Home Journals Progress in Chemistry
Progress in Chemistry

Abbreviation (ISO4): Prog Chem      Editor in chief: Jincai ZHAO

About  /  Aim & scope  /  Editorial board  /  Indexed  /  Contact  / 
Chemistry: A Century of Life-Special Edition

The Photodissociation and Photoionization Dynamics of Some Important Small Molecules

  • Min Cheng , 1, 3, * ,
  • Lijuan Zhang , 2, * ,
  • Xiling Xu , 1, 3, * ,
  • Hong Gao 1, 3 ,
  • Weijun Zheng 1, 3
Expand
  • 1 Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, China
  • 2 Shangdong University of Aeronautics, Binzhou 256600, China
  • 3 University of Chinese Academy of Sciences, Beijing 100049, China
* e-mail: (Min Cheng);
(Lijuan Zhang);
(Xiling Xu)

Received date: 2024-11-19

  Revised date: 2024-12-12

  Online published: 2024-12-20

Abstract

The study of microscopic mechanisms of photodissociation and photoionization on small molecules is the major focus in the field of molecular reaction dynamics, which is important from both theoretical and practical aspects. It not only can reveal the physicochemical nature of the interaction between molecules and light, but also can help to understand and eventually regulate the chemical reaction process at the quantum level. This paper systematically reviews the research accomplishments achieved by Academician Zhu Qihe’s group in this field over the years. By utilizing the home-made photofragment translational spectrometers, they have comprehensively explored the photodissociation processes and revealed the microscopic reaction mechanisms for a series of halogenated hydrocarbons in the A band, by measuring the translational energies and spatial angular distributions of the photofragments. They have also investigated the geometric configurations, vibrational spectra, transition energies and ionization energies of benzene derivatives in different electronic states, by using the home-made resonance-enhanced multi-photon ionization and mass-analyzed threshold ionization spectrometers combined with quantum chemical calculations. They summarized the influences of multi-halogen effects, substituent effects and conformational isomerism effects on molecular properties and spectroscopy, supplying important information on the characteristics of excited and ionic states of molecules. These achievements not only deepen our understanding of the microscopic mechanism of chemical reactions, but also provide an important theoretical basis for their practical applications in the fields of atmospheric chemistry, environmental chemistry, biochemistry and material sciences.

Cite this article

Min Cheng , Lijuan Zhang , Xiling Xu , Hong Gao , Weijun Zheng . The Photodissociation and Photoionization Dynamics of Some Important Small Molecules[J]. Progress in Chemistry, 2024 , 36(12) : 1830 -1848 . DOI: 10.7536/PC241110

1 Introduction

Photon-molecule interactions, in addition to exciting molecules to excited states, can also lead to photodissociation or photoionization of the molecules. In-depth studies on the photodissociation and photoionization processes of molecules are very important in the field of molecular reaction dynamics.
Molecular photodissociation is one of the key reaction types in chemical reactions. In the process of molecular photodissociation, the parent molecule first absorbs one or more photons to transition to an excited state, and then dissociates into two or more fragments. These fragments are mostly radicals or atoms, with radical fragments having a certain internal energy, exhibiting high reactivity, and serving as the initiating reactants for many photochemical reactions. Molecular photodissociation is also known as a semi-collision reaction. The bimolecular (atomic) elementary reaction involves two molecules first colliding and temporarily binding together to form a transition state structure, which then decomposes into two new molecules; the process of collision, binding, and decomposition in molecular photodissociation is similar to the latter part of bimolecular reactions. Therefore, studying the dynamics of molecular photodissociation not only helps in understanding the characteristics of the electronic excited states of molecules and the dynamic processes after excitation but also plays a significant role in gaining a deeper understanding of the mechanisms of bimolecular reactions[1]. By investigating the dynamics of multi-channel photodissociation reactions, the branching ratios of products can be related to reaction conditions, especially the energy of the dissociating photons, thereby providing important theoretical basis for controlling chemical reactions. Studying the dynamics of molecular photodissociation is also of great significance in industrial and agricultural production and atmospheric environmental fields. For example, chlorofluorocarbons (CFCs), which are a very important class of refrigerant molecules in human production and daily life, upon photodissociation produce halogen atoms that lead to ozone depletion in the stratosphere. The photodissociation of halocarbons released in the ocean also exacerbates the destruction of the ozone layer, with the photodissociation cross-sections of halogenated compounds affecting the photochemical parameters and theoretical models of ozone layer destruction[2-4]; researching the photodissociation of halocarbon molecules has important guiding significance for addressing atmospheric pollution and protecting the ozone layer. Molecular photodissociation is a critical reaction in atmospheric chemistry, where solar ultraviolet radiation can cause many atmospheric molecules to undergo photodissociation, facilitating various chemical reactions in the atmosphere. Molecular photodissociation also plays a crucial role in areas such as the utilization of solar energy.
Photofragment Translational Spectroscopy is a very important method for studying the dynamics of molecular photodissociation reactions[5]. This method can accurately measure the translational energy distribution and spatial angular distribution of photodissociation fragments, obtaining various dissociation channels and channel branching ratios of photodissociation reactions, the quantum state populations and available energy distributions of fragments in each channel. Further analysis can reveal the influence of non-adiabatic interactions such as quantum state interference, resonance, and conical intersections on the molecular photodissociation process, playing a significant role in deeply understanding the physicochemical mechanisms of molecular photodissociation. The team led by Academician Qihong Zhu was one of the first research groups in China to carry out studies on the dynamics of molecular photodissociation. They independently developed several photofragment translational spectroscopy instruments and used these devices to conduct a series of photodissociation dynamics studies on small molecules, achieving research results with important scientific significance.
When molecules absorb photons and transition to an excited state, in addition to the possible photodissociation process mentioned above, direct ionization or further ionization after absorbing additional photons may also occur. The study of molecular photoionization spectra can reveal detailed information about molecules in excited states and ionic states, making it an important area of research in the field of molecular reaction dynamics. Research on molecular photoionization not only promotes the development of traditional fundamental disciplines such as physics, chemistry, biology, and astronomy, but also holds broad prospects in many applied disciplines. Studying the microscopic mechanisms of molecular photoionization can accurately measure the energy levels of molecular excited states and precisely describe their structures, obtaining detailed information such as electronic structure, nuclear vibrations, and rotations; it can also uncover the dynamic processes of interactions between molecules and light, achieving precise state-to-state dynamic studies of chemical reactions[6-7].
Molecular electronic excited states exhibit different characteristics from the molecular ground state in terms of configuration, energy, and polarity, which endow them with higher chemical reactivity. By utilizing methods such as UV-Vis absorption spectroscopy, resonance-enhanced multiphoton ionization (REMPI) spectroscopy[8-9], and laser-induced fluorescence (LIF) spectroscopy[10], information about molecular electronic excited states can be obtained. REMPI spectroscopy is one of the commonly used methods for studying excited state spectra. Its principle involves molecules absorbing appropriately tuned laser photon energy to resonantly transition to a specific excited state, then absorbing one or more photons to undergo ionization. By measuring the generated ion signals, high-resolution, high-signal-to-noise ratio spectra of the electronic excited states can be obtained. In processes such as flame combustion, interstellar medium evolution, life activities, and plasma reactions, highly chemically reactive ionic species play a crucial role. In-depth research on ionic state spectra can provide key information such as spatial configuration and energy level structure, which helps us to better understand the important roles ions play in chemical reaction networks. Spectroscopic techniques for studying molecular ionic states include zero-kinetic-energy electron spectroscopy (ZEKE)[11], photoelectron spectroscopy (PES)[12], and mass-resolved threshold ionization spectroscopy (MATI)[13]. The team led by Academician Zhu Qihui has independently built REMPI and MATI spectrometers and employed a variety of IR-UV combined spectroscopic techniques to conduct a series of studies on the photoionization micro-mechanisms of benzene derivatives and their clusters, achieving significant progress.
This paper will systematically introduce the achievements made by Academician Zhu Qihuo's team in the research on the microscopic mechanisms of photodissociation and photoionization of some important small molecules. The main contents include:
● Study on the Photodissociation Dynamics of Important Small Molecules
Development of a Photofragment Translational Spectrometer
Study on the Photodissociation of Halomethanes
Study on the Photodissociation of Haloethanes
Studies on the photodissociation of other halocarbons
● Research on the photoionization dynamics of important small molecules
REMPI/MATI and IR-UV spectroscopy experimental setup
REMPI and MATI Spectroscopy Studies of Benzene Derivatives
Study on Conformational Isomerism of Benzene Derivatives
Research on the IR/UV Laser Spectroscopy of Benzene Derivatives and Their Clusters
On this basis, this paper summarizes and prospects the related research work.

2 Investigation of Photodissociation Dynamics of Important Small Molecules

2.1 Development of a Photofragment Translational Spectrometer

In 1985, Academician Zhu Qihao and others successfully developed the first domestic molecular beam rotatable laser photolysis fragment translational energy spectrometer (Figure 1)[14-15], which passed the evaluation by the Chinese Academy of Sciences in 1986 and became one of the top ten science and technology news stories in China for that year, winning the First Prize of Science and Technology Progress of the Chinese Academy of Sciences in 1987. The instrument is mainly composed of a beam source chamber, a reaction chamber, a detection system, a vacuum system, an electronic measurement system, and a control system. It measures the translational energy of photolysis fragments by recording the time-of-flight (TOF) of the fragments to the detector and can also measure the spatial angular distribution of the fragments, thereby obtaining key kinetic information on molecular photodissociation. The use of a rotatable beam source is the core and key technology of this instrument.
图1 分子束可转动式激光裂解碎片平动能谱仪实物图(其中左为朱起鹤先生、右为黄寿龄先生)和原理图[15]

Fig. 1 The picture (left: Prof. Qihe Zhu; right: Prof. Shouling Huang) and schematic diagram of photofragment translational spectrometer with rotatable pulsed molecular beam [15]

In 2006, Academician Zhu Qihui's team successfully developed a micro laser photofragment translational spectroscopy instrument[16](Figure 2). This instrument is characterized by low cost, simple and compact structure, high velocity resolution (~1.2%), good signal-to-noise ratio, and does not require a very high vacuum level; its core technology is short-distance flight (the flight distance of fragment ions is only 25~50 mm). In the experiment, a cylindrical grid with a curvature radius equal to the Newton sphere radius of the arriving ions was used to receive the ions to be measured, which enhanced the detection signal intensity and further improved the velocity resolution. Depending on the characteristics of the research system, this spectrometer can operate in three modes: no accelerating electric field, weak accelerating electric field, and pulsed weak electric field acceleration, and it can use single, dual, or triple laser beams, making it suitable for the study of photodissociation dynamics of ground-state or vibrationally excited parent molecules.
图2 微型激光光解碎片平动能谱仪无场工作模式原理图[16]

Fig. 2 The principle scheme of mini-TOF photofragment translational spectrometer working under the accelerated electric field free model [16]

In 2016, they successfully developed a low-voltage accelerated ion velocity imaging small photofragment translational energy spectrometer by adopting a scheme of weak electric field acceleration for short-distance flight[17](Figure 3). The core technology lies in the ingenious combination of the design concept of low-voltage acceleration and short-distance flight with the traditional ion lens focusing method. A lower voltage (30~150 V) was used instead of the traditional high voltage (650~4000 V) to accelerate and focus the photofragment ions, with the ion flight distance being only 12 cm. This instrument is characterized by its compact structure, resistance to interference, ease of operation, and high velocity resolution (~0.8%). Low-voltage acceleration allows the backward photofragment ions to have a longer turnaround time, meaning that the Newton sphere of the ions extends significantly along the TOF axis, allowing slices of the same pulse width to achieve better results and obtain higher-resolution images. Subsequently, they employed a tunable high-resolution vacuum ultraviolet laser generated through rare gas two-photon resonance four-wave mixing as the light source for the instrument, further expanding its research scope.
图3 低电压加速离子速度成像式小型光解碎片平动能谱仪工作原理图[17]

Fig. 3 The schematic diagram of mini time-sliced ion velocity map imaging photofragment translational spectrometer using low voltage acceleration [17]

2.2 Photodissociation Studies of Halomethanes

Using the high-performance photofragment translational spectroscopy instrument developed by themselves, the Zhu Qihui team has carried out a series of studies on the photo-dissociation dynamics of small gas-phase molecules, with halogenated hydrocarbons being the most representative. Methyl iodide CX3I (X = H, D, F), as a quasi-linear molecule, serves as a model system for studying the photo-dissociation dynamics of polyatomic molecules. In the past few decades, its photo-dissociation in the A absorption band (generally located in the ultraviolet region) has received considerable attention from both experimental and theoretical research[18-34].
In the A band, methyl iodide CH3I molecule has two photodissociation channels:
CX3I $\xrightarrow{\mathrm{~h} \nu}$ CX3 (v) + I*(2P1/2) ( I* channel )
CX3 (v) + I(2P3/2) ( I channel )
CX3I molecules undergo σ*$\leftarrow$n transition after absorbing photons in the A band, which leads to the dissociation of the C-I bond. The electronic states involved in the absorption and dissociation in the A band mainly include the repulsive states 3Q1, 3Q0, and 1Q1, with the 3Q0$\leftarrow$X transition being the main absorption in the A band[35]. Among these, the absorption from the ground state to the 1Q1 and 3Q1 states is a vertical transition, leading to direct dissociation products of CX3 and I(2P3/2), known as the ground-state I channel; the absorption from the ground state to the 3Q0 state is a parallel transition, leading to direct dissociation products of CX3 and I*(2P1/2), referred to as the excited-state I* channel. There exists a conical intersection between the repulsive potential energy surfaces of 3Q0 and 1Q1; after the molecule initially transitions to these two states upon photon absorption, it may experience crossing (Curve crossing, abbreviated as cc, as shown in Figure 4) between the potential energy surfaces during the dissociation process. Therefore, there are five photodissociation pathways for CX3I in the A band:
图4 CX3I光解的势能曲线示意图

Fig. 4 The scheme for the involved potential energy curves in the photodissociation of CX3I

Researchers often use the one-dimensional Landau-Zener formula to evaluate the dynamic behavior in this crossing region, and the energy Ec of the conical intersection can be estimated from the experimentally measured traversal probability (Pcc)[22,36]. After CX3I absorbs a photon and breaks the bond, the remaining energy (available energy Eavl) will be distributed between the translational energy of the fragments and the internal energy of CX3 (Eint). Due to the high symmetry of the fragment CX3, it is generally believed that the most likely vibrations to be excited during the bond-breaking process are the umbrella vibration and the symmetric stretching vibration of C-X, while its rotational excitation is relatively weak. Through the study of the vibrational state-resolved photodissociation fragment translational energy spectrum, key dynamic information such as the vibration modes, vibrational quantum state populations, the distribution ratio of internal and translational energies, and the spatial angular distribution of the products can be determined, which plays an important role in deeply understanding the physicochemical mechanisms of molecular photodissociation processes.

2.2.1 Methyl Iodide (CH3I)3I)的光解离研究

CH3I is the simplest system in the CX3I series of molecules, and its photodissociation process has received the most attention from experimental and theoretical studies. The CH3 group in the parent molecule has a trigonal pyramidal structure, but the equilibrium configuration of the CH3 radical fragment is planar, with its umbrella vibration (Umbrella, ν2) being considered the most likely and strongly excited vibration during photodissociation. Due to the lack of vibrationally resolved results, early experimental studies had disagreements over the highest populated vibrational state of the CH3 fragment produced by the photodissociation of CH3I. In 1985, Qihuo Zhu et al.[14] used a self-developed rotatable laser photolysis molecular beam translational energy spectrometer to conduct a photodissociation experiment on CH3I at 248 nm, obtaining a time-of-flight spectrum (TOF) for the iodine atom fragments (Figure 5). This spectrum showed two distinct peaks corresponding to the I* and I dissociation channels, and the quantum yield Φ(I*) was measured. Analysis indicated that in the I* channel, the highest populated peak of the CH3 umbrella vibration excitation was at v2 = 2, while in the I channel, it was at v2 = 4. The experiment also measured the C-I bond dissociation energy of the CH3I molecule as D0 = 56 ± 1 kcal/mol. In subsequent experiments, spectra with better signal-to-noise ratios were obtained, and the ratio of internal energy excitation of the fragments to the available energy in the photodissociation process was determined, calculating an anisotropy parameter β = 1.85 for the spatial angular distribution of the photodissociation fragments, further confirming that both the I* and I channels at 248 nm originated from the parallel transition dissociation of the 3Q0 state[3739]. In 1986, Mr. Qihuo Zhu visited Professor Yuan Tseh Lee's laboratory at the University of California, Berkeley, and repeated the photodissociation experiment of CH3I, acquiring vibrationally resolved translational energy spectra of the photodissociation fragments. It was experimentally confirmed that at 248 nm photodissociation, the highest populated vibrational excitation peak of the CH3 fragment in the I* channel was v2 = 0, and in the I channel, it was v2 = 1; moreover, weak excitation of the C-H symmetric stretching vibration (ν1) of the CH3 radical was observed (the research results were not publicly published). Subsequent experimental and theoretical studies supported this conclusion[24,25,40-41].
图5 CH3I在248 nm光解时碎片碘离子飞行时间谱图[14]

Fig. 5 The TOF spectrum of I+ for the photodissociation of CH3I at 248 nm[14]

Using a self-developed high-resolution miniature laser photolysis fragment translational energy spectrometer, Zhu Qihao et al.[4243] systematically carried out the study of photodissociation dynamics of CH3I across the entire A band. At 277, 280, 297, and 304 nm in the middle and red end of the A band[42], vibrational-resolved photofragment translational energy spectra (Figure 6) were obtained, yielding a series of photodissociation kinetic parameters (Table 1). Within the 277-304 nm range, only the excitation of the umbrella vibration of the CH3 radical was observed in the I* channel, with its highest populated peak always remaining at v2 = 0. Although the internal energy of CH3 decreases as the photon energy of the dissociation laser decreases, the ratio of internal energy to available energy Eint/Eavl remains approximately 0.03, essentially unchanged. In the I channel, where the available energy is much greater than in the I* channel, weaker excitation of the symmetric stretching vibration of CH (ν1) was observed, with the maximum reaching v1 = 1. At shorter wavelengths such as 277.87 and 279.71 nm, the highest populated peak of CH3 vibrational excitation is at (v1 = 0, v2 = 1). At 298.23 and 304.67 nm, the highest populated peak of CH3 vibrational excitation is at (v1 = 0, v2 = 0). In this wavelength range, when the symmetric stretching vibration of CH is excited, the highest populated peak of the umbrella vibration is always at (v1 = 1, v2 = 0). As the wavelength increases, the available energy decreases, leading to a decrease in the internal energy of the CH3 fragment, and the ratio Eint/Eavl drops from around 0.09 at 277 nm to around 0.06 at 304 nm. The total population of v1 = 1 decreases from around 0.20 at 277 nm to around 0.06 and 0.03 at 304 nm. At the vertical transition of 304.67 nm (3Q1 channel), the internal energy of the CH3 fragment is slightly lower than that of the corresponding parallel transition (1Q1 $\leftarrow$ 3Q0 channel). By measuring the anisotropy parameters of the spatial angular distribution of photofragments, it was found that within the 277-304 nm range, the I* channel all comes from contributions via the 3Q0 potential surface; for dissociation wavelengths < 280 nm, the I channel all comes from contributions via the 1Q1 $\leftarrow$ 3Q0 dissociation pathway, and for dissociation wavelengths > 280 nm, the contribution from 3Q1 gradually increases, consistent with the composition of the A band absorption spectrum. At 277 and 304 nm, the branching ratios, I* quantum yield Φ(I*), and potential curve crossing probability Pcc for each channel were also measured, as shown in Table 1. According to the one-dimensional Landau-Zener formula, the energy Ec of the conical intersection point between the 1Q1 and 3Q0 potential surfaces was calculated to be approximately 32 740 cm−1.
图6 CH3I在277 nm光解时碎片平动能谱图[42]

Fig. 6 The photofragment translational spectra of CH3I at 277 nm[42]

表1 CH3I光解的动力学参数

Table 1 The dynamical parameters for the photodissociation of CH3I

λ(nm) Pathway Channel Fraction Φ(I*) Pcc Eint/Eavl β Ref.
225 3Q0 I* 0.09 0.12 0.129 1.25 43
3Q0$\leftarrow$1Q1 I* 0.03 0.08 0.104
1Q1 I 0.34 0.134 0.84
1Q1$\leftarrow$3Q0 I 0.54 0.86 0.159
248 3Q0 I* 0.74 0.74 0.26 0.125 1.85 14, 37-39
1Q1$\leftarrow$3Q0 I 0.26 0.173
277 3Q0 I* 0.59 0.59 0.028 1.93 42
1Q1$\leftarrow$3Q0 I 0.41 0.41 0.087 1.91
279.71 1Q1$\leftarrow$3Q0 I 0.088 1.92 42
281.73 3Q0 I* 0.029 1.92 42
295.91 3Q0 I* 0.030 1.92 42
298.23 1Q1$\leftarrow$3Q0 I 0.065 1.70 42
304 3Q0 I* 0.05 0.05 0.029 1.88 42
1Q1$\leftarrow$3Q0 I 0.74 0.94 0.062 1.35
3Q1 I 0.21 0.057
In the blue end of the A band of CH3I, its photodissociation exhibits significant differences from the middle and red ends. Zhu Qihao et al.[43]measured the translational energy spectrum of photodissociation fragments at 225 nm, which showed good vibrational state resolution in both I and I* channels, through which the vibrational state population of CH3fragments was obtained. The experiment found that regardless of whether in the I* or I channel, the symmetric stretching vibration of C-H in the CH3fragment was strongly excited. In the I* channel, the ν 1vibration of the CH3fragment was maximally excited to v 1 = 1; in the I channel, the ν 1vibration of the CH3fragment was maximally excited to v 1 = 2. The peaks on the spectra exhibited a clear clustering phenomenon, indicating that the umbrella vibration of CH3is not easily populated to v 2 = 4 and above; only the I channel from the 1Q1$\leftarrow$3Q0dissociation path did not show obvious clustering. The ratio of internal energy to available energy in the I* channel is significantly lower than that in the I channel. It was also found that the internal energy excitation of CH3generated by parallel transitions initially excited to the 3Q0state is higher than that generated by vertical transitions initially excited to the 1Q1state, possibly because the 3Q0state has a relatively low energy, and when the parent molecule is excited to the 3Q0state, the umbrella and C-H stretching vibrations of the CH3group are more strongly excited, exhibiting a certain degree of vibrational adiabatic process. The branching ratios for the four dissociation pathways 3Q0, 3Q0 1Q1, 1Q1, and 1Q1 3Q0were measured as 0.090.030.340.54, with the quantum yield of I* Φ(I*) = 0.12, β = 1.25 for the I* channel, and β = 0.84 for the I channel. The experimental crossing probability P ccfor 1Q1 3Q0was 0.86, and for 3Q0 1Q1it was 0.08, proving that the one-dimensional Landau-Zener equation is not suitable for predicting the crossing probabilities between potential energy surfaces during the photodissociation of CH3I at the blue end of the A band, as the one-dimensional Landau-Zener equation predicts equal P ccfor the crossings of 3Q0 1Q1and 1Q1 3Q0. The P cc= 0.08 for 3Q01Q1indicates that it is difficult for CH3I molecules in the 1Q1state to cross over to the 3Q0state.
Zhu Qihao et al. also observed the photodissociation signal of CH3I parent molecules in a vibrationally excited state, which were excited by thermal excitation leading to C—I stretching vibrations (ν3ʹ). It was found that the proportion of photodissociation signals from the ν3ʹ vibration of the parent molecule gradually increased from 277 to 304 nm, with the thermal spectral effect of the I channel being lower than that of the corresponding I* channel; and after the vibrationally excited parent molecule absorbs a photon, its 1Q1 ← 3Q0 crossing probability Pcc decreases[42]. The photodissociation of vibrationally excited parent molecules often exhibits significant differences compared to that of ground-state parent molecules[44-46], making it an important subject in state-to-state reaction dynamics and a crucial foundation for controlling chemical reactions in a favorable direction. In atmospheric and environmental chemistry, under natural conditions, there are many reactant molecules populating vibrational excited states, where the impact of initial vibrational excited state molecular photodissociation is more pronounced. To further study the photodissociation dynamics of vibrationally excited CH3I molecules, Zhu Qihao et al. used infrared lasers to pre-excite the C-H symmetric stretching vibration to v1ʹ = 1, and then studied its photodissociation dynamics in the ultraviolet region. While previous researchers mostly focused on the influence of bond-breaking vibrational excitation on photodissociation, Zhu Qihao et al. chose to explore other dimensions of vibrational excitation, which is of great significance for understanding the transfer and transformation of energy during the photodissociation process.
At 277 nm[47], by detecting the iodine atoms and CH3 fragment signals from the photodissociation of CH3I (v1ʹ = 1), it was found that the impact of initial v1ʹ = 1 vibrational excitation on the photodissociation cross section of CH3I is relatively small, with most of the initial v1ʹ=1 vibrational energy being retained in the CH3 fragments (Figure7). The umbrella vibrational excitation of CH3 (v1 = 1, v2) is similar to that of the ground state CH3I molecule when it undergoes photodissociation near 277 nm; and compared to the experimental results of the photodissociation of the ground state CH3I at shorter wavelengths (225, 248 nm), it can be clearly confirmed that the v1 = 1 vibrational excitation of the CH3 fragment originates from the retention of the initial vibrational excitation, rather than due to an increase in dissociation energy. At this point, the initial vibrational excitation plays a role akin to a "spectator" in the photodissociation, and the process can be considered approximately vibrationally adiabatic. However, at 304 nm, the photodissociation of CH3I (v1ʹ = 1) is significantly different from that at 277 nm[48]. By detecting the iodine atom fragments produced during the photodissociation, it was found that the photodissociation cross section of CH3I (v1ʹ = 1) is about 2.67 times larger than that of the ground state methyl iodide, showing a significant enhancement. The quantum yield Φ(I*) = 0.24 for the photoproduct I* is much higher than that for the photodissociation of the ground state methyl iodide; its 1Q13Q0 crossing probability Pcc = 0.72 is lower than the Pcc value for the photodissociation of the ground state methyl iodide at this wavelength[42,48]. By detecting the methyl and iodine fragments, the vibrational excitation modes and vibrational state populations of the CH3 (ν1, ν2) fragments produced during the photodissociation of CH3I (v1ʹ = 1) were determined; it was found that in the I channel, the proportion of CH3 fragments excited to v1 = 1 is much higher than in the I* channel, indicating that in the I* channel, the initial vibrational excitation of the parent molecule is more difficult to retain in the fragments. Comparing the photodissociation data at 277 nm, it was found that at low photon energy photodissociation, the initial vibrational excitation of the parent molecule is more difficult to retain in the dissociation fragments. Whether at 277 or 304 nm, the 1Q13Q0 crossing probability Pcc in the photodissociation of CH3I (v1ʹ = 1) is lower than that in the photodissociation of the ground state CH3I molecule, which also leads to a higher quantum yield Φ(I*) for the dissociation product I*. Currently, both experimental and theoretical studies on the photodissociation of vibrationally excited CH3I molecules are still limited and not systematic, making it impossible to comprehensively and accurately explain the kinetic mechanisms, thus further in-depth research is needed.
图7 CH3I(v1ʹ=1)在277 nm光解时检测CH3(v1=1, v2=0)获得的平动能谱图[47]

Fig. 7 The photofragment translational spectra of CH3I(v1ʹ=1) via the detection of CH3(v1=1, v2=0) at 277 nm[47]

Zhu Qihet al also studied the photodissociation of the iodomethane dimer (CH3I)2 to produce (CH3)2I + I, finding that the translational energy of the produced fragment products was very small, revealing that van der Waals forces play an important role in the photodissociation process[49].

2.2.2 Photodissociation Study of I)3I)的光解离研究

When the H atom in CH3I is replaced by F, the molecular symmetry remains unchanged. CF3I, during photodissociation at the A band, involves potential energy surfaces and dissociation pathways consistent with those of CH3I. However, the equilibrium configuration of the dissociation fragment CF3 radical in its ground state is a trigonal pyramidal structure, which significantly differs from the planar structure of the CH3 radical in its ground state; as F atoms are much heavier than H atoms, the vibrational modes and frequencies of the CF3 radical also differ greatly from those of CH3. Based on theoretical calculations by Clary[33] and Bowman et al.[50], this paper identifies the vibration of the CF3 radical at 701 cm−1 as the symmetric stretching (ν1) mode of CF, and the vibration at 1086 cm−1 as the umbrella ν2 mode of the CF3 fragment (this identification may be opposite to that in some literature, so readers should distinguish carefully). Clary[33] and van Veen et al.[32] predicted through theoretical calculations that, due to the presence of vibrational interactions on the excited state potential energy surface during the A-band photodissociation of CF3I, the CF3 fragment would exhibit significant dual-mode vibrational excitation (ν1 and ν2), but experimental studies have never been able to observe this phenomenon.
Zhu Qihao et al. conducted a photodissociation study of CF3I at 248 nm using a molecular beam rotatable laser photolysis fragment translational energy spectrometer[15], obtaining vibrationally resolved photofragment translational energy spectra in the I* channel. By identifying the spectra, they concluded that the peak vibrational population of the CF3 fragment is at v1 = 5, and based on this, calculated the C—I bond dissociation energy D0 = 53.8 ± 0.6 kcal/mol, thus determining the ratio of internal energy to available energy Eint/Eavl for the photofragments in the I* channel. They also investigated the photofragment translational energy spectra in the I* channel at 281.73 nm[51], identifying the peak vibrational population of the CF3 fragment at v1 = 2, with its Eint/Eavl being approximately 0.21. The spectra also showed a peak suspected to originate from bimodal vibrational excitation of the CF3 fragment, but due to the limitations of spectral resolution and signal-to-noise ratio, it could not be clearly confirmed.
To further explore the dual-mode vibrational excitation of CF3 fragments, Zhu Qihao et al. used a high-resolution micro laser photolysis fragment translational energy spectrometer to systematically study the photodissociation dynamics of CF3I across the entire A band[52-54] (Table2). At 248 nm[53], high-vibration-state-resolved I* channel photofragment translational energy spectra (Figure8) were obtained, where the spectral peaks could be clearly divided into two alternating groups with stronger and slightly weaker intensities. The energy interval ΔEV between the higher peaks was about 700 cm−1, which matches the vibration frequency of the ν1 symmetric stretching mode of the CF3 radical, and was identified as the (v1 = 0 ~ 7, v2 = 0) vibrational states of the CF3 fragment, with the highest vibrational population peak at (v1 = 4, v2 = 0). The slightly weaker group of spectral peaks, located in the middle between two (v1, v2 = 0) peaks, had an energy gap consistent with the difference in vibration frequencies for (v1, v2 = 1), and was identified as the (v1 = 0 ~ 5, v2 = 1) states of the CF3 fragment. This was the first clear experimental observation of dual-mode vibrational excitation of the CF3 radical during the photodissociation of CF3I. In the I* channels at 266 and 277 nm, two alternating groups of vibrational spectral peaks, also attributed to the dual-mode vibrational excitation of the CF3 fragment, were observed; the highest vibrational population peaks were at (v1 = 2, v2 = 0) for 266 nm photolysis and at (v1 = 1, v2 = 0) for 277 nm. As the wavelength increased and photon energy decreased, the population of v2 = 1 vibrational excitation significantly decreased. During photolysis at 248, 266, and 277 nm, the ratios of internal energy to available energy Eint/Eavl were 0.220, 0.145, and 0.108, respectively, and the ratios of ν2 umbrella vibrational excitation ∑P(v1, v2 = 1)/∑P(v1, v2) were 0.41, 0.35, and 0.28, all of which rapidly decreased as the available energy decreased. Due to resolution limitations, it is not ruled out that the ν2 umbrella vibration of the CF3 fragment might be excited to even higher quantum states. When the photolysis photon energy was further reduced to around 304 nm[52], many closely spaced spectral peaks were observed on the translational energy spectrum in the I* channel. Pressure experiments revealed that these peaks originated from the ν1 symmetric stretching vibrational excitation (v2 = 0) of CF3 fragments produced by the photolysis of vibrationally ground state or C-I stretching vibrationally excited CF3I (v3ʹ = 0, 1, 2) molecules, with no clear ν2 vibrational excitation of the CF3 fragment observed. At this point, the highest vibrational population peak of the CF3 fragment was at v1 = 1. The anisotropy parameter β of the I* product angular distribution approached 2, indicating direct dissociation via the 3Q0 potential energy surface.
表2 CF3I光解的动力学参数

Table 2 The dynamical parameters for the photodissociation of CF3I

λ(nm) Pathway Channel Fraction Φ(I*) Pcc Eint/Eavl β Ref.
238 3Q0 I* 0.664 0.738 0.266 1.70 54
3Q0$\leftarrow$1Q1 I* 0.074 0.294
1Q1 I 0.178 -0.04
1Q1$\leftarrow$3Q0 I 0.084 0.112
248 3Q0 I* 0.220 1.85 53
266 3Q0 I* 0.90 0.90 0.145 1.86 53
1Q1$\leftarrow$3Q0 I 0.07 0.07 0.202 1.03
3Q1 I 0.03
277 3Q0 I* 0.87 0.87 0.108 1.86 53
1Q1$\leftarrow$3Q0 I 0.086 0.09 0.154 0.98
3Q1 I 0.044
281.73 3Q0 I* 0.21 51
304 3Q0 I* 0.06 0.06 0.12 1.69 52, 57
1Q1$\leftarrow$3Q0 I 0.15 0.71 0.18 -0.45
3Q1 I 0.79 0.15
For the photodissociation I channel, no significant signal was observed at 248 nm[53]. This is mainly because the photon energy at 248 nm is far from the absorption centers of the 1Q1 and 3Q1 states, and CF3I molecules cannot absorb photons to transition to these two states; while excitation to the 3Q0 state results in fast photodissociation, making the probability of dissociation through the 1Q13Q0 transition very small. In the I channel photodissociation at 266 and 277 nm, experiments have measured the translational energy spectra of photofragments with vibrational state resolution[53]. The spectra reveal that in the I channel, the bimodal vibration of the photofragment CF3 is also simultaneously excited, with the highest vibrational peak populations being (v1 = 5, v2 = 0) for 266 nm photodissociation and (v1 = 3, v2 = 0) for 277 nm photodissociation. The ratios of internal energy to available energy Eint/Eavl are 0.202 and 0.154, respectively, and the excitation ratios of the umbrella mode vibration ∑P(v1, v2 = 1)/∑P(v1, v2) are 0.47 and 0.46, respectively. These kinetic parameter values are all higher than those in the I* channel, due to the larger available energy in the I channel. Near 304 nm, the I channel has both a direct dissociation path from the 3Q1 state (vertical transition) and a dissociation path via the 1Q13Q0 transition (parallel transition). Translational energy spectra with good vibrational state resolution were obtained on both dissociation paths[52], showing strong excitation of the ν1 symmetric stretching vibration of the CF3 fragment, without noticeable bimodal vibration excitation of the CF3 fragment. On these two dissociation paths, the highest vibrational state population of the CF3 fragment is v1 = 1; however, the ratio of internal energy to available energy Eint/Eavl for the 1Q13Q0 dissociation path is slightly higher than that for direct dissociation from the 3Q1 state, consistent with CH3I photodissociation at 304 nm[42]. For the I channel photodissociation from 266 to 304 nm, the anisotropy parameters β of the product spatial angular distribution are 1.03, 0.98, and -0.45, indicating that the proportion of products from direct dissociation of the 3Q1 state gradually increases as the photon energy decreases. At 277 nm photodissociation, the quantum yield Φ(I*) of I* was measured to be 0.87, with branching ratios for the three dissociation paths 3Q0, 1Q13Q0, and 3Q1 being 0.870.0860.044, and the crossing probability Pcc for the potential energy surface 1Q13Q0 being 0.09; at 304 nm photodissociation[57], Φ(I*) = 0.06, with branching ratios for the dissociation paths 3Q0, 1Q13Q0, and 3Q1 being 0.060.150.79, and the crossing probability Pcc for 1Q13Q0 being 0.71.
图8 CF3I在248 nm光解的平动能谱图[53]

Fig. 8 The photofragment translational spectra of CF3I at 248 nm[53]

In the blue end of the A absorption band, there are four photodissociation pathways for CF3I molecules, namely, the I* channel via 3Q0 and 3Q01Q1, and the I channel via 1Q1 and 1Q13Q0. Zhu Qihao et al. conducted a photodissociation study of CF3I at 238 nm[54]; only in the 3Q0 dissociation pathway was a vibrationally resolved translational energy spectrum of photofragments obtained; the signals from the other three pathways were too weak to yield spectra with high resolution. In the I* channel of the 3Q0 dissociation pathway, bimodal vibrational excitation of the CF3 fragment was also observed, with the highest vibrational state population being (v1 = 6, v2 = 0), Eint/Eavl = 0.266, and ∑P(v1, v2=1)/∑P(v1, v2) = 0.48. They measured the branching ratios of the four dissociation pathways, and the anisotropy parameter β values for the angular distributions of products in the I* and I channels were 1.70 and −0.04, respectively. The crossing probability Pcc of the potential energy surface 1Q13Q0 was found to be 0.112, while that for 3Q01Q1 was 0.294, showing a significant difference; this revealed that when dealing with the very important non-adiabatic interactions between potential energy surfaces at conical intersections, the direction of crossing should be considered as an important factor.

2.2.3 Methyl Chloride Iodide (I-CH)2Cl)的光解离研究

When one H in the CH3I molecule is replaced by another halogen atom, the C3v symmetry of the dihalo I-CH2X molecule is disrupted, and the dissociation dynamics also show significant differences. After dissociation, the eccentric equilibrium configuration of the CH2X fragment differs from the planar configuration of the methyl group or the trigonal pyramidal configuration of the CF3 radical. The most easily excited vibrational mode may no longer be the umbrella vibration or symmetric stretching vibration, and under the influence of bond-breaking recoil, the rotational excitation of the fragments will also be more strongly induced. Dihalo-methanes have important implications in atmospheric chemistry and marine chemistry[3,55], but there are relatively few studies on their gas-phase photodissociation dynamics.
Zhu Qihao et al.[37,39] used a molecular beam rotatable laser photolysis fragment translational energy spectrometer to conduct a study on the photodissociation of I-CH2Cl molecules at 248 nm, observing the C—I bond breaking process. The quantum yield Φ(I*) = 0.70 for the I* channel was measured, and the ratio of internal energy to available energy Eint/Eavl for the photofragments in both the I* and I channels was approximately 0.69. However, due to resolution limitations, it was not possible to obtain vibrationally resolved photofragment translational energy spectra, making it impossible to identify the vibrational excitation modes of the CH2Cl fragments.
Using a high-resolution miniature laser photofragment translational spectroscopy, they conducted the photodissociation study of I-CH2Cl molecules at 277 and 304 nm[56], obtaining photofragment translational energy spectra with clear vibrational state resolution (Figure 9). Through the analysis of well-resolved I* channel spectra, it was identified that the maximum vibrational populations of the CH2Cl fragment during photodissociation at 277 and 304 nm were v = 3 and 2, respectively, and it was calculated that the products had very high rotational excitation (average ER/ET ≈ 0.72, with rotational excited states reaching up to J = 95 and 79), and the energy distribution ratio among translational, rotational, and vibrational energies was about 0.48:0.34:0.18. By the energy spacing of vibrational peaks on the spectrum and considering different rotational excitations, it was confirmed that the C—Cl stretching vibration of the CH2Cl fragment was the most likely vibrational mode to be excited during the photodissociation process. In the I channel, the ratio of product rotational to translational energy ER/ET was slightly less than in the I* channel, with the energy distribution ratio among translational, rotational, and vibrational energies being approximately 0.45:0.31:0.24. A large number of vibrational peaks appeared on the translational energy spectrum, attributed to the excitation of C—Cl stretching vibrations from the CH2Cl fragment. Unlike in the I* channel, the vibrational peaks on the I channel's translational energy spectrum clearly showed components from two dissociation pathways (Figure 9), and their contributions changed with different photon energies. The component with smaller translational energy (left side of the spectrum) was assigned to the I channel produced by direct dissociation after transitioning to the 3A' state upon absorbing a photon, while the component with larger translational energy (right side of the spectrum) was assigned to the I channel produced after transitioning to the 4A' state and then dissociating through the conical intersection 5A'$\leftarrow$4A'. When the photon energy decreased from 277 to 304 nm, more I-CH2Cl molecules absorbed photons and transitioned to the lower-energy 3A' state, leading to a higher population of the component with smaller translational energy on the translational energy spectrum. It was found that the anisotropy parameter β values for both I* and I channel fragment spatial angular distributions were close to 2, indicating that both reaction channels were mainly activated via parallel transitions. The quantum yields Φ(I*) for the I* channel during photodissociation at 277 and 304 nm were measured to be 0.36 and 0.26, respectively. As the photon energy for photodissociation decreased, the speed at which excited-state molecules passed through the potential crossing region decreased, and the time increased, resulting in more excited-state molecules crossing from 4A' to 5A' to produce ground-state I, and simultaneously, more molecules being excited to the 3A' state, which directly produces the ground-state I; all these factors would lead to a decrease in Ф(I*). Compared to the photodissociation of CX3I, the photodissociation of dihalomethanes shows significant differences, and the current research has not systematically revealed its photodissociation dynamics mechanism, requiring further in-depth exploration combining experiments and theory.
图9 I-CH2Cl在304 nm光解的平动能谱图[56]

Fig. 9 The photofragment translational spectra of I-CH2Cl at 304 nm[56]

2.3 Study on the Photodissociation of Halogenated Ethanes

Ethyl iodide (C2H5I) is one of the simplest iodinated alkanes, and its excited state potential energy curve involved in A-band photodissociation is similar to that of CH3I. Wilson et al.[18] once studied the photodissociation of C2H5I using a molecular beam laser photolysis product spectrometer, but due to the low resolution of the instrument, they were unable to distinguish between the I* and I channels on the TOF spectrum. Zhu Qihao et al. used a self-developed rotatable molecular beam laser photofragment translational energy spectrometer to study the photodissociation of C2H5I at 248 nm[38-39,58], achieving separation of the I* and I channels on the TOF spectrum (Figure10), and measuring a quantum yield Φ(I*) of 0.70 (Table3). The research found that both photodissociation channels follow the impulse model mechanism, with their dissociations occurring as rapid processes along the repulsive potential energy surface of the C—I bond, meaning that after the C—I bond absorbs light energy, it dissociates before this energy can be randomly distributed among other degrees of freedom within the molecule. Based on the fragment TOF spectra and conservation of energy, the internal energy distributions of the C2H5 radicals in the two photodissociation channels were obtained (Table3). In the I* channel, besides the umbrella vibration (540 cm-1) being excited, the stretching vibration of the C—C bond (1138 cm-1) was also likely excited; in the I channel, apart from the aforementioned vibrations, the torsional (193 cm-1) and wagging (1175 cm-1) vibrations, which belong to the asymmetric vibration modes, were also possibly excited in the C2H5 radical.
图10 在248 nm下C2H5I光解碎片I+的飞行时间谱[39]

Fig. 10 The TOF spectrum of photofragments I+ from photodissociation of C2H5I at 248 nm[39]

表3 248 nm下部分碘代烃光解离的动力学参数

Table 3 The dynamical parameters for the photodissociation of partial iodohydrocarbon at 248 nm

Reaction Channels Φ(I*) Eint(R)/Eavl Ref.
C2H5I→C2H5 + I* 0.70 0.32 38
C2H5 + I 0.39
n-C3H7I→n-C3H7 + I* 0.62 0.49 38
n-C3H7 + I 0.54
i-C3H7I→i-C3H7 + I* 0.49 0.63 38
i-C3H7 + I 0.64
n-C4H9I→n-C4H9 + I* 0.31 0.70 39
n-C4H9 + I 0.74
t-C4H9I→t-C4H9 + I 0.76 39
n-C5H11I→n-C5H11 + I* 0.67 0.77 59
n-C5H11 + I 0.76
CH2=CHI→CH2=CH + I* 0.57 0.31 60
CH2=CH + I 0.41
Later, Zhu Qihao et al.[61] studied the photodissociation dynamics of C2H5I near 280 and 304 nm, obtaining partially vibrationally resolved translational energy spectra (Figure11), where the vibrational peaks were assigned to the excitation of the umbrella vibrational mode (540 cm-1) of the C2H5 radical. The dissociation energy D0 of the C—I bond in C2H5I was measured to be 53.4 ± 0.7 kcal/mol. At 281.73 and 304.02 nm for the I* channel, the ratio of internal energy to available energy Eint/Eavl was 0.221 and 0.224, respectively; at 279.71 and 304.67 nm for the I channel, Eint/Eavl was 0.252 and 0.259, respectively (Table4). Compared with the photodissociation results at 248 nm, it can be found that when the available energy decreases, the proportion of internal energy allocated to the ethyl fragment Eint/Eavl becomes smaller.
图11 在281.73 nm下光解C2H5I时I*通道中I+的飞行时间谱[61]

Fig. 11 The TOF spectrum of I+ from the photodissociation of C2H5I at 281.73 nm for the I* channel[61]

表4 279’305 nm范围部分碘代烃光解的动力学参数

Table 4 The dynamical parameters for the photodissociation of partial iodohydrocarbon at 279~305 nm

λ/nm Channels Eint/Eavl β Ref
281.73 C2H5 + I* 0.221 - 61
304.02 0.224 -
279.71 C2H5 + I 0.252 - 61
304.67 0.259 -
281.73 C2F5 + I* 0.52 1.70 16
304.02 0.50 1.64
279.71 C2F5 + I 0.60 1.25 16
304.67 0.55 0.88
281.73 n-C3H7 + I* 0.48 1.68 62
304.02 0.49 ’2.00
279.71 n-C3H7 + I 0.52 ’2.00 62
304.67 0.52 1.57
281.73
304.02
i-C3H7 + I* 0.61 1.72 62
0.65 1.75
279.71 i-C3H7 + I 0.62 1.32 62
304.67 0.49 1.31
Compared to C2H5I, perfluoroiodoethane (C2F5I) has been less studied. Zhu Qihao et al.[16] used a high-resolution micro-laser photolysis fragment translational energy spectrometer to study the photodissociation of C2F5I near 280 and 304 nm. In the I* channel, they obtained vibrationally resolved photofragment translational energy spectra of the C2F5 radical (Figure 12), and assigned the vibrational peaks to the excitation of the CF2 wag mode (ν11 = 366 cm-1). For photodissociation at 281.73 and 304.02 nm in the I* channel, Eint/Eavl were measured as 0.52 and 0.50, respectively; for photodissociation at 279.71 and 304.67 nm in the I channel, Eint/Eavl were 0.60 and 0.55, respectively (Table 4). The internal energy excitation of the C2F5 fragment during the photodissociation of C2F5I is very high and much higher than that of C2H5I, consistent with the conclusion drawn from the comparison between CF3I and CH3I photodissociations[42,52-53]. This indicates that after F atoms replace H atoms, more available energy is distributed to the hydrocarbon fragments during the photodissociation of iodoalkanes. The anisotropy parameter β values of the photofragment angular distribution in the I* channel are all close to 2, indicating that the I* channel mainly originates from the contribution of 3Q0 state dissociation. However, in the I channel at 279.71 and 304.67 nm, β was measured as 1.25 and 0.88, respectively (Table 4), suggesting that both the direct dissociation via the 3Q1 state and the dissociation through the 1Q13Q0 pathway contribute to the I channel; and as the dissociation wavelength increases, the proportion of direct dissociation via the 3Q1 state becomes larger, but within this wavelength range, the contribution of the 1Q13Q0 pathway remains dominant.
图12 在281.73 nm下C2F5I光解时I*通道中I+的飞行时间谱[16]

Fig. 12 The TOF spectrum of I+ from the photodissociation of C2F5I at 281.73 nm for the I* channel[16]

2.4 Photodissociation Studies of Other Halocarbons

Zhu Qihao et al.[38,63] used a molecular beam rotatable laser photolysis fragment translational energy spectrometer to study the photodissociation of propyl iodide (C3H7I) at 248 nm. The obtained spectra clearly distinguished the two photodissociation channels, I* and I, for both n-C3H7I and i-C3H7I. The quantum yields Φ(I*) were measured to be 0.62 and 0.49, respectively. They also obtained the ratio of internal energy to available energy Eint/Eavl (Table3). They further investigated the photodissociation behavior of n-C3H7I and i-C3H7I near 280 and 304 nm[62], obtaining partially resolved vibrational state translational energy spectra for both I* and I channels. The spectra showed that during the photodissociation of n-C3H7I, the RCH2 deformation vibration (frequency 530 cm−1) was most easily excited, while in the case of i-C3H7I, the HC(CH3)2 out-of-plane bending vibration (frequency 364 cm−1) was the most easily excited vibrational mode. The distribution of available energy after the photodissociation of n-C3H7I and i-C3H7I was determined (Table4), and it was found to be similar to Eint/Eavl at 248 nm. Compared with n-C3H7I, a higher proportion of internal energy was allocated during the photodissociation of i-C3H7I. For n-C3H7I, the β values measured in both I and I* channels were close to 2 within this wavelength range, indicating that both photodissociation channels originated from the 1Q13Q0 and 3Q0 dissociation pathways. For i-C3H7I, the β value in the I* channel was close to 2, suggesting a primary origin from direct 3Q0 dissociation; however, the β value in the I channel was smaller, indicating that part of the I channel signal came from the contribution of direct 3Q1 dissociation.
Using a molecular beam rotatable laser photolysis fragment translational energy spectrometer, Zhu Qihui et al. studied the photodissociation mechanisms of larger-sized iodobutanes (n-C4H9I) and tert-iodobutane (t-C4H9I)[39], as well as n-iodopentane (n-C5H11I)[59] at 248 nm. They measured the quantum yields Φ(I*) for the I* channel during the photodissociation of n-C4H9I and n-C5H11I to be 0.31 and 0.67, respectively, and obtained the ratio of the internal energy of the dissociation fragments to the available energy (see Table 3). The experimental results of the photodissociation of these iodinated alkanes indicate that as the size of the alkyl group increases, a larger proportion of the available energy is converted into the internal energy of the alkyl group during the photodissociation process. This is mainly due to the increase in the size of the alkyl group, which leads to an increase in the number of vibrational degrees of freedom, resulting in more internal energy being excited. The quantum yield Φ(I*) for the I* channel and the branching ratios of each dissociation pathway of the iodinated alkanes are primarily controlled by the transition intensities from the ground state to the 3Q0, 1Q1, and 3Q1 states upon photon absorption, as well as the crossing probabilities of the 3Q0 and 1Q1 state potential energy curves during the dissociation process. The speed and direction with which the excited iodinated alkane molecules cross the intersection region are important factors affecting the crossing probability.
In addition to alkyl iodides, they also used a molecular beam rotatable laser photodissociation fragment translational energy spectrometer to study the photodissociation of vinyl iodide (CH2=CHI) at 248 nm[60]. Similar to the photodissociation of alkyl iodides, the photodissociation of CH2=CHI has two channels: I*(2P1/2) and I(2P3/2). The experimental quantum yield for the I* channel Φ(I*) = 0.57, and the C-I bond dissociation energy D0 = 61.9 ± 1.0 kcal/mol, revealing the energy distribution during the dissociation process (Table3). The research results confirmed that the photodissociation of CH2=CHI is a rapid process occurring on the repulsive potential energy surface along the C—I bond. In the I* channel, the bending and out-of-plane bending vibrational modes ν7 (918 cm-1), ν9 (783 cm-1), and ν8 (827 cm-1) of the CH2=CH fragment are most likely to be excited; in the ground state I channel, in addition to the excitation of the above modes, the C—C stretching vibrational mode (1670 cm-1) may also be excited.
They also investigated the photodissociation dynamics of 3-C3H5Br and 1-C3H5Br in bromopropene at 193nm[64]. Based on the Br+ time-of-flight spectra obtained from experiments and energy conservation analysis, it was found that the dissociation of 3-C3H5Br and 1-C3H5Br differs from that of iodoalkanes and cannot be explained by the impulsive model; there is a competition between their dissociation and isomerization caused by hydrogen transfer.
The Qihui Zhu team systematically studied the photodissociation of molecules such as halogenated hydrocarbons in the A band, obtained their photodissociation rules, and proposed their microscopic reaction mechanisms, providing theoretical basis for controlling air pollution and protecting the ozone layer.

3 Study on the Photoionization Dynamics of Important Small Molecules

3.1 REMPI/MATI and IR-UV Spectroscopy Experimental Setup

The Zhu Qihao team independently built an experimental setup for resonance-enhanced multiphoton ionization spectroscopy and mass-resolved threshold ionization spectroscopy based on a linear time-of-flight mass spectrometer. The device mainly consists of the following parts: internal chamber system (source chamber, ionization chamber, free flight zone, and ion detection zone), vacuum system, laser system, timing control system, and signal acquisition system (Figure 13). Later, the instrument was improved by adopting laser desorption sampling, enabling REMPI/MATI spectral studies of high-boiling, low-thermal-stability molecules. Schematic diagrams of several resonance two-photon ionization (R2PI) spectra and MATI spectra are shown in Figure 14. Infrared lasers were also introduced to achieve several infrared-ultraviolet double-resonance IR/UV spectroscopic techniques, including: IR-UV spectroscopy for studying vibrations of monomers and clusters in the electronic ground state (S0 state), autoionization-detected infrared spectroscopy (ADIR) for studying vibrations of monomers in the ionic electronic ground state (D0 state), and infrared photodissociation spectroscopy (IRPD) for studying vibrations of clusters in the D0 state (Figure 15). Based on these experimental setups and combined with theoretical calculations, the Zhu Qihao team conducted a series of photoionization spectroscopic studies on benzene derivative molecules. Important spectroscopic data such as transition energies (E1) from the S0 state to the first electronically excited state (S1 state) and ionization energies (IE) for some key benzene derivative molecules discussed in this paper are shown in Table 5.
图13 REMPI/MATI光谱实验装置示意图

Fig. 13 The schematic diagram of REMPI/MATI spectroscopy experimental device

图14 几种REMPI光谱和MATI光谱的原理示意图

Fig. 14 The principles of REMPI and MATI spectroscopy

图15 IR-UV、ADIR和IRPD光谱原理示意图

Fig. 15 The principles of IR-UV, ADIR, and IRPD spectroscopy

表5 若干典型苯衍生物分子的第一电子激发态跃迁能(E1)和电离能(IE)汇总表

Table 5 Summary of electronic transition energies (E1) and ionization energies (IE) for several typical benzene derivative molecules

Molecules E1 (cm−1) IE (cm−1) Ref.
cis m-fluorostyrene 34403 - 65
trans m-fluorostyrene 34663 - 65
p-methylstyrene 34276 - 66
p-fluoroanisole 35149 - 67
cis p-methoxystyrene 33242 - 68
trans p-methoxystyrene 33324 - 68
p-chloroanisole 34859 - 69
cis 3-chloro-4-fluoroanisole 34703 67349 70
trans 3-chloro-4-fluoroanisole 34747 67595 70
cis 3-chlorostyrene 33766 69701 71
trans 3-chlorostyrene 34061 69571 71
cis m-aminostyrene 30937 61278 72
trans m-aminostyrene 31140 61495 72
3,5-difluoroanisole 37595 70096 73
cis 3-chloro-5-fluoroanisole 36468 69720 74
trans 3-chloro-5-fluoroanisole 36351 69636 74
cis 3-fluoro-N-methylaniline 33816 61742 75, 76
trans 3-fluoro-N-methylaniline 34023 61602 75, 76
cis 4-chloro-3-fluoroanisole 35443 67585 77
trans 4-chloro-3-fluoroanisole 35326 67324 77
trans 2-fluoro-N-methylaniline 34010 61101 78
cis 3-chloro-N-methylaniline 33003 61531 79
trans 3-chloro-N-methylaniline 32886 61625 79
cis 2-methoxypyridine 36098 69379 80
cis 2-N-methylaminopyridine 32456 62518 80
trans 2-N-methylaminopyridine 32188 62709 80

3.2 REMPI and MATI Spectroscopy Studies of Benzene Derivatives

The derivatives of benzene are synthetic units for many pharmaceuticals, industrial and agricultural products, and new materials, and they are also important research subjects in the fields of environmental chemistry and biochemistry. The spectral characteristics of the electronic states and ionic states of these compounds have attracted increasing attention from researchers. Compared to the relative electronic ground state, the spectroscopic studies of molecules in electronically excited states and ionic states are relatively fewer, especially for larger benzene derivative systems with more atoms.
Zhu Qihet alconducted monochromatic resonance two-photon ionization (1C-R2PI) spectroscopy experiments on styrene and meta-fluorostyrene[66][65]using a self-constructed spectrometer, accurately determining the first electronic excited state transition energy (E1), and obtaining their vibrational spectra in the S1state. Combining theoretical calculations, they discussed the structures of these molecules in the S0and S1states, identified the 1C-R2PI spectral lines, and analyzed the influence of substituents on the molecular E1. The results showed that in the S1state, the interaction between the substituent and the benzene ring is relatively stronger compared to the electronic ground state, while the effect of the vinyl C=C bond is relatively weaker; the stretching vibration frequency of the substituent with the benzene ring increases relative to the electronic ground state, whereas the C=C stretching vibration frequency of the vinyl group decreases relatively; the E1of both para-methylstyrene and meta-fluorostyrene are red-shifted compared to styrene.
The 1C-R2PI spectroscopy technique was employed to study the vibrational spectrum of p-fluoroanisole[67] in the S1 state. Combining quantum chemical theoretical calculations, the molecular structure was analyzed, and the excited-state spectra were identified and interpreted. The substituent effects of different substituents in the S1 state were also analyzed. The E1 value of p-fluoroanisole was measured to be 35 149 cm-1, which is a red shift of 1234 cm-1 compared to anisole. In conjunction with theoretical calculations, the active vibrational modes of p-fluoroanisole in the S1 state obtained from the experiment were identified and simulated. It was found that in the S1 state, due to a higher degree of p-π conjugation, the C—O bond connecting the oxygen atom to the benzene ring is shortened, exhibiting partial double-bond characteristics. By comparing the transition energies of benzene derivatives and their corresponding para-fluorinated compounds, it was discovered that para-fluorination leads to a greater reduction in energy in the S1 state than in the electronic ground state and ionic ground state. Therefore, the transition energy from the ground state to the S1 state undergoes a red shift, while the transition energy from the S1 state to the ionic ground state experiences a blue shift.
Theoretical calculations have revealed that the molecular skeleton of p-methoxystyrene[68] is planar in both the ground state and the excited state, with two stable rotamers: cis (cis) and trans (trans). Whether in the S0 or S1 state, the trans isomer is more stable than the cis isomer. In the S1 state, the interaction between the substituent and the benzene ring is stronger than in the S0 state, and the interaction between the vinyl group and the benzene ring is stronger than that between the methoxy group and the benzene ring; the C4—O bond and the C1—Cα bond exhibit partial double-bond character, while the Cα=Cβ bond length elongates as the conjugation degree of the C1—Cα increases. The 1C-R2PI spectra yield E1 values for the cis and trans isomers at 33,242 and 33,324 cm−1, respectively. The vibrational modes of the two isomers in the S1 state obtained from the spectral experiments were identified in conjunction with theoretical calculations.
Using 2-color resonance two-photon ionization (2C-R2PI) spectroscopy experiments combined with theoretical calculations, they studied the effects of substituent effects, isomer effects, and isotope effects on the molecular properties of 3-chloro-4-fluoroanisole (3C4FA)[70]. Through 1C-R2PI and 2C-R2PI spectral measurements, the E1 and IE of cis-3C4FA were found to be 34,703 and 67,349 cm−1, respectively; whereas for trans-3C4FA, the E1 and IE were 34,747 and 67,595 cm−1, respectively. Combining theoretical calculations, the R2PI spectral peaks of chlorine isotope substitution and cis-trans isomers of 3C4FA were identified, revealing that the impact of isomer effects on the transition energy, ionization energy, and vibrational frequencies of 3C4FA is greater than that of isotope effects.
Regarding the study on the effects of substituent and conformational isomerism on the properties of meta-aminostyrene[72], it was found that the E1 values for cis and trans meta-aminostyrene are 30,937 and 31,141 cm-1 respectively, showing a redshift compared to para-aminostyrene; this is contrary to the trend observed in most meta- and para-disubstituted benzene derivatives regarding the energy of the first electronic excited state. The active vibrational modes in the spectrum mainly involve in-plane vibrations of the benzene ring in the S1 state, with the relative position of the substituents affecting the vibration frequencies. Experimentally determined ionization energies (IE) for cis and trans meta-aminostyrene were 61,278 and 61,495 cm−1 respectively, and the energy differences between the S0, S1, and D0 states for the cis and trans isomers were found to be 30, 234, and 247 cm-1. It was discovered that the cis isomer is more stable than the trans isomer in the S0, S1, and D0 states.
For meta-chlorostyrene[71], they used REMPI and MATI spectroscopy to measure the E1 of cis and trans meta-chlorostyrene as 33,766 and 34,061 cm-1, respectively, with adiabatic ionization energies being 69,701 and 69,571 cm-1, respectively. The energy differences between the cis and trans isomers in the S0, S1, and D0 states were found to be 218, 513, and 88 cm-1, respectively, indicating that the cis isomer is more stable than the trans isomer in the S0, S1, and D0 states. The vibrational modes observed in the REMPI and MATI spectra mainly involve vibrations within the plane of the benzene ring and those associated with the substituent.
The molecular configurations and excited state vibrational spectra of cis and trans 3-chloro-5-fluoro-anisole (3C5FA) in different electronic states were studied using REMPI spectroscopy experiments and theoretical calculations[74]. The E1 values for cis- and trans-3C5FA were found to be 36,468 and 36,351 cm-1, respectively. Analysis of the spectral data revealed that the conformational isomer effect on the frequencies of certain vibrational modes in the S1 state is greater than the isotopic effect, with the latter mainly depending on the extent of Cl atom involvement in the vibrational mode. Using the photoelectron imaging efficiency (PIE) experimental method, they measured the ionization energies of cis and trans-3C5FA to be 69,720 and 69,636 cm-1, respectively. They also found that the cumulative effect of dihalogen substitution on the E1 and ionization energy of 3C5FA deviates somewhat from the additive rule for multiple substituents summarized by previous studies, but this additive rule still has some reference value for predicting the E1 and ionization energy of complex multi-substituted derivatives.
The photoionization spectra of 3,5-difluoroanisole (3,5-DFA) and its Ar cluster[73] have been studied. The E1 value of the 3,5-DFA molecule was experimentally determined to be 37,595 cm−1, which is blue-shifted by 933, 776, and 1212 cm-1 compared to cis- and trans-m-fluoroanisole and anisole, respectively; the ionization energy of 3,5-DFA is 70,096 cm-1. It was found that the experimental values of the blue shift in E1 and ionization energy for 3,5-DFA compared to anisole are 497 and 324 cm-1 larger than the predicted blue shifts based on additivity rules. Through theoretical calculations, it was discovered that in the 3,5-DFA···Ar cluster, the Ar atom is located above the plane of the benzene ring and tends to favor the OCH3 side. This finding demonstrates that the Ar atom binds to 3,5-DFA through a weak dispersion interaction, with little effect on the intrinsic properties of the 3,5-DFA molecule, also explaining why the E1 value of the cluster is only red-shifted by 9 cm−1 compared to the monomer.
Research has also been conducted on the excited state and ionic state structures and properties of meta-fluoro-N-methyl aniline (3FNMA) and its argon clusters[75-76]. Experimental measurements have accurately determined the E1 and ionization energy values for cis- and trans-3FNMA, obtaining information such as molecular configurations and vibrational spectra in the S1 and D0 states. The impact of substituent effects on the relative stability and vibrational behavior of the cis and trans isomers in different electronic states was further summarized. Using three experimental methods, namely REMPI spectroscopy, PIE curves, and MATI spectroscopy, the cluster binding energies of 3FNMA···Ar1 in different electronic states were obtained, and theoretical calculations were combined to determine their configurations[81].

3.3 Conformational Isomerism Study of Benzene Derivatives

Conformational isomerism refers to a type of stereoisomerism in organic molecules where the spatial arrangement of atoms or groups of atoms varies due to the rotation around single bonds. This phenomenon is widely present in biological systems and plays a crucial role. Common benzene derivatives such as anisole, phenol, styrene, aniline, and N-methyl aniline, often exhibit conformational isomerism because the C—O, C—C, and C—N single bonds can rotate. The relative stability of different conformers can be inferred through theoretical calculations and spectral data.
Zhu Qihet al. studied the conformational isomerism of benzene derivatives, represented by anisole and N-methyl aniline derivatives. Theoretical calculations showed that 2-fluoro-N-methyl aniline (2FNMA)[78], 3-chloro-N-methyl aniline (3ClNMA)[79], and 4-chloro-3-fluoro anisole (4Cl3FA)[77] exist in both cis and trans conformations in the S0, S1, and D0 states. However, only the photoionization spectrum of the trans-2FNMA conformation was observed experimentally, with theoretical calculations providing potential energy curves for both conformations; for 3ClNMA and 4Cl3FA, the photoionization spectra of both cis and trans conformations were observed. The precise data of E1 and IE for each conformation of these molecules were obtained through REMPI and MATI spectroscopy, yielding their vibrational characteristics in the S1 and D0 states; further, combined with theoretical calculations, the effects of substituent and conformational isomerism on the properties of these molecules in the ground state, excited state, and ionic state were summarized.
Using REMPI/MATI spectroscopy experiments combined with theoretical calculations, they also investigated the conformational selectivity and conformational isomerization caused by excitation and ionization of 2-methoxypyridine (2MOP) and 2-methylaminopyridine (2NMP)[80]. Through spectroscopic experiments, E1 and IE values for the cis-2MOP, cis, and trans-2NMP conformations were obtained, as well as their configurations, vibrational spectra in the S1 and D0 states, etc. (see Figure 16, 17). They used natural bond orbital, reduced density gradient (RDG) analysis, and donor-acceptor orbital interactions to explain the causes of conformational selectivity of 2MOP and 2NMP in various electronic states, finding that the relative stability of the s-cis/e-cis conformation is simultaneously controlled by exchange repulsion and hyperconjugation.
图16 (a) 2MOP和(b) 2NMP (c = cis, t = trans)的REMPI光谱[80]

Fig. 16 The REMPI spectra of (a) 2MOP and (b) 2NMP (c = cis, t = trans)[80]

图17 cis 2NMP的MATI光谱[80]

Fig. 17 The MATI spectra of cis 2NMP[80]

3.4 IR/UV Laser Spectroscopy of Benzene Derivatives and Their Clusters

Zhu Qihet al. used REMPI/MATI spectroscopy combined with IR-UV, ADIR, and IRPD methods to study the spectra of 3-fluoro-N-methyl aniline (3FNMA) and 3FNMA···Ar1 clusters, obtaining spectral information for these systems in the S0 and D0 states[76]. They analyzed the effects of substituent and conformational isomer on the spectral behavior of different vibrational modes of 3FNMA, and explained these using conjugation and rehybridization effects.
Using the IR-UV dual-resonance spectroscopy method, the NH and CH stretching vibration information of 3FNMA monomer and 3FNMA···Ar1 cluster in the S0 state was obtained (Figure 18); using the ADIR spectroscopy method, the NH stretching vibration information of 3FNMA monomer in the D0 state was acquired; using the IRPD spectroscopy method, the NH and CH stretching vibration information of 3FNMA···Ar1 cluster in the D0 state was obtained. Since the effect of Ar atoms in the cluster on the vibrations of 3FNMA can be neglected, the IRPD method can serve as a "messenger" technique, using the IRPD spectrum of 3FNMA···Ar1 cluster to indirectly obtain the missing CH vibration information in the ADIR spectrum of 3FNMA monomer (Figure 19).
图18 (a) cis 3FNMA和(b) cis 3FNMA···Ar1团簇在S0态的IR-UV双共振光谱,(c) cis 3FNMA的IR的理论模拟光谱[76]

Fig. 18 The IR-UV double resonance spectrum of (a) cis 3FNMA and (b) cis 3FNMA···Ar1 cluster in the S0 state, (c) theoretically simulation IR spectrum of cis 3FNMA[76]

图19 (a)cis 3FNMA在D0态的ADIR谱,(b)cis 3FNMA···Ar1团簇在D0态的IRPD光谱.(c)cis 3FNMA在D0态的IR理论模拟光谱[76]

Fig. 19 (a) The ADIR spectrum of cis 3FNMA and (b) the IRPD spectrum cis 3FNMA···Ar1 cluster in the D0 state. (c) The theoretically simulation IR spectrum of cis 3FNMA[76]

By comparing the IR spectral experimental results of cis and trans isomers, it was found that the CH stretching vibration of 3FNMA exhibits a certain isomer effect, and this effect decreases in the D0 state compared to the S0 state. The CH stretching vibration spectra of both isomers show varying degrees of peak splitting due to the Fermi resonance effect. By comparing the IR spectra of 3FNMA in the S0 and D0 states, it was observed that ionization has distinctly different effects on the NH and CH stretching vibrations. The former shows a redshift (~80 cm-1) in frequency and an increase in IR absorption intensity in the D0 state; whereas the latter experiences a blueshift (30~60 cm-1) in frequency and a significant decrease in IR absorption intensity in the D0 state. The spectral characteristics of the 3FNMA molecule indicate the presence of conjugation and rehybridization effects within the molecule.
The Qihui Zhu team adopted IR/UV laser spectroscopy technology combined with REMPI and MATI spectroscopy methods, obtaining more comprehensive spectral information of molecules in different electronic states experimentally, and extended the research objects to clusters, providing a powerful experimental method for deeply exploring important scientific issues such as cluster configurations and intermolecular interactions.

4 Summary and Prospects

Mr. Zhu Qihewas one of the pioneers in the field of molecular reaction dynamics research in China. This article only reviews Mr. Zhu Qihew's research work in the field of micro-dynamics of photodissociation and photoionization of important small molecules. Mr. Zhu Qihewas uniquely innovative in the development of scientific instruments. He was adept at utilizing innovative principles to miniaturize and simplify complex scientific instruments, which could improve instrument performance while reducing costs and making them easy to operate and maintain. This not only opened up new ideas for the development and innovation of scientific instruments but also provided valuable insights for the cultivation of talent in scientific instrument development.
Mr. Zhu Qihew and his team have delved deeply into the photodissociation dynamics of halogenated hydrocarbons, systematically studying the photodissociation processes of a series of halogenated hydrocarbons, obtaining their photodissociation rules, and proposing microscopic reaction mechanisms, which has promoted the development of molecular photodissociation dynamics theory. They have also provided important model parameters and theoretical bases for fields such as air pollution control and environmental protection. Their research also shows that the study of photodissociation dynamics of halogenated hydrocarbons is very important, with its reaction mechanisms being extremely complex, worthy of in-depth investigation.
From the perspective of photodissociation mechanisms, there is still a need to advance research into ultrafast time-resolved dynamics and quantum state-selective excited state dissociation dynamics. First, conical intersections of potential energy surfaces are widely present in the photodissociation processes of halocarbons and are one of the core elements affecting molecular photodissociation dynamics. However, most current studies still rely on indirectly inferring information about the conical intersection regions of potential energy surfaces through experimental measurements of reaction products combined with theoretical calculations. Fortunately, the emergence of technologies such as attosecond lasers and ultrafast electron diffraction has made it possible to directly probe the dynamic processes in the conical intersection regions[82-83]. Applying these ultrafast experimental techniques to the study of halocarbon photodissociation and deeply exploring the conical intersections between potential energy surfaces are of great significance for revealing the physicochemical essence of molecular photodissociation and understanding the complex non-adiabatic interactions between electronic states. Secondly, current research on the photodissociation of halocarbons mainly focuses on ground-state molecules, and there is a significant lack of work that deeply investigates the photodissociation of parent molecules in vibrationally excited states at the state-to-state level. Using methods such as infrared lasers to selectively excite specific vibrational-rotational modes of the parent molecule, and then further studying its photodissociation[47-48], can provide information on the photodissociation cross-sections of parent molecules excited by different vibrational-rotational modes, understand the coupling effects between different vibrational-rotational modes and photodissociation coordinates, elucidate the relationship between the initial vibrational-rotational state of the parent molecule and the internal energy excitation of the photodissociation products, and determine the impact of different vibrational-rotational modes on the quantum yield of the photodissociation products. This will help to truly achieve a deep understanding of the mechanisms of chemical reactions at the quantum state level.
From the perspective of practical application models for halocarbon photodissociation, more systematic research on a wider range of molecules and wavelengths is still needed to ultimately establish a comprehensive database, which can play an important role in industrial production and atmospheric environmental fields. For the photodissociation of ground-state methyl iodide in the A band, both experimental and theoretical studies are relatively thorough, able to well explain its microscopic reaction mechanism; however, experimental and theoretical studies on photodissociation at higher absorption bands are still lacking. Solar radiation reaching the Earth's atmosphere also has a strong distribution at shorter wavelengths, which plays a significant role in the dissociation and reactions of pollutant molecules. However, relevant research data is currently insufficient. For the photodissociation of more complex halocarbons, current experimental and theoretical studies are not yet systematic. Therefore, more systematic and comprehensive experimental and theoretical studies on the photodissociation dynamics of various halocarbons across the entire solar ultraviolet radiation spectrum will help in establishing accurate and complete models to assess their impact on the Earth's atmosphere and environment[2-3].
In the research on the photoionization spectra of important small molecules, Mr. Zhu Qiheng and his team conducted experimental and theoretical studies on the excited state and ionic state vibrational spectra of a series of benzene derivative molecules and their clusters, obtaining detailed information on the geometric structure, vibrational frequencies, excitation energies, and ionization energies of molecules in different electronic states. They deeply explored the effects and regularities of substituent effects, conformational isomer effects, and isotope effects on the spectral behavior of molecules and clusters. Through the study of these benzene derivative model molecules, it helps to better understand the structures and physicochemical properties of complex systems such as aromatic compounds or functional material molecules, which has significant reference value for exploring the functions of biomolecules, proposing new photochemical reaction mechanisms, developing and improving new functional materials, and many other fields. Further in-depth research and expansion in this direction are expected to achieve important breakthroughs.
From the perspective of experimental techniques, for heterocycles containing multiple heteroatoms, complex side chains, and aromatic compounds with multiple substituents, especially biomolecules, due to their large molecular weights, high boiling points, or poor thermal stability, the study of the excited states and ionic state spectra of these molecules still poses challenges. Firstly, it is necessary to improve the experimental sample introduction system; for example, laser desorption and other technologies can be used to increase the sample introduction amount without destroying the sample molecules, thereby achieving a higher concentration of sample molecules for subsequent spectral studies; secondly, the mass spectrometry resolution and detection sensitivity need to be improved to enable precise photoionization spectroscopy studies of high-molecular-weight samples[84-85]. For systems where the excited state lifetime is short and the Franck-Condon factor for electronic excitation transitions is low, leading to weak photoionization spectrum signal intensity, technologies such as picosecond lasers and vacuum ultraviolet lasers, which have high photon energy, can be used to directly ionize and obtain the photoionization mechanisms of these molecules[84-86]. From an application perspective, with the continuous development of technology and theory, photoionization spectroscopy will play an increasingly important role in fields such as environmental monitoring, materials, and biomedicine. For instance, photoionization spectroscopy technology can be used for real-time monitoring of pollutants in the atmosphere, water quality, and other environments, providing crucial data support for environmental protection; real-time in-situ measurement of the spectral characteristics of biomolecules during the photoionization process can reveal their molecular structures and interaction mechanisms. Additionally, the rapid development of artificial intelligence technology has provided strong support for research in the field of photoionization spectroscopy. For example, by training algorithms to process large-scale spectral data, the efficiency and accuracy of spectral analysis can be significantly improved, helping researchers better understand the spectral information of complex reaction systems.
The field of photodissociation and photoionization of small molecules has broad development prospects and significant application value, requiring continuous in-depth exploration and research by scientists. A deep understanding of the microscopic physicochemical mechanisms of molecular photodissociation and photoionization processes will play a crucial role in achieving the "ultimate goal" of chemical reaction control.
[1]
Schinke R. Photodissociation dynamics: spectroscopy and fragmentation of small polyatomic molecules. Cambridge [England]: Cambridge University Press, 1993.

[2]
Molina M J, Rowland F S. Nature, 1974, 249(5460): 810.

[3]
Vogt R, Sander R, von Glasow R, Crutzen P J. J. Atmos. Chem., 1999, 32(3): 375.

[4]
Chen H Y, Lien C Y, Lin W Y, Lee Y T, Lin J J. Science, 2009, 324(5928): 781.

[5]
Ashfold M N R, Lambert I R, Mordaunt D H, Morley G P, Western C M. J. Phys. Chem., 1992, 96(7): 2938.

[6]
Herzberg G. Molecular Spectra and Molecular Structure. Van Nostrand Reinhold: New York:1950.

[7]
Wu Z K, Tang A Q. Monograph on Molecular Spectroscopy. Shandong Science and Technology Press: Jinan, 1999.

(吴征铠, 唐敖庆, 分子光谱学专论. 山东科学技术出版社: 济南, 1999.)

[8]
Boesl U, Neusser H J, Schlag E W. Z. Für Naturforschung A, 1978, 33(12): 1546.

[9]
Zandee L, Bernstein R B, Lichtin D A. J. Chem. Phys., 1978, 69(7): 3427.

[10]
Zare R N. Annual Rev. Anal. Chem., 2012, 5: 1.

[11]
Müller-Dethlefs K, Schlag E W. Annu. Rev. Phys. Chem., 1991, 42: 109.

[12]
Turner D W, Al Jobory M I. J. Chem. Phys., 1962, 37(12): 3007.

[13]
Zhu L C, Johnson P. J. Chem. Phys., 1991, 94(8): 5769.

[14]
Zhu Q H, Huang S L, et al. Acta Phys.-Chim. Sin., 1985, 1: 211.

(朱起鹤, 黄寿令, 等. 物理化学学报, 1985, 1: 211.)

[15]
Wang X, Tian Z X, Shi T J, Shi X H, Yang D L, Zhu Q H. Chem. Phys. Lett., 2003, 380(5-6): 600.

[16]
Yu Z J, Xu X L, Cheng M, Yu D, Du Y K, Zhu Q H. J. Chem. Phys., 2009, 131(4): 044323.

[17]
Qi W K, Jiang P, Lin D, Chi X P, Cheng M, Du Y K, Zhu Q H. Rev. Sci. Instrum., 2018, 89(1): 013101.

[18]
Riley S J, Wilson K R. Faraday Discuss. Chem. Soc., 1972, 53: 132.

[19]
Sparks R K, Shobatake K, Carlson L R, Lee Y T. J. Chem. Phys., 1981, 75(8): 3838.

[20]
Eppink A T J B, Parker D H. J. Chem. Phys., 1998, 109(12): 4758.

[21]
Eppink A T J B, Parker D H. J. Chem. Phys., 1999, 110(2): 832.

[22]
Li G S, Shin Y K, Hwang H J. J. Phys. Chem. A, 2005, 109(41): 9226.

[23]
Rubio-Lago L, García-Vela A, Arregui A, Amaral G A, Bañares L. J. Chem. Phys., 2009, 131(17): 174309.

[24]
Guo H, Schatz G C. J. Chem. Phys., 1990, 93(1): 393.

[25]
Guo H. J. Chem. Phys., 1992, 96(4): 2731.

[26]
Alekseyev A B, Liebermann H P, Buenker R J, Yurchenko S N. J. Chem. Phys., 2007, 126(23): 234102.

[27]
Alekseyev A B, Liebermann H P, Buenker R J. J. Chem. Phys., 2007, 126(23): 234103.

[28]
Evenhuis C R, Manthe U. J. Phys. Chem. A, 2011, 115(23): 5992.

[29]
Person M D, Kash P W, Butler L J. J. Chem. Phys., 1991, 94(4): 2557.

[30]
Felder P. Chem. Phys., 1990, 143(1): 141.

[31]
Aguirre F, Pratt S T. J. Chem. Phys., 2003, 118(3): 1175.

[32]
Van Veen G N A, Baller T, De Vries A E, Shapiro M. Chem. Phys., 1985, 93(2): 277.

[33]
Clary D C. J. Chem. Phys., 1986, 84(8): 4288.

[34]
Li G, Hwang H J. J. Chem. Phys., 2006, 124(24): 244306.

[35]
Gedanken A, Rowe M D. Chem. Phys. Lett., 1975, 34(1): 39.

[36]
Godwin F G, Paterson C, Gorry P A. Mol. Phys., 1987, 61(4): 827.

[37]
Huang Y H. Master's Dissertation of Graduate University of Chinese Academy of Sciences, 1988.

(黄玉惠. 中国科学院研究生院硕士论文, 1988.)

[38]
Zhu Q H, Cao J R, Wen Y, Zhan, J M, Zhong X, Huang Y H, Fang W Q, Wu X J. Chem. Phys. Lett., 1988, 144(5,6): 486.

[39]
Cao J R, Huang Y H, Yang D L, Gao Z, Fang W Q, Wu X J, Zhu Q H. Chin. J. Chem. Phys., 1990, 3: 235.

(曹建如, 黄玉惠, 杨达林, 高振, 方万全, 武小军, 朱起鹤. 化学物理学报, 1990, 3: 235.)

[40]
Suzuki T, Kanamori H, Hirota E. J. Chem. Phys., 1991, 94(10): 6607.

[41]
Hammerich A D, Manthe U, Kosloff R, Meyer H D, Cederbaum L S. J. Chem. Phys., 1994, 101(7): 5623.

[42]
Cheng M, Yu Z J, Hu L L, Yu D, Dong C W, Du Y K, Zhu Q H. J. Phys. Chem. A, 2011, 115(7): 1153.

[43]
Hu L L, Zhou Z M, Dong C W, Zhang L J, Du Y K, Cheng M, Zhu Q H. J. Chem. Phys., 2012, 137(14): 144302.

[44]
Hause M L, Yoon Y H, Crim F F. J. Chem. Phys., 2006, 125(17): 174309.

[45]
Hause M L, Yoon Y H, Cas, A S, Crim F F. J. Chem. Phys., 2008, 128(10): 104307.

[46]
Epshtein M, Portnov A, Rosenwaks S, Bar I. J. Chem. Phys., 2011, 134(20): 201104.

[47]
Hu L L, Zhou Z M, Dong C W, Zhang L J, Du Y K, Cheng M, Zhu Q H. J. Phys. Chem. A, 2013, 117(21): 4352.

[48]
Zhou Z M, Hu L L, Dong C W, Zhang L J, Liu S, Du Y K, Cheng M, Zhu Q H. Sci. China Chem., 2014, 57(6): 902.

[49]
Li R J, Zhong Q H, Kong F A, Zhu Q H. Chin. Chem. Lett., 1992, 3(12): 989.

[50]
Bowman J M, Huang X C, Harding L B, Carter S. Mol. Phys., 2006, 104(1): 33.

[51]
Tian Z X, Bi W B, Deng H D, Wang X, Tang Z C, Zhu Q H. Chem. Phys. Lett., 2004, 400(1-3): 15.

[52]
Yu Z J, Cheng M, Xu X L, Yu D, Du Y K, Zhu Q H. Chem. Phys. Lett., 2010, 488(4-6): 158.

[53]
Lin D, Hu L L, Liu S, Qi W K, Cheng M, Du Y K, Zhu Q H. J. Phys. Chem. A, 2016, 120(49): 9682.

[54]
Lin D, Cheng M, Du Y K, Zhu Q H. Chem. J. Chin. Univ.-Chin., 2018, 39: 1713.

(林丹, 程敏, 杜宜奎, 朱起鹤. 高等学校化学学报, 2018, 39:1713.)

[55]
Varner R K, Zhou Y, Russo R S, Wingenter O W, Atlas E, Stroud C, Mao H, Talbot R, Sive B C. J. Geophys. Res.-Atmos., 2008, 113: D10303.

[56]
Cheng M, Lin D, Hu L L, Du Y K, Zhu Q H. Phys. Chem. Chem. Phys., 2016, 18(4): 3165.

[57]
Yu Z J. Doctoral Dissertation of Graduate University of Chinese Academy of Sciences, 2010.

(余紫钧. 中国科学院研究生院博士论文, 2010. )

[58]
Cao J R, Wen Y, Zhang J M, Gu H G, Zhong X, Fang W Q, Duan S X, Wu X J, Zhu Q H. Acta Phys.-Chim. Sin., 1988, 4: 256.

(曹建如, 温晔, 张建明, 顾好刚, 钟宪, 方万全, 段素香, 武小军, 朱起鹤. 物理化学学报, 1988, 4: 256.)

[59]
Tian R J, Li R J, Kong F A, Zhu Q H. Chin. J. Chem. Phys., 1994, 7: 407.

(田如江, 李润君, 孔繁敖, 朱起鹤. 化学物理学报, 1994, 7: 407.)

[60]
Cao J R, Zhang J M, Zhong X, Huang Y H, Fang W Q, Wu X J, Zhu Q H. Chem. Phys., 1989, 138: 377.

[61]
Bi W B, Xu X L, Huang J G, Xiao D Q, Zhu Q H. Sci. China Ser. B Chem., 2007, 50(4): 476.

[62]
Xu X L, Yu Z J, Bi W B, Xiao D Q, Yu D, Du Y K, Zhu Q H. J. Phys. Chem. A, 2008, 112(9): 1857.

[63]
Huang Y H, Cao J R, Wen Y, Zhong X, Zhang J M, Fang W Q, Wu X J, Zhu Q H. Acta Phys.-Chim. Sin., 1987, 3: 337.

(黄玉惠, 曹建如, 温晔, 钟宪, 张建明, 方万全, 武小军, 朱起鹤. 物理化学学报, 1987, 3: 337.)

[64]
Zhu W S, Zhao X S, Han D G, Li R J, Zhong Q H, Zhu Q H. Chem. Phys. Lett., 1993, 204: 538.

[65]
Huang J G, Xiao D Q, Bi W B, Xu X L, Gao Z, Zhu Q H. Zhang C H. J. Mol. Struct., 2006, 794(1-3): 320.

[66]
Huang J G, Xiao D Q, Bi W B, Xu X L, Gao Z, Zhu Q H, Zhang C H. Spectrochim. Acta Part A Mol. Biomol. Spectrosc., 2007, 66(2): 371.

[67]
Xiao D Q, Yu D, Xu X L, Yu Z J, Du Y K, Gao Z, Zhu Q H, Zhang C H. J. Mol. Struct., 2008, 882(1-3): 56.

[68]
Xiao D Q, Yu D, Xu X L, Yu Z J, Du Y K, Gao Z, Zhu Q H. Zhang C H. J. Mol. Struct., 2009, 918(1-3): 154.

[69]
Yu D, Dong C W, Cheng M, Hu L L, Du Y K, Zhu Q H, Zhang C H. J. Mol. Spectrosc., 2011, 265(2): 86.

[70]
Yu D, Dong C W, Zhang L J, Cheng M, Hu L L, Du Y K, Zhu Q H, Zhang C H. J. Mol. Struct., 2011, 1000(1-3): 92.

[71]
Dong C W, Zhang L J, Liu S, Hu L L, Cheng M, Du Y K, Zhu Q H, Zhang C H. J. Mol. Spectrosc., 2013, 292: 35.

[72]
Dong C W, Zhang L J, Liu S, Hu L L, Cheng M, Du Y K, Zhu Q H, Zhang C H. J. Mol. Struct., 2014, 1058: 205.

[73]
Zhang L J, Dong C W, Cheng M, Hu L L, Du Y K, Zhu Q H, Zhang C H. Spectrochim. Acta Part A Mol. Biomol. Spectrosc., 2012, 96: 578.

[74]
Zhang L J, Yu D, Dong C W, Cheng M, Hu L L, Zhou Z M, Du Y K, Zhu Q H, Zhang C H. Spectrochim. Acta Part A Mol. Biomol. Spectrosc., 2013, 104: 235.

[75]
Zhang L J, Liu S, Dong C W, Cheng M, Du Y K, Zhu Q H, Zhang C H. J. Mol. Spectrosc., 2014, 296: 28.

[76]
Zhang L J, Liu S, Cheng M, Du Y K, Zhu Q H. J. Phys. Chem. A, 2016, 120(1): 81.

[77]
Liu S, Dai W S, Lin D, Cheng M, Du Y K, Zhu Q H. J. Mol. Spectrosc., 2017, 338: 15.

[78]
Liu S, Dai W S, Zhang L J, Cheng M, Du Y K, Zhu Q H. J. Mol. Struct., 2017, 1146: 138.

[79]
Liu S, Zhang L J, Dai W S, Cheng M, Du Y K, Zhu Q H. J. Mol. Spectrosc., 2017, 336: 12.

[80]
Dai W S, Liu S, Zhang Z, Chi X P, Cheng M, Du Y K, Zhu Q H. Phys. Chem. Chem. Phys., 2018, 20(9): 6211.

[81]
Zhang L J, Li D Z, Cheng M, Du Y K, Zhu Q H. Spectrochim. Acta Part A Mol. Biomol. Spectrosc., 2017, 183: 177.

[82]
Kobayashi Y, Chang K F, Zeng T, Neumark D M, Leone S R. Science, 2019, 365(6448): 79.

[83]
Yang J, Zhu X L, Wolf T J A, Li Z, Nunes J P F, Coffee R, Cryan J P, Gühr M, Hegazy K, Heinz T F, Jobe K, Li R K, Shen X Z, Veccione T, Weathersby S, Wilkin K J, Yoneda C, Zheng Q, Martinez T J, Centurion M, Wang X J. Science, 2018, 361(6397): 64.

[84]
Park S M, Kwon C H. J. Phys. Chem. Lett., 2023, 14(42): 9472.

[85]
Li X, Gao X H, Li W K, Yang T, Zhang D D, He L H, Luo S Z, Zhao S F, Ding D J. Phys. Rev. A, 2024, 109: 013103.

[86]
Kallos I S, Bar I, Baraban J H. J. Phys. Chem. Lett., 2024, 15(9): 2639.

Outlines

/