Network


Latest external collaboration on country level. Dive into details by clicking on the dots.

Hotspot


Dive into the research topics where Timothy W. Marin is active.

Publication


Featured researches published by Timothy W. Marin.


Journal of Physical Chemistry Letters | 2014

Charge Trapping in Photovoltaically Active Perovskites and Related Halogenoplumbate Compounds

Ilya A. Shkrob; Timothy W. Marin

Halogenoplumbate perovskites (MeNH3PbX3, where X is I and/or Br) have emerged as promising solar panel materials. Their limiting photovoltaic efficiency depends on charge localization and trapping processes that are presently insufficiently understood. We demonstrate that in halogenoplumbate materials the holes are trapped by organic cations (that deprotonate from their oxidized state) and Pb(2+) cations (as Pb(3+) centers), whereas the electrons are trapped by several Pb(2+) cations, forming diamagnetic lead clusters that also serve as color centers. In some cases, paramagnetic variants of these clusters can be observed. We suggest that charge separation in the halogenoplumbates resembles latent image formation in silver halide photography. Electron and hole trapping by lead clusters in extended dislocations in the bulk may be responsible for accumulation of trapped charge observed in this photovoltaic material.


Journal of Physical Chemistry A | 2010

Carbonate Radical Formation in Radiolysis of Sodium Carbonate and Bicarbonate Solutions up to 250 °C and the Mechanism of its Second Order Decay

Kyle Haygarth; Timothy W. Marin; Ireneusz Janik; Kotchaphan Kanjana; Christopher M. Stanisky; David M. Bartels

Pulse radiolysis experiments published several years ago (J. Phys. Chem. A, 2002, 106, 2430) raised the possibility that the carbonate radical formed from reaction of *OH radicals with either HCO(3)(-) or CO(3)(2-) might actually exist predominantly as a dimer form, for example, *(CO(3))(2)(3-). In this work we re-examine the data upon which this suggestion was based and find that the original data analysis is flawed. A major omission of the original analysis is the recombination reaction *OH + *CO(3)(-) --> HOOCO(2)(-). Upon reanalysis of the published data for sodium bicarbonate solutions and analysis of new transient absorption data we are able to establish the rate constant for this reaction up to 250 degrees C. The mechanism for the second-order self-recombination of the carbonate radical has never been convincingly demonstrated. From a combination of literature data and new transient absorption experiments in the 1-400 ms regime, we are able to show that the mechanism involves pre-equilibrium formation of a C(2)O(6)(2-) dimer, which dissociates to CO(2) and peroxymonocarbonate anion: *CO3(-)+*CO3(-)<-->C2O6(2-)-->CO2+O2COO(2-) *CO3(-) reacts with the product peroxymonocarbonate anion, producing a peroxymonocarbonate radical *O2COO(-), which can also recombine with the carbonate radical: *CO3(-)+CO4(2-)-->*CO4(-)+CO3(2-) *CO3(-)+CO4(-)-->C2O7(2-).


Journal of Physical Chemistry B | 2011

Radiation and radical chemistry of NO3(-), HNO3, and dialkylphosphoric acids in room-temperature ionic liquids.

Ilya A. Shkrob; Timothy W. Marin; Sergey D. Chemerisov; James F. Wishart

Hydrophobic room-temperature ionic liquids (ILs) are considered as possible replacements for molecular diluents for nuclear separations, as well as the basis of new separations processes. Such applications may put the solvents both in high radiation fields and in contact with aqueous raffinate containing 1-6 M HNO(3). In this study, we address the effect of the extracted nitrate and nitric acid on the radiation chemistry of hydrophobic ILs composed of 1-alkyl-3-methylimidazolium cations (and closely related systems). We demonstrate that the nitrate anion competes with the solvent cation as an electron scavenger, with most of the primary radical species converted to NO(3)(•2-) and NO(2)(•) that initiate a complex sequence of radical reactions. In hydrophobic ILs equilibrated with 3 M HNO(3), nearly all electrons released by the ionizing radiation are converted to NO(2)(•). While the reductive pathway is strongly affected by the nitrate and there is also some N-O bond scission via direct excitation, the extent of interference with the oxidative pathway is relatively small; the cation damage is not dramatically affected by the presence of nitrate as most of the detrimental radiolytic products are generated via the oxidative pathway. These results are contrasted with the behavior of dialkylphosphoric acids (a large class of extraction agents for trivalent metal ions). We demonstrate that IL solvents protect these dialkylphosphoric acids against radiation-induced dealkylation.


Chemical Physics Letters | 2003

Reaction of OH* radicals with H2 in sub-critical water

Timothy W. Marin; Charles D. Jonah; David M. Bartels

Abstract The rate constant for the reaction of hydroxyl radicals (OH ) with hydrogen in aqueous solutions has been measured from 200 to 350 °C by competition kinetics using nitrobenzene as a competing OH * scavenger. Measurements below 250 °C agree with previous results in other laboratories. At higher temperatures, the rate constant undershoots an extrapolation of the Arrhenius plot and actually decreases in value above 275 °C. At 350 °C, the measured rate constant is more than a factor of 5 below the Arrhenius extrapolation. We propose an explanation based largely on the hydrophobic solvation properties of the H 2 molecule.


Journal of Physical Chemistry B | 2013

Radiation Stability of Cations in Ionic Liquids. 2. Improved Radiation Resistance through Charge Delocalization in 1-Benzylpyridinium

Ilya A. Shkrob; Timothy W. Marin; Jasmine L. Hatcher; Andrew R. Cook; Tomasz Szreder; James F. Wishart

Hydrophobic room-temperature ionic liquids (ILs) hold promise as replacements for molecular diluents for processing of used nuclear fuel as well as for the development of alternative separations processes, provided that the solvent can be made resistant to ionizing radiation. We demonstrate that 1-benzylpyridinium cations are uniquely suited as radiation resistant cations due to the occurrence of charge delocalization in both their reduced and oxidized forms in the ILs. It is suggested that the excess electron and hole in the latter ILs are stabilized through the formation of π-electron sandwich dimers that are analogous to the well-known dimer radical cations of aromatic molecules. This charge delocalization dramatically reduces the yield of fragmentation by deprotonation and the loss of benzyl arms, thereby providing a synthetic path to radiation resistant ILs that are suitable for nuclear fuel processing.


Journal of Physical Chemistry B | 2013

Radiation Stability of Cations in Ionic Liquids. 1. Alkyl and Benzyl Derivatives of 5-Membered Ring Heterocycles

Ilya A. Shkrob; Timothy W. Marin; Huimin Luo; Sheng Dai

In order to use hydrophobic room temperature ionic liquids (ILs) as diluents in nuclear separations for advanced fuel cycles, it is desirable to reduce the breakdown of the constituent ions caused by ionizing radiation. In this series, we survey radiation stability for different classes of organic cations used to formulate ILs. While radiolysis of 1-alkyl-3-methylimidazolium cations has been extensively studied, there have not been complementary studies of 1-benzyl derivatives of these cations nor organic cations that are derived from 5-membered ring heterocycles other than imidazole, such as 1,2,4-triazole and thiazole. In part 1, we establish the fragmentation pathways for such cations and quantify product yields for 2.5 MeV electron beam radiolysis of these aromatic cations. Radiolytic reduction of 1-benzyl cations derived from imidazole and 1,2,4-triazole is shown to cause the elimination of benzyl radicals from their electron adducts, whereas this elimination does not occur in the thiazole derivatives due to stabilization of the excess electron as a dimer radical cation. No such elimination occurs in the corresponding 1-alkyl derivatives, but there is significant C-N and C-C bond fragmentation in the aliphatic arms. As such bond dissociation reactions are irreversible, there is significant loss of 1-alkyl cations during the radiolysis. For 1-benzyl derivatives, this electronic excitation causes fragmentation of the C-N bonds in the benzyl arms with the release of the corresponding base and the benzyl carbocation that can subsequently attack this base or add to another cation. Such systems exhibit more predictable fragmentation patterns and yield well-defined products; some of the systems also exhibit increased radiation resistance. The C-N bond fragmentation in the reduced cations can be further suppressed through the use of appropriate electron scavengers, including acids and aromatic imide anions. The observed trends are rationalized using density functional theory calculations, and the implications of these results for the design of IL diluents are examined.


Journal of Physical Chemistry A | 2013

Photo- and radiation-chemistry of halide anions in ionic liquids.

Ilya A. Shkrob; Timothy W. Marin; Robert A. Crowell; James F. Wishart

One- and two- photon excitation of halide anions (X(-)) in polar molecular solvents results in electron detachment from the dissociative charge-transfer-to-solvent state; this reaction yields a solvated halide atom and a solvated electron. How do such photoreactions proceed in ionic liquid (IL) solvents? Matrix isolation electron paramagnetic resonance (EPR) spectroscopy has been used to answer this question for photoreactions of bromide in aliphatic (1-butyl-1-methylpyrrolidinium) and aromatic (1-alkyl-3-methyl-imidazolium) ionic liquids. In both classes of ILs, the photoreaction (both 1- and 2-photon) yields bromine atoms that promptly abstract hydrogen from the alkyl chains of the IL cation; only in concentrated bromide solutions (containing >5-10 mol % bromide) does Br2(-•) formation compete with this reaction. In two-photon excitation, the 2-imidazolyl radical generated via the charge transfer promptly eliminates the alkyl arm. These photolytic reactions can be contrasted with radiolysis of the same ILs, in which large yield of BrA(-•) radicals was observed (where A(-) is a matrix anion), suggesting that solvated Br(•) atoms do not occur in the ILs, as such a species would form three-electron σ(2)σ(*1) bonds with anions present in the IL. It is suggested that chlorine and bromine atoms abstract hydrogen faster than they form such radicals, even at cryogenic temperatures, whereas iodine mainly forms such bound radicals. These XA(-•) radicals convert to X2(•-) radicals in a reaction with the parent halide anion. Ramifications of these observations for photodegradation of ionic liquids are discussed.


Journal of Physical Chemistry B | 2011

Hydrogen-Bonding Interactions and Protic Equilibria in Room-Temperature Ionic Liquids Containing Crown Ethers

Timothy W. Marin; Ilya A. Shkrob; Mark L. Dietz

Nuclear magnetic resonance (NMR) spectroscopy has been used to study hydrogen-bonding interactions between water, associated and dissociated acids (i.e., nitric and methanesulfonic acids), and the constituent ions of several water-immiscible room-temperature ionic liquids (ILs). In chloroform solutions also containing a crown ether (CE), water molecules strongly associate with the IL ions, and there is rapid proton exchange between these bound water molecules and hydronium associated with the CE. In neat ILs, the acids form clusters differing in their degree of association and ionization, and their interactions with the CEs are weak. The CE can either promote proton exchange between different clusters in IL solution when their association is weak or inhibit such exchange when the association is strong. Even strongly hydrophobic ILs are shown to readily extract nitric acid from aqueous solution, typically via the formation of a 1:1:1 {H(3)O(+)•CE}NO(3)(-) complex. In contrast, the extraction of methanesulfonic acid is less extensive and proceeds mainly by IL cation-hydronium ion exchange. The relationship of these protic equilibria to the practical application of hydrophobic ILs (e.g., in spent nuclear fuel reprocessing) is discussed.


Journal of Physical Chemistry B | 2013

Radiation stability of cations in ionic liquids. 3. Guanidinium cations.

Ilya A. Shkrob; Timothy W. Marin; Jason R. Bell; Huimin Luo; Sheng Dai

Due to their superb structural versatility, guanidinium cations find increasing use as constituent ions in room temperature ionic liquids (ILs). This versatility allows fine-tuning of hydrophobicity, which is an important concern for the use of ILs as diluents for metal ion separations. However, the presence of six C-N bonds in such cations poses a question, whether the guanidinium based ILs can be considered as diluents for nuclear separations, given that the radiation emitted by the decaying radionuclides can break these relatively weak bonds over the use cycle of the solvent. As nothing is presently known about the radiolytic stability of the guanidinium cations, we addressed this question using 2-dialkylamino-1,3-dimethylimidazolidine based cations (R = Et, Pr, and Bu) as a representative model for the entire class of such cations, and assessed their stability in 2.5 MeV electron beam radiolysis. Electron paramagnetic resonance spectroscopy, nuclear magnetic resonance spectroscopy, and mass spectrometry have been used to establish chemical mechanisms for radiation damage in guanidinium cations. Our conclusion is that radiation stability of these cations is not significantly different from that of more familiar aliphatic and aromatic IL cations. In fact, these cations yield exceptionally stable radicals, and fragmentation occurs only in their radiolytically generated excited states. The predominant chemical pathway for the cation decomposition is the elimination of their aliphatic arms, with radiolytic yields of 0.65 to 1.06 to 1.46 per 100 eV from R = Et to R = Bu, respectively. The total loss of the parent cation was estimated as 2.62, 1.65, and 1.98 species per 100 eV. While this attrition is not negligible, it is comparable to other organic cations that have fewer fissile C-N bonds. Many of the products are either modified guanidinium ions or protonated bases that are not expected to interfere with radionuclide separations.


Separation Science and Technology | 2011

Extraction and Reductive Stripping of Pertechnetate from Spent Nuclear Fuel Waste Streams

Ilya A. Shkrob; Timothy W. Marin; Dominique C. Stepinski; George F. Vandegrift; John V. Muntean; Mark L. Dietz

An approach directed at rapid sequestration and disposal of technetium-99 from UREX (uranium extraction) liquid waste streams is presented. This stream is generated during reprocessing of light-water-reactor spent fuel to recycle the actinides and separate fission products for waste disposal. U and Tc are co-extracted from a nitric acid solution using tri-n-butylphosphate in dodecane, so that Tc(VII) is present in the strip solution after the actinide separations. The goal is to separate uranyl from the pertechnetate in this U-Tc stream and then sequester Tc in the metallic form. Our approach is based on reductive stripping of pertechnetate either from aqueous solution (for column extractions) or organic solvents (for liquid-liquid extractions). In both of these methods, metallic zinc in the presence of formic acid serves as a reducing agent, and 99Tc is recovered as a co-precipitate of Zn(II) hydroxide and hydrous Tc(IV) oxide, with a Zn:Tc ratio between 1:1 and 2:1 mol/mol. This solid residue can be reduced to a Zn-Tc alloy by high temperature (500–700°C) hydrogenation, and the resulting heterophase alloy can be added to a metallic Fe-Zr-Mo waste form that is processed at 1600°C, with subsequent loss of Zn by evaporation. Alternatively, Zn and Tc can be separated and 99Tc sequestered as NH4TcO4 for further reduction to Tc(0) metal. The aqueous Zn reduction process removes ∼90% of 99Tc per cycle. The nonaqueous Zn reduction in 1:1 methanol – formic acid removes 60–70% of 99Tc per cycle, depending on the extracting agent (such as a tetraalkylammonium nitrate). The extracting agent is recycled in the process. The pertechnetate is extracted from the aqueous phase into 1,2-dichloroethane, which is removed by evaporation and reused. The residue is either calcined and steam reformed to Tc(0) or processed by the nonaqueous Zn reduction method. These methods can be used not only to remove the pertechnetate from the U-Tc product stream, but also to sequester the pertechnetate from aqueous waste streams generated through the processes described in this paper, thereby closing the cycle. The same approaches can be used to close the 99Tc cycle for other methods that are currently being developed at Los Alamos and Argonne National Laboratories.

Collaboration


Dive into the Timothy W. Marin's collaboration.

Top Co-Authors

Avatar

Ilya A. Shkrob

Argonne National Laboratory

View shared research outputs
Top Co-Authors

Avatar

David M. Bartels

Argonne National Laboratory

View shared research outputs
Top Co-Authors

Avatar

Charles D. Jonah

Argonne National Laboratory

View shared research outputs
Top Co-Authors

Avatar

James F. Wishart

Brookhaven National Laboratory

View shared research outputs
Top Co-Authors

Avatar

Ireneusz Janik

University of Notre Dame

View shared research outputs
Top Co-Authors

Avatar

David M. Bartels

Argonne National Laboratory

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Daniel P. Abraham

Argonne National Laboratory

View shared research outputs
Top Co-Authors

Avatar

Huimin Luo

Oak Ridge National Laboratory

View shared research outputs
Researchain Logo
Decentralizing Knowledge