<<

Pure Appl. Chem., Vol. 85, No. 10, pp. 1991–2004, 2013. http://dx.doi.org/10.1351/PAC-CON-12-12-01 © 2013 IUPAC, Publication date (Web): 13 August 2013 A new reaction mechanism of Claisen rearrangement induced by few-optical-cycle pulses: Demonstration of nonthermal chemistry by femtosecond vibrational spectroscopy*

Izumi Iwakura1,2,3,‡, Atsushi Yabushita4, Jun Liu2, Kotaro Okamura2, Satoko Kezuka5, and Takayoshi Kobayashi2

1Innovative Use of Light and Materials/Life, PRESTO, JST, 4-1-8 Honcho, Kawaguchi, Saitama, 332-0012, Japan; 2University of Electro-Communications, 1-5-1 Chofugaoka, Chofu, Tokyo 182-8585, Japan; 3Faculty of Engineering, Kanagawa University, 3-27-1 Rokkakubashi, Yokohama 221-8686, Japan; 4Department of Electrophysics, National Chiao-Tung University, Hsinchu 300, Taiwan; 5Department of Applied Chemistry, School of Engineering, Tokai University, 1117 Kitakaname, Hiratsuka-shi, Kanagawa 259-1292, Japan

Abstract: Time-resolved vibration spectroscopy is the only known way to directly observe reaction processes. In this work, we measure time-resolved vibration spectra of the Claisen rearrangement triggered and observed by few-optical-cycle pulses. Changes in molecular structure during the reaction, including its transition states (TSs), are elucidated by observ- ing the transient changes of molecular vibration wavenumbers. We pump samples with visi- ble ultrashort pulses of shorter duration than the molecular vibration period, and with photon energies much lower than the minimum excitation energy of the sample. The results indicate that the “nonthermal Claisen rearrangement” can be triggered by visible few-optical-cycle pulses exciting molecular vibrations in the electronic ground state of the sample, which replaces the typical thermal Claisen rearrangement.

Keywords: laser spectroscopy; reaction mechanisms; rearrangements.

INTRODUCTION The Claisen rearrangement is one of the most popular sigmatropic rearrangements in organic chemistry. Along with the Cope rearrangement, it is known for its high stereoselectivity and is extremely useful in organic synthesis. The Claisen rearrangement was first reported by Claisen [1] for allyl aryl (or vinyl) , spurring the development of various other reactions [2–4]. In 1938, it was found that allyl vinyl generates allyl acetaldehyde upon heating at 255 °C [5], which was later found to be generated via [3,3]-sigmatropic rearrangement (Fig. 1). Following the first kinetic study of the Claisen rearrange- ment in 1950 [6], various investigations of its reaction mechanism have been reported. The Claisen rearrangement is thought to proceed through a six-membered transition state (TS) by a supra-supra facial reaction following Woodward–Hoffmann rules [7] and frontier orbital theory [8]. Experimental

*Pure Appl. Chem. 85, 1919–2004 (2013). A collection of invited papers based on presentations at the 21st International Conference on Physical Organic Chemistry (ICPOC-21), Durham, UK, 9–13 September 2012. ‡Corresponding author

1991 1992 I. IWAKURA et al.

Fig. 1 Claisen rearrangement of allyl vinyl ether and three proposed reaction mechanisms. stereochemical outcomes [9–12] and theoretical calculations [13–15] implicated the six-membered chair-form TS. However, understanding of the mechanism of the Claisen rearrangement in greater detail has remained elusive (Fig. 1) [4,14–25]. In one proposed mechanism, the reaction progresses along a synchronous concerted pathway via an aromatic-like TS. Another possible mechanism involves the reaction progressing by an asynchronous concerted pathway in which either C1–C6 bond formation or C4–O bond cleavage occurs in advance of the other. An asynchronous reaction in which the C1–C6 bond forms first may proceed through a 1,4-diyl-like TS, whereas that in which the C4–O bond breaks first may proceed through a bis-allyl-like TS. As the simplest example of the Claisen rearrangement, allyl vinyl ether was studied by quantum chemical calculations and the kinetic isotope effect. In this work, the Claisen rearrangement was studied directly by observing the reaction process through time-resolved vibration spectroscopy using a pulsed laser with few optical cycles [26,27]. Direct observation of the Claisen rearrangement has not been performed by any other methods to date. Here, we observed molecular structural changes during the “nonthermal Claisen rearrangement” process in the electronic ground state, including the TS, as changes in the instantaneous wavenumber of molecular vibrations. The reaction mechanism of the nonthermal Claisen rearrangement in the elec- tronic ground state is elucidated. The observed dynamics of the reaction process will provide important information to clarify the reaction mechanism of the thermal Claisen rearrangement process, even though the reaction processes of the two kinds of Claisen rearrangement may differ.

EXPERIMENTAL Visible few-optical-cycle pulses To generate a broadband intense laser pulse for a visible few-optical-cycle broadband pulse, we have used a non-collinear optical parametric amplifier (NOPA) described elsewhere [26]. In short, the pump source of the NOPA is a Ti:sapphire regenerative amplifier (Spectra-Physics, Spitfire; 150 μJ, 100 fs, 5 kHz at 805 nm). Visible white light generated by self-phase modulation in a sapphire plate was ampli- fied in the double-pass NOPA having a broadband spectrum extending from 525 to 725 nm. Its time duration was compressed with a main compressor, resulting in the pulse duration of 5 fs, which is nearly Fourier-transform limited. The polarizations of the pump pulse and the probe pulse were parallel to each other. The focal areas of the pump pulse and the probe pulse were 100 and 75 μm2, respectively.

Ultraviolet few-optical-cycle pulses Second harmonic generation (SHG) of a Ti:sapphire regenerative amplifier (Coherent, Legend Elite- USP; 2.5 mJ, 35 fs, 1 kHz at 800 nm) was focused into a hollow-core fiber. Argon gas filled in the hol- low-core fiber has broadened spectral width to generate a few-optical-cycle broadband ultraviolet (UV) pulse [27]. The broadened spectrum coming out from the hollow-core fiber was extending from 360 to 440 nm. The main compressor compressed the pulse duration as 8 fs. The polarizations of the pump

© 2013, IUPAC Pure Appl. Chem., Vol. 85, No. 10, pp. 1991–2004, 2013 Directly observing reaction process including TS 1993 pulse and the probe pulse were parallel to each other. The focal areas of the pump pulse and the probe pulse were 100 and 75 μm2, respectively.

Sample cell Neat liquid of allyl vinyl ether and allyl phenyl ether were used as samples. The time-resolved signal was measured storing the sample in a liquid cell with a 1-mm optical path length. When the UV pulse irradiates at a fixed point, UV light absorption by the sample causes thermal accumulation and bub- bling. Therefore, we have translated the sample cell continuously in the plane perpendicular to the probe beam to avoid the thermal accumulation and bubbling. The cell was translated in a near-circular and octagon-like trajectory with the speed of ca. 5 mm/s, thus, the sample was displaced ca. 10 μm during the pulse period of 1 ms suppressing the sample photodamage to minimum. All of the measurements were performed at room temperature of 295 ± 1 K.

Pump–probe measurement A polychromator (300 grooves/mm, 500 nm blazed for visible pulses, and 300 nm blazed for UV pulses) has spectrally dispersed the probe spectrum coupling probe wavelength components to 128 ava- lanche photodiodes via a 128-channel bundled fiber. Lock-in amplifiers of the same number were uti- lized at all avalanche photodiodes to improve the signal-to-noise ratio.

Theoretical calculation The theoretical calculation was performed by the Gaussian 03 program [28] optimizing geometry by the B3LYP/6-311+G** method and basis set. Calculations were performed without assuming symme- try using 5d functions for the d orbital. For all of the obtained structures, frequencies were calculated at the same level. Calculation results have confirmed that all the frequencies were real for the ground states and one imaginary frequency existed for the TS. Vectors of the imaginary frequencies directed the reaction mode and intrinsic reaction coordinate (IRC) calculations have confirmed that the obtained TSs were on the saddle points of the energy surface between the reactant and the product.

RESULTS AND DISCUSSION Claisen rearrangement of allyl vinyl ether Allyl vinyl ether has an absorption band located at shorter than 220 nm (Fig. 2a) that cannot be excited by either one- or two-photon absorption of few-optical-cycle pulses in the visible range from 525 to 725 nm (Fig. 2b). Therefore, a visible few-optical-cycle pulse excites molecular vibrations in allyl vinyl ether only in the electronic ground state [29,30]. Figure 2c shows an averaged real-time trace of ΔA obtained by pump–probe measurement over 16 probe channels with a delay time range of –100 to 3000 fs. In the pump–probe measurement using visible few-optical-cycle pulses, the observed signal reflects the wave packet motion; i.e., fine oscilla- tion in the real-time trace of ΔA reflects the transformation in periodic molecular structure caused by molecular vibrations. Therefore, spectrogram analysis [31] of the real-time traces was performed using a Blackman window function with a full width at half-maximum (FWHM) of 400 fs. Spectrogram analysis shows time evolution of the vibration modes after the laser pulse irradiation. Figure 3b shows the calculated spectrogram, where x- and y-axes show reaction time and instan- taneous vibrational wavenumber, respectively, and pseudocolor reflects the power of fast Fourier trans- form (FFT). In the spectrogram, the molecular vibrations observed just after visible-pulse excitation were caused by the reactant allyl vinyl ether, as shown in the following assignment. The observed

© 2013, IUPAC Pure Appl. Chem., Vol. 85, No. 10, pp. 1991–2004, 2013 1994 I. IWAKURA et al.

Fig. 2 (a) Transmittance spectrum of allyl vinyl ether. (b) Spectrum of visible few-optical-cycle pulses. (c) Real- time traces of absorbance changes (ΔA) of allyl vinyl ether averaged over 16 probe channels in the delay time range between –100 and 3000 fs [30]. The probe wavelength regions of the eight traces are 700–725, 675–700, 650–675, 625–650, 600–625, 575–600, 550–575, and 525–550 nm, respectively, from top to bottom.

Fig. 3 (a) Raman spectrum of allyl vinyl ether. (b) Spectrogram of nonthermal Claisen rearrangement of allyl vinyl ether induced by visible few-optical-cycle pulses [30]. (c) Raman spectrum of allyl acetaldehyde.

ν wavenumbers can be assigned to C=C stretching vibrations ( C=C) of the allyl and vinyl groups –1 δ –1 (1650 cm ), C–H2 deformation vibration ( C–H ) of the methylene group (1500 cm ), C–H deforma- δ 2 –1 tion vibrations ( C–H) of the vinyl and allyl groups (1320 and 1290 cm , respectively), and C–O–C ν –1 symmetric stretching vibration ( s C–O–C) of the ether group (900 cm ). These observed wavenumbers agree well with the Raman spectrum of the electronic ground state of allyl vinyl ether (Fig. 3a), which confirms that the pump–probe observations closely reflect the molecular vibration dynamics of the elec- tronic ground state of allyl vinyl ether.

© 2013, IUPAC Pure Appl. Chem., Vol. 85, No. 10, pp. 1991–2004, 2013 Directly observing reaction process including TS 1995

Molecular structure deformation during the reaction process along the reaction coordinate changes the vibrational wavenumber of the wave packet vibrating along the coordinate perpendicular to the reaction one. There are 3N-7 coordinates perpendicular to the reaction coordinate in a molecule composed of N atoms with 3N-6 normal modes. Therefore, when the molecule stays between the two species of reactant, the intermediates, and the product, the vibrational wavenumbers in a spectrogram should change. If the molecule stays as the reactant, intermediate, or product, the vibrational wave - numbers should not shift during the lifetime of the relevant state. In the following, we discuss the real- time frequency shifts of the molecular vibration wavenumbers observed in the spectrogram of allyl vinyl ether. The disappearance of the two modes of the δ (1500 cm–1) and ν (900 cm–1) bands at CH2 s C–O–C a delay of around 800 fs shows that the C4–O bond is weakened or broken in the first stage of the reac- ν –1 tion. The signal from C=C of the vinyl and allyl groups, observed at 1650 cm just after visible pulse irradiation, was separated into a blue-shifted mode moving toward 1690 cm–1 and a red-shifted mode heading toward 1570 cm–1 in the probe-delay region from 500 to 800 fs. These wavenumber shifts also suggest that the C4–O bond is weakened, because electron transfer from the to the causes the electronic density in the vinyl and allyl groups to increase and decrease, respectively. After the C4–O bond weakens, electrons are transferred in the opposite direction from the vinyl group to the allyl group to form a weak C1–C6 bond. This transfer decreases the electronic density along 1 2 ν –1 the C =C bond in the vinyl group, resulting in a red shift of C=C from 1690 to 1580 cm . In the allyl 5= 6 ν group, the electronic density along the C C bond increased, inducing the observed blue shift of C=C from 1570 to 1580 cm–1, and the C4–C5 single bond changed into a C4=C5 double bond. Thus, the three C=C bonds in the vinyl and allyl groups (C1=C2, C4=C5, and C5=C6) become equivalent at a delay of about 1500 fs, all exhibiting a band at 1580 cm–1. The appearance of a band at around 1580 cm–1 indi- ν cates that aromatic-like C=C bonds were formed because the aromatic C=C in benzene appears at 1585 cm–1 [32]. This implies that the generated intermediate does not have the perfect C6 symmetry of a benzene ring, but instead has a six-membered structure with aromatic C=C bonds, which is also sup- δ ν –1 δ ported by the appearance of C–H and s C–C–C at 1190 and 1000 cm , respectively ( C–H of benzene is observed at 1180 cm–1 [32]). Finally, C4–O bond cleavage and C1–C6 bond formation are thought to proceed simultaneously (i.e., in a synchronous concerted process) to generate the product, allyl acetaldehyde. Therefore, new bands that appeared at a delay of 2000 fs reflect the formation of the product. These new bands at 1750, 1150, and 1030 cm–1 can be assigned to the C=O stretching vibra- ν ν tion ( C=O), C–C–C asymmetric stretching vibration ( as C–C–C), and C–C–C symmetric stretching ν vibration ( s C–C–C), respectively, of allyl acetaldehyde. The wavenumbers of these new modes corre- late well with the Raman spectrum of allyl acetaldehyde (Fig. 3c) that was synthesized by the oxidation of 4-penten-1-ol. We also performed a pump–probe experiment using the synthesized product (allyl acetaldehyde) to compare with the measurements obtained for the reactant. No wavenumber shift was observed in the pump–probe measurements of the product, indicating that no reaction occurred. NMR spectra of allyl vinyl ether before and after the pump–probe experiment further confirmed the formation of allyl acetaldehyde. We performed a pump–probe experiment with a small amount of allyl vinyl ether in a glass cell (10 mm3). The NMR spectrum of allyl vinyl ether after the measurement indicated the presence of allyl acetaldehyde with a fraction of 1 % w/w (Fig. 4). The quantum yield of the photoinduced process was estimated to be about 0.01. As described above, time-resolved observa- tion of the conformational changes during the [3,3]-sigmatropic rearrangement of allyl vinyl ether was performed by observing shifts in molecular vibrations. The [3,3]-sigmatropic rearrangement of allyl vinyl ether in the electronic ground state was triggered by irradiation with visible few-optical-cycle pulses, and here this reaction in the electronic ground state is called nonthermal Claisen rearrangement. In general, three possible mechanisms are suggested to explain the Claisen rearrangement. One of the proposed mechanisms is a synchronous concerted pathway reaction via an aromatic-like TS (yel- low line 1 in Fig. 5). Two other possible mechanisms are stepwise pathway reactions via a bis-allyl-like

© 2013, IUPAC Pure Appl. Chem., Vol. 85, No. 10, pp. 1991–2004, 2013 1996 I. IWAKURA et al.

Fig. 4 NMR spectrum of a sample after neat allyl vinyl ether was irradiated with visible few-optical-cycle pulses during time-resolved measurement [30].

Fig. 5 Map of alternative routes for the Claisen rearrangement [34]. 1: Synchronous concerted reaction pathway via an aromatic-like TS. 2: Stepwise reaction pathway via a bis-allyl-like TS. 3: A stepwise reaction pathway via a 1-4-diyl-like TS. 4: Mechanism proposed in this work [33].

TS in which C4–O bond cleavage takes place in the first stage of the reaction (green dotted line 2 in Fig. 5) or a 1-4-diyl-like TS in which C1–C6 bond formation takes place in the first stage of the reac- tion (purple dotted line 3 in Fig. 5). In this work, a new possible mechanism for the following three- stage pathway is proposed for the Claisen rearrangement using the data obtained by exciting allyl vinyl ether using visible few-optical-cycle pulses [30,33]. First, the C4–O bond is weakened to generate a bis- allyl-like intermediate. Formation of a weak C1–C6 bond then generates an aromatic-like intermediate. Finally, C4–O bond cleavage and C1–C6 bond formation occur simultaneously to generate allyl acetaldehyde (red line 4 in Fig. 5). We then compared the observed nonthermal Claisen rearrangement process in the electronic ground state with the standard Claisen rearrangement process. Geometric optimization of TSs and IRCs were calculated using B3LYP/6-311+G** (Fig. 6). In the reactant, the C1=C2 bond in the vinyl group 5 6 ν is longer than the C =C bond in the allyl group, so C=C of the vinyl group appears at lower wave - number than that of the allyl. In the first stage of the reaction, the molecular structure of allyl vinyl ether changes from its normal chain structure to a ring structure. Along with this structural change, the C4–O ν bond length increases (green line 1 in Fig. 6) causing a blue shift of s C–O–C followed by its disap- 1 2 ν pearance. Associated with this change, the C =C bond length decreases (blue line 3 in Fig. 6) and C=C ν of the vinyl group exhibits a blue shift. In contrast, C=C of the allyl group exhibits a red shift as the length of the C5=C6 bond increases (pink line 4 in Fig. 6). In the next stage of the reaction, along with formation of a weak C1–C6 bond, the C1=C2 bond ν 1 2 length increases, which induces a red shift of C=C of the vinyl group. The bond orders of C =C and C5=C6 bonds become similar because the wavenumbers of C1=C2 and C5=C6 stretching modes are almost equal. In addition, the C4–C5 bond of the allyl group (orange line 5 in Fig. 6) also becomes

© 2013, IUPAC Pure Appl. Chem., Vol. 85, No. 10, pp. 1991–2004, 2013 Directly observing reaction process including TS 1997

Fig. 6 IRC calculation at the B3LYP/6-311+G** level. The six curves show bond lengths of (1) C4–O, (2) C1–C6, (3) C1=C2 of the vinyl group, (4) C5=C6 of the allyl group, (5) C4–C5 of the allyl group, and (6) calculated IRC energy of the reaction pathway. In each molecular model, C, O, and H are dark, light, and small spheres, respectively. equivalent to the C1=C2 and C5=C6 bonds. The lengths of all three C=C bonds were 1.39 Å, showing that they can be assigned as aromatic C=C bonds. After this process, the C1=C2 and C5=C6 bonds of the vinyl and allyl groups, respectively, become longer and their signals disappear because they become Raman inactive. In contrast, the C4–C5 bond of the allyl group shortens and forms a new C=C bond. The change in bond length calculated by IRC is consistent with the results of pump–probe exper- iments using visible few-optical-cycle pulses. The pump–probe experiment suggests a three-stage path- way, whereas the IRC calculation suggests an asynchronous concerted process. It should be clarified in a future study if the difference originates from the different reaction trigger, reaction in solution vs. the gas phase, or for some other reason.

Claisen rearrangement of allyl phenyl ether The [3,3]-sigmatropic rearrangement of allyl vinyl ether is known to be a thermally allowed process. That is, it is a photochemically forbidden process following Woodward–Hoffmann rules. Even though it is photochemically forbidden, the rearrangement occurs using visible few-optical-cycle pulses. To confirm that a photochemically forbidden process can triggered using visible few-optical-cycle pulses, we tried to observe the Claisen rearrangement process of allyl phenyl ether. When allyl phenyl ether is heated, [3,3]-sigmatropic rearrangement (thermally allowed Claisen rearrangement) in the electronic ground state occurs to generate ortho-substituted following the mechanism shown in Fig. 7a. This rearrangement is thought to proceed through the six-membered TS by a supra-supra facial reac- tion to generate a keto-intermediate. Instability of the keto-intermediate leads to subsequent keto-enol tautomerization, generating ortho-substituted phenol. Therefore, it is expected that the carbonyl stretch- ing mode of a keto-intermediate is observed at the beginning of the reaction. As the keto-intermediate undergoes keto-enol tautomerization, the carbonyl stretching mode disappears. In contrast, Kharasch et al. [35] reported in 1952 that photoirradiation triggered the rearrange- ments of allyl phenyl and benzyl phenyl ethers in their electronic excited states. The mechanism of photo chemical Claisen rearrangement is shown in Fig. 7b [36]. Under photoirradiation, the electronic • • excited state of allyl phenyl ether forms a pair of radical intermediates (PhO and CH2CHCH2 ), and

© 2013, IUPAC Pure Appl. Chem., Vol. 85, No. 10, pp. 1991–2004, 2013 1998 I. IWAKURA et al.

Fig. 7 (a) Thermally allowed Claisen rearrangement of allyl phenyl ether. (b) Photochemical Claisen rearrangement of allyl phenyl ether. radical reactions generate the parent phenol as well as ortho- and para-substituted . Therefore, both phenoxy and allyl radicals can be observed. To compare both cases (photochemically forbidden and allowed), we tried to selectively induce the photochemical and thermally allowed Claisen rearrangement of allyl phenyl ether using UV and vis- ible few-optical-cycle pulses, respectively. Allyl phenyl ether has an absorption band at shorter than 400 nm (Fig. 8). Therefore, irradiation of allyl phenyl ether with UV pulses of 360 to 440 nm triggers electronic excitation, which is followed by photochemical Claisen rearrangement in the electronic excited state. Meanwhile, irradiation of allyl phenyl ether with visible pulses of 525 to 725 nm does not trigger electronic excitation. Therefore, the visible pulse triggers coherent molecular vibrations in the electronic ground state that should be followed by Claisen rearrangement in the electronic ground state.

Fig. 8 Absorption spectrum of APE (1) and spectrum of visible few-optical-cycle pulses (2) and spectrum of UV few-optical-cycle pulses (3) [37].

First, pump–probe measurement of allyl phenyl ether using UV pulses was performed to observe the photochemical Claisen rearrangement. Time-resolved vibrational spectra were obtained by sliding- window Fourier transformation with a Blackman window function with a FWHM of 400 fs (Fig. 9). Following irradiation with UV pulses, we observed only the molecular vibrations of the reactant allyl ν phenyl ether. Signals were assigned as the C=C stretching vibration ( C=C) of the allyl group (1670 cm–1), C–H deformation vibration (δ ) of the methylene group (1482 cm–1), C–H deforma- 2 C–H2

© 2013, IUPAC Pure Appl. Chem., Vol. 85, No. 10, pp. 1991–2004, 2013 Directly observing reaction process including TS 1999

Fig. 9 Spectrogram of photochemical Claisen rearrangement of allyl phenyl ether induced by UV few-optical cycle pulses [37].

δ –1 ν tion vibration ( C–H) of the allyl group (1320 cm ), C–O–C symmetric stretching vibration ( s C–O–C) –1 ν of the ether group (1080 cm ), and C6 symmetric stretching vibration ( s Ph) of the phenyl group –1 –1 ν (1007 cm ). The highest band at 1000 cm ( s Ph) was separated into two modes. One exhibited a red –1 ν shift toward 800 cm , and the other did not move. The C=C of the allyl group disappeared at a delay of about 1700 fs. At the same time, the allyl radical was generated, and a new band consistent with the ν –1 ν C–C stretching vibration ( C–C) of an allyl radical appeared at 1510 cm . Moreover, s C–O–C observed at 1080 cm–1 just after UV-pulse irradiation, exhibited a gradual blue shift as the C–O group formed, ν –1 and a new band ( C–O) at 1480 cm appeared at a delay of about 2000 fs. These results also indicate that the active radical species were generated after about 2000 fs. Thus, irradiation with few-optical- cycle pulses induces photoreaction by excitation of the electronic state if the laser spectrum overlaps with the absorption band of the sample [37]. Pump–probe measurement of allyl phenyl ether using visible pulses was then performed to observe thermally allowed Claisen rearrangement. The transient wavenumber changes of molecular vibration modes were analyzed by calculating spectrogram (see Fig. 10). Molecular vibrational modes just after visible pulse irradiation are consistent with the reactant allyl phenyl ether. The signals δ CH2 (1420 cm–1), δ (1230 cm–1), ν (1030 cm–1), and ν (1000 cm–1) disappeared after a delay CH2 s C–O–C s Ph of about 700 fs. This implies that the C4–O bond is either weakened or broken in the first stage of the ν –1 reaction. Next, C=C of the phenyl group, observed at 1595 cm just after visible-pulse irradiation, exhibited a red shift toward 1500 cm–1 in the probe-delay region from 500 to 750 fs. This indicates that a six-membered structure with aromatic C=C bonds formed in the second stage of the reaction. The ν –1 C=C of the allyl group at 1650 cm just after visible-pulse irradiation exhibited a red shift to 1580 cm–1, which also supports the formation of a six-membered structure with aromatic C=C bonds. –1 ν A new band that appeared at about 1750 cm after 1 ps delay can be assigned to C=O, which verifies that a keto-intermediate was generated in the third stage of the reaction. Instability of this keto-inter- mediate leads to keto-enol tautomerization as the final stage of the reaction. Therefore, after 2000 fs ν delay, C=O disappeared and the phenol product was formed. The observed changes in wavenumber indicate that the reaction pathway of allyl phenyl ether [38] after visible-pulse irradiation is equivalent to that of allyl vinyl ether. Comparison of Figs. 9 and 10 reveals the following differences. When allyl phenyl ether was irra- ν ν ν diated with visible pulses, s Ph disappeared and C=O appeared. In contrast, the s Ph did not disappear

© 2013, IUPAC Pure Appl. Chem., Vol. 85, No. 10, pp. 1991–2004, 2013 2000 I. IWAKURA et al.

Fig. 10 Spectrogram of nonthermal Claisen rearrangement in the electronic ground state of allyl phenyl ether induced by visible few-optical-cycle pulses [38].

ν under UV-pulse irradiation, and C=O did not appear. These differences imply that [3,3]-sigmatropic rearrangement of allyl phenyl ether occurs under visible few-optical-cycle pulses. In addition, photo- allowed Claisen rearrangement occurs using UV few-optical-cycle pulses. These results certainly show that visible few-optical-cycle pulses induced nonthermal Claisen rearrangement in the electronic ground state of allyl phenyl ether. Nonthermal Claisen rearrangement in the electronic ground state without electronic excitation has the possibility of being either a thermal reaction or a third type of reac- tion triggered by a novel scheme that is different from both photo- or thermal reactions.

“Nonphoto nonthermal Claisen rearrangement” and thermal Claisen rearrangement Nonthermal Claisen rearrangement is different from thermal Claisen rearrangement in the following three ways. (i) Thermal Claisen rearrangement is reported to proceed with an activation energy of 20–40 kcal/mol [14,16,39–42], whereas nonthermal Claisen rearrangement proceeded in the electronic ground state under irradiation with a visible pulse (525–725 nm) with a bandwidth of just 5200 cm–1 = 15 kcal/mol. (ii) Cleavage of the C–O bond took 700–800 fs in the spectrograms of both allyl vinyl ether (Fig. 3) and allyl phenyl ether (Fig. 10), which is much longer than the time suggested in the trajectory calculation. (iii) The keto-enol tautomerization observed in Fig. 10 occurs much faster than that in typ- ical chemical reactions. To elucidate why (i) the reaction proceeds with such a low activation energy under visible ultra- short pulse irradiation, and (ii) cleavage of the C–O bond takes longer than expected, we studied the pump power dependence of the real-time trace (Fig. 11a) and Fourier power spectrum (Fig. 11c) of allyl vinyl ether in the reaction triggered by visible few-optical-cycle pulse irradiation. The traces of real-time vibrational amplitude show oscillations at a delay of 50–80 fs (see Fig. 11a), which has maximum amplitude with a pump pulse of 150 nJ (see Fig. 11b). If the excitation is caused by a three-photon absorption process, the Fourier power spectrum should be proportional to the cubic power of the pump power. However, the Fourier power spectrum exhibited maximum inten- sity when the sample was pumped with a pulse of 150 nJ (see Fig. 11c). The observed pump power

© 2013, IUPAC Pure Appl. Chem., Vol. 85, No. 10, pp. 1991–2004, 2013 Directly observing reaction process including TS 2001

Fig. 11 Pump power dependence of (a) the real-time trace of allyl vinyl ether in the reaction triggered by visible few-cycle-pulse irradiation and (b) ΔA averaged in the delay region between –20 and 20 fs. (c) Fourier power spectrum of allyl vinyl ether in the reaction triggered by visible few-cycle-pulse irradiation. dependency can be explained by assuming that when the pump power is higher than 150 nJ, allyl vinyl ether starts to be ionized by a nonlinear process, which suppresses the nonthermal reaction. The reason why the ionization is thought to proceed is as follows. Ionization generally occurs when a molecule is irradiated by light with an intensity of ~1014 W/cm2 [43,44]. When the average laser power is 150 μW, its pulse energy is 150 nJ (=150 μW/kHz). Considering the pulse energy, a pulse duration of 5 fs, and focus area of 100 μm2, the power density of the laser pulse can be calculated as 0.3 × 1014 W/cm2 (=150 nJ/100 μm2/5 fs). Therefore, it is thought that ionization starts to occur when the sample is pumped by a pulse of 150 nJ. We propose the following hypothesis to explain why nonthermal Claisen rearrangement and thermally allowed Claisen rearrangement have the three differences mentioned at the beginning of this section. (i) The reason the nonthermal Claisen rearrangement proceeds with much lower activation energy than that for thermal Claisen rearrangement can be explained by the two possible mechanisms explained below. The first possible mechanism is as follows. The impulsive excitation excites only Raman active modes in the sample molecules. Because of low cross-section of the Raman process, a small portion of the molecules are excited by the few-optical-cycle pulse. Raman active modes excited by pulsed excitation cause nonlinear interaction by dynamic mode coupling [45], result- ing in the excitation of C–O stretching. In the case of incoherent thermal excitation, even though thermal energy is widely distributed over all of the molecules, the excited molecules have differ- ent small vibrational amplitudes with random phases. Under thermal excitation, high wave - number vibration modes can be pumped by increasing system temperature. Such excitation gen- erates a small population of high wavenumber modes together with a high population of low wavenumber modes following the Boltzmann distribution. Meanwhile, under coherent pulsed excitation, the Raman excitation populates states equally in the broadband laser bandwidth of 0.7 eV (=5200 cm–1). Therefore, even though the low cross-section of the Raman process gener- ates less total population than the case of thermal excitation, the pulsed excitation can efficiently excite high wavenumber vibration modes. The dynamic mode coupling in the excited molecule results in excitation of other vibration modes. The selection of vibration mode is controlled by the quantum probability in each excited molecule; i.e., the C–O stretching mode will be excited in some molecules, which triggers nonthermal Claisen rearrangement. Furthermore, this nonthermal distribution is formed in a single molecule within a few femtoseconds nonresonantly when a non- linear process takes place by the combination of the two spectral components with a frequency difference corresponding to the vibration wavenumber. Therefore, by exciting the C–O stretching mode accidentally, C–O bond dissociation can take place, allowing the reaction to proceed with an activation energy much lower than that for thermal Claisen rearrangement.

© 2013, IUPAC Pure Appl. Chem., Vol. 85, No. 10, pp. 1991–2004, 2013 2002 I. IWAKURA et al.

The second possible mechanism is as follows. The amplitude of the oscillation at 50–80 fs delay shown in Fig. 11a is proportional to the pump intensity when the ionization of allyl vinyl ether does not proceed (when the pump power is in the range of 90–100 nJ), which indicates that electrons are shaken by the pump pulse. It is thought that electrons of allyl vinyl ether are shaken in the direction of their dipole moments by the electronic field of the visible pulse when allyl vinyl ether is irradiated with a high-intensity visible pulse with an ultrashort pulse width that is much narrower than the period of molecular vibration. Thus, the electrons attached to C and O are driven to oscillate with the nuclei along the direction of the C–O bond axis to break this bond, which causes nonthermal Claisen rearrangement to proceed with a much lower activation energy than that of thermally allowed Claisen rearrangement. (ii) The reason why cleavage of the C–O bond does not occur instantaneously within a few vibra- tional periods of C–O stretching but at longer delay time of about 700–800 fs is considered to be as follows. Shaking of electrons together with C and O nuclei occurs by dynamic mode coupling between Raman active modes and the C–O stretching mode. Activated Raman vibrations take about several hundred femtoseconds to relax. It is not known how shaking of electrons by an elec- tric field of the ultrashort visible pulse changes the reaction pathway. Thus, the mechanism of nonthermal Claisen rearrangement induced by the irradiation of coherent ultrashort visible pulse may differ from that of the thermal Claisen rearrangement under incoherent thermal excitation, resulting in delayed cleavage of the C–O bond. (iii) The reason why the keto-enol tautomerization observed in Fig. 10 occurs much faster than in common chemical reactions can be explained as follows. In the spectrogram of allyl phenyl ether (Fig. 10), almost no vibration modes appear at longer delay than 2 ps, showing that most vibra- tional coherence is lost. In the spectrogram shown in Fig. 10, keto-enol tautomerization proceeds 2.5 ps after excitation, which indicates that the small fraction of molecules in the sample that retain vibrational coherence may have isomerized even though the remaining major fraction of the molecules maintain their original keto structure.

CONCLUSION In this work, we demonstrated that nonthermal Claisen rearrangement replaces thermally allowed Claisen rearrangement in the electronic ground state. Even though it is not clear how similar these Claisen rearrangements are, this work confirms that Claisen rearrangement proceeds in the electronic ground state of a sample when molecular vibrations are excited by pumping with an ultrashort pulse that is much shorter than the molecular vibration period, with photon energy that is much lower than the minimum excitation energy.

ACKNOWLEDGMENTS The authors are grateful to the Information Technology Center of the University of Electro- Communications for their support of the DFT calculations. I. Iwakura acknowledges JSPS for Grant- in-Aid for Young Scientists (A) KAKENHI Grant Number 24685005.

REFERENCES 1. L. Claisen. Chem. Ber. 45, 3157 (1912). 2. F. E. Ziegler. Chem. Rev. 88, 1423 (1988). 3. U. Nubbemeyer. Synthesis 961 (2003). 4. A. M. Martín Castro. Chem. Rev. 104, 2939 (2004). 5. C. D. Hurd, M. A. Pollack. J. Am. Chem. Soc. 60, 1905 (1938). 6. F. W. Schuler, G. W. Murphy. J. Am. Chem. Soc. 72, 3155 (1950).

© 2013, IUPAC Pure Appl. Chem., Vol. 85, No. 10, pp. 1991–2004, 2013 Directly observing reaction process including TS 2003

7. R. Hoffmann, R. B. Woodward. Acc. Chem. Res. 1, 17 (1968). 8. K. Fukui. Acc. Chem. Res. 4, 57 (1971). 9. W. v. E. Doering, W. R. Roth. Tetrahedron 18, 67 (1962). 10. P. Vittorelli, T. Winkler, H.-J. Hansen, H. Schmid. Helv. Chim. Acta 51, 1457 (1968). 11. H.-J. Hansen, H. Schmid. Tetrahedron 30, 1959 (1974). 12. J. J. Gajewski, J. Jurayj, D. R. Kimbrough, M. E. Gande, B. Ganem, B. K. Carpenter. J. Am. Chem. Soc. 109, 1170 (1987). 13. R. L. Vance, N. G. Rondan, K. N. Houk, H. F. Jensen, W. T. Borden, A. Komornicki, E. Winner. J. Am. Chem. Soc. 110, 2314 (1988). 14. O. Wiest, K. A. Black, K. N. Houk. J. Am. Chem. Soc. 116, 10336 (1994). 15. H. Hu, M. N. Kobrak, C. Xu, S. Hammes-Schiffer. J. Phys. Chem. A 104, 8058 (2000). 16. J. G. Hill, P. B. Karadakov, D. L. Cooper. Theor. Chem. Acc. 115, 212 (2006). 17. M. J. S. Dewar, E. F. Healy. J. Am. Chem. Soc. 106, 7127 (1984). 18. M. J. S. Dewar, C. Jie. J. Am. Chem. Soc. 111, 511 (1989). 19. J. J. Gajewski, N. D. Conrad. J. Am. Chem. Soc. 101, 2747 (1979). 20. L. Kupczyk-Subotkowska Jr., W. H. Saunders, H. J. Shine, W. Subotkowski. J. Am. Chem. Soc. 116, 7088 (1994). 21. M. M. Davidson, I. H. Hillier. J. Phys. Chem. 99, 6748 (1995). 22. J. J. Gajewski. Acc. Chem. Res. 30, 219 (1997). 23. V. Aviyente, H. Y. Yoo, K. N. Houk. J. Org. Chem. 62, 6121 (1997). 24. H. Y. Yoo, K. N. Houk. J. Am. Chem. Soc. 119, 2877 (1997). 25. M. P. Meyer, A. J. DelMonte, D. A. Singleton. J. Am. Chem. Soc. 121, 10865 (1999). 26. A. Baltuska, T. Fuji, T. Kobayashi. Opt. Lett. 27, 306 (2002). 27. J. Liu, Y. Kida, T. Teramoto, T. Kobayashi. Opt. Expr. 18, 4672 (2010). 28. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, J. A. Montgomery Jr., T. Vreven, K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, 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, P. Y. Ayala, K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C. Gonzalez, J. A. Pople. Gaussian 03, (Revision D.02), Gaussian, Inc., Wallingford CT. (2004). 29. I. Iwakura, A. Yabushita, T. Kobayashi. J. Am. Chem. Soc. 131, 688 (2009). 30. I. Iwakura, A. Yabushita, T. Kobayashi. Chem. Lett. 39, 374 (2010). 31. M. J. J. Vrakking, D. M. Villeneuve, A. Stolow. Phys. Rev. A 54, R37 (1996). 32. G. Herzberg. Infrared and Raman Spectra, Van Nostrand Reinhold, New York (1945). 33. I. Iwakura. Phys. Chem. Chem. Phys. 13, 5546 (2011). 34. The scale in Fig. 5 cannot be discussed. Figure 5 was just shown as a schematic figure to emphasize that the spectrogram analysis has elucidated that the aromatic TS appears after impulsive excitation even though the bond order of C4–O3 starts to decrease earlier than the increase of the bond order of C1–C6. 35. M. S. Kharasch, G. Stampa, W. Nudenberg. Science 116, 309 (1952). 36. F. Galindo. J. Photochem. Photobiol. C: Photochem. Rev. 6, 123 (2005). 37. I. Iwakura, A. Yabushita, J. Liu, K. Okamura, T. Kobayashi. Phys. Chem. Chem. Phys. 14, 9696 (2012).

© 2013, IUPAC Pure Appl. Chem., Vol. 85, No. 10, pp. 1991–2004, 2013 2004 I. IWAKURA et al.

38. I. Iwakura, A. Yabushita, T. Kobayashi. Chem. Phys. Lett. 501, 567 (2011). 39. S. Yamabe, S. Okumoto, T. Hayashi. J. Org. Chem. 61, 6218 (1996). 40. B. Gómez, P. K. Chattaraj, E. Chamorro, R. Contreras, P. Fuentealba. J. Phys. Chem. A 106, 11227 (2002). 41. J. J. Gajewski, N. D. Conrad. J. Am. Chem. Soc. 101, 6693 (1979). 42. C. J. Burrows, B. K. Carpenter. J. Am. Chem. Soc. 103, 6983 (1981). 43. S. M. Hankin, D. M. Villeneuve, P. B. Corkum, D. M. Rayner. Phys. Rev. Lett. 84, 5082 (2000). 44. S. M. Hankin, D. M. Villeneuve, P. B. Corkum, D. M. Rayner. Phys. Rev. A 64, 013405 (2001). 45. T. Kobayashi, A. Shirakawa, H. Matsuzawa, H. Nakanishi. Chem. Phys. Lett. 321, 385 (2000).

© 2013, IUPAC Pure Appl. Chem., Vol. 85, No. 10, pp. 1991–2004, 2013