 1448 1491   1294 1252 1268   1413 1372
8a a1   1637 1587 1635   1577 1542 1507   1665 1613
20a a1   3191 3121 3224   3201 3133 3212   3199 3128
2 a1   3212 3143 3248   3223 3155 3266   3217 3148
10a a2   249 239 247   194 189 183 201 218 209
16a a2   597 573 588   762 766 519 546 565 541
17a a2   893 850 873   1043 1107 788   920 880
17b b1   124 119 125 25 31 41 51 97 98 94
10b b1   275 264 273 198 215 214 176 280 289 277
16b b1   530 510 521 404 428 407 351 439 434 415
4 b1   717 684 675   805 779 649 649 729 700
11 b1   792 754 772   640 632 548 839 837 802
5 b1   968 922 944   979 946 781   1000 954
15 b2   212 204 208 63 64 67 154 213 221 212
6b b2   509 494 501 443 451 463 439 367 386 369
3 b2   555 534 543   511 501 492 489 539 522
7b b2   1020 990 1014   979 958 948 982 1018 986
9b b2   1179 1088 1167   1137 1128 1133   1110 1087
18b b2   1266 1226 1262   1208 1171 1194   1303 1259
14 b2   1322 1331 1424   1380 1393 1456   1354 1316
19b b2   1503 1456 1495   1441 1382 1381   1548 1500
8b b2   1626 1577 1623   1477 1451 1680   1381 1370
20b b2   3206 3137 3242   3207 3139 3215   3214 3145


The quantum chemical calculations predict a S1(π* ← π) transition into the first excited state for 1,3,2-DCFB. However, due to σ*-orbitals localized on the Halogen–Carbon bond, it is possible for modes of appropriate symmetry to induce coupling to (πσ*)-states. In accordance with the assumption of a vibronic coupling effect we prevalently observed a substantial decrease in frequency comparing the S0 state and the S1 state of the neutral for formal forbidden modes with strong influence of halogen atoms on their displacement pattern. For the modes 15, 6b, 3 (b2 symmetry) and 16a (a2 symmetry) we observed such a decrease.

The decrease in frequency for the modes 15, 6b, 3 is explainable in accordance with the Herzberg–Teller (HT) effect. However, the frequency lowering of the symmetry allowed mode 17b2 could not be described in terms of the HT effect in a sufficient way. It seems that the distortion along 17b blurs the symmetry differences between the (πσ*)- and (ππ*)-states which would be an indication for a pseudo Jahn–Teller effect.

5.2. 1,3,2-DFCB

In 1,3,2-DFCB the vibration 15 appears in the REMPI spectrum as a progression-forming and largely frequency lowered mode. Remarkably, the progression reveals a negative anharmonicity. The occurrence of the 15 as a b2 symmetric mode is explained in terms of the Herzberg–Teller effect. Besides the mode 15, also the modes 6a and 9a exhibit a largely lowered frequency in the first excited state S1. Both modes are likely to contribute in pseudo Jahn–Teller distortion in going from the electronic ground to excited state or from excited state to ionic ground state, respectively. This is seen from the different MATI spectra as well as the different calculated structures given in Tables 3–8. The MATI spectrum via the electronic origin exhibits a harmonic, three-membered progression of mode 9a (9a1, 9a2, 9a3), whereby the band 9a1 is almost as intense as the 0–0 transition. Such a shift of the Franck–Condon maximum is a clear indication for a distortion along the eigenvector of this mode during ionization.
Table 3 B3LYP/TZVPP calculated geometries of 1,3,2-DCFB

image file: c4ra12873g-u1.tif

B3LYP/TZVPP C 2v C 2 C 2v
Bond length [Å] S0 S1 D0
C1–C2 1.391 1.364 1.438
C2–C3 1.391 1.445 1.438
C3–C4 1.389 1.403 1.373
C4–C5 1.389 1.370 1.407
C5–C6 1.389 1.370 1.407
C6–C1 1.389 1.380 1.373
C1–Cl7 1.739 2.358 1.698
C2–F8 1.334 1.325 1.290
C3–Cl9 1.739 1.700 1.698
C4–H10 1.080 1.080 1.081
C5–H11 1.081 1.082 1.082
C6–H12 1.080 1.085 1.081
[thin space (1/6-em)]
Bond angles [°]
C1–C2–C3 120.125 123.523 122.001
C2–C3–C4 119.972 119.825 118.077
C3–C4–C5 119.648 117.525 119.488
C4–C5–C6 120.635 120.436 122.869
C5–C6–C1 119.648 123.643 119.488
C6–C1–C2 119.972 115.049 118.077
Cl7–C1–C2 119.336 131.346 118.378
C1–C2–F8 119.938 121.023 118.996
C4–C3–Cl9 120.692 120.051 123.541
[thin space (1/6-em)]
Dihedral angle [°]
Cl7–C1–C6–H12 0 −0.018 0
Cl7–C1–C2–F8 0 0.016 0
Cl9–C3–C4–H10 0 −0.005 0


Table 4 BP86/TZVPP calculated geometries of 1,3,2-DCFB

image file: c4ra12873g-u2.tif

BP86/TZVPP C 2v C 2 C 2v
Bond length [Å] S0 S1 D0
C1–C2 1.400 1.377 1.444
C2–C3 1.400 1.445 1.444
C3–C4 1.396 1.411 1.382
C4–C5 1.395 1.379 1.411
C5–C6 1.395 1.434 1.411
C6–C1 1.396 1.395 1.382
C1–Cl7 1.740 2.325 1.701
C2–F8 1.341 1.335 1.300
C3–Cl9 1.740 1.708 1.701
C4–H10 1.088 1.088 1.089
C5–H11 1.089 1.090 1.090
C6–H12 1.088 1.094 1.089
[thin space (1/6-em)]
Bond angles [°]
C1–C2–C3 120.007 123.861 121.791
C2–C3–C4 120.019 120.100 118.195
C3–C4–C5 119.613 117.257 119.447
C4–C5–C6 120.728 120.361 122.865
C5–C6–C1 119.613 124.461 119.477
C6–C1–C2 120.019 113.959 118.195
Cl7–C1–C2 119.198 130.690 118.269
C1–C2–F8 119.938 121.023 118.996
C4–C3–Cl9 120.783 119.621 123.535
[thin space (1/6-em)]
Dihedral angle [°]
Cl7–C1–C6–H12 0 −0.030 0
Cl7–C1–C2–F8 0 0.038 0
Cl9–C3–C4–H10 0 −0.006 0


Table 5 CC2/cc-pVTZ calculated geometries of 1,3,2-DCFB

image file: c4ra12873g-u3.tif

BP86/TZVPP C 2v C 2
Bond length [Å] S0 S1
C1–C2 1.395 1.432
C2–C3 1.395 1.432
C3–C4 1.393 1.426
C4–C5 1.393 1.421
C5–C6 1.393 1.421
C6–C1 1.393 1.426
C1–Cl7 1.725 1.719
C2–F8 1.333 1.326
C3–Cl9 1.725 1.719
C4–H10 1.080 1.079
C5–H11 1.081 1.082
C6–H12 1.080 1.079
[thin space (1/6-em)]
Bond angles [°]
C1–C2–C3 119.992 121.970
C2–C3–C4 120.078 118.706
C3–C4–C5 119.618 118.910
C4–C5–C6 120.617 122.532
C5–C6–C1 119.618 118.910
C6–C1–C2 120.078 118.706
Cl7–C1–C2 119.105 118.817
C1–C2–F8 120.004 119.015
C4–C3–Cl9 110.817 120.760
[thin space (1/6-em)]
Dihedral angle [°]
Cl7–C1–C6–H12 0 −17.842
Cl7–C1–C2–F8 0 17.222
Cl9–C3–C4–H10 0 −17.842


Table 6 B3LYP/TZVPP calculated geometries of 1,3,2-DFCB

image file: c4ra12873g-u4.tif

B3LYP/TZVPP C 2v C 2v C 2v
Bond length [Å] S0 S1 D0
C1–C2 1.393 1.365 1.440
C2–C3 1.393 1.365 1.440
C3–C4 1.384 1.444 1.365
C4–C5 1.389 1.384 1.409
C5–C6 1.389 1.384 1.409
C6–C1 1.384 1.444 1.365
C1–F7 1.340 1.326 1.309
C2–Cl8 1.729 2.361 1.665
C3–F9 1.340 1.326 1.309
C4–H10 1.081 1.081 1.081
C5–H11 1.081 1.080 1.082
C6–H12 1.081 1.081 1.081
[thin space (1/6-em)]
Bond angles [°]
C1–C2–C3 117.528 111.693 118.393
C2–C3–C4 121.919 126.328 120.838
C3–C4–C5 118.997 118.962 118.528
C4–C5–C6 120.640 117.726 122.876
C5–C6–C1 118.997 118.962 118.528
C6–C1–C2 117.528 126.328 120.838
F7–C1–C2 118.843 120.164 117.338
C1–C2–Cl8 121.236 124.155 120.803
C4–C3–F9 119.237 113.508 121.823
[thin space (1/6-em)]
Dihedral angle [°]
F7–C1–C6–H12 0 0.027 0
F7–C1–C2–Cl8 0 −0.204 0
F9–C3–C4–H10 0 −0.032 0


Table 7 BP86/TZVPP calculated geometries of 1,2,3-DFCB

image file: c4ra12873g-u5.tif

BP86/TZVPP C 2v C 2v C 2v
Bond length [Å] S0 S1 D0
C1–C2 1.401 1.378 1.446
C2–C3 1.401 1.378 1.446
C3–C4 1.391 1.446 1.374
C4–C5 1.396 1.391 1.414
C5–C6 1.396 1.391 1.414
C6–C1 1.391 1.446 1.374
C1–F7 1.348 1.337 1.318
C2–Cl8 1.730 2.332 1.671
C3–F9 1.348 1.337 1.318
C4–H10 1.089 1.089 1.089
C5–H11 1.089 1.087 1.090
C6–H12 1.089 1.089 1.089
[thin space (1/6-em)]
Bond angles [°]
C1–C2–C3 117.514 110.326 118.508
C2–C3–C4 121.885 127.177 120.727
C3–C4–C5 118.999 118.870 118.543
C4–C5–C6 120.718 117.579 122.956
C5–C6–C1 118.999 118.870 118.543
C6–C1–C2 117.514 127.177 120.727
F7–C1–C2 118.763 119.276 117.392
C1–C2–Cl8 121.242 124.836 120.740
C4–C3–F9 119.349 113.548 121.870
[thin space (1/6-em)]
Dihedral angle [°]
F7–C1–C6–H12 0 0.041 0
F7–C1–C2–Cl8 0 −0.191 0
F9–C3–C4–H10 0 −0.032 0


Table 8 CC2/cc-pVTZ calculated geometries of 1,2,3-DFCB

image file: c4ra12873g-u6.tif

B3LYP/TZVPP C 2v C 2v
Bond length [Å] S0 S1
C1–C2 1.396 1.416
C2–C3 1.396 1.416
C3–C4 1.387 1.431
C4–C5 1.394 1.407
C5–C6 1.394 1.407
C6–C1 1.387 1.431
C1–F7 1.339 1.337
C2–Cl8 1.716 1.825
C3–F9 1.339 1.337
C4–H10 1.080 1.825
C5–H11 1.081 1.337
C6–H12 1.080 1.083
[thin space (1/6-em)]
Bond angles [°]
C1–C2–C3 117.753 110.391
C2–C3–C4 121.786 125.833
C3–C4–C5 118.972 120.624
C4–C5–C6 120.732 116.095
C5–C6–C1 118.972 120.624
C6–C1–C2 121.786 125.833
F7–C1–C2 118.680 118.508
C1–C2–Cl8 121.124 118.556
C4–C3–F9 119.534 115.651
[thin space (1/6-em)]
Dihedral angle [°]
F7–C1–C6–H12 0 0.912
F7–C1–C2–Cl8 0 46.945
F9–C3–C4–H10 0 −0.912


A similar situation is found in the MATI spectrum via 9a1. Here, in contradiction to the Δv = 0 propensity rule, the transition into the S16a1 state is the most favorable while the vertical transition into the 9a1 is scarcely observable. Nevertheless the mode 9a is strongly present in overtones (9a16a1, 9a111, 9a26a1). The geometry found by the TDDFT calculations (Fig. 6a; Tables 6 and 7) support the experimental findings in predicting a C2v-symmetric structure distorted along the 6a. The CC2 calculation suggests a 17b-like distorted structure (Fig. 6b; Table 8).


image file: c4ra12873g-f6.tif
Fig. 6 Excited state geometry of 1,3,2-DFCB obtained by different quantum mechanical methods. (a) TDDFT, (b) CC2.

The CC2 calculation suggests a 17b-like distorted structure (Fig. 6b). Considering the vanishing low activity of 17b in the MATI spectra, the suggestion seems incorrect. Nevertheless the substantial decrease in frequency for 6a1 and 9a1 comparing the D0 state and the S1 state is accurately reproduced by all quantum chemical methods employed. Comparing experimental and calculated results for 1,3,2-DFCB, the unusual, striking deviation (up to 49%) for a1-symmetrical S1-frequencies was particularly noticeable. A poor reproduction of the molecular equilibrium structure in the S1 (wrong minimum on the PES) could be one possible explanation.

A contrary indication is the fact that we find very similar frequencies for widely differing structures predicted by TDDFT (planar structure) and CC2 method (out-of-plane-structure). The planar structure resembles the experimental observations well, but since all frequency analyses were performed in harmonic approximation, a strong anharmonicity for this particular vibrations could be another explanation for this findings.

6. Conclusion

Comparing the results from two compounds presented in this paper with results obtained from different isomeric trichloro-benzenes,1 the increasing number of fluorine atoms lead to a progressive decrease in some of the observed frequencies in going from S0 to S1 and from D0 to S1, respectively. This is especially true for modes with a strong fluorine participation in their vibrational pattern like 17b (cf.Table 1). We interpreted this as a strong indication for an out-of-plane distortion during excitation or a replanarisation during ionization along the eigenvector of those modes. This phenomenon could be attributed to an above, bounding S(πσ*) state which is stabilized by an increasing number of fluorine atoms.12 It can be concluded that the fluorine atoms contribute a significant share in form of σ* ← π character to the transition.

For 1,3,2-DCFB the quantum chemical calculations gave excitation energies (EE) of 5.06 eV (B3LYP/TZVPP), 4.71 eV (BP86/TZVPP) and 4.44 eV (CC2/cc-pVTZ). The CC2 value reproduces the experimental value of 4.52 eV with a deviation of 0.08 eV best.

The experimental value for the ionization energy (IE) of 9.32 eV is underestimated by both DFT methods with 9.00 eV (B3LYP/TZVPP) and 8.97 eV (BP86/TZVPP). (The coupled cluster method CC2 is not suited for the ionic species!). In this case the frequency analysis performed by the CC2 method showed to be most appropriate to reproduce the experimental frequencies.

For 1,3,2-DFCB the quantum chemical calculations gave excitation energies (EE) of 5.23 eV (B3LYP/TZVPP), 4.90 eV (BP86/TZVPP) and 3.88 eV (CC2/cc-pVTZ). The TDDFT value reproduces the experimental value of 4.64 eV best, whereas BP86 underestimates and B3LYP overestimates the experimental value. Both DFT methods with 9.03 eV (B3LYP/TZVPP) and 9.02 eV (BP86/TZVPP) underestimate the experimental value for the ionization energy (IE) of 9.38 eV. Contrary to the 1,3,2 DCFB the frequency analysis performed by the DFT methods results in a better reproduction of the experimental frequencies for the 1,3,2 DFCB.

Furthermore the S1 ← S0 electronic excitation energies (EE) and D0 ← S0 adiabatic ionization energies (IE) could be determined very exactly:

EE(1,3,2-DFCB) = 36[thin space (1/6-em)]460 ± 2 cm−1, IE(1,3,2-DFCB) = 75.242 ± 6 cm−1;

EE(1,3,2-DCFB) = 37[thin space (1/6-em)]449 ± 2 cm−1, IE(1,3,2-DCFB) = 75.627 ± 6 cm−1.

References

  1. F. Witte, R. Mikko, F. Gunzer and J. Grotemeyer, Mass analyzed threshold ionization (mati) spectroscopy of trichlorobenzenes via different intermediate vibrational states in the s1 state, Int. J. Mass Spectrom., 2011, 306, 129–137,  DOI:10.1016/j.ijms.2010.10.002.
  2. L. Yuan, C. Li, J. Lin, S. Yang and W. Tzeng, Mass analyzed threshold ionization spectroscopy of o-fluorophenol and o-methoxyphenol cations and influence of the nature and relative location of substituents, Chem. Phys., 2006, 323, 429–438,  DOI:10.1016/j.chemphys.2005.10.004.
  3. K. Müller-Dethlefs, M. Sander and E. W. Schlag, Two-colour photoionization resonance spectroscopy of NO: complete separation of rotational levels of NO+ at the ionization threshold, Chem. Phys. Lett., 1984, 112, 291–294,  DOI:10.1016/0009-2614(84)85743-7.
  4. L. Zhu and P. Johnson, Mass analyzed threshold ionization spectroscopy, J. Chem. Phys., 1991, 94(8), 5769–5771,  DOI:10.1063/1.460460.
  5. K. Walter, U. Boesl and E. W. Schlag, Molecular ion spectroscopy: resonance-enhanced multiphoton dissociation spectra of the fluorobenzene cation, Chem. Phys. Lett., 1989, 162(4, 5), 261–268,  DOI:10.1016/0009-2614(89)87041-1.
  6. G. Reiser, D. Rieger, T. G. Wright, K. Müller-Dethlefs and E. W. Schlag, Zero-kinetic-energy (ZEKE) photoelectron spectroscopy of p-difluorobenzene via different intermediate vibrational levels in the s1 state, J. Phys. Chem., 1993, 97, 4335–4343,  DOI:10.1021/j100119a015.
  7. A. Gaber, M. Riese and J. Grotemeyer, Mass analyzed threshold ionization spectroscopy of o-, m-, and p-dichlorobenzenes. Influence of the chlorine position on vibrational spectra and ionization energy, J. Phys. Chem. A, 2008, 112, 425–434,  DOI:10.1021/jp074802t.
  8. A. Gaber, M. Riese and J. Grotemeyer, Detailed analysis of the cation ground state of three dichlorobenzenes by mass analyzed threshold ionization spectroscopy, Phys. Chem. Chem. Phys., 2008, 10, 1168–1176,  10.1039/B715496H.
  9. C. Weickhardt, R. Zimmermann, K. W. Schramm, U. Boesl and E. W. Schlag, Laser mass spectrometry of the di-, tri- and tetrachlorobenzenes: isomer-selective ionization and detection, Rapid Commun. Mass Spectrom., 1994, 8, 381–384,  DOI:10.1002/rcm.1290080508.
  10. C. H. Kwon and M. S. Kim, One-photon mass-analyzed threshold ionization spectroscopy of 1,3,5-trifluorobenzene: the Jahn–Teller effect and vibrational analysis for the molecular cation in the ground electronic state, J. Chem. Phys., 2004, 121(6), 2622–2629,  DOI:10.1063/1.1765655.
  11. S. Sato, K. Ikeda and K. Kimura, Zeke photo electron spectroscopy and ab initio force-field calculation of 1,2,4,5-tetrafluorobenzene, J. Electron. Spectrosc. Relat. Phenom., 1998, 88–91, 137–142,  DOI:10.1016/S0368-2048(97)00256-9.
  12. M. Z. Zgierski, T. Fujiwara and E. C. Lim, Photophysics of aromatic molecules with low-lying pi sigma * states: fluorinated benzenes, J. Chem. Phys., 2005, 122(14), 144312,  DOI:10.1063/1.1873752.
  13. C. H. Kwon and M. S. Kim, Vacuum ultraviolet mass-analyzed threshold ionization spectroscopy of hexafluorobenzene: the Jahn–Teller effect and vibrational analysis, J. Chem. Phys., 2004, 120(24), 11578,  DOI:10.1063/1.1753258.
  14. Y. Tsuchiya, K. Takazawa, M. Fujii and M. Ito, Electronic spectra of o-, m-, p-difluorobenzene cations: striking similarity in vibronic coupling between the neutral molecule and its cation, J. Phys. Chem., 1992, 96(1), 99–104,  DOI:10.1021/j100180a022.
  15. F. Gunzer and J. Grotemeyer, New features in the mass analysed threshold ionization (MATI) spectra of alkyl benzenes, Phys. Chem. Chem. Phys., 2002, 24, 5966–5972,  10.1039/b208283g.
  16. F. Gunzer and J. Grotemeyer, Enhancing of the signal-to-noise ratio in MATI spectra, Int. J. Mass Spectrom., 2003, 228, 921–931,  DOI:10.1016/S1387-3806(03)00195-7.
  17. R. Ahlrichs, M. Bär, M. Häser, H. Horn and C. Kölmel, Electronic structure calculations on workstation computers: the program system turbomole, Chem. Phys. Lett., 1989, 162, 165–169,  DOI:10.1016/0009-2614(89)85118-8.
  18. M. v. Arnim and R. Ahlrichs, Geometry optimization in generalized natural internal coordinates, J. Chem. Phys., 1999, 111, 9183–9190,  DOI:10.1063/1.479510.
  19. M. Häser and R. Ahlrichs, Improvements on the direct SCF method, J. Comput. Chem., 1989, 10, 104–111,  DOI:10.1002/jcc.540100111.
  20. O. Treutler and R. Ahlrichs, Efficient molecular numerical integration schemes, J. Chem. Phys., 1995, 102, 346–354,  DOI:10.1063/1.469408.
  21. (a) A. D. Becke, Density functional calculations of molecular bond energies, J. Chem. Phys., 1986, 84, 4524–4529,  DOI:10.1063/1.450025; (b) A. Becke, Density-functional exchange-energy approximation with correct asymptotic behavior, Phys. Rev. A: At., Mol., Opt. Phys., 1988, 38, 3098,  DOI:10.1103/PhysRevA.38.3098; (c) F. Weigend, M. Häuser, H. Patzelt and R. Ahlrichs, RI-MP2: optimized auxiliary basis sets and demonstration of efficiency, Chem. Phys. Lett., 1998, 294, 143–152,  DOI:10.1016/S0009-2614(98)00862-8; (d) F. Weigend, A. Köhn and C. Hättig, Efficient use of the correlation consistent basis sets in resolution of the identity MP2 calculations, J. Chem. Phys., 2002, 116, 3175–3183,  DOI:10.1063/1.1445115.
  22. G. Varsanyi and S. Szoke, Vibrational Spectra of Benzene Derivatives, Academic Press, New York, London, 1969 Search PubMed.
  23. E. B. Wilson, The normal modes and frequencies of vibration of the regular plane hexagon model of the benzene molecule, J. Phys. Rev., 1934, 41, 706–714,  DOI:10.1103/PhysRev.45.706.
  24. J. Green, D. Harrison and W. Kynaston, Vibrational spectra of benzene derivatives-XI 1,3,5- and 1,2,3-trisubstituted compounds, Spectrochim. Acta, Part A, 1971, 27(6), 793–806,  DOI:10.1016/0584-8539(71)80158-7.
  25. M. Mohraz, J. Maier and E. Heilbronner, He(Iα) and He(IIα) photoelectron spectra of fluorinated chloro- and bromo-benzenes, J. Electron Spectrosc. Relat. Phenom., 1980, 19, 429–446,  DOI:10.1016/0368-2048(80)80063-6.

This journal is © The Royal Society of Chemistry 2015