Ultrafast Deep-Ultraviolet Laser Ionization Mass Spectrometry

Jul 30, 2018 - Target-Triggered Catalytic Hairpin Assembly-Induced Core–Satellite Nanostructures for High-Sensitive “Off-to-On” SERS Detection o...
0 downloads 0 Views 865KB Size
Subscriber access provided by University of Rochester | River Campus & Miner Libraries

Article

Ultrafast Deep-ultraviolet Laser Ionization Mass Spectrometry Applicable to Identify Phenylenediamine Isomers Haiming Wu, Chengqian Yuan, Hanyu Zhang, Guanhua Yang, Chaonan Cui, Mengzhou Yang, Wensheng Bian, Hongbing Fu, Zhixun Luo, and Jiannian Yao Anal. Chem., Just Accepted Manuscript • DOI: 10.1021/acs.analchem.8b03167 • Publication Date (Web): 30 Jul 2018 Downloaded from http://pubs.acs.org on August 1, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 8 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

Ultrafast Deep-ultraviolet Laser Ionization Mass Spectrometry Applicable to Identify Phenylenediamine Isomers Haiming Wu,1,2 Chengqian Yuan,1,2 Hanyu Zhang,1,2 Guanhua Yang,1,2 Chaonan Cui,1,2 Mengzhou Yang,1,2 Wensheng Bian,2,3 Hongbing Fu,4 Zhixun Luo,1,2,* Jiannian Yao1,2 1

State Key Laboratory for Structural Chemistry of Unstable and Stable Species, CAS Key Laboratory of Photochemistry, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, P. R. China.

2

University of Chinese Academy of Sciences, Beijing 100049, P. R. China.

3

State Key Laboratory of Molecular Reaction Dynamics, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, P. R. China.

4

Department of Chemistry, Capital Normal University, Beijing 100048, P. R. China. *Correspondence. Email: [email protected]

ABSTRACT: The application of low-fragmentation mass spectrometry to identify chemicals has been recognized of particular importance in chemistry, biomedicine and materials science. Utilizing customized all-solid-state picoseconds-pulsed deep-ultraviolet (DUV) laser, here we present new advances into photoionization mass spectrometry. The DUV laser ionization mass spectrometry (DUV-LIMS) takes on very clean spectra pertaining to minimized structure relaxation and fragmentation under the ultrafast ionization process. Typical DUV-LIMS applications are illustrated not only for small organic molecules, but also for long-chain unsaturated hydrocarbons and the clusters of benzene. The unique advantages of DUV-LIMS enable us to detect and determine confusable organic compound mixture indicating promising applications. Interestingly, DUV-LIMS is also found applicable to identify phenylenediamine isomers. An in-depth analysis of reaction dynamics is provided showing how hydrogen-atom-transfer (HAT) initiates the distinguishable photodissociation of phenylenediamines under near-resonant excitation. In particular, orthophenylenediamine (OPD) finds remarkable dehydrogenation product with comparable intensity as the molecular ion peak, which is associated with the quantum tunnelling tautomers providing new subjects to study intramolecular noncovalent interactions.

With development trend of precise chemistry which has raised new challenges to atomic-resolved analytical techniques, low-fragmentation mass spectrometry makes a way for chemical identification simply by measuring the ions abundance relating to their mass-to-charge ratios.1-3 Among the soft-ionization techniques, proper laser ionization mass spectrometry (LIMS) takes advantages of its matrix-free, interference-free and highly-controllable procedures. For LIMS investigations, a wavelength-tunable light source is generally expected,4,5 to acquire high photoionization efficiency and low fragmentation probability by means of single-photon ionization (SPI) or resonance-enhanced multiphoton ionization (REMPI).6,7 However, the inconvenient instrument maintenance of tunable light sources impedes the widespread use of LIMS in the present circumstances; also it is difficult to tune a right wavelength if the test samples include multiple components or just a mass distribution (e.g., atomic/molecular clusters) having different values of ionization potentials. Does high-efficient LIMS have to be attained by tunable laser? A comprehensive insight into photoexcitation, photoionization and photodissociation may be necessary to solve this question. A sketch in Figure 1 depicts the correlation between light-matter interaction energy with respect to time evolution on a basis of Franck-Condon diagram and Jablonsky molecular energy level diagram.1 It is shown that the transition from

molecule ground state to the ionic state (i.e., cation radical M+ ) is often correlated with dissociation depending on the residual energies after photoionization and the dynamic fragment appearance energy (i.e., the amount of necessary energy to be transferred to the neutral M for the detection of the aim fragment ion). Also, the fragment appearance probability of detectable ions bears novel correlation with the rate constants. Considering that photo-absorption is under femtosecond scale while photo-induced dissociation could be within nanosecond, ultrashort laser pulses in the picosecond (rather than a conventional nanosecond regime) can largely prevent the molecules from relaxation or fragmentation prior to the formation of molecule ions.1-3 In other words, a picosecond ultrashort pulse could be more important than whether or not a tunable long-pulse light source to obtain low-fragmentation mass spectra. .

On the other hand, a single-photon or two-photon process is well-known to exhibit higher ionization efficiency and lower fragmentation comparing with a multi-photon process. This is because the magnitude order of multiphoton absorption cross section (σn) is significantly smaller than that of single photon process; simultaneously, the multiphoton absorption transition probability (Wn) is proportional to the n-order photoabsorption cross section and luminous flux, i.e., Wn ∝ σn. In,8 simply according to the time-dependent perturbation theory.

ACS Paragon Plus Environment

1

Analytical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

In view of this, a larger laser flux is required to make up for the significantly decreased σn, which increases the laser thermal effect and readily gives rise to unwanted relaxation and fragmentation.1-3 Considering that ~90% compounds bear absorption in deep ultraviolet (DUV) region, a proper DUV laser is necessary for high-efficient ionization. Previously there have been a few light sources available for such investigations, including electron synchrotrons,9,10 VUV lamps and laser.11,12 Recently, the newly-developed ps-pulsed DUV laser found unique advantages in bandwidth, beam quality/coherence and photon flux (~1015 photons/second).13,14 Beneficial from the largely improved sensitivity and resolution, DUV-LIMS is believed to be a truly general method and low-cost technique in laboratory available to identify organic molecules.15

Figure 1. Illustration of light-matter interaction energy with respect to time scale, displaying transition from the neutral molecule ground state to the ionic state (molecular ion M+ .) along with the subsequent dissociation/fragmentation where the dynamic fragment appearance energy (the amount of energy need to be transferred to the neutral M to allow for the detection of the fragment ion) is addressed. The inset on the right bottom corner shows the correlation between the rate constants of ions dissociations and the fraction of detectable ions. The abbreviation of AE refers to fragment appearance energy.

In addition to chemical identification, significant advances of mass spectrometry have been previously made with crucial applications in genomics and proteomics,16-18 where the protein structures can be determined when both the mass/charge (m/z) ratio and the precursor/product formulae are in agreement.19 The availability of mass spectrometry to identify chemical structure is thus highly important although this is not its necessary task. Here we report an interesting finding to identify phenylenediamine isomers by taking full advantages of the newly-developed DUV-LIMS technique. This is important considering that such molecular isomers bear similar properties and are often blended in the products of chemical synthesis.20 Insights into chemical selectivity regarding positional isomers is also a significant topic in molecule science attracting extensive researches for methods of chemical separation and purification.21,22 METHODS Experimental Methods. The experiments were performed on a customized instrument of reflection time-of-flight mass spectrometer (Re-TOF-MS, Figure S1, ESI) based on the all-solidstate DUV laser system (177.3 nm, 15.5 ps, 15 μJ, 10 Hz) by

Page 2 of 8

second harmonic generation of ps-pulsed 355nm laser with help of a KBBF-CaF2 prism coupled device. The powder samples of β-carotene (97%, Sigma-Aldrich), 1,5diaminonaphthalene (1,5-DAN, 99%, Sigma-Aldrich), para-, ortho- and meta-phenylenediamine (PPD, OPD, MPD, 99+% purity, Acros Organics) were put in a customized quartz container and heated to a certain proper temperature (80 oC for βcarotene, 50 oC for 1,5-DAN, 38 oC for PPD, 35 oC and 40 oC for OPD and MPD, respectively) that could produce enough vapor to form a stable molecular beam for photoionization under 355nm, 118nm and 177.3nm laser, respectively (Figure S2, ESI). Computational Methods. All the optimization, frequency and energy calculations were carried out on a basis of density functional theory (DFT) embedded within the Gaussian 09 program package.23 Geometries of all species were fully optimized at the unrestricted b3lyp/6-311++g(d,p) level of theory.24,25 All transition states (TS) structures were checked and confirmed by intrinsic reaction coordinate (IRC)26 calculations. Natural population analysis (NPA)27 was performed to reveal the charge distribution changes when phenylenediamine molecules are ionized. To estimate the amount of charge transfer during ionization process, natural bond orbital (NBO)28 analysis of neutral phenylenediamine isomers were also performed. The NBO orbitals were plotted via VMD and Multiwfn software.29 RESULTS AND DISCUSSION Regarding to the sensitivity and ionization efficiency of DUV-LIMS, we have performed a comprehensive study of different organic molecules. As typical results, Figure 2a and 2b display the mass spectra of para-phenylenediamine (PPD) and 1,5-diaminonapthalene (1,5-DAN), where the molecule ion peaks dominate the two spectra. The low-fragmentation advantage of DUV-LIMS not only lies in its 7-eV singlephoton energy, but also the ultrafast ps-pulse and high photon flux (one or two orders of magnitude comparing with usual vacuum ultra-violet (VUV) laser from gas cells). The ultrashort pulse duration reduces the time of interaction between DUV light and the molecules, which minimizes the fragmentation probability and hinders the subsequent photo-induced reactions.

Figure 2. DUV-LIMS of (a) para-phenylenediamine (PPD), (b) 1,5-diaminonapthalene (1,5-DAN), (c) a mixture of PPD and 1,5-

ACS Paragon Plus Environment

2

Page 3 of 8 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry DAN (1:1 mass ratio mixed evenly and thoroughly prior to use in the evaporation source) at ~5 μJ energy power. As a comparison, a 355nm-LIMS showing confusing fragments of the mixture of PPD and 1,5-DAN at ~10 mJ energy power (d).

In view of the DUV-LIMS low-fragmentation spectra, we have further prepared a sample by mixing the two solid samples together, endeavoring to explore the capability to identify multi-components with distinguishable fingerprints. The mixture sample of PPD and 1,5-DAN was prepared with a mass ratio of 1:1 by grinding method. As is shown in Figure 2c, the molecule ion peaks of PPD and 1,5-DAN are solely observed, which is in stark contrast with the undistinguishable fragments ionized by the 355nm laser (Figure 2d). In general, LIMS experiments based on usual UV multiphoton ionization processes render serious fragmentation hence often difficult to recognize the actual chemicals. In contrast, the lowfragmentation and matrix-free DUV-LIMS bears advantages to unambiguously identify multi-components from a mixture. In the past decades, considerable efforts have been made to devise methods to understand the behavior of polymer and biosystem chains. It is desirable of direct observation of the interaction units and conformation-specific pathways so as to identify and control the functionals of long-chain polymers and biosystems in a realistic way.30,31 However, long-chain molecules readily dissociate under laser radiation or just in solvent interactions due to confining vibrational relaxation. For example, β-carotene undergoes serious fragmentation under the normal 355nm laser ionization process (Figure 3Ab), leading to deficient molecular information from usual LIMS analysis. The 355nm laser (3.5-eV single-photon energy) gives rise to serious fragmentation as it allows transition to excited states associating with vibrionic couplings and chemical bond fissions in the photoinduced evolution of excited molecules.32 In contrast to essential fragmentation under 355nm laser ionization (even if laser power is fully weakened, Figure S4, ESI), DUV-LIMS enables to measure the molecular weight, as shown in Figure 3A-a, verifying the aforementioned general principles that transient high intensity of DUV laser significantly increases the ionization efficiency and reduces the fragmentation probability.

Figure 3. (A) LIMS of β-carotene ionized by the 177.3nm DUV laser at ~8 μJ (a) and 355nm laser at ~10 mJ (b). (B) DUV-LIMS of benzene clusters obtained via Ar buffer gas.

The high photoionization efficiency of DUV-LIMS has also been found to enable applications to probe organic molecular clusters which are formed by weak intermolecular interactions and readily dissociate under usual laser radiation. As a result, Figure 3B presents the DUV-LIMS results of neutral benzene clusters generated by pulsed supersonic beam expansion. A clean distribution of [C6H6]n series demonstrates that, the intrinsic dissociation of benzene molecule individuals or clusters formed by week inter-molecular interactions can be completely eliminated under an ultrafast DUV laser ionization process. It is worth noting that, among these clusters (n=125), the medium-sized ones [C6H6]4-6 display dominant intensity indicating enhanced stability of these clusters in contrast to the monomer and other counterparts. The identification of such molecule clusters provides important fundamentals for studies of intermolecular interactions. Recently, uprising research interest has been addressed on non-natural amino acids and amine molecules which form a variety of secondary structures. The impetus for such efforts is invariably devoted to understand conformational behavior and to develop biomolecular systems and materials with novel properties.33,34 Taking DUV-LIMS advantages, we address here the finding of novel selectivity for phenylenediamine isomers. Figure 4 presents the DUV-LIMS of orthophenylenediamine (OPD), meta-phenylenediamine (MPD) and the aforementioned PPD. As is shown, the mass spectra for all the three isomers are dominated by the molecule ion peak (C6N2H8+ .). However, interestingly for OPD, there is a remarkable dehydrogenation product (C6N2H7+) aside the molecular ion peak; in comparison, MPD also displays the parent ion peak, but there are fruitful fragment peaks pertaining to C-C bond cleavage (more details in Figure S6-10, ESI). To exclude the diversity of the three isomers being not selectively from collision-induced dissociation procedures, we have also employed the VUV 118nm laser (10.5-eV single-photon energy), to run the same experiments but found that VUV gives rise to undistinguishable parent ion peaks for the three isomers (Figure S5, ESI). Exclusion experiments including thermogravimetric analysis etc. reveal that the selectivity of DUV-LIMS of these isomers is determined by the altered photoionization processes (Figure S3, ESI).

Figure 4. DUV-LIMS of PPD (a), OPD (b) and MPD (c) at ~1 μJ energy power respectively.

ACS Paragon Plus Environment

3

Analytical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

For investigations of photoionization and reactivity, it is important for a precise control of the initial state of the parent ions with exact amount of energy and specific conformations imparted in the system.35-37 To unravel the interesting selectivity of phenylenediamine isomers by the DUV laser, simply we have checked the ionization energies (IEs), HOMO and LUMO energies, and molecular electrostatic potentials (MEP) for both neutral and ionic isomers of these three molecules (Figure S11-12, Table S1-2, ESI). As indicated by the changes of frontier orbitals and charge distributions on N atoms, the first ionization for the three isomers are all associated with the amino groups. Nevertheless, PPD bears ~7 eV ionization energy (~6.87 eV from NIST database)38 pertaining to DUV single-photon ionization process of high efficiency; comparatively, MPD and OPD have slightly larger IEs than 7 eV (ESI), thus undergo two-photons ionization for DUV excitation along with slightly different residual energies. This is consistent with the observation of different mass abundance of the parent ions for the three isomers. In brief, the photoionization process can be attained via single or multiple photons excitation, written as, 

   →    ∗ →    •   

(1)

Why only OPD takes on remarkable dehydrogenation? To understand this finding, we have checked out the varying potential energies with N-H bond dissociation (or, equivalently, H elimination) for the ionic phenylenediamine isomers, as shown in Figure 5a-c. Also noted is that, the H atom removal accompanies with a smaller threshold level (~4.1 eV for OPD, in contrast to ~5.9 eV and ~6.0 eV for PPD and MPD, respectively), indicating facile dehydrogenation for OPD. It was also noted that the donor-acceptor charge-transfer interaction for BDN-H2→BD*C-C is weaker (0.14 eV) in OPD than in the other two isomers (0.17 eV), and this value further decreases at the vertical ionization (0.07 eV), indicating accelerate fracture of the N-H bond, as summarized to be, 

Page 4 of 8

Figure 5. (a/b/c) Varying potential energies associated with the H elimination pathway for PPD, MPD and OPD respectively, via first-principles calculations at the b3lyp/6-311++g(d,p) level of theory. Red and navy lines indicate the asymptotic values of H elimination. Insets are the NBO donor-acceptor overlaps between N and H atoms of the nascent molecule in comparison with that after vertical ionization. BD and BD* refer to bonding orbitals and antibonding orbitals respectively. (d) Potential energy profile along the isomerization path of OPD. The atoms in dark gray, blue and light gray refer to C, N, H atom, respectively.

Besides the interesting dehydrogenation diversity of the three isomers, there are still differences for a few other fragment peaks. For example, for MPD, there is a onefold fragment peak appearing at 91 amu corresponding to the fragmentation of a NH3 molecule removal. The deamination readily occurs with thermodynamic-favorable processes in view of the NH3 molecule stability and a low-energy fragment ion C6NH5+ . . 

   →    ∗ →         

(2)

In particular, it is important to note that the dehydrogenization product of OPD involves resonating structures (tunneling tautomers) pertaining to H atom tunneling effect through a 0.69 eV energy barrier (Figure 5d), which may result in measurable quantum tunneling splitting. The quantum tunneling tautomerism under proper laser excitation creates a dynamic condition profiting to the stability of OPD dehydrogenization product,39-41 which is similar to malonaldehyde,40 an important benchmark system for the study of intramolecular proton transfer. The non-covalent interactions between H and N in such simple aromatic amine molecules caused by proton transfer would be an interesting future topic for those working in tunneling splitting measurement.42,43

   →    ∗ →   •     

(3)

Further, we found that hydrogen atom transfer (HAT) facilitates the NH3 removal. Figure 6A depicts the pathways of the three isomers in rendering deamination. It is noteworthy that the transition state energy values for deamination from MPD (3.01 eV) is smaller than that from PPD (3.64 eV) and OPD (3.41 eV). Also, the final dissociation of NH3 from MPD costs a relatively smaller activation energy of 1.92 eV (1.97 eV for OPD and 2.20 eV for PPD) in our calculations. This is coincident with the experimental observation that MPD shows notable fragment C6NH5+ . (amu=91) in contrast to PPD and OPD. Also, from the NBO analysis (Figure 6B), although the three isomers display slightly different donor-acceptor chargetransfer interactions around the dominant LPN→BD*C-C, for MPD the energies largely decrease at the vertical ionization profiting to C-N bond dissociation.

ACS Paragon Plus Environment

4

Page 5 of 8 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

Figure 6. (a) The fragment pathways towards the NH3 removal initiated by the HAT from the benzene ring to the amidogen group. (b) NBO donor-acceptor overlaps between the N and C atoms of the nascent molecule and that after vertical ionization. LP and BD* refer to lone pairs and antibonding orbitals respectively. The atoms in dark gray, blue and light gray refer to C, N, H, respectively.

It is additionally notable that the two amidogen groups in the three phenylenediamine isomers may also donate hydrogen atom to the benzene ring leading to enlarged neighboring C-C bond lengths (Figure 7), hence beneficial to the C-C bond cleavage along with the removal of a triatomic molecule CNH,44 

   →    ∗ →   •      (4) The transition state energy for HAT, C-C bond cleavage and recombination, and the CNH removal process from MPD (3.04 eV, 2.62 eV, 3.11 eV, 2.51 eV) are smaller than that from PPD and OPD (more details see Figure S16). This is also consistent with the observation that both MPD and OPD display a fragment peak at 80/81 amu suggesting secondary fragment appearance based on the parent ion (m/z=108) or the dehydrogenation counterpart (m/z=107).

process; also, DUV-LIMS is available for identification of phenylenediamine isomers. Among the three isomers, PPD takes on a well-resolved mass spectrum with a dominant parent ion peak articulating the high efficiency of DUV singlephoton ionization; in comparison, OPD displays a remarkable dehydrogenation counterpart aside the parent ion peak profiting from the hydrogen atom tunneling effect; MPD accompanies with hydrogen atom transfer followed by NH3 and CNH dissociation. Such selectively is associated with the novel DUV excitation for the three isomers which bear ionization potentials close to 7 eV. On this basis, we propose that nearresonance photoexcitation (not necessarily tunable precise resonance) is available to identify structure isomers simply by laser ionization mass spectrometry. Comparing with the general method to identify positional isomers by NMR which relies on specific isotopic nuclei, this near-resonant photoionization technique has its advantages of without involving a necessary sample preparation step of isotopic labeling (e.g., 13 C and 15N) prior to structural characterization. Besides, DUV-LIMS could be also available to produce small amount of ammonia from remote control.45,46 In particular, remarkable dehydrogenation products are obtained, suggesting an experimental method to produce quantum tunneling tautomer, which paves a way for measuring intramolecular (weak noncovalent) forces.

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website. Additional details of experimental and calculation methods, thermogravimetric analysis, DUV-LIMS, 118nm-LIMS, 355nm-LIMS, DFT-calculated details (PDF)

AUTHOR INFORMATION Corresponding Author *E-mail: [email protected]

Author Contributions The manuscript was written through contributions of all authors.

Notes The authors declare no competing financial interests.

ACKNOWLEDGMENT Figure 7. The proposed reaction pathways for HAT from the amino group to the benzene ring giving rise to the fragment ion peak of C5NH7+ .. The atoms in dark gray, blue and light gray refer to C, N, and H, respectively.

CONCLUSION In all, utilizing ultrafast all-solid-state DUV laser technique, we make essential improvement to LIMS with high ionization efficiency and quick ionization rate prior to structure relaxation and fragmentation. Typical DUV-LIMS applications are shown herewith for small organic molecules, long-chain hydrocarbon substance, multi-component mixture and benzene clusters. DUV-LIMS not only takes advantages for the detection and determination of confusable organic compound especially those readily dissociative under usual laser ionization

We thank Dr. Haitao Ma for his friendly discussion. This work is financially supported by the "National Project Development of Advanced Scientific Instruments Based on Deep Ultraviolet Laser Source" (No. Y31M0112C1), and by Key Research Program of Frontier Sciences (CAS, Grant QYZDB-SSW-SLH024) and the National Natural Science Foundation of China (Grant No. 21722308). Z. Luo acknowledges the National Thousand Youth Talents Program.

REFERENCE (1) Gross, J. Mass Spectrometry; Springer: Heidelberg, 2013. (2) Hoffmann, E. d.; Stroobant, V. Mass Spectrometry: Principles and Applications; John Wiley & Sons Ltd: Chichester, 2007. (3) McLuckey, S. A.; Wells, J. M. Mass Analysis at the Advent of the 21st Century. Chem. Rev. (Washington, DC, U. S.) 2001, 101, 571606.

ACS Paragon Plus Environment

5

Analytical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(4) Cool, T.; McIlroy, A.; Qi, F.; Westmoreland, P.; Poisson, L.; Peterka, D.; Ahmed, M. Photoionization Mass Spectrometer for Studies of Flame Chemistry with a Synchrotron Light Source. Rev. Sci. Instrum. 2005, 76, 094102. (5) Swinbanks, D. Synchrotron Light-Sources-International Interest. Nature 1990, 344, 8-8. (6) Kruth, C.; Czech, H.; Sklorz, M.; Passig, J.; Ehlert, S.; Cappiello, A.; Zimmermann, R. Direct Infusion Resonance-Enhanced Multiphoton Ionization Mass Spectrometry of Liquid Samples under Vacuum Conditions. Anal. Chem. 2017, 89, 10917-10923. (7) Zandee, L.; Bernstein, R. Resonance-Enhanced Multi-Photon Ionization and Fragmentation of Molecular-Beams-NO, I2, benzene, and butadiene. J. Chem. Phys. 1979, 71, 1359-1371. (8) Qian, Y.; Fang, L.; Li, S.; Zheng, H.; Wu, C. The Introduction of Experimental Study Method in MPID Mass Spectra. Chin. J. Chem. Phys. 1992, 5, 399-404. (9) Li, Y.-Y.; Qi, F. Recent Applications of Synchrotron VUV Photoionization Mass Spectrometry: Insight into Combustion Chemistry. Acc. Chem. Res. 2009, 43, 68-78. (10) Guo, H.-J.; Zhang, L.-D.; Deng, L.-L.; Jia, L.-Y.; Pan, Y.; Qi, F. Vacuum Ultraviolet Photofragmentation of Sarcosine: Photoionization Mass Spectrometric and Theoretical Insights. J. Phys. Chem. A 2010, 114, 3411-3417. (11) Zhu, Z.-X.; Wang, J.; Qiu, K.-Q.; Liu, C.-Y.; Qi, F.; Pan, Y. Note: A Novel Vacuum Ultraviolet Light Source Assembly with Aluminum-Coated Electrodes for Enhancing the Ionization Efficiency of Photoionization Mass Spectrometry. Rev. Sci. Instrum. 2014, 85, 046110. (12) Hua, L.; Wu, Q.-H.; Hou, K.-Y.; Cui, H.-P.; Chen, P.; Wang, W.-G.; Li, J.-H.; Li, H.-Y. Single Photon Ionization and Chemical Ionization Combined Ion Source Based on a Vacuum Ultraviolet Lamp for Orthogonal Acceleration Time-of-Flight Mass Spectrometry. Anal. Chem. 2011, 83, 5309-5316. (13) Chen, J.; Luo, Z.-X.; Fu, H.-B.; Yao, J.-N. Photoreactions of Porphyrins Initiated by Deep Ultraviolet Single Photons. J. Phys. Chem. A 2017, 121, 4626-4632. (14) Yuan, C.-Q.; Liu, X.-H.; Zeng, C.-H.; Zhang, H.-Y.; Jia, M.Y.; Wu, Y.-S.; Luo, Z.-X.; Fu, H.-B.; Yao, J.-N. All-Solid-State Deep Ultraviolet Laser for Single-Photon Ionization Mass Spectrometry. Rev. Sci. Instrum. 2016, 87, 024102. (15) Luo, Z.-X. Mass Spectrometry, Aliofkhazraei, M., Ed.; InTech, 2017. (16) Selbach, M.; Schwanhausser, B.; Thierfelder, N.; Fang, Z.; Khanin, R.; Rajewsky, N. Widespread Changes in Protein Synthesis Induced by MicroRNAs. Nature 2008, 455, 58-63. (17) Aebersold, R.; Mann, M. Mass Spectrometry-Based Proteomics. Nature 2003, 422, 198-207. (18) Ho, Y.; Gruhler, A.; Heilbut, A.; Bader, G. D.; Moore, L.; Adams, S.-L.; Millar, A.; Taylor, P.; Bennett, K.; Boutilier, K.; Yang, L.; Wolting, C.; Donaldson, I.; Schandorff, S.; Shewnarane, J.; Vo, M.; Taggart, J.; Goudreault, M.; Muskat, B.; Alfarano, C.; Dewar, D.; Lin, Z.; Michalickova, K.; Willems, A. R.; Sassi, H.; Nielsen, P. A.; Rasmussen, K. J.; Andersen, J. R.; Johansen, L. E.; Hansen, L. H.; Jespersen, H.; Podtelejnikov, A.; Nielsen, E.; Crawford, J.; Poulsen, V.; Sørensen, B. D.; Matthiesen, J.; Hendrickson, R. C.; Gleeson, F.; Pawson, T.; Moran, M. F.; Durocher, D.; Mann, M.; Hogue, C. W. V.; Figeys, D.; Tyers, M. Systematic Identification of Protein Complexes in Saccharomyces Cerevisiae by Mass Spectrometry. Nature 2002, 415, 180-183. (19) Zhou, L.; Bosscher, M.; Zhang, C.-S.; Özçubukçu, S.; Zhang, L.; Zhang, W.; Li, C. J.; Liu, J.-Z.; Jensen, M. P.; Lai, L.-H.; He, C. A Protein Engineered to Bind Uranyl Selectively and with Femtomolar Affinity. Nat. Chem. 2014, 6, 236-241. (20) Prival, M. J.; Mitchell, V. D.; Gomez, Y. P. Mutagenicity of a New Hair Dye Ingredient-4-Ethoxy-M-Phenylenediamine. Science 1980, 207, 907-908. (21) Hauser, A.; Penkert, M.; Hackenberger, C. P. R. Chemical Approaches to Investigate Labile Peptide and Protein Phosphorylation. Acc. Chem. Res. 2017, 50, 1883-1893.

Page 6 of 8

(22) Letokhov, V. S. Use of Lasers to Control Selective Chemical Reactions. Science 1973, 180, 451-458. (23) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, revision A.1; Gaussian, Inc.: Wallingford, CT. 2009. (24) Stephens, P. J.; Devlin, F. J.; Chabalowski, C. F.; Frisch, M. J. Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force-Fields. J. Phys. Chem. 1994, 98, 11623-11627. (25) Santos, A. F. L. O. M.; da Silva, M. A. V. R. Diaminobenzenes: An Experimental and Computational Study. J. Phys. Chem. B 2011, 115, 4939-4948. (26) Gonzalez, C.; Schlegel, H. B. An Improved Algorithm for Reaction Path Following. J. Chem. Phys. 1989, 90, 2154-2161. (27) Reed, A. E.; Weinstock, R. B.; Weinhold, F. Natural Population Analysis. J. Chem. Phys. 1985, 83, 735-746. (28) Glendening, E. D.; Landis, C. R.; Weinhold, F. Natural Bond Orbital Methods. WIREs Compu Mol Sci 2012, 2, 1-42. (29) Lu, T.; Chen, F.-W. Multiwfn: A Multifunctional Wavefunction Analyzer. J. Comput. Chem. 2012, 33, 580-592. (30) Russell, T. P.; Deline, V. R.; Dozier, W. D.; Felcher, G. P.; Agrawal, G.; Wool, R. P.; Mays, J. W. Direct Observation of Reptation at Polymer Interfaces. Nature 1993, 365, 235-237. (31) Joh, N. H.; Min, A.; Faham, S.; Whitelegge, J. P.; Yang, D.; Woods, V. L.; Bowie, J. U. Modest Stabilization by Most HydrogenBonded Side-Chain Interactions in Membrane Proteins. Nature 2008, 453, 1266-1270. (32) Ashfold, M. N. R.; Cronin, B.; Devine, A. L.; Dixon, R. N.; Nix, M. G. D. The role of πσ* Excited States in the Photodissociation of Heteroaromatic Molecules. Science 2006, 312, 1637-1640. (33) You, M.-X.; Lyu, Y.-F.; Han, D.; Qiu, L.-P.; Liu, Q.-L.; Chen, T.; Wu, C.-C. S.; Peng, L.; Zhang, L.-Q.; Bao, G.; Tan, W.-H. DNA Probes for Monitoring Dynamic and Transient Molecular Encounters on Live Cell Membranes. Nat. Nanotech. 2017, 12, 453-459. (34) Hossain, M. A.; Wade, J. D. Novel Methods for the Chemical Synthesis of Insulin Superfamily Peptides and of Analogues Containing Disulfide Isosteres. Acc. Chem. Res. 2017, 50, 2116-2127. (35) McDonald, M.; McGuyer, B. H.; Apfelbeck, F.; Lee, C. H.; Majewska, I.; Moszynski, R.; Zelevinsky, T. Photodissociation of Ultracold Diatomic Strontium Molecules with Quantum State Control. Nature 2016, 535, 122-126. (36) Kling, M.; Siedschlag, C.; Verhoef, A.; Khan, J.; Schultze, M.; Uphues, T.; Ni, Y.; Uiberacker, M.; Drescher, M.; Krausz, F.; Vrakking, M. J. Control of electron localization in molecular dissociation. Science 2006, 312, 246-248. (37) Park, S. T.; Kim, S. K.; Kim, M. S. Observation of Conformation-Specific Pathways in the Photodissociation of 1Iodopropane Ions. Nature 2002, 415, 306-308. (38) http://webbook.nist.gov/chemistry/ie-ser/. (39) Shen, Z.-T.; Ma, H.-T.; Zhang, C.-F.; Fu, M.-K.; Wu, Y.-N.; Bian, W.-S.; Cao, J.-W. Dynamical Importance of Van Der Waals Saddle and Excited Potential Surface in C(1D)+D2 ComplexForming Reaction. Nat. Commun. 2017, 8, 14094. (40) Wu, F.; Ren, Y.-H.; Bian, W.-S. The Hydrogen Tunneling Splitting in Malonaldehyde: A Full-Dimensional Time-Independent Quantum Mechanical Method. J. Chem. Phys. 2016, 145, 074309.

ACS Paragon Plus Environment

6

Page 7 of 8 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry (41) Ren, Y.-H.; Bian, W.-S. Mode-Specific Tunneling Splittings for a Sequential Double-Hydrogen Transfer Case: An Accurate Quantum Mechanical Scheme. J. Phys. Chem. Lett. 2015, 6, 18241829. (42) Koch, M.; Pagan, M.; Persson, M.; Gawinkowski, S.; Waluk, J.; Kumagai, T. Direct Observation of Double Hydrogen Transfer via Quantum Tunneling in a Single Porphycene Molecule on a Ag(110) Surface. J. Am. Chem. Soc. 2017, 139, 12681-12687. (43) Hill, C. M.; Kim, J.; Bodappa, N.; Bard, A. J. Electrochemical Nonadiabatic Electron Transfer via Tunneling to Solution Species through Thin Insulating Films. J. Am. Chem. Soc. 2017, 139, 61146119. (44) Rakitzis, T. P.; van den Brom, A. J.; Janssen, M. H. M. Directional Dynamics in the Photodissociation of Oriented Molecules. Science 2004, 303, 1852-1854. (45) Khusnutdinova, J. R.; David, Y. B.; Milstein, D. Direct Deamination of Primary Amines by Water to Produce Alcohols. Angew. Chem., Int. Ed. 2013, 52, 1-5. (46) Poyner, R.; Anderson, M.; Bandarian, V.; Cleland, W.; Reed, G. Probing Nitrogen-Sensitive Steps in the Free-Radical-Mediated Deamination of Amino Alcohols by Ethanolamine Ammonia-Lyase. J. Am. Chem. Soc. 2006, 128, 7120-7121.

ACS Paragon Plus Environment

7

Analytical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 8

for TOC only

ACS Paragon Plus Environment

8