I.
INTRODUCTION
Photochemical reactions are the result of concerted electronic-nuclear motion induced by light-matter interaction. Gaining a fundamental understanding of these coupled dynamics by combining ultrafast spectroscopy with first principles calculations is the driving force behind a fast growing field of research.1,2 For example, the UV photochemistry of halogen-containing molecules involves ultrafast carbon-halogen bond cleavage, forming radical species that are short-lived and highly reactive. Here, direct detection of transient species during the photoinduced dissociation of bromoform using core-to-valence transitions is demonstrated, which provides a sensitive probe of the evolving molecular electronic structure and nuclear geometry, with ultrafast time resolution.
The photochemistry of halogen-containing species has significant practical implications.3 For instance, bromoform is recognized as a key contributor to polar ozone depletion4 and is the largest source of bromine in the atmosphere due to a combination of increased absorption at longer wavelengths and considerable natural abundance.5 Increasing bromination of alkanes leads to an increase in absolute absorption cross section, new absorption bands, and a strong red-shift of the first absorption band as illustrated in Fig. 1(a), which shows gas-phase absorption spectra of CH3Br, CH2Br2, CHBr3, and CBr4.6 Absorption at longer wavelengths extends the photochemistry to lower altitudes.
FIG. 1.
(a) Absorption spectra of bromomethanes: CH3Br (dashed), CH2Br2 (dotted), CHBr3 (solid), and CBr4 (dot-dash).6 (b) Absorption spectrum of CHBr3 (left, Gillotay et al.) and calculated oscillator strengths (right, Peterson and Francisco) for low-lying states of predominantly singlet and triplet character.29,30
Detailed knowledge of the UV photochemistry of haloalkanes, for example the CH3I photochemistry as summarized by Gardiner et al.,7 provides important building blocks for single-photon UV dissociation studies of gas-phase dihaloalkanes (CH2I2,8–11 CH2ICl,12–15 CH2BrCl,16–18 and CH2BrI19–22) and trihaloalkanes (CHBrCl223 and CHI2Cl24). Generally, the photochemistry of the low-lying electronically excited states of haloalkanes arises from the promotion of an electron from a nonbonding pair of electrons on the halogen atom (X) to a σ* antibonding orbital (
FIG. 2.
(a) Correlation diagram connecting the low-lying electronically excited states to photoproducts. Adiabatic (dashed) and diabatic (solid) correlations are shown. (b) Schematic adiabatic potential energy curves along the reaction coordinate.
Several product channels are energetically accessible following UV excitation31
(I)
(II)
(III)
(IV)
Broadly, two different viewpoints exist on the outcome of single-photon UV dissociation of bromoform within the first absorption band. Either only C–Br fission by channel (I) occurs, or both channel (I) and Br2 elimination by channel (III) are active. Br2 elimination was first suggested by Xu et al., based on the detection of CHBr2+, CHBr+, and CBr+ products using vacuum ultraviolet (VUV) ionization time-of-flight mass spectrometry and imaging.32 Bayes et al. measured the Br atom yield between 266 and 344 nm by monitoring atomic fluorescence. While a unity quantum yield was observed at λ > 300 nm, the yield decreases to 0.76 ± 0.03 at 266 nm, which was interpreted as an indication that additional dissociation channels may be active at shorter wavelengths.33 Huang et al. detected Br2 via B–X band absorption following photolysis of bromoform at 248 nm using cavity ringdown spectroscopy.34 Contrasting these observations, multiple compelling pieces of evidence that only C–Br fission occurs upon 248 nm single-photon excitation have been presented by Zou et al.35 Using frequency modulated transient absorption to directly detect CHBr near 10841 cm−1 (1.344 eV), an upper limit for the CHBr yield of ∼0.003 was estimated. Second, the power dependence of Br2 products in photofragment translational spectroscopy experiments was shown to be quadratic (1.9 ± 0.1) with the photolysis laser intensity, consistent with a multiphoton origin of the Br2 fragments. Molecular elimination in polyhaloalkanes following multiphoton excitation is well documented.36–38Upon photolysis, polyhaloalkanes can form stable isomers containing a carbon-halogen-halogen bond linkage, as was first demonstrated by Maier et al. employing cryogenic matrices.39,40 Formation of the isomer has been attributed to a recombination mechanism, where the environment confines the fragments and dissipates excess energy, eventually leading to the reassociation of the fragments into a local minimum isomer conformation.41 An alternative mechanism has been proposed, whereby ultrafast photoisomerization proceeds through a conical intersection that connects the lowest-lying singlet excited state of the parent molecule to the ground electronic state, forming the isomer en-route to dissociation.42 Subsequent ultraviolet-visible (UV-vis) transient absorption studies of gas-phase CHBr3 following 250 nm photolysis detected an absorption feature that forms and decays on timescales of 50 fs and 85 fs, respectively.43 It was hypothesized that the intermediate absorption corresponds to isomer formation.43 The proposed isomerization mechanism challenges the conventional understanding of haloalkane photochemistry, where a one-dimensional model of C–X bond extension is sufficient to capture the essential details of the dissociation rather than the full 3N-6 degrees of freedom that are, in principle, involved.7
Here, a UV pump–extreme UV (XUV) probe femtosecond transient absorption spectroscopy study of the 268 nm induced dissociation of CHBr3 is presented. Experimental spectra are complemented by high-level ab initio calculations of XUV spectral fingerprints of transient molecular species obtained from excited-state molecular dynamics (MD) simulations. The XUV spectra track the formation of singly occupied molecular orbitals (SOMO) in the photoexcitation step and the evolution of transient electronic configurations throughout the C–Br bond fission, eventually leading to the formation of radical Br and CHBr2 products. A global fit of the experimental XUV spectra indicates the formation and decay of an absorption feature on timescales of 40 ± 20 fs and 85 ± 10 fs, respectively. Comparison to first principles calculations reveals that the changes in the transient absorption spectra reflect the different characteristic timescales for motion of the light C and H atoms and the heavy Br atoms. Within the first 40 fs, distortion from
II.
EXPERIMENTAL AND THEORETICAL METHODS
A.
Femtosecond XUV transient absorption spectroscopy
Experiments are performed using a femtosecond XUV transient absorption setup that has been described in detail previously.44 High-harmonic generation (HHG) produces a quasicontinuum of XUV by focusing 4.0 mJ, 35 fs, 1 kHz, 804 nm pulses with a 500 mm focal length lens into a semi-infinite gas cell containing 300 Torr of neon. The gas load is removed from the HHG propagation path using a differential pumping stage between the exit of the HHG cell and a layer of adhesive aluminum tape mounted 20 mm further downstream. The fundamental near-infrared (NIR) beam drills submillimeter diameter holes into the exit of the HHG cell and the aluminum tape. The volume between the two pinhole apertures is evacuated by a dry scroll vacuum pump. The XUV and NIR beams then enter a turbomolecular-pumped chamber where the NIR is rejected by a 200 nm thick aluminum foil. The transmitted XUV light impinges on a toroidal mirror that focuses the beam into an ∼4 mm long sample cell mounted inside a turbomolecular-pumped chamber. The sample cell consists of a Polytetrafluoroethylene (PTFE) capillary with ∼400 μm apertures on opposing faces through which both the UV pump beam and the XUV probe beam enter and exit the cell.45 The sample cell is mounted on an XYZ translation stage and is resistively heated to ∼80 °C. Bromoform (Sigma Aldrich, 99%) is gently heated to 40 °C to increase the vapor pressure and delivered to the sample cell using stainless steel tubing. The vacuum chamber that houses the sample cell is isolated from the other vacuum chambers using 200 nm thick aluminum foils to prevent contamination of optics. The turbomolecular-pumped detector chamber houses an XUV spectrometer consisting of a variable line spaced grating (∼1200 lines/mm) and an X-ray CCD camera.
Static absorption spectra are recorded by measuring the XUV spectrum with the gas on/off. Transient absorption spectra are measured by introducing a pump pulse into the sample cell at an ∼2° angle relative to the XUV using a drilled-through mirror to transmit the XUV beam and reflect the pump beam. Downstream from the sample cell, a copper gasket with a small hole is used to transmit the XUV beam into the detector chamber and to dump the pump beam. In the pump arm, up to 4.0 mJ of NIR are used to generate UV pump pulses by frequency-tripling the fundamental. Second harmonic generation (SHG) is performed in a 0.1 mm thick, 29.2° β-barium borate (BBO) crystal to produce 402 nm pulses. The 804 nm and 402 nm beams are separated using a dichroic mirror. A half-wave plate rotates the NIR polarization to vertical for type-I mixing; a second dichroic mirror recombines the 804 nm and 402 nm beams collinearly in a 0.1 mm thick 44.3° BBO to produce up to 300 μJ pulses of UV light at 268 nm. The polarization of both the UV and the HHG beams is horizontal in the laboratory frame. The UV pulse energy is controlled using a combination of a half waveplate and a thin film polarizer to attenuate the fundamental NIR. No UV pulse compression is applied. Four high-reflective UV mirrors remove residual fundamental and 402 nm light prior to focusing the 268 nm beam into the sample cell with a 500 mm focal length lens. The UV beam is focused to a spot diameter of ∼100 μm (FWHM). An optical chopper that intercepts the pump beam with a 50% duty cycle acts as the master clock for the experiment. The rising and falling edges trigger CCD exposures at 8 Hz to alternately record pump-on and pump-off spectra. The time delay between each pump and probe pulse is varied between −500 fs and +10 000 fs using a computer-controlled delay stage, with positive Δt defined as the pump pulse preceding the probe pulse.
The UV-induced change in absorbance is calculated for each time delay
(1)
whereSpectral and temporal calibration of the setup is performed by filling the sample cell with xenon gas at pressures of ≪1 bar. The instrument response function (IRF) is determined by recording transient absorption spectra of xenon, whereby the intense UV pump pulse causes a ponderomotive shift of the absorption lines associated with np ← 4d inner-valence to Rydberg excitations. A Gaussian fit to the time-dependent
B.
Ab initio calculations
Excited-state MD simulations are performed by iteratively calculating the potential energy surface (PES) gradient using time-dependent density functional theory (TDDFT) and propagating the nuclei to follow the dissociation. XUV absorption spectra are calculated for a subset of nuclear configurations resulting from the MD simulations. The theoretical methodology follows the approach outlined for 1,3-cyclohexadiene.2 Note that in contrast to minimum energy pathways, the trajectories used here consider the momenta of the nuclei.
The initial conditions are obtained from a Boltzmann sampling of the Born-Oppenheimer molecular dynamics (BOMD) of the ground state molecule with a 300 K Nosé–Hoover thermostat as implemented in Q-Chem.46,47 The time step of the BOMD is 40 a.u. (∼0.96 fs) and the thermostat characteristic response time is 12 fs. Snapshots separated from one another by at least 48 fs are sampled as the starting coordinates and velocities to perform the excited-state MD with the fewest-switches surface-hopping (FSSH) algorithm.48 TDDFT surface-hopping dynamics are initiated on the lowest energy triplet (T1) excited-state PES. The Velocity-Verlet algorithm49 with a time step of 8 a.u. (∼0.19 fs) is used to evolve the nuclear coordinates, and the derivative-couplings between the lowest 8 triplet states are calculated to determine surface-hopping probabilities at each time step.
The trajectories are sampled every 40 a.u. (∼0.96 fs) to obtain the nuclear coordinates and the active PESs to calculate X-ray absorption spectra (XAS). Calculation of near-edge X-ray absorption fine structure (NEXAFS) spectra as implemented in NWChem50 involves a two-step approach. First, a reference valence excited state is modeled using the maximum overlap method (MOM);51,52 the resultant DFT eigenvectors and eigenvalues are then used in a restricted energy window (REW)53 linear-response TDDFT54 calculation to obtain the corresponding absorption spectrum. All calculations are carried out using the def2-tzvpd55 basis-set and the PBE056 hybrid-DFT functional.
III.
RESULTS
The transient absorption spectra shown in Fig. 3(a) are obtained by integrating the measured
FIG. 3.
(a) UV pump–XUV probe transient absorption spectra at the bromine
The time-dependent behavior of the XUV transient absorption spectra is explored in more detail in Fig. 3(b) by inspecting key spectral regions using the same colors as marked by color-coded arrows in Fig. 3(a). The circular data points in Fig. 3(b) have been derived by integrating the time-dependent signals across 0.1–0.5 eV wide spectral regions and their respective maxima are normalized to unity. Solid curves are the result of a fit procedure described in the following section. Depletion of the parent CHBr3 population, as indicated by the Br core
IV.
ANALYSIS
A.
Global fit procedure
In order to analyze the dynamic trends more quantitatively, the time-resolved spectra are decomposed into transient spectral contributions using a global fit procedure based on the standard least squares algorithm. The left panel in Fig. 4(a) shows a false-color representation of the measured
(2)
where the exponent T denotes the transpose of a matrix. This approach does not impose any constraints on the spectral shape related to a particular state and different kinetic models can be rapidly compared. A variety of different kinetic models have been evaluated to describe the data. A 1-step model clearly cannot accurately reproduce the observed dynamics. A 2-step model, however, is found to be sufficient to capture the dynamics (see the supplementary material for details). The heuristic approach to achieve the best description of the data with the least amount of free fit parameters leads to a time matrix composed of three states/components with time-dependent populations(3)
(4)
(5)
Here,FIG. 4.
Global fit analysis of the time-resolved XUV absorption spectrum of CHBr3. (a) Measured (left) and global fit (right) data as a function of XUV absorption energy and pump-probe time delay. (b) The time basis set is defined by a three-state model where each state is sequentially populated, while the rate constants connecting the three states are free fit parameters. (c) Decomposed spectra.
The result of the fit procedure is shown in the right panel of Fig. 4(a). The time-dependent behavior of the fit and the experimental data are compared in detail in Fig. 3(b). The fit provides a very good description of the measured
The spectra associated with the three components are largely overlapping but exhibit clear differences regarding the intensities, positions, and shapes of their absorption features. In the energy range corresponding to the bleach of the parent molecule, the components
B.
Spin-orbit branching
The detection of both Br and Br* photoproducts following 268 nm excitation indicates that multiple potential energy surfaces participate in the dissociation of bromoform. Spin-orbit interaction can lead to the breakdown of the adiabatic approximation and coupling between otherwise orthogonal states. The dynamics for both Br and Br* appear indistinguishable, possibly due to our 110 fs IRF. Two methods are used to independently determine the spin–orbit branching. The intensity ratio is first obtained by fitting the asymptotic Br and Br* absorption peaks with Lorentzian functions, and then these are converted to amplitude by either (1) using the ab initio reduced transition dipole matrix elements given by Loh and Leone for the parallel polarization of pump and probe pulses used here,59 or (2) using the reported experimental absorption cross sections for the Br 2D5/2 ← 2P3/2, 2D3/2 ← 2P3/2 and Br*
2D3/2 ← 2P1/2 transitions of 45.5 Mb, 5.4 Mb, and 54.3 Mb, respectively.60 For either method, a consistent Br/Br* spin–orbit branching ratio of 3.7 ± 0.5 is found. Alternatively, the quantum yield of Br* is found as Φ(Br*) = 0.21 ± 0.04, where
(6)
for a pump intensity of 1 TW/cm2. A slightly larger value of Φ(Br*) = 0.23 ± 0.04 is obtained at 5 TW/cm2 but the two values agree within their uncertainties. A smaller spin–orbit branching ratio of 2.3 [or larger Φ(Br*) = 0.3] has been reported by comparison of translational energy distributions for Br and Br* to those of CHBr2 following nanosecond photolysis.32 However, the authors noted that the comparison for slow moving fragments was problematic.C.
Theoretical calculations
Ab initio calculations by Peterson and Francisco suggest that excitation to the lowest-lying singlet state, Ã1A2, is unimportant at the red edge of the absorption band.29 First, the transition is dipole forbidden in
Excited-state dynamics of bromoform are investigated by propagating trajectories on the T1 electronic surface using TDDFT with fewest-switches surface hopping (FSSH) and calculating the XUV absorption spectra for a subset of transient molecular configurations, as described in Sec. II A. Figure 5 compares the measured and calculated XUV absorption spectra, demonstrating remarkable agreement especially in capturing the dynamical evolution. Measured [Fig. 5(a)] and calculated [Fig. 5(b)] false-color maps show the spectral changes during the first few hundred femtoseconds using a similar color coding to that in Fig. 4. The TDDFT spectra are calibrated by a single energy offset (+0.5 eV) to agree with the experimental parent CHBr3 absorption peak position. Note that no spin–orbit splitting of the core Br(3d) manifold is included in the calculations. Instead, it is included after-the-fact by adding a replica of the computed spectrum that is shifted by the atomic spin–orbit splitting of ∼1 eV and rescaled by
FIG. 5.
(a) Measured and (b) calculated TDDFT transient absorption as a function of XUV energy and pump-probe time delay. Negative-going signal (blue) near 70 eV corresponds to decreased absorption due to depletion of the parent molecule concentration by the pump pulse. Positive-going signal (red) corresponds to absorption features. The theoretical spectra have been broadened spectrally and temporally by convolution with Gaussian functions to match the experimental resolution. (c) Theoretical transient absorption without temporal broadening, dotted lines emphasizes the bifurcation of a spectral feature (see the text for details). Integration of (a)–(c) over pump-probe time delays specified in legend results in the spectra in panels (d)–(f), respectively. In (g), a schematic molecular orbital and XUV transition energy diagram is shown, tracing the evolution of the valence electronic occupation and structure.
V.
DISCUSSION
XUV transient absorption at the Br(3d) edge probes the photoinduced chemistry from the viewpoint of the Br atoms since only transitions from well-localized Br(3d) inner-shell orbitals to valence orbitals exhibit notable oscillator strengths. A schematic molecular orbital and XUV transition energy diagram is shown in Fig. 5(g) for the ground state molecule and the photoexcited triplet state T1. Within this simplified picture, UV excitation transforms the originally fully occupied HOMO and unoccupied LUMO orbitals into singly occupied SOMO1 and SOMO2 orbitals. SOMO1 is derived from the Br(4p) orbitals that originally contribute to the HOMO in the parent molecule and asymptotically become atomic orbitals in the departing Br atom. Thus, XUV transitions into SOMO1 offer particularly detailed insight into the evolution of the electronic structure and the nuclear dynamics connecting the molecule to the isolated atom. SOMO2 develops localized character on the CHBr2 fragment providing a complementary viewpoint of changes in the molecular fragment. Throughout this work, shaded three-dimensional isosurfaces are rendered using an isovalue of 0.025 e where positive orbital values are colored in yellow and negative orbital values in green.
In the following, we provide a detailed discussion of the connections between changes in the transient absorption spectra and the orbitals involved in the transitions, providing deep insight into the coupled electronic-nuclear dynamics throughout the dissociation of CHBr3. Different stages of the dissociation are analyzed by comparing the
FIG. 6.
Decomposed spectra from the global fit of UV pump–XUV probe transient absorption spectra of bromoform recorded at the bromine
A.
Fast spectral evolution near t0
Transient spectra associated with very short delays are shown in Fig. 6(a). The fastest appearing spectrum,
The second absorption feature in the
B.
Spectral evolution at intermediate and long delays
At intermediate delays, corresponding to the
The experimental spectra associated with long delays, corresponding to the
C.
Contributions to changes in transient XUV absorption spectra
The spectra presented above illustrate that inner-shell transient absorption is sensitive to changes in both electronic structure and atomic configurations. The experimental signatures are amenable to comparison with high-level ab initio calculations, providing a rigorous approach to distinguish photochemical reaction mechanisms. Generally, inner-shell absorption energies correspond to the total energy differences between initial and final state electronic configurations. Predicting these differences is often challenging due to the breakdown of the single active electron picture, i.e., principally the response of all electronic orbitals in an atom or molecule needs to be taken into account. In many cases, however, it is instructive to discuss the transition energies in a simplified picture of core- and valence-orbital energies with the inclusion of an electron-hole interaction term, often referred to as exciton binding energy. This approximation is possible when the transition involves states that are dominated by a single orbital. This picture is illustrated in Fig. 7 for two types of transitions into the SOMO1 valence orbital, one from a 3d orbital on the Br atom that is departing the molecule [Fig. 7(a)] and the other for transitions from the 3d orbitals in the molecular CHBr2 fragment [Fig. 7(b)]. Note that in this particular example, the molecule is already slightly distorted from
FIG. 7.
Calculated binding energies of the 3dz2 orbital for each Br atom and SOMO1 for two time steps: following T1 electronic excitation (Δt = 0 fs) and after nuclear motion/dissociation (Δt = 105 fs). Isosurfaces of the T1 orbital electron densities are shown. Note that in this particular trajectory, the molecule is slightly distorted from
TABLE I.
Valence-excited core 3dz2 energies, core 3dz2-SOMO1 difference energies, calculated transition energies (EXAS), and electron-hole pair interaction energies for transitions originating from different Br atoms (all energies in electron-volt). No rigid shift of EXAS to match experimental transition energies is applied.
Δt = 0 fs | Δt = 105 fs | Δ(t105 − t0) | ||
---|---|---|---|---|
SOMO1 | −6.28 | −5.73 | 0.55 | |
Brdissoc | Core 3dz2 | −73.73 | −73.45 | +0.28 |
ΔEcore-SOMO1 | 67.45 | 67.72 | +0.27 | |
EXAS | 64.38 | 63.93 | −0.45 | |
e-h+ | 3.07 | 3.79 | +0.72 | |
BrA | Core 3dz2 | −73.60 | −73.83 | −0.23 |
ΔEcore-SOMO1 | 67.32 | 68.10 | +0.78 | |
EXAS | 66.03 | 67.03 | +1.00 | |
e-h+ | 1.29 | 1.07 | −0.22 | |
BrB | Core 3dz2 | −73.54 | −73.42 | +0.12 |
ΔEcore-SOMO1 | 67.26 | 67.69 | +0.43 | |
EXAS | 65.38 | 66.78 | +1.40 | |
e-h+ | 1.88 | 0.91 | −0.97 |
Within the picture illustrated in Fig. 7, the XUV transition energies correspond to the differences between core- and valence-orbital energies minus the electron-hole binding energies of the XUV-excited states. For example, 3d–SOMO1 transitions localized on the departing Brdissoc atom redshift by ∼0.5 eV during the dissociation [Fig. 7(a)], while corresponding transitions involving BrA and BrB in the remaining molecular fragment blueshift by ≥1 eV [Fig. 7(b)], leading to the bifurcation of inner-shell absorption signals near ∼66 eV illustrated in Fig. 5(c). The different signs of the shifts originate from the change in electron-hole binding energy. We note that an alternative and equivalent description to changes in Br(3d) inner-shell absorption energies during dissociation has been presented recently,61 where the valence-excited (initial state) and (valence+)core-excited (final state) potentials of a molecule (HBr) along the interatomic coordinate are first calculated and subsequently the absorption strengths are determined. Changes to the electronic configuration of both the valence and core orbitals that result from inner-shell absorption are included in the core-excited potential. Here, however, we choose a different representation to illustrate the physical underpinnings of the transition energies and strengths associated with the localization of SOMO1 into Brdissoc(4p) as the dissociation proceeds. Note that both before and after dissociation the energy gap between the Brdissoc(3d)/CHBr2(3d) orbitals and SOMO1 in the valence-excited T1 state is consistently ∼67.5 eV (Table I), yet the transitions appear at distinct energies and show different trends owing to the different magnitude and sign of the electron-hole interaction term. Viewed from each Br atom, core-excitation produces a valence electron-3d hole pair whose spatial extent depends on the evolving valence orbitals. The sensitivity of the probe to ongoing changes to the valence chemical environment, in terms of absorption energy and strength, is governed by both the energy difference between, and overlap of, the initial and final states.
The results presented in Fig. 7 and Table I also demonstrate that different inner-shell spectroscopy techniques may exhibit different degrees of sensitivity with respect to a particular chemical transformation. For the specific case of UV-induced dissociation in bromoform, our calculations indicate that the Br(3d) core-level shifts of each distinct Br atom differ by only ∼0.1 to ∼0.3 eV across the dynamical range presented, whereas the XAS transition energies exhibit variations from a 0.45 eV redshift to a 1.4 eV blueshift. Intuitively, one might expect the opposite: that core-level shifts would be systematically larger than corresponding shifts in XAS peak positions, due to cancelation of shifts in the same direction for both the core-level and core-excited molecular orbital binding energies. Clearly, this is not the case here. Energetics, however, are not the only aspect to consider. While photoionization is always an allowed process at sufficiently high photon energies, XAS is subject to more restrictive selection rules and requires spatial overlap of the core- and valence-orbitals participating in a transition. At Δt = 0, all localized 3d electron densities overlap with the delocalized SOMO1 orbital. At Δt = 105 fs, however, only the Brdissoc 3d orbitals are in the vicinity of the, now localized, SOMO1 while its overlap with the BrA/BrB 3d orbitals vanishes, leading to rapid loss of the high-energy branch in the bifurcated Br(3d)–SOMO1 signal in Fig. 5(c).
D.
Trajectories and isomerization pathways
Having established the ability of the ab initio XUV spectra to capture the temporal evolution of the experimental XUV spectra, we now analyze the underlying nuclear dynamics in greater detail by inspection of the MD trajectories from which the spectra are obtained. Figure 8 shows the distances between the departing atom (Brdissoc) and each remaining Br atom (BrA/BrB) in the CHBr2 fragment for 113 simulated TDDFT trajectories (blue traces) initiated on the T1 electronic surface. The diagonal represents a symmetrical pathway where the BrA and BrB atoms remain equidistant to the leaving Brdissoc atom throughout the dissociation. As BrA and BrB are experimentally indistinguishable, trajectories are plotted such that they appear predominantly above the diagonal. The trajectories are available in the supplementary material. The pictographs in the lower half of Fig. 8 illustrate the atomic displacements during the early stages of a particular trajectory, starting with the initial
FIG. 8.
Correlation between the distance of the departing atom (Brdissoc) and each remaining Br atom of the CHBr2 fragment (labeled BrA/BrB). The diagonal represents a pathway with the least isomerization, where the BrA and BrB atoms remain equidistant to the leaving atom. The TDDFT trajectory calculations initiated on the T1 electronic surface (blue) show that dissociation is most-often equidistant, with a Gaussian distribution about the diagonal. 2 of 113 trajectories approach the vicinity of the minimum energy iso-CHBr3 geometry on the T1 surface (green circle) and the reported MS-CASPT2 iso-CHBr3 S0 geometry (red circle).43 The MS-CASPT2 trajectory (dashed red) falls within the distribution of symmetrical trajectories.43
The path taken by the MS-CASPT2 trajectory reported by Mereshchenko et al. is overlaid in Fig. 8 by a red dotted line; while being somewhat asymmetric, it falls within the distribution of direct C–Br fission TDDFT trajectories.43 Of the 113 TDDFT trajectories calculated here, two trajectories did approach the relaxed iso-CHBr3 geometry minimum shown as a green circle for the T1 surface. This T1 TDDFT isomer minimum geometry is similar to the S0 MS-CASPT2 minimum reported in Ref. 43, marked here as a red circle. The ensemble of TDDFT trajectories exhibit a Gaussian distribution about the diagonal, quantified in Fig. S6, i.e., symmetrical dissociation is the most probable outcome. While the exact branching will depend on the accuracy of the PES, isomerization appears to be strongly disfavored in the 268 nm photochemistry of isolated bromoform molecules.
The time scale for isomerization in the trajectory calculations, a rearrangement that requires significant motion of the heavy Br atoms, is much longer than the characteristic time constants on which the XUV spectra evolve, 40 ± 20 fs and 85 ± 10 fs, as obtained from a global fit. For example, the trajectory that passes through the iso-CHBr3 minimum conformation in Fig. 8, reaches this point after 230 fs. In order to test the prediction of the MD simulations that iso-forming trajectories are strongly suppressed, their expected spectroscopic fingerprints are compared with experimental observations. Figure 9 shows the experimental XUV absorption spectrum at a pump–probe delay of 230 fs (black) recorded at a relatively low pump intensity of 1 TW/cm2 to ensure that single-photon features dominate the spectrum. The measurement is compared to calculated XUV spectra for trajectories sampling either the iso-CHBr3 conformation (red, iso traj.) or following an approximately symmetric dissociation (blue, sym. traj.). XUV spectra for the relaxed iso-CHBr3 minimum conformation (green dot, Fig. 8) are also shown (green, iso min.). The XUV spectra obtained for the isomer minimum energy conformation are consistent with the spectra obtained from the isomer-sampling trajectory and only differ in the relative intensities of specific features. Note that the Br* and Br+ channels are not included in the trajectory calculations, and they are therefore absent in the calculated XUV absorption spectra. The calculated iso-CHBr3 spectra predict characteristic absorption features arising from 3d–SOMO1 transitions of the Br2-like C–Br–Br linkage at ∼68 ± 1 eV. No such feature is observed in the experiment. Instead, the measured spectrum is consistent with the XUV spectrum derived from a symmetrical dissociation pathway. The spectral fingerprints of specific molecular conformations, therefore, confirm that photodissociation of CHBr3 following the isomerization pathway is insignificant compared to the symmetric breakup channel.
FIG. 9.
The experimental XUV absorption spectrum at a pump–probe delay of 230 fs (black) is compared to TDDFT XUV absorption spectra at the same time delay. The red spectrum corresponds to a trajectory that samples the formation of the iso-CHBr3 conformation (approximately red dot, Fig. 8), the blue spectrum to a typical trajectory following symmetric dissociation. The calculated XUV spectrum for the relaxed iso-CHBr3 T1 minimum geometry is also shown (green) (green dot, Fig. 8). The calculated iso-CHBr3 spectrum has additional absorption features near 67–68 eV arising from the molecular bonding character of the C–Br–Br linkage. The time evolution of the calculated spectra for symmetric dissociation is consistent with that observed in the experimental spectra.
Returning to the conflicting reports of Br2 as a primary photoproduct of bromoform, we have spectroscopically determined that no Br2 forms in a collision free environment on timescales <10 ps, i.e., no excited state directly dissociates to form Br2. The trajectory calculations are consistent with this observation, leading solely to C–Br fission, no other product channels were obtained. While it is possible that unimolecular dissociation on S0 could form Br2 over longer time scales, the molecular channel (III) is unfavorable as the radical channel (I) lies 1.14 eV lower in energy. Kalume et al. indicate that only 4% branching to molecular products at 248 nm is anticipated.62 Branching to form molecular products from unimolecular dissociation on S0 following photolysis at 268 nm would be even less favorable. To explain the reduced quantum yield of channel (I) as reported in Ref. 33, experiments probing radiative relaxation may prove valuable.
VI.
CONCLUSIONS
Unprecedented insight into the 268 nm-induced photodissociation of bromoform is reported. The element-specific nature of inner-shell spectroscopy is used to monitor the evolution of the valence electron environment of the different Br atoms from a well-localized perspective. A global fit model is used to decompose the XUV spectral evolution. Transformation of the initially populated excited state into photoproducts can be described by two sequential steps characterized by transition timescales of 40 ± 20 fs and 85 ± 10 fs. However, despite the ability of the 3-state model to accurately describe the experimental data, we caution that sampling continuous spectral changes with a finite experimental time-resolution may suggest the appearance of intermediate species but does not necessarily provide the most accurate description of the underlying physics. Here, the dynamics are investigated using excited-state MD simulations initiated on the T1 surface from which XUV absorption spectra are calculated. Analysis of the trajectories shows a dominant propensity for a continuous C–Br extension along the
SUPPLEMENTARY MATERIAL
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
© 2019 Author(s). This work is published under http://creativecommons.org/licenses/by/4.0/ (the “License”). Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.
Abstract
UV pump–extreme UV (XUV) probe femtosecond transient absorption spectroscopy is used to study the 268 nm induced photodissociation dynamics of bromoform (CHBr3). Core-to-valence transitions at the Br(3d) absorption edge (∼70 eV) provide an atomic scale perspective of the reaction, sensitive to changes in the local valence electronic structure, with ultrafast time resolution. The XUV spectra track how the singly occupied molecular orbitals of transient electronic states develop throughout the C–Br bond fission, eventually forming radical Br and CHBr2 products. Complementary ab initio calculations of XUV spectral fingerprints are performed for transient atomic arrangements obtained from sampling excited-state molecular dynamics simulations. C–Br fission along an approximately
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer