Next Article in Journal
Human Vital Signs Signal Monitoring and Repairment with an Optical Fiber Sensor Based on Deep Learning
Next Article in Special Issue
Laser Scanning Method for Time-Resolved Measurements of Wavefront Distortion Introduced by Active Elements in High-Power Laser Amplifiers
Previous Article in Journal
A Retrospective Study: Are the Multi-Dips in the THz Spectrum during Laser Filamentation Caused by THz–Plasma Interactions?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dissociative Ionization of the CHBr2Cl Molecule in 800 nm and 400 nm Femtosecond Laser Fields

1
School of Information Engineering, Yancheng Institute of Technology, Yancheng 224051, China
2
State Key Laboratory of Precision Spectroscopy, School of Physics and Electronic Science, East China Normal University, Shanghai 200062, China
*
Author to whom correspondence should be addressed.
Photonics 2024, 11(8), 706; https://doi.org/10.3390/photonics11080706
Submission received: 8 July 2024 / Revised: 26 July 2024 / Accepted: 28 July 2024 / Published: 29 July 2024
(This article belongs to the Special Issue New Perspectives in Ultrafast Intense Laser Science and Technology)

Abstract

:
The dissociative ionization of CHBr2Cl molecules in femtosecond laser fields at 800 nm and 400 nm is investigated to enhance the comprehension of ultrafast dynamics phenomena. The kinetic energy distribution of the resulting ions following photo-dissociation is analyzed using time-of-flight mass spectrometry in combination with DC-sliced ion velocity map imaging. The findings from the experimental study indicate that the presence of low kinetic energy components is attributed to the dissociative ionization processes of CHBr2Cl molecules. The complexity of individual dissociation pathways remains unaffected by the laser fields but is determined by factors such as bond energy, ionization energy of neutral groups, and charge distribution. In the case of 400 nm laser fields, distinct elimination channels enable CHBr2Cl+ ions to circumvent the transition state, leading to the formation of BrCl+ and Br2+ fragments.

1. Introduction

The advancement of ultrafast laser technologies has transformed the laser into a potent instrument for investigating the rapid dissociation dynamics of molecules [1,2,3,4,5]. When a laser interacts with matter, ionization emerges as the predominant physical phenomenon. At sufficiently high laser intensities, molecules can undergo electron loss via multiphoton and above-threshold ionization (MPI/ATI), even when the energy of a single photon is lower than the ionization potential (IP). The interaction of the laser field with the molecular structure alters the potential barrier, thereby influencing the Coulomb potential between the molecular nucleus and the valence electron as the laser intensity escalates. The valence electron can surmount the barrier and transition to a continuum state through tunneling ionization (TI). As the laser intensity further rises, the energy barrier of the bound state can be completely suppressed below the energy level of the valence electron, facilitating the easy removal of the valence electron from the molecular system, a process known as over-the-barrier ionization (OTBI). Ionization is typically the initial step preceding dissociation processes, which may take hundreds of femtoseconds or even as much as a picosecond to occur. Consequently, ionization precedes the onset of dissociation in the femtosecond laser field, a phenomenon referred to as dissociative ionization (DI). In the presence of a femtosecond laser field with an intensity exceeding 1014 W/cm2, multi-electron dissociative ionization can occur, leading to the rupture of chemical bonds within molecules due to the internal Coulomb repulsive energy of the multi-charged parent ions, a phenomenon termed Coulomb Explosion (CE) [6,7,8,9].
The photo-dissociation of halogenated hydrocarbons has been the subject of numerous investigations over the past few decades. Ibuki et al. measured the photo-absorption and fluorescence excitation spectra of CHBr2Cl in the 106–200 nm region using synchrotron radiation. The CHCl radical was observed, and the lifetime of the excited state was first determined to be 7.01 ± 0.25 μs [10]. Zheng et al. have reported the transient resonance Raman spectra for the photoproducts produced after the ultraviolet excitation of CHBr2Cl in a cyclohexane solution. The transient resonance Raman spectrum observed following ultraviolet excitation was primarily attributable to the iso-CHBrCl–Br and iso-CHClBr–Br species, according to their findings [11]. Yang et al. have revealed the principal direct estimations of the transitional items in the multiphoton photo-dissociation response of CHBr2Cl at 266 nm. The multiphoton photo-dissociation mechanism was confirmed based on these results [12]. Muthiah et al. have utilized cavity ring-down absorption spectroscopy (CRDS) to investigate the photo-dissociation of CHBr2Cl at 248 nm. The photo-dissociation of CHBr2Cl mainly produces Br2 fragments, and the production of BrCl is limited by the high transition state energy barrier, which makes its experimental detection difficult [13]. While UV nanosecond laser fields are commonly used in halogenated hydrocarbon studies, limited attention has been given to ultraviolet femtosecond laser fields generated through second harmonic production of 800 nm laser fields. Halogenated hydrocarbons, characterized by easy dissociation in UV laser beams and low ionization energy, may ionize and fragment into smaller particles under weak UV femtosecond lasers. Hence, the examination of the dissociative ionization process of halogenated hydrocarbons under femtosecond laser fields at wavelengths of 800 and 400 nm, corresponding to single photon energies of 1.55 and 3.10 eV, respectively, has the potential to reveal the rapid dynamics of CHBr2Cl molecules.
In this study, we investigated the dissociative ionization process of the CHBr2Cl molecule under femtosecond laser fields with wavelengths of 800 nm and 400 nm. We analyzed the kinetic energy distribution of the photo-dissociation product ions. The results suggest that CHBr2Cl molecules can undergo single ionization and dissociate into fragment ions with low kinetic energy. The ionization energy of the fragments influences the charge distribution of the parent ions, thereby affecting the emergence of the dissociation channel. Furthermore, a specific route observed under the 400 nm laser field allows the molecular ion to bypass the transition state, facilitating bond cleavage and the production of BrCl+ and Br2+ ions.

2. Materials and Methods

The experimental setup utilized in the study is described in more detail in previous literature [14,15]. In brief, femtosecond laser pulses from Spectra-Physics were employed, featuring a central wavelength of 800 nm, a repetition rate of 1 kHz, and a pulse duration of 100 fs. These pulses were focused into the reaction chamber using a 400 mm lens. A pulsed valve from General Valve, Parker, operating at a repetition rate of 100 Hz and a duration of 130 μs, was utilized to introduce a gaseous sample of CHBr2Cl seeded with argon (1:10) into the reaction chamber at a pressure of 5.0 × 10−9 mbar. The interaction between the supersonic molecular beam and the femtosecond pulse occurred at the center of the ion lens’s repeller and extractor plates. Detection of the fragment ion cloud was achieved using micro-channel plates, a phosphor screen, and a mapping lens, with the resulting time-of-flight (TOF) mass spectra recorded by a photomultiplier tube (H7732-11, Hamamatsu, Shizuoka, Japan) connected to a digital oscilloscope (LeCroy Wave Pro). Momentum images were captured using a charge-coupled device camera (PI-MAXII, Princeton Instrument, Trenton, NJ, USA) with a time resolution of 5 ns. The laser’s polarization vector was oriented perpendicular to the ion flight direction. The intensity within the reaction region was calibrated by the Ar2+/Ar+ yield ratio [16]. A digital delay/pulse generator (DG645, Stanford Instrument, Sunnyvale, CA, USA) was employed for precise timing sequence control.
The dissociative ionization process was verified using Gaussian 16 software [17]. Optimization of fragment geometries and calculation of zero-point correction (ZPE) at the B3LYP level were performed with a 6-31G* basis set [18,19]. Single-point energies were computed using the CCSD(T) method with the cc-pVTZ basis set [20,21]. Vibrational frequencies were calculated to validate transition states, and the intrinsic reaction coordinate (IRC) calculation method was employed to determine the pathway of least energy from the transition state to the corresponding minima.

3. Results

The time-of-flight (TOF) mass spectrum of the CHBr2Cl molecule under femtosecond laser irradiation at 800 nm and 400 nm, with intensities of 1.0 × 1014 and 4 × 1013 W/cm2, respectively, is depicted in Figure 1. The abundance of Br+ ions in each spectrum is standardized to one. The experimental results reveal the predominant formation of Cl+, CHCl+, Br+, CHBr+, CHBrCl+, and CHBr2+ ions. Notably, Figure 1a exclusively displays the ion CHBrCl2+, potentially attributed to the elevated intensity of the 800 nm laser fields at 2.5 × 1014 W/cm2. The generation of fragment ions may be attributed to multi-electron dissociative ionization of the parent molecules. Figure 1b illustrates the parent ion CHBr2Cl+ alongside the fragment ions BrCl+ and Br2+, indicating the occurrence of distinct reaction pathways, which will be elaborated upon in subsequent sections.
In addition to identifying the ion species, interesting discoveries about ion yields can also be made by further exploration of time-of-flight mass spectra. It is generally believed that in the femtosecond laser action to the molecular system, when the laser intensity is low, the main occurrence of multiphoton ionization produces low valence state molecular ions, and then dissociates to produce fragment ions, while the laser intensity enhancement will occur in the field ionization, then produce high valence molecular ions and then the Coulomb explosion. As depicted in Figure 2, at lower intensities in both 800 nm and 400 nm laser fields, the yields of product fragment ions are relatively low and comparable. However, as laser intensity escalates, discernible differences in yield growth emerge, particularly with the most rapid increases observed in Br and Cl ions. The ion yields resulting from dissociation pathways involving the C–Br bond are observed to be greater than those involving the C–Cl bond. This observation indicates that the dissociation process is not significantly influenced by the laser wavelength but is likely influenced by the energy of the molecular bond, which will be analyzed in detail in the next section.
In addition to analyzing time-of-flight mass spectra, the investigation extends to the examination of kinetic energy release (KER) of fragment ions under varying laser conditions. Figure 3 displays the normalized velocity distribution of fragment ions along with the DC-sliced images. The red curve corresponds to data obtained using an 800 nm laser field, while the blue curve represents data acquired with a 400 nm laser field. The velocity distribution exhibits a multi-peaked structure for some of the fragments, indicating the existence of multiple reaction pathways. Typically, the low kinetic energy component of the fragments results from the dissociation of single-charged parent ions, whereas the high kinetic energy component is attributed to the Coulomb explosion process of multi-charged parent ions. Notably, it has been observed to reduce the yields of the heavy fragments CHCI+, CHBr+, CHBrCl+, and CHBr2+ with high KER in the 400 nm laser filed. This reduction is linked to the limited laser intensity of up to 4 × 1013 W/cm2 in 400 nm laser fields, rendering the Coulomb explosion effect negligible. The high kinetic energy component observed in the Br ion data in Figure 3(c′) is proposed to stem from a sequential dissociation process rather than a Coulomb explosion, as no corresponding ion with a high kinetic energy component has been identified. Additionally, fragments of BrCl+ and Br2+ ions have only been detected in Figure 3(g′,h′) when irradiated by 400 nm laser fields.

4. Discussion

4.1. Dissociation of Single-Charged Parent Ion CHBr2Cl+ in 800 and 400 nm Femtosecond Laser Fields

Given the 100 femtosecond (full width at half maximum) duration of our laser pulse, it is reasonable to expect that the CHBr2Cl molecule will undergo ionization to form a molecular ion, followed by the subsequent fragmentation of the molecular ion as its bonds are broken. This section will concentrate on the dissociative ionization mechanisms, specifically the disintegration of the singly charged parent ion CHBr2Cl+ under 800 and 400 nm femtosecond laser fields. The dissociative process results in the creation of two components: a neutral fragment and ions. Therefore, the potential pathways for fragment ion dissociation in both laser fields are delineated. It is important to note that the initial four pathways mentioned involve the cleavage of a single chemical bond. Channels (1) and (2) involve the rupture of a C–Cl bond, with the positive charge being attributed to either the CHBr2+ or Cl+ fragment. Channels (3) and (4) entail the breaking of a C–Br bond, with the positive charge being assigned to either the CHClBr+ or Br+ fragment. The final two pathways mentioned involve dissociation across multiple chemical bonds. Channel (5) involves the cleavage of a C–Br bond and a C–Cl bond, assigning the positive charge to the CHBr+ fragment. Channel (6) involves the breaking of two C–Br bonds, with the positive charge being assigned to the CHCl+ fragment.
CHBr2Cl+ → Cl+ + CHBr2
CHBr2Cl+ → CHBr2+ + Cl
CHBr2Cl+ → Br+ + CHBrCl
CHBr2Cl+ → CHBrCl+ + Br
CHBr2Cl+ → CHBr+ + Br+ Cl
CHBr2Cl+ → CHCl+ + 2Br
For the problem in the previous section, we performed some calculations to explain. Initially, the geometry of the CHBr2Cl molecule in its ground state was optimized. Subsequently, one electron was removed, and the structure was re-optimized to generate the ground-state CHBr2Cl+ ions. The vibrational frequency was then determined. The compliance constant matrix was derived using a compliance program [22]. A chemical bond is deemed weaker if the diagonal element in the compliance constant matrix has a higher value. Consequently, the reciprocals of these diagonal elements, referred to as relaxed force constants, can serve as indicators of bond strength. As indicated in Table 1, the C–Cl bond in CHBr2Cl+ is notably stronger than the C–Br bond. Thus, the primary dissociation reactions predominantly entail the cleavage of the C–Br bond.
In order to investigate the dissociation processes more effectively, the Gaussian 16 software package was employed to calculate all six channels. For simplification in the analysis, the ground state of the CHBr2Cl molecule was designated as the zero-energy reference point. Consequently, the appearance energies of the dissociation channels were determined as the energy differences between the fragments and the CHBr2Cl molecule. The energy values associated with the dissociation pathways are presented in Table 2, with detailed explanations of appearance energy and available energy provided in reference [4]. It is evident that all KER values of the fragment ions are lower than the calculated available energy, indicating that the fragment ions with low-KER components are likely products of the dissociation processes of the singly charged molecular ions CHBr2Cl+.
The appearance energy in Table 2 indicates a noticeable disparity between channels (1) and (2), as well as between channels (3) and (4). This difference may be attributed to variations in charge distribution among the groups and discrepancies in ionization energy levels among the neutral groups. Gaussian 16 was utilized to examine the natural bond orbital (NBO) charge distribution of CHBr2Cl+ ions, as illustrated in Figure 4. The positive charge is distributed between the Cl atom and the CHBr2 group, with charges of +0.228 and +0.772, respectively. In the event of the C–Cl bond breaking, either Cl or CHBr2 will release the “extra electron” to form channel (1) or channel (2). The ionization energies of the fragments Cl and CHBr2 are measured at 14.37 and 8.12 eV, respectively. This indicates that CHBr2 is more inclined to release the “extra electron” to produce the CHBr2+ ion, resulting in channel (2) having a notably lower appearance energy than channel (1). A similar investigation was conducted on the dissociation channel of the C–Br bond. Br and CHBrCl exhibit charges of +0.477 and +0.523, respectively, with associated ionization energies of 13.09 and 8.19 eV. This analysis suggests that the ionization energy of fragments significantly impacts the appearance energy of the dissociation channel, thereby explaining the lower appearance energy of channel (4) in comparison to channel (3).

4.2. The Special Dissociation Channels for CHBr2Cl in 400 nm Laser Fields

The generation of BrCl+ and Br2+ ions in 400 nm laser field, as depicted in Figure 1, indicates the presence of varied dissociation pathways. To delve deeper into these pathways, Gaussian 16 was employed for computational analysis. Figure 5 illustrates the elimination process, which can be divided into three sequential stages: firstly, the neutral parent molecule loses an electron to form the parent ion; secondly, the parent ion transitions via a conformation change to an immediate compound; and finally, the intermediate compound facilitates bond cleavage, resulting in the formation of distinct fragments.
The detailed structural modifications observed throughout the entire reaction process are of particular interest. Initially, the molecule loses one electron, transitioning into an ionic state. Despite the loss of one electron, the length of the C–Br bond remains relatively constant, while the distance between the two Br atoms decreases from 3.2 to 2.8 Å. Additionally, the Br–C–Br bond angle decreases from 111.6° to 90.6°. This alteration is attributed to the electron of the CHBr2Cl molecule’s highest occupied molecular orbital (HOMO), which is shared by both Br atoms. For channel (7), the elimination reaction progresses from the ground state of singly charged molecular ions to the transition state, where both C–Br bonds initially lengthen slightly. However, the separation between the Br atoms diminishes rapidly, along with the Br–C–Br bond angle. Subsequently, the transition state (TS_Br2+) advances along the reaction pathway towards the metastable intermediate (IS_Br2+). Throughout this progression, either the C–Br bond elongates until rupture or the Br–Br bond distance stabilizes. Ultimately, the C–Br bond breaks, resulting in the formation of Br2+ and CHCl. For channel (8), the angle of the Br–C–Cl bond experiences a significant decrease. As the transition state (TS_BrCl+) is formed, there is a rapid reduction in the distance between the Br–Cl atoms, leading to stabilization, while the C–Cl bond elongates further to create a metastable intermediate (IS_BrCl+). Finally, the C–Br bond is cleaved to release the desired ions.
CHBr2Cl+ → Br2+ + CHCl
CHBr2Cl+ → BrCl+ + CHBr
Channels (7) and (8) are uniquely favored in the 400 nm laser field due to the higher excitation efficiency of the transition state in comparison to the 800 nm laser field. This presents an opportunity for coherent control through the manipulation of laser field parameters, such as wavelength, to selectively stimulate particular dissociation channels.

5. Conclusions

The dissociative ionization of CHBr2Cl molecules in 800 and 400 nm laser fields was investigated utilizing time-of-flight mass spectra and DC-sliced velocity ion imaging technology. In our experiments the dissociation processes are independent of the laser field intensities. They are rather influenced by the bond energy, neutral group ionization energy, and the charge distribution. In a 400 nm laser field, we find additional sequential dissociation channels that form BrCl+ and Br2+ fragment ions. Our findings demonstrate that by changing the wavelength of the laser fields, specific ultrafast photo-dissociation channels can be selectively triggered in molecular systems.

Author Contributions

B.L. completed the experimental work and draft, and Z.L. finished part of theoretical calculation. The original experiment idea was proposed and experiments were designed by B.L. and B.L. submitted the final manuscript for publication. All authors have read and agreed to the published version of the manuscript.

Funding

This work has been partially supported by The Scientific Research Project for the Introduction of Talent of Yancheng Institute of Technology (No. xjr2021069), and The Natural Science Foundation of the Jiangsu Higher Education Institutions of China (No. 24KJB140020).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

Thanks to Yan Yang and Zhenrong Sun of East China Normal University for the experimental setup and technical support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hertel, I.; Radloff, W. Ultrafast dynamics in isolated molecules and molecular clusters. Rep. Prog. Phys. 2006, 69, 1897. [Google Scholar] [CrossRef]
  2. Brinks, D.; Hildner, R.; Van Dijk, E.M.; Stefani, F.D.; Nieder, J.B.; Hernando, J.; van Hulst, N.F. Ultrafast dynamics of single molecules. Chem. Soc. Rev. 2014, 43, 2476–2491. [Google Scholar] [CrossRef]
  3. Cornaggia, C. Electronic dynamics of charge resonance enhanced ionization probed by laser-induced alignment in C2H2. J. Phys. B At. Mol. Opt. Phys. 2016, 49, 19LT01. [Google Scholar] [CrossRef]
  4. Wu, H.; Zhang, S.; Zhang, J.; Yang, Y.; Deng, L.; Jia, T.; Wang, Z.; Sun, Z. Observation of hydrogen migration in cyclohexane under an intense femtosecond laser field. J. Phys. Chem. A 2015, 119, 2052–2057. [Google Scholar] [CrossRef] [PubMed]
  5. Lai, W.; Guo, C. Selective charge asymmetric distribution in heteronuclear diatomic molecules in strong laser fields. Phys. Rev. A 2015, 92, 013402. [Google Scholar] [CrossRef]
  6. Wu, Z.; Wu, C.; Liu, X.; Deng, Y.; Gong, Q.; Song, D.; Su, H. Double ionization of nitrogen from multiple orbitals. J. Phys. Chem. A 2010, 114, 6751–6756. [Google Scholar] [CrossRef] [PubMed]
  7. Hatherly, P.; Stankiewicz, M.; Codling, K.; Frasinski, L.J.; Cross, G.M. The multielectron dissociative ionization of molecular iodine in intense laser fields. J. Phys. B At. Mol. Opt. Phys. 1994, 27, 2993. [Google Scholar] [CrossRef]
  8. McKenna, J.; Suresh, M.; Srigengan, B.; Williams, I.D.; Bryan, W.A.; English, E.M.L.; Stebbings, S.L.; Newell, W.R.; Turcu, I.C.E.; Smith, J.M.; et al. Ultrafast ionization study of N2 in intense linearly and circularly polarized laser fields. Phys. Rev. A 2006, 73, 043401. [Google Scholar] [CrossRef]
  9. Zhang, J.; Li, Z.; Yang, Y. Multi-ionization of the Cl2 molecule in the near-infrared femtosecond laser field. RSC Adv. 2020, 10, 332–337. [Google Scholar] [CrossRef]
  10. Ibuki, T.; Hiraya, A.; Shobatake, K. Vacuum ultraviolet absorption spectra and photodissociative excitation of CHBr2Cl and CHBrCl2. J. Chem. Phys. 1992, 96, 8793–8798. [Google Scholar] [CrossRef]
  11. Zheng, X.; Lee, C.W.; Li, Y.-L.; Fang, W.-H.; Phillips, D.L. Transient resonance Raman spectroscopy and density functional theory investigation of iso-CHBr2Cl and iso-CCl3Br photoproducts produced following ultraviolet excitation of CHBr2Cl and CCl3Br. J. Chem. Phys. 2001, 114, 8347–8356. [Google Scholar] [CrossRef]
  12. Yang, S.-X.; Hou, G.-Y.; Dai, J.-H.; Chang, C.-H.; Chang, B.-C. Spectroscopic Investigation of the Multiphoton Photolysis Reactions of Bromomethanes (CHBr3, CHBr2Cl, CHBrCl2, and CH2Br2) at Near-Ultraviolet Wavelengths. J. Phys. Chem. A 2010, 114, 4785–4790. [Google Scholar] [CrossRef] [PubMed]
  13. Muthiah, B.; Kasai, T.; Lin, K.-C. Probing BrCl from photo-dissociation of CH2BrCl and CHBr2Cl at 248 nm using cavity ring-down spectroscopy. Phys. Chem. Chem. Phys. 2021, 23, 6098–6106. [Google Scholar] [CrossRef]
  14. Zhang, J.; Yang, Y.; Li, Z.; Sun, H.; Zhang, S.; Sun, Z. Channel-resolved multiorbital double ionization of molecular Cl2 in an intense femtosecond laser field. Phys. Rev. A 2018, 98, 043402. [Google Scholar] [CrossRef]
  15. Zhang, J.; Yang, Y.; Li, Z.; Sun, Z. Dissociative ionization of CH2Br2 in 800 and 400 nm femtosecond laser fields. Chem. Phys. Lett. 2017, 685, 151–156. [Google Scholar] [CrossRef]
  16. Guo, C.; Li, M.; Nibarger, J.P.; Gibson, G.N. Single and double ionization of diatomic molecules in strong laser fields. Phys. Rev. A 1998, 58, R4271–R4274. [Google Scholar] [CrossRef]
  17. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16 Rev. C.01; Gaussian: Wallingford, CT, USA, 2016. [Google Scholar]
  18. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785. [Google Scholar] [CrossRef] [PubMed]
  19. Woon, D.E.; Dunning, T.H., Jr. Gaussian basis sets for use in correlated molecular calculations. III. The atoms aluminum through argon. J. Chem. Phys. 1993, 98, 1358–1371. [Google Scholar] [CrossRef]
  20. Pople, J.A.; Head-Gordon, M.; Raghavachari, K. Quadratic configuration interaction. A general technique for determining electron correlation energies. J. Chem. Phys. 1987, 87, 5968–5975. [Google Scholar] [CrossRef]
  21. Truhlar, D.G. Basis-set extrapolation. Chem. Phys. Lett. 1998, 294, 45–48. [Google Scholar] [CrossRef]
  22. Brandhorst, K.; Grunenberg, J. Efficient computation of compliance matrices in redundant internal coordinates from Cartesian Hessians for nonstationary points. J. Chem. Phys. 2010, 132, 184101. [Google Scholar] [CrossRef]
Figure 1. The time-of-flight mass spectrum of the molecule CHBr2Cl irradiated by (a) 1.0 × 1014 W/cm2, 800 nm, and (b) 4 × 1013 W/cm2, 400 nm laser fields, respectively.
Figure 1. The time-of-flight mass spectrum of the molecule CHBr2Cl irradiated by (a) 1.0 × 1014 W/cm2, 800 nm, and (b) 4 × 1013 W/cm2, 400 nm laser fields, respectively.
Photonics 11 00706 g001
Figure 2. The yield of product fragment ions varies with laser power at (a) 800 nm, and (b) 400 nm laser fields, respectively.
Figure 2. The yield of product fragment ions varies with laser power at (a) 800 nm, and (b) 400 nm laser fields, respectively.
Photonics 11 00706 g002
Figure 3. The normalized velocity distribution of the fragment ions, (a) Cl+, (b) CHCl+, (c) Br+, (d) CHBr+, (e) CHBrCl+, (f) CHBr2+ in 800 nm laser fields and (a′) Cl+, (b′) CHCl+, (c′) Br+, (d′) CHBr+, (e′) CHBrCl+, (f′) CHBr2+, (g′) BrCl+, and (h′) Br2+ in 400 nm laser fields, respectively. The circle represents the experiment data, the green dashed lines represent the fitting peaks, and the red and blue solid lines represent the simulated distribution in 800 nm and 400 nm laser fields, respectively. The peak values of the correlating kinetic energy release (KER) are labeled in eV. The right inset shows the DC-sliced image of the corresponding fragment ions, and the double arrows show the direction of laser polarization.
Figure 3. The normalized velocity distribution of the fragment ions, (a) Cl+, (b) CHCl+, (c) Br+, (d) CHBr+, (e) CHBrCl+, (f) CHBr2+ in 800 nm laser fields and (a′) Cl+, (b′) CHCl+, (c′) Br+, (d′) CHBr+, (e′) CHBrCl+, (f′) CHBr2+, (g′) BrCl+, and (h′) Br2+ in 400 nm laser fields, respectively. The circle represents the experiment data, the green dashed lines represent the fitting peaks, and the red and blue solid lines represent the simulated distribution in 800 nm and 400 nm laser fields, respectively. The peak values of the correlating kinetic energy release (KER) are labeled in eV. The right inset shows the DC-sliced image of the corresponding fragment ions, and the double arrows show the direction of laser polarization.
Photonics 11 00706 g003aPhotonics 11 00706 g003b
Figure 5. Ab initio calculated dissociation pathways of the singly charged CHBr2Cl+ ions.
Figure 5. Ab initio calculated dissociation pathways of the singly charged CHBr2Cl+ ions.
Photonics 11 00706 g005
Figure 4. Natural bond orbital (NBO) charge distribution of CHBr2Cl+ ions.
Figure 4. Natural bond orbital (NBO) charge distribution of CHBr2Cl+ ions.
Photonics 11 00706 g004
Table 1. The compliance matrix element and relaxed force constant of the singly charged ion CHBr2Cl+, calculated at the B3LYP/6-31G* level.
Table 1. The compliance matrix element and relaxed force constant of the singly charged ion CHBr2Cl+, calculated at the B3LYP/6-31G* level.
BondCompliance Matrix Diagonal ElementRelaxed Force
Constant (mdyn/Å)
C–H0.185.56
C–Cl0.283.57
C–Br(1)0.711.41
C–Br(2)0.711.41
Table 2. The kinetic energy released by fragment ions, the appearance energy, and the available energy of the corresponding channel.
Table 2. The kinetic energy released by fragment ions, the appearance energy, and the available energy of the corresponding channel.
ChannelEappear (eV)KER (eV)Eavail (eV)
800 nm400 nm800 nm400 nm
(1)17.580.270.051.021.02
(2)11.320.110.081.081.08
(3)15.670.260.181.382.93
(4)10.770.170.050.081.63
(5)15.240.210.120.260.26
(6)14.740.250.140.760.76
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, B.; Li, Z. Dissociative Ionization of the CHBr2Cl Molecule in 800 nm and 400 nm Femtosecond Laser Fields. Photonics 2024, 11, 706. https://doi.org/10.3390/photonics11080706

AMA Style

Liu B, Li Z. Dissociative Ionization of the CHBr2Cl Molecule in 800 nm and 400 nm Femtosecond Laser Fields. Photonics. 2024; 11(8):706. https://doi.org/10.3390/photonics11080706

Chicago/Turabian Style

Liu, Botong, and Zhipeng Li. 2024. "Dissociative Ionization of the CHBr2Cl Molecule in 800 nm and 400 nm Femtosecond Laser Fields" Photonics 11, no. 8: 706. https://doi.org/10.3390/photonics11080706

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop