Evan Perez’s research while affiliated with University of Utah and other places

What is this page?


This page lists works of an author who doesn't have a ResearchGate profile or hasn't added the works to their profile yet. It is automatically generated from public (personal) data to further our legitimate goal of comprehensive and accurate scientific recordkeeping. If you are this author and want this page removed, please let us know.

Publications (21)


Structural Determination of Zn2+, Cu2+, and Fe2+ Complexed with Glutathione by IRMPD Spectroscopy and Complimentary ab Initio Calculations
  • Article

November 2024

·

5 Reads

Physical Chemistry Chemical Physics

Samantha K. Walker

·

Amanda R. Bubas

·

Brandon C. Stevenson

·

[...]

·

Peter B. Armentrout

Glutathione is a biologically abundant and redox active tripeptide that serves to protect cells from oxidative stress and rid the body of toxic heavy metals. The present study examines the...





Structures and Chemical Rearrangements of Benzoate Derivatives Following Gas Phase Decarboxylation

September 2022

·

19 Reads

·

5 Citations

Journal of the American Society for Mass Spectrometry

Decarboxylation of carboxylate ions in the gas phase provides a useful window into the chemistry displayed by these reactive carbanion intermediates. Here, we explore the species generated by decarboxylation of two benzoate derivatives: 2-formylbenzoate (2FBA) and 2-benzoylbenzoate (2BBA). The nascent product anions are transferred to a cryogenic ion trap where they are cooled to ∼15 K and analyzed by their pattern of vibrational bands obtained with IR photodissociation spectroscopy of weakly bound H2 molecules. The structures of the quenched species are then determined by comparison of these spectra with those predicted by electronic structure calculations for local minima on the potential energy surface. The 2-phenide carbanion generated by decarboxylation of 2FBA occurs in two isomeric forms that differ in the orientation of the formyl group, both of which yield a very large (∼110 cm-1) redshift in the stretching frequency of the H2 molecule attached to the anionic carbon center. Although calculated to be a local minimum, the analogous 2-phenide species could not be isolated upon decarboxylation of 2BBA. Rather, the anionic product adopts a ring-closed structure, indicating efficient nucleophilic attack on the pendant phenyl group by the nascent phenide. The barrier for ring closing is evaluated with electronic structure calculations.


ANIONIC REARRANGEMENTS FOLLOWING DECARBOXYLATION OF BENZOPHENONE DERIVATIVES WITH CRYOGENIC IR SPECTROSCOPY
  • Conference Paper
  • Full-text available

June 2022

·

16 Reads

Download


of the vibrational spectrum and band assignments in H2‐tagged HOx. The blue νsym(O-C-O) and purple νasym(O-C-O) represent the symmetric and asymmetric stretches of the oxalate anion nominal −CO2⁻, respectively, with the dashed vertical arrow indicating the νasym(O-C-O) fundamental of the free carboxylate group on propionate for reference.[12] The diffuse band highlighted in the orange box (νboundO-H) is derived from the OH stretch of the oxalate anion, with a representative structure shown in the insets. The grey labels νooPOHbend and νiPOHbend are the out‐of‐plane and in‐plane OH bend, respectively. The green label ν(C=O) indicates the carbonyl stretch of the oxalate anion. The spectrum is adapted with permission from Ref. [19] Copyright 2015, American Chemical Society.
Evolution of bands assigned to HOx upon complexation to [M2dcpm2]²⁺ (M=Cu, Ag). A) Vibrational spectrum of H2‐tagged HOx, reproduced and modified from reference with permission.[19] The labels and color scheme are the same as in Figure 1. B–C) N2‐tagged vibrational spectra of [CuCu]⁺ and [AgAg]⁺, with peak labels from C1–C12 and A1–A16, respectively, for ease of reference in text. The assignments for these peaks are shown in Tables 1 and 2, see below. The * on A7 in C) indicates the blueshifted νasym(O-C-O) of the [AgAg]⁺ complex with the HOx adduct bound in a side‐on fashion (sμ or smμ). The sμHb binding motif is displayed inset in a dashed square. D) IR‐IR hole burning spectra probing at C1 (blue, 3551 cm⁻¹) and C2 (green, 3467 cm⁻¹) positions. The dip spectra separate the heterogeneous spectrum to reveal patterns of two isomers, both with the HOx adduct bound in an end‐on fashion (eb), but one with a free‐OH (Hf) and the other with a bound‐OH (Hb) (blue and green traces, respectively). The ebHf and ebHb binding motifs are shown inset, boxed in blue and green rectangles. Vertical dashed lines show the spectral shifts of various HOx vibrational transitions upon complexation with [M2dcpm2]²⁺.
Lowest three structures calculated for the [CuCu]⁺ complex. Atoms are coded by color (brown for Cu, orange for P, red for O, grey for C, white for H); isomers are coded for their structures as follows: e: end‐on, s: side‐on, b: bidentate, m: monodentate, μ: one O‐atom coordinates between two metal, Hf: free OH, Hb: bonded OH. For a better overview, H atoms of the cyclohexyl groups are omitted. Relative energies are given for each isomer. More isomers are shown in Figure S1. Level of theory: PBE0/def2TZVP.
Comparison of the experimental [CuCu]⁺ spectrum with those calculated for the three lowest energy isomers recovered in DFT calculations shown in Figure 3. The calculated stick spectra (red) are convoluted with a Lorentzian curve (black, fwhm=10 cm⁻¹). Level of theory: PBE0/def2TZVP; frequency scaled: 0.961 (900–2200 cm⁻¹), 0.946 (2600–3800 cm⁻¹).
Calculated low‐lying structures of [AgAg]⁺. Atoms are coded by color (silver for Ag, orange for P, red for O, grey for C, white for H); isomers are coded for their structures as follows: e: end‐on, s: side‐on, b: bidentate, m: monodentate, μ: one O‐atom coordinates between two metal atoms, Hf: free OH, Hb: bonded OH. For a better overview, H atoms of cyclohexyl group are omitted. Relative energies are given for each isomer. More isomers are shown in Figure S1. Level of theory: PBE0/def2TZVP.

+4

On the Hydrogen Oxalate Binding Motifs onto Dinuclear Cu and Ag Metal Phosphine Complexes

October 2021

·

74 Reads

·

3 Citations

We report the binding geometries of the isomers that are formed when the hydrogen oxalate ((CO2)2H=HOx) anion attaches to dinuclear coinage metal phosphine complexes of the form [M1M2dcpm2(HOx)]⁺ with M=Cu, Ag and dcpm=bis(dicyclohexylphosphino)methane, abbreviated [MM]⁺. These structures are established by comparison of isomer‐selective experimental vibrational band patterns displayed by the cryogenically cooled and N2‐tagged cations with DFT calculations of the predicted spectra for various local minima. Two isomeric classes are identified that feature either attachment of the carboxylate oxygen atoms to the two metal centers (end‐on docking) or attachment of oxygen atoms on different carbon atoms asymmetrically to the metal ions (side‐on docking). Within each class, there are additional isomeric variations according to the orientation of the OH group. This behavior indicates that HOx undergoes strong and directional coordination to [CuCu]⁺ but adopts a more flexible coordination to [AgAg]⁺. Infrared spectra of the bare ions, fragmentation thresholds and ion mobility measurements are reported to explore the behaviors of the complexes at ambient temperature.


Determination of the SmO + bond energy by threshold photodissociation of the cryogenically cooled ion

October 2021

·

22 Reads

·

22 Citations

The SmO⁺ bond energy has been measured by monitoring the threshold for photodissociation of the cryogenically cooled ion. The action spectrum features a very sharp onset indicating a bond energy of 5.596 {plus minus} 0.004 eV. This value, when combined with the literature value of the samarium ionization energy, indicates that the chemi-ionization reaction of atomic Sm with atomic oxygen is endothermic by 0.048 {plus minus} 0.004 eV, which has important implications on the reactivity of Sm atoms released into the upper atmosphere. The SmO⁺ ion was prepared by electrospray ionization followed by collisional breakup of two different precursors and characterized by the vibrational spectrum of the He-tagged ion. The UV photodissociation threshold is similar for the 10 K bare ion and the He tagged ion, which rules out the possible role of metastable electronically excited states. Reanalysis and remeasurement of previous reaction kinetics experiments that are dependent on D0(SmO⁺) are included, bringing all experimental results in accord.



Citations (9)


... Article likely due to oxidation of NA. 20 Yet, a simultaneous growth of a negative absorption band at 1704 cm −1 , also attributed to the stretching mode of C�O functional groups, indicates a loss of mass due to fractionation or decarboxylation of the complex sample, leading to VOC and CO 2 formation. 33,55,56 Slight positive absorptions at 1404 and 1370 cm −1 , corresponding to the bending modes of aldehydic C−H and O−H product functional groups, further support these observations. 20 These simultaneous processes of oxidation and decarboxylation offset one another, resulting in no measurable mass change in the QCM. Figure 7B shows the vibrational spectroscopy results of 1:5 m-DOM/NA thin film, where distinct features of oxygenated product formation are identified. ...

Reference:

Photooxidation of Nonanoic Acid by Molecular and Complex Environmental Photosensitizers
Structures and Chemical Rearrangements of Benzoate Derivatives Following Gas Phase Decarboxylation
  • Citing Article
  • September 2022

Journal of the American Society for Mass Spectrometry

... eV lower than the experiment, whereas MP2 is about 0.5 eV too high. The D 0 (SmO + ) values given by MP2 and CCSD(T) are about 0.4 eV under the experimental value of Lachowicz et al. 7 Combined, these discrepancies lead to CCSD(T) reproducing the enthalpy of reaction 4 ΔrH 4 better than the MP2 method (∼0.25 vs ∼0.59 e V lower than the experiment, respectively). The prediction of the energy difference between the ground state and the first excited state of Sm + , E(Sm + , 6 F), is accurately predicted by both MP2 and CCSD(T). ...

Determination of the SmO + bond energy by threshold photodissociation of the cryogenically cooled ion
  • Citing Article
  • October 2021

... The À OH peak (3450 cm À 1 ) of the chiral ligand disappeared in the Raman spectrum, confirming its participation in the reaction. [11] The powder X-ray diffraction patterns showed two clear peaks at 36.1 and 44.7 eV ( Figure S4a). This indicates that chiral Ag 2 Se:Nd/Yd/Er NPs had a body-centered cubic phase structure (JCPDS No. 27-0619). ...

On the Hydrogen Oxalate Binding Motifs onto Dinuclear Cu and Ag Metal Phosphine Complexes

... In this Feature Article, the advantages of these techniques are discussed. In defining the scope of this perspective, IR PD techniques-such as infrared multiphoton photodissociation (IRMPD) [18][19][20][21][22][23][24][25] and infrared photodissociation (IRPD) [26][27][28][29][30] -and m/z selected photoelectron (PE) spectroscopy [31][32][33][34] are not included in detail, with the exception of select examples. The interested reader is directed to these aforelisted references. ...

Characterization of the non-covalent docking motif in the isolated reactant complex of a double proton-coupled electron transfer reaction with cryogenic ion spectroscopy
  • Citing Article
  • June 2020

... Jestilä J. S. et. al [53] showed that such intermediates have different structures and length of M-O bonds. It was shown that the stronger interactions of Li, Na, and K with the oxygens in the oxalate fragment could lead to the elongation of the C-C bond, and, probably, its faster rupture during FeC 2 O 4 decomposition. ...

Characterization of the alkali metal oxalates (MC 2 O 4 − ) and their formation by CO 2 reduction via the alkali metal carbonites (MCO 2 − )
  • Citing Article
  • March 2020

Physical Chemistry Chemical Physics

... [33] By coupling it with infrared photodissociation (IRPD) spectroscopy, we can characterize the structure of the intermediates in a controlled environment free from external interference. [34][35][36][37][38] This approach has already yielded molecular-level insights into reactive intermediates in electrocatalytic O 2 and CO 2 reduction reactions. [32,39,40] Here, we apply VESI-MS to investigate the reaction mechanism of homogeneous electrocatalytic ammonia oxidation by [Ru II (NH 3 ...

Integration of High-Resolution Mass Spectrometry with Cryogenic Ion Vibrational Spectroscopy

Journal of the American Society for Mass Spectrometry

... The process is reminiscent of the fragmentation pathway by loss of an acetyloxyl radical from a gaseous uranyl complex, thus undergoing reduction from U VI O 2 2+ to U V O 2 + [60]. The formation of reduced species after CID activation has indeed been reported in the breakdown behavior of deprotonated uranyl-containing complexes [61][62][63], while the corresponding protonated species showed no tendency to undergo reduction reactions [62]. ...

Formation of [U V OF 4 ] − by collision-induced dissociation of a [U VI O 2 (O 2 )(O 2 C-CF 3 ) 2 ] − precursor
  • Citing Article
  • November 2017

International Journal of Mass Spectrometry

... More recent experiments by our group have demonstrated that the 2-D, linear ion trap (LIT) can provide access to fragmentation pathways and reactions not observed in earlier studies with 3-D ion traps [51][52][53][54][55]. For example, our past studies of the dissociation behavior of gas-phase actinyl complexes using a 3-D ion trap were complicated by high yields of product ions obviously generated by collisions with background H 2 O. Relatively high levels of H 2 O in the ion trap (ca. 10 −6 Torr) create hydrated ions or lead to charge reduction reactions that form products such as [U VI O 2 (OH)] + and [U V O 2 ] + and larger hydrated complexes containing these cations [6,7,43]. ...

Collision-induced dissociation of uranyl-methoxide and uranyl-ethoxide cations: Formation of UO2 H(+) and uranyl-alkyl product ions
  • Citing Article
  • July 2016

Rapid Communications in Mass Spectrometry