Josef Michl

Find an error

Name: Michl, Josef
Organization: University of Colorado , USA
Department: Department of Chemistry and Biochemistry
Title: Professor(PhD)

TOPICS

Co-reporter:Jan Merna, Petr Vlček, Victoria Volkis, and Josef Michl
Chemical Reviews 2016 Volume 116(Issue 3) pp:771
Publication Date(Web):January 13, 2016
DOI:10.1021/acs.chemrev.5b00485
After a brief survey of conventional radical polymerization of alkenes, we review their Li+ catalyzed radical polymerization and their controlled radical polymerization. Emphasis is on homopolymerization, but related copolymerization of less activated monomers is mentioned as well.
Co-reporter:Cecile C. Givelet; Paul I. Dron; Jin Wen; Thomas F. Magnera; Matibur Zamadar; Klára Čépe; Hiroki Fujiwara; Yue Shi; Michael R. Tuchband; Noel Clark; Radek Zbořil
Journal of the American Chemical Society 2016 Volume 138(Issue 20) pp:6676-6687
Publication Date(Web):April 1, 2016
DOI:10.1021/jacs.5b12050
Proving the structures of charged metallacages obtained by metal ion coordination-driven solution self-assembly is challenging, and the common use of routine NMR spectroscopy and mass spectrometry is unreliable. Carefully determined diffusion coefficients from diffusion-ordered proton magnetic resonance (DOSY NMR) for six cages of widely differing sizes lead us to propose a structural reassignment of two molecular cages from a previously favored trimer to a pentamer or hexamer, and another from a trimer to a much higher oligomer, possibly an intriguing tetradecamer. In the former case, strong support for the reassignment to a larger cage is provided by an observation of a slow reversible transformation of the initially formed cage into a smaller but spectrally very similar one upon dilution. In the latter case, freeze-fracture transmission electron micrographs demonstrate that at least some of the solutions are colloidal, and high-resolution electron transmission and atomic force microscopy images are compatible with a tetradecamer but not a trimer. Comparison of solute partial molar volumes deduced from measurement of solution density with volumes anticipated from molecular models argues strongly against the presence of large voids (solvent vapor bubbles) in cages dissolved in nitromethane. The presence of bubbles was previously proposed in an attempt to account for the bilinear nature of the Eyring plot of the rate constant for pyridine ligand edge exchange reaction in one of the cages and for the unusual activation parameters in the high-temperature regime. An alternative interpretation is proposed now.
Co-reporter:Paul I. Dron;Ke Zhao;Ji&x159;í Kaleta;Yongqiang Shen;Jin Wen;Richard K. Shoemaker;Charles T. Rogers
Advanced Functional Materials 2016 Volume 26( Issue 31) pp:5718-5732
Publication Date(Web):
DOI:10.1002/adfm.201600437

A new class of rod-shaped strongly dipolar molecular rotors for insertion into channels of hexagonal tris(o-phenylenedioxy)cyclotriphosphazene (TPP) has been examined. Seven different 3,6-disubstituted pyridazines and one singly 3-substituted system have been prepared and studied by solid-state nuclear magnetic resonance (NMR), X-ray powder diffraction, and dielectric spectroscopy. NMR and X-ray diffraction both show that all but one of these molecular rotors form hexagonal bulk inclusion compounds with TPP. In-plane lattice parameters for the hexagonal phases increase with the size of the end group, which also controls the energy barriers for rotation of the pyridazine dipole. The barriers range from ≈4 kcal mol−1 for small or flexible end groups to less than 0.7 kcal mol−1 for 3-methylbicyclo[1.1.1]pent-1-yl end groups after annealing to 235 °C, and an interpretation of these differences is offered. Computer modeling of the relaxed TPP channels followed by density functional calculation of the environment for one of the rotors provides quantitative agreement with the observed barrier. The systems with the lowest rotational barriers show signs of collective behavior, discussed in terms of antiferroelectric intrachannel and ferroelectric interchannel dipole–dipole interactions. A Curie temperature of 22 K is deduced for 3,6-diadamant-1′-ylpyridazine, but no ordered dielectric phases are found. Conclusions have been drawn for improved rotor design.

Co-reporter:Filip Šembera, Jan Plutnar, Alexander Higelin, Zbyněk Janoušek, Ivana Císařová, and Josef Michl
Inorganic Chemistry 2016 Volume 55(Issue 8) pp:3797-3806
Publication Date(Web):March 28, 2016
DOI:10.1021/acs.inorgchem.5b02678
The anionic nitriles 1-R-12-NC–CB11H10– (R = H, CH3, I, COOH), 12-NC-1-H–CB11Me10–, and 12-NC-1-H–CB11F10– were prepared, and three of them were examined for complex formation with (Et3P)2Pt(II) and (Et3P)2Pd(II). Several stable internally charge-compensated zwitterionic complexes were obtained and characterized. RI-BP86/SV(P) calculations suggest that their dipole moments exceed 20 D. An attempt to measure the dipole moments in solution failed due to insufficient solubility in solvents of low polarity.
Co-reporter:Dr. Zdenek Havlas; Josef Michl
Israel Journal of Chemistry 2016 Volume 56( Issue 1) pp:96-106
Publication Date(Web):
DOI:10.1002/ijch.201500054

Abstract

The choice of chromophores and of their mutual geometrical arrangement for optimized singlet fission (SF) rates are considered. The electronic matrix element that enters the Fermi golden rule for the rate of SF is worked out algebraically for a simple model, but the density of states factor is not analyzed here. The model treats only the highest occupied and lowest unoccupied orbitals of the partners. It provides an approximate formula that requires only the knowledge of the expansion coefficients of these orbitals and of overlap integrals between atomic orbitals on the partners to obtain an estimate of the electronic matrix element. An illustrative application to a pair of ethylene molecules suggests that favored geometries will be those in which one of the AOs on the first ethylene overlaps with both AOs on the second ethylene, while the other AO on the first ethylene overlaps with at least one, and preferably both, AOs of the second ethylene as little as possible.

Co-reporter:Dr. Lubomír Pospí&x161;il;Dr. Ji&x159;í Kaleta; Josef Michl
ChemElectroChem 2016 Volume 3( Issue 2) pp:332-336
Publication Date(Web):
DOI:10.1002/celc.201500332

Abstract

In toluene containing LiCB11(CH3)12 as the supporting electrolyte, the oxidation potential of H2 on a platinized Pt electrode is a strong function of the composition of a basic buffer, lithium 2,2,6,6-tetramethylpiperidinate / 2,2,6,6-tetramethylpiperidine, and can approach the onset of the reduction of Li+. This suggests that only a moderate voltage might be needed to use H2 as a surrogate for Li and to take advantage of the specific chemical reactivity of the latter.

Co-reporter:Eva Kaletová; Anna Kohutová; Jan Hajduch; Jiří Kaleta; Zdeněk Bastl; Lubomír Pospíšil; Ivan Stibor; Thomas F. Magnera
Journal of the American Chemical Society 2015 Volume 137(Issue 37) pp:12086-12099
Publication Date(Web):September 1, 2015
DOI:10.1021/jacs.5b07672
Treatment of cleaned gold surfaces with dilute tetrahydrofuran or chloroform solutions of tetraalkylstannanes (alkyl = methyl, ethyl, n-propyl, n-butyl) or di-n-butylmethylstannyl tosylate under ambient conditions causes a self-limited growth of disordered monolayers consisting of alkyls and tin oxide. Extensive use of deuterium labeling showed that the alkyls originate from the stannane and not from ambient impurities, and that trialkylstannyl groups are absent in the monolayers, contrary to previous proposals. Methyl groups attached to the Sn atom are not transferred to the surface. Ethyl groups are transferred slowly, and propyl and butyl rapidly. In all cases, tin oxide is codeposited in submonolayer amounts. The monolayers were characterized by ellipsometry, contact angle goniometry, polarization modulated IR reflection absorption spectroscopy, X-ray photoelectron spectroscopy, and electrochemical impedance spectroscopy with ferrocyanide/ferricyanide, which revealed a very low charge-transfer resistance. The thermal stability of the monolayers and their resistance to solvents are comparable with those of an n-octadecanethiol monolayer. A preliminary examination of the kinetics of monolayer deposition from a THF solution of tetra-n-butylstannane revealed an approximately half-order dependence on the bulk solution concentration of the stannane, hinting that more than one alkyl can be transferred from a single stannane molecule. A detailed structure of the attachment of the alkyl groups is not known, and it is proposed that it involves direct single or multiple bonding of one or more C atoms to one or more Au atoms.
Co-reporter:Akin Akdag, Abdul Wahab, Pavel Beran, Lubomír Rulíšek, Paul I. Dron, Jiří Ludvík, and Josef Michl
The Journal of Organic Chemistry 2015 Volume 80(Issue 1) pp:80-89
Publication Date(Web):November 10, 2014
DOI:10.1021/jo502004r
The synthesis of covalent dimers in which two 1,3-diphenylisobenzofuran units are connected through one phenyl substituent on each is reported. In three of the dimers, the subunits are linked directly, and in three others, they are linked via an alkane chain. A seventh new compound in which two 1,3-diphenylisobenzofuran units share a phenyl substituent is also described. These materials are needed for investigations of the singlet fission process, which promises to increase the efficiency of solar cells. The electrochemical oxidation and reduction of the monomer, two previously known dimers, and the seven new compounds have been examined, and reversible redox potentials have been compared with results obtained from density functional theory. Although the overall agreement is satisfactory, some discrepancies are noted and discussed.
Co-reporter:Marco Cipolloni
The Journal of Physical Chemistry C 2015 Volume 119(Issue 16) pp:8805-8820
Publication Date(Web):March 19, 2015
DOI:10.1021/acs.jpcc.5b01960
We examine the fluorescence anisotropy of rod-shaped guests held inside the channels of tris(o-phenylenedioxy)cyclotriphosphazene (TPP) host nanocrystals, characterized by powder X-ray diffraction and solid state NMR spectroscopy. We address two issues: (i) are light polarization measurements on an aqueous colloidal solution of TPP nanocrystals meaningful, or is depolarization by scattering excessive? (ii) Can measurements of the rotational mobility of the included guests be performed at low enough loading levels to suppress depolarization by intercrystallite energy transfer? We find that meaningful measurements are possible and demonstrate that the long axis of molecular rods included in TPP channels performs negligible vibrational motion.
Co-reporter:Jin Wen; Zdenĕk Havlas
Journal of the American Chemical Society 2014 Volume 137(Issue 1) pp:165-172
Publication Date(Web):December 5, 2014
DOI:10.1021/ja5070476
Singlet fission offers an opportunity to improve solar cell efficiency, but its practical use is hindered by the limited number of known efficient materials. We look for chromophores that satisfy the desirable but rarely encountered adiabatic energy conditions, E(T2) – E(S0) > E(S1) – E(S0) ≈ 2[E(T1) – E(S0)], and are small enough to permit highly accurate calculations. We provide a rationale for the use of captodative biradicaloids, i.e., biradicals stabilized by direct interaction between their radical centers, which carry both an acceptor and a donor group. A computation of vertical excitation energies of 14 structures of this type by time-dependent density functional theory (TD-DFT) yielded 11 promising candidates. The vertical excitation energies from S0 and T1 were recalculated by complete-active-space second-order perturbation theory (CASPT2), and five of the compounds met the above energy criteria. Their adiabatic excitation energies from the S0 into the S1, S2, T1, and T2 excited states were subsequently calculated, and three of them look promising. For 2,3-diamino-1,4-benzoquinone, adiabatic E(T1) and E(S1) energies were close to optimal (1.12 and 2.23 eV above the S0 ground state, respectively), and for its more practical N-peralkylated derivative they were even lower (0.63 and 1.06 eV above S0, respectively). PCM/CASPT2 results suggested that the relative energies can be further tuned by varying the polarity of the environment.
Co-reporter:Lukáš Kobr, Ke Zhao, Yongqiang Shen, Richard K. Shoemaker, Charles T. Rogers, and Josef Michl
Crystal Growth & Design 2014 Volume 14(Issue 2) pp:559-568
Publication Date(Web):January 24, 2014
DOI:10.1021/cg4013608
A rod-shaped molecular rotor consisting of a p-terphenyl shaft attached to p-carborane whose antipodal position carries a dipolar 2,3-dichlorophenyl rotator forms an inclusion compound with hexagonal tris-o-phenylenedioxycyclotriphosphazene (TPP). Results of solid-state NMR spectroscopy, X-ray powder diffraction, dielectric loss spectroscopy, and density functional theory calculations lead us to propose that the whole molecule inserts into the TPP channels, with the rotator located in the outermost surface layer. Although the placement and alignment of the dipoles at the surface appear favorable, the sample does not exhibit collective behavior even at 7 K, presumably due to the relatively large barrier to rotation (∼8.6 kcal/mol). In incompletely annealed samples of the inclusion compound, some of the rotators protrude outside the surface and have a rotational barrier of ∼3.4 kcal/mol. In the inclusion compound of an analog in which the rotator is replaced with a methyl group, some of the methyl substituents are located inside the surface layer of TPP and others protrude above it.
Co-reporter:Matthew K. MacLeod and Josef Michl
The Journal of Physical Chemistry A 2014 Volume 118(Issue 45) pp:10538-10553
Publication Date(Web):July 11, 2014
DOI:10.1021/jp504805y
Excited singlet state structures believed to be responsible for the Franck–Condon-allowed and the strongly Stokes-shifted (blue) emissions in linear permethylated oligosilanes (SinMe2n+2) have been found and characterized with time-dependent density functional (TD-DFT) methods for chain lengths 4 ≤ n ≤ 16. For chain lengths with n > 7, the S1 relaxed structures closely resemble the S0 equilibrium structures where all valence angles are tetrahedral and all backbone dihedral angles are transoid. At chain lengths with n < 8 more strongly distorted structures with one long Si–Si bond built from silicon 3p orbitals are encountered. The large Stokes shift is due more to a large destabilization of the ground state than the relaxation in the S1 excited state. For n = 7, both types of minima were located, exactly reproducing the borderline between the large-radius and the small-radius self-trapped excitons known from experiments.
Co-reporter:Gregg S. Kottas, Thierry Brotin, Peter F. H. Schwab, Kamal Gala, Zdeněk Havlas, James P. Kirby, John R. Miller, and Josef Michl
Organometallics 2014 Volume 33(Issue 13) pp:3251-3264
Publication Date(Web):June 13, 2014
DOI:10.1021/om400403j
The known (tetraphenyl-η4-cyclobutadiene)-η5-cyclopentadienylcobalt (1) and a series of its new substituted derivatives have been prepared. The electronic states of a few representatives have been characterized by absorption and magnetic circular dichroism. Time-dependent density functional theory has been used to arrive at spectral assignments for several prominent low-energy bands. The absorption spectra of the radical ions of 1 have also been recorded.
Co-reporter: Hayato Tsuji;Dr. Heather A. Fogarty; Masahiro Ehara; Ryoichi Fukuda;Dr. Deborah L. Casher; Kohei Tamao; Hiroshi Nakatsuji; Josef Michl
Chemistry - A European Journal 2014 Volume 20( Issue 30) pp:9431-9441
Publication Date(Web):
DOI:10.1002/chem.201403495

Abstract

Unlike π-electron chromophores, the peralkylated n-tetrasilane σ-electron chromophore resembles a chameleon in that its electronic spectrum changes dramatically as its silicon backbone is twisted almost effortlessly from the syn to the anti conformation (changing the SiSiSiSi dihedral angle ω from 0 to 180°). A combination of UV absorption, magnetic circular dichroism (MCD), and linear dichroism (LD) spectroscopy on conformationally controlled tetrasilanes 19, which cover fairly evenly the full range of angles ω, permitted a construction of an experimental correlation diagram for three to four lowest valence electronic states. The free chain tetrasilane n-Si4Me10 (10), normally present as a mixture of three enantiomeric conformer pairs of widely different angles ω, has also been included in our study. The spectral trends are interpreted in terms of avoided crossings of 1B with 2B and 2A with 3A states, in agreement with SAC-CI calculations on the excited states of 17 and conformers of 10.

Co-reporter:Luká&x161; Kobr;Ke Zhao;Yongqiang Shen;Richard K. Shoemaker;Charles T. Rogers
Advanced Materials 2013 Volume 25( Issue 3) pp:443-448
Publication Date(Web):
DOI:10.1002/adma.201203294
Co-reporter:Malgorzata Mucha ; Eva Kaletová ; Anna Kohutová ; Frank Scholz ; Elizabeth S. Stensrud ; Ivan Stibor ; Lubomír Pospíšil ; Florian von Wrochem
Journal of the American Chemical Society 2013 Volume 135(Issue 15) pp:5669-5677
Publication Date(Web):February 11, 2013
DOI:10.1021/ja3117125
Treatment of a gold surface with a solution of C18H37HgOTs under ambient conditions results in the formation of a covalently adsorbed monolayer containing alkyl chains attached directly to gold, Hg(0) atoms, and no tosyl groups. It is stable against a variety of chemical agents. When the initial deposition is performed at a positive applied potential and is followed by oxidative electrochemical stripping, the mercury can be completely removed, leaving a gold surface covered only with alkyl chains. The details of the attachment structure are not known. The conclusions are based on infrared spectroscopy, X-ray and UV photoelectron spectroscopy, ellipsometry, contact angle goniometry, differential pulse polarography, and measurements of electrode blocking and electrochemical admittance.
Co-reporter:Dr. Ji&x159;í Kaleta;Dr. Akin Akdag; Raül Crespo; Mari-Carmen Piqueras; Josef Michl
ChemPlusChem 2013 Volume 78( Issue 9) pp:1174-1183
Publication Date(Web):
DOI:10.1002/cplu.201300219

Abstract

The trideuteriomethylation of BH vertices in CB11H12 and its derivatives with CD3OTf (OTf=triflate, trifluoromethanesulfonate) yields a mixture of BCD3 and BCHD2 substitution products, thus demonstrating the intermediacy of a species with a long enough lifetime for hydrogen scrambling between the boron vertex and the methyl substituent. No such scrambling is observed if CD3OTf is used to methylate toluene. According to density functional theory calculations, the intermediate in BH vertex methylation is a three-center bonded σ adduct of a methyl cation to the BH bond and the proton scrambling occurs via a transition structure containing a distorted square-pyramidal methane attached axially to a “naked” boron vertex. The subsequent proton or deuteron loss is presently not understood in detail. A general comparison of electrophilic substitution on closo-boranes and arenes is provided and similarities as well as differences are discussed. A recalculation of the optimized geometry of the CB11Me12. radical produced a second Jahn–Teller distorted minimum and resulted in a somewhat improved agreement between calculated and measured proton hyperfine coupling constants.

Co-reporter:Lukáš Kobr, Ke Zhao, Yongqiang Shen, Kateřina Polívková, Richard K. Shoemaker, Noel A. Clark, John C. Price, Charles T. Rogers, and Josef Michl
The Journal of Organic Chemistry 2013 Volume 78(Issue 5) pp:1768-1777
Publication Date(Web):July 10, 2012
DOI:10.1021/jo3009897
We examine the insertion of two dipolar molecular rotors as guests into a host, tris(o-phenylenedioxy)cyclotriphosphazine (TPP, 1), using differential scanning calorimetry, solid-state NMR, powder X-ray diffraction, and dielectric spectroscopy. The rotors are 1-(4′-n-pentylbiphenyl-4-yl)-12-(2,3-dichlorophenyl)-p-dicarba-closo-dodecaborane and 1,12-bis(2,3-dichlorophenyl)-p-dicarba-closo-dodecaborane. Both enter the bulk even though their nominal diameter exceeds the nominal channel diameter and although a closely related rotor, 1-n-hexadecyl-12-(2,3-dichlorophenyl)-p-dicarba-closo-dodecaborane, is known to produce a surface inclusion compound. Rotational barriers of 5.4–9.3 kcal/mol were found for the dichlorophenyl rotator contained within the TPP channel. Clearly, van der Waals diameters in themselves do not suffice to predict TPP channel entry. It is suggested that the efficacy of the p-carborane stopper is reduced by the presence of the two relatively bulky adjacent benzene rings, which help to stretch the channel, and by the axial direction of its axis, which prevents the attached rotator from contributing to the stopping action.
Co-reporter:Frank Scholz, Eva Kaletová, Elizabeth S. Stensrud, William E. Ford, Anna Kohutová, Malgorzata Mucha, Ivan Stibor, Josef Michl, and Florian von Wrochem
The Journal of Physical Chemistry Letters 2013 Volume 4(Issue 16) pp:2624-2629
Publication Date(Web):July 22, 2013
DOI:10.1021/jz4011898
n-Alkyl self-assembled monolayers can be directly attached to gold through C–Au bonds by the deposition of organomercury salts on gold substrates, as shown here using n-butylmercury and n-octadecylmercury tosylate derivatives. The Hg atoms, which are codeposited during this process, are removed by thermal annealing at 95 °C, resulting in alkyl monolayers having a significantly enhanced thermal stability compared with alkanethiol monolayers, however, a lower degree of conformational order. The monolayer properties are elucidated by X-ray photoemission and IR spectroscopy, STM, ellipsometry, and contact-angle goniometry.Keywords: anchor group; carbon−Au bond; n-alkyl; photoemission spectroscopy; self-assembled monolayer;
Co-reporter:Lukáš Kobr ; Ke Zhao ; Yongqiang Shen ; Angiolina Comotti ; Silvia Bracco ; Richard K. Shoemaker ; Piero Sozzani ; Noel A. Clark ; John C. Price ; Charles T. Rogers
Journal of the American Chemical Society 2012 Volume 134(Issue 24) pp:10122-10131
Publication Date(Web):June 1, 2012
DOI:10.1021/ja302173y
We describe an approach to regular triangular arrays of dipolar molecular rotors based on insertion of dipolar rotator carrying shafts as guests into channels of a host, tris(o-phenylenedioxy)cyclotriphosphazene (TPP). The rotor guests can either enter the bulk of the host or stay at or near the surface, if a suitable stopper is installed at the end of the shaft. Differential scanning calorimetry, solid-state NMR, and powder X-ray diffraction were used to examine the insertion of a dipolar rotor synthesized for the purpose, 1-n-hexadecyl-12-(2,3-dichlorophenyl)-p-dicarba-closo-dodecaborane, and it was found that it forms a surface inclusion compound. Rotational barriers from 1.2 to 9 kcal/mol were found by dielectric spectroscopy and were attributed to rotors inserted into the surface to different degrees, some rubbing the surface as they turn.
Co-reporter:Akin Akdag ; Zdeněk Havlas
Journal of the American Chemical Society 2012 Volume 134(Issue 35) pp:14624-14631
Publication Date(Web):August 9, 2012
DOI:10.1021/ja3063327
Of the five small biradicaloid heterocycles whose S1, S2, T1, and T2 adiabatic excitation energies were examined by the CASPT2/ANO-L-VTZP method, two have been found to meet the state energy criterion for efficient singlet fission and are recommended to the attention of synthetic chemists and photophysicists.
Co-reporter:Jiří Kaleta, Ján Tarábek, Akin Akdag, Radek Pohl, and Josef Michl
Inorganic Chemistry 2012 Volume 51(Issue 20) pp:10819-10824
Publication Date(Web):September 20, 2012
DOI:10.1021/ic301236s
The syntheses of all 16 CB11(CH3)n(CD3)12–n• radicals with 5-fold substitution symmetry are described. The variation in the width of their broad and featureless electron paramagnetic resonance signals as a function of the deuteriation pattern is used to deduce the relative values of the average hyperfine coupling constants aH of the hydrogen atoms in the ipso (1), ortho (2–6), meta (7–11), and para (12) methyl groups, aH(i):aH(o):aH(m):aH(p) = (0.18 ± 0.09):(0.71 ± 0.02):(1.00 ± 0.03):(0.52 ± 0.05), and these can be compared with ratios expected from a B3LYP/EPRII calculation, 0.04:0.55:1:0.51.
Co-reporter:Abdul Wahab ; Brian Stepp ; Christos Douvris ; Michal Valášek ; Jan Štursa ; Jiřı́ Klı́ma ; Mari-Carmen Piqueras ; Raül Crespo ; Jiřı́ Ludvı́k
Inorganic Chemistry 2012 Volume 51(Issue 9) pp:5128-5137
Publication Date(Web):April 18, 2012
DOI:10.1021/ic2026939
Cyclic voltammetry of 31 icosahedral carborane anions 1-X-12-Y-CB11Me10– at a Pt electrode in liquid SO2 revealed a completely reversible one-electron oxidation even at low scan rates, except for the anions with Y = I, which are oxidized irreversibly up to a scan rate of 5.0 V/s, and the anion with X = COOH and Y = H, whose oxidation is irreversible at scan rates below 1.0 V/s. Relative reversible oxidation potentials agree well with RI-B3LYP/TZVPP,COSMO and significantly less well with RI-BP86/TZVPP,COSMO or RI-HF/TZVPP,COSMO calculated adiabatic electron detachment energies. Correlations with HOMO energies of the anions are nearly as good, even though the oxidized forms are subject to considerable Jahn–Teller distortion. Except for the anion with X = F and Y = Me, the oxidation potentials vary linearly with substituent σp Hammett constants. The slopes (reaction constants) are ∼0.31 and ∼0.55 V for positions 1 and 12, respectively.
Co-reporter:Miroslav Dudič, Ivana Císařová, and Josef Michl
The Journal of Organic Chemistry 2012 Volume 77(Issue 1) pp:68-74
Publication Date(Web):November 22, 2011
DOI:10.1021/jo202016v
A simple synthesis of a hexadehydrotribenzo[a,e,i][12]annulene by insertion of acetylene into an open-chain diiodo precursor under Sonogashira coupling conditions has been developed and used to prepare a rigid three-armed star connector for testing as a building block for a two-dimensional hexagonal hydrogen-bonding array.
Co-reporter:Victoria Volkis, Richard K. Shoemaker, and Josef Michl
Macromolecules 2012 Volume 45(Issue 23) pp:9250-9257
Publication Date(Web):November 15, 2012
DOI:10.1021/ma301850c
In the presence of a nonoxidizing radical initiator, azo-tert-butane, and a high concentration of LiCB11(CH3)12, isobutylene undergoes thermal or light-induced radical polymerization to b-PIB, a highly branched polymer of modest molecular weight (mostly a few thousand and up to ∼25 000 g/mol based on GPC with polystyrene standards). The structure of b-PIB was elucidated by NMR spectroscopy of a low-molecular-weight fraction. The polymer is branched on every carbon atom of the main chain; one chain end carries an isobutenyl group, and the other carries a tert-butyl group originating in the initiator. The branches are short segments of l-PIB (linear polyisobutylene), on the average composed of five IB units. A mechanism of formation if this dendrimer-like structure is proposed.
Co-reporter:Alexandr Prokop, Jaroslav Vacek, and Josef Michl
ACS Nano 2012 Volume 6(Issue 3) pp:1901
Publication Date(Web):February 2, 2012
DOI:10.1021/nn300003x
Friction in molecular rotors is examined by classical molecular dynamics simulations for grid-mounted azimuthal dipolar molecular rotors, whose rotation is either allowed to decay freely or is driven at GHz frequencies by a flow of rare gas or by a rotating electric field. The rotating parts (rotators) are propeller-shaped. Their two to six blades consist of condensed aromatic rings and are attached to a deltahedral carborane hub, whose antipodal carbons carry [n]staffane axles mounted on a square molecular grid. The dynamic friction constant η has been derived in several independent ways with similar results. Analysis of free rotation decay yields η as a continuous exponentially decreasing function of rotor frequency. The calculated dependence of friction torque on frequency resembles the classical macroscopic Stribeck curve. Its relation to rotational potential energy barriers and the key role of the rate of intramolecular vibrational redistribution (IVR) of energy and angular momentum from rotator rotation to other modes are considered in two limiting regimes. (i) In the strongly overdamped regime, rotation is much slower than IVR, and effective friction can be expressed through potential barriers to rotation. (ii) In the strongly underdamped regime, rotation is much faster than IVR, whose rate then determines friction.Keywords: friction; intramolecular vibrational redistribution; molecular dynamics; molecular rotors; potential energy barriers
Co-reporter:Alexandre G. L. Olive ; Kamil Parkan ; Cecile Givelet
Journal of the American Chemical Society 2011 Volume 133(Issue 50) pp:20108-20111
Publication Date(Web):November 3, 2011
DOI:10.1021/ja209051t
Supramolecular self-assembly using weak interactions under quasi-equilibrium conditions has provided easy access to very complex but often quite fragile molecules. We now show how a labile structure obtained from reversible transition-metal-directed self-assembly of rods and connectors serves as a template that can be converted into a sturdy structure of identical topology and similar geometry. The process consists of Cu(I)-catalyzed replacement of all rods or connectors terminated with pyridines for analogues terminated with ethynyls, converting dative N→Pt+ bonds into covalent C–Pt bonds. The procedure combines the facility and high yield of reversible self-assembly with the robustness of covalent synthesis.
Co-reporter:Victoria Volkis ; Christos Douvris
Journal of the American Chemical Society 2011 Volume 133(Issue 20) pp:7801-7809
Publication Date(Web):May 4, 2011
DOI:10.1021/ja111659u
A solution of a mechanistic puzzle is reported: upon initiation with air at 25 °C or with di-tert-butyl peroxide at 80 °C, isobutylene (IB) polymerizes at 1 atm in weakly coordinating solvents containing 10 wt % LiCB11(CH3)12 to a mixture of highly branched (b-PIB) and linear (l-PIB) polyisobutylene. The former polymer is separable by solvent extraction and is identical with the b-PIB that is produced from IB as a sole product under similar conditions under nonoxidizing radical initiation with azo-tert-butane. The latter polymer differs from standard l-PIB in that it carries a carborane anion attached at the chain end. The molecular weight of b-PIB ranges up to 26 000, and that of l-PIB, up to 50 000. Evidence is presented that the concurrent polymerization of IB to b-PIB and l-PIB is launched by an initial oxidation of the CB11(CH3)12– anion to a neutral radical CB11(CH3)12•. This radical is proposed to subsequently transfer a methyl radical to IB, thus launching the formation of b-PIB by the radical mechanism while leaving behind the borenium ylide CB11(CH3)11, which is a strong Lewis acid and induces simultaneously the formation of l-PIB by the cationic mechanism.
Co-reporter:Matthew G. Fete ; Zdeněk Havlas
Journal of the American Chemical Society 2011 Volume 133(Issue 11) pp:4123-4131
Publication Date(Web):February 22, 2011
DOI:10.1021/ja111035k
Cs salts of four of the title anions were prepared by fluorination of salts of partly methylated (n = 11, 10) or partly methylated and partly iodinated (n = 6, 5) CB11H12− anions. The CH vertex is acidic, and in the unhindered anion with n = 6 it has been alkylated. Neat Cs+[1-H−CB11(CF3)11]− is as treacherously explosive as Cs+[CB11(CF3)12]−, but no explosions occurred with the salts of the other three anions. BL3YP/6-31G* gas-phase electron detachment energies of the title anions are remarkably high, 5-8 eV. Treated with NiF3+ in anhydrous liquid HF at −60 °C, anions with n = 11 or 10 resist oxidation, whereas anions with n = 6 or 5 are converted to colored EPR-active species, presumably the neutral radicals [HCB11(CF3)nF11-n]•. These are stable for hours at −60 °C after extraction into cold perfluorohexane or perfluorotri-n-butylamine solutions. On warming to −20 °C in a Teflon or quartz tube, the color and EPR activity disappear, and the original anions are recovered nearly quantitatively, suggesting that the radicals oxidize the solvent.
Co-reporter:Hua Mei, Christos Douvris, Victoria Volkis, Phillip Hanefeld, Nicole Hildebrandt, and Josef Michl
Macromolecules 2011 Volume 44(Issue 8) pp:2552-2558
Publication Date(Web):March 25, 2011
DOI:10.1021/ma101926c
In nonpolar media LiCB11Me12 catalyzes the radical copolymerization of ethyl acrylate with isobutylene. The reaction is initiated by azo-tert-butane at 75−80 °C or with UV-irradiated di-tert-butyl peroxide at room temperature. Nonalternating copolymers with novel somewhat branched structures were produced and characterized by NMR. Isobutylene incorporation (up to 56 mol %) depends on the comonomer ratio, and a copolymerization mechanism is proposed.
Co-reporter:Wade A. Braunecker, Akin Akdag, Byron A. Boon, and Josef Michl
Macromolecules 2011 Volume 44(Issue 6) pp:1229-1232
Publication Date(Web):February 8, 2011
DOI:10.1021/ma102825r
Co-reporter:Justin C. Johnson ; Arthur J. Nozik
Journal of the American Chemical Society 2010 Volume 132(Issue 46) pp:16302-16303
Publication Date(Web):November 2, 2010
DOI:10.1021/ja104123r
Direct observation of triplet absorption and ground-state depletion upon pulsed excitation of a polycrystalline thin solid film of 1,3-diphenylisobenzofuran at 77 K revealed a 200 ± 30% triplet yield, which was attributed to singlet fission.
Co-reporter:Michal Valášek, Jan Štursa, Radek Pohl, and Josef Michl
Inorganic Chemistry 2010 Volume 49(Issue 22) pp:10255-10263
Publication Date(Web):October 8, 2010
DOI:10.1021/ic101235e
We report the syntheses of several [1-R-CB11-Me11]− and [1-R-12-R′-CB11-Me10]− anions (R, R′ = alkyl) and the solubilities of their lithium salts in cyclohexane. These solutions are of interest as Lewis acid catalysts. The new anions are not directly accessible by methylation with methyl triflate because of intervening triflyloxy substitution on one or more boron vertices. The difficulty has been circumvented in two ways. Either (i) an iodo substituent is first introduced into position 12, permitting a clean decamethylation, and then replaced with a methyl by reaction with trimethylaluminum or (ii) the offending triflyloxy substituents are replaced with methyls by reaction with trimethylaluminum.
Co-reporter:Eric C. Greyson, Brian R. Stepp, Xudong Chen, Andrew F. Schwerin, Irina Paci, Millicent B. Smith, Akin Akdag, Justin C. Johnson, Arthur J. Nozik, Josef Michl, and Mark A. Ratner
The Journal of Physical Chemistry B 2010 Volume 114(Issue 45) pp:14223-14232
Publication Date(Web):December 21, 2009
DOI:10.1021/jp909002d
Singlet exciton fission, a process that converts one singlet exciton to a pair of triplet excitons, has the potential to enhance the efficiency of both bulk heterojunction and dye-sensitized solar cells and is understood in crystals but not well understood in molecules. Previous studies have identified promising building blocks for singlet fission in molecular systems, but little work has investigated how these individual chromophores should be combined to maximize triplet yield. We consider the effects of chemically connecting two chromophores to create a coupled chromophore pair and compute how various structural choices alter the thermodynamic and kinetic parameters likely to control singlet fission yield. We use density functional theory to compute the electron transfer matrix element and the thermodynamics of fission for several promising chromophore pairs and find a trade-off between the desire to maximize this element and the desire to keep the singlet fission process exoergic. We identify promising molecular systems for singlet fission and suggest future experiments.
Co-reporter:Annika Be Dr.
Chemistry - A European Journal 2009 Volume 15( Issue 34) pp:8504-8517
Publication Date(Web):
DOI:10.1002/chem.200901521

Abstract

The effects of σ-electron delocalization on optical properties of saturated linear chains of permethylated oligosilanes are strongly conformation dependent. We analyze the origin of the conformational dependence of the energies of molecular orbitals and of electronic excitations in simple intuitively understandable terms by using a first-order approximation to the Hückel version of the “Ladder C” model. The analysis is supported by comparison with results of numerical calculations from time-dependent density functional theory, which agree well with experiment. To facilitate the comparison, a simple procedure has been developed that defines the overall and local fractional σ and π characters of a backbone molecular orbital and a fractional overall and local σσ* and σπ* characters of an excited state for any conformation of a linear chain.

Co-reporter:Mari Carmen Piqueras, Raül Crespo and Josef Michl
The Journal of Physical Chemistry A 2008 Volume 112(Issue 50) pp:13095-13101
Publication Date(Web):September 27, 2008
DOI:10.1021/jp804677v
Time-dependent density functional theory (TD-DFT/B3LYP(AC)/cc-pVTZ/cc-pVTZ/6-311G//MP2/cc-pVTZ/cc-pVTZ/6-31G**) has been used to compute vertical excitation energies and oscillator strengths of the six low-lying excited states of four peralkylated disilanes, hexamethyldisilane (1), hexa-tert-butyldisilane (2), 1,6-disila[4.4.4]propellane (3), and 1,7-disila[5.5.5]propellane (4). The results provide an accurate interpretation of the reported UV absorption spectra of 1−4 in solution, and for 1 also in the gas phase up to 62 000 cm−1. The excellent agreement of the calculated with the available experimental energies and oscillator strengths, and with magnetic circular (MCD) and linear (LD) dichroism, gives us confidence that the method will be useful for dependable interpretation of the electronic spectra of longer oligosilanes. Although the disilane chromophore finds itself in quite different environments in 1−4, its fundamental characteristics remain the same, with one important exception. In all four compounds, the first valence excited state is due to an electron promotion from the σ1 HOMO to the π1* orbital, and the second valence excited state to a promotion from the σ1 HOMO to the σ1* orbital. Surprisingly, however, it is only in 2, which has an extraordinarily long SiSi bond, that the terminating σ1* orbital is the σ*(SiSi) antibond, as anticipated, and the σσ* transition has the expected very high oscillator strength. In 1, 3, and 4, the σ*(SiSi) antibonding orbital is high in energy and does not play any role in low-energy excitations. Instead, the terminating orbital of the σ1σ1* excitation is represented by Si−alkyl antibonds, combined symmetrically with respect to rotation around the SiSi axis and antisymmetrically with respect to operations that interchange the two Si atoms. The common assumption that the characteristic intense σσ* transitions of longer peralkylated oligosilanes extrapolate to the lowest σσ* transition in common peralkylated disilanes is incorrect, and only the weak σπ* transitions extrapolate simply.
Co-reporter:J. Vacek;J. Michl
Advanced Functional Materials 2007 Volume 17(Issue 5) pp:730-739
Publication Date(Web):15 MAR 2007
DOI:10.1002/adfm.200601225

The development of artificial surface-mounted molecular rotors has benefited from theoretical guidance by molecular dynamics simulations. After a brief survey of the origins of the project, the present understanding of the way in which these simple molecular machines operate is reviewed.

Co-reporter:Masahiro Ehara Assoc.  ;Shuhei Fukawa;Donald E. David Dr.;Hiroshi Nakatsuji ;Evgueni Z. Pinkhassik Dr.;Michael D. Levin Dr.;Marcin Apostol Dr.
Chemistry – An Asian Journal 2007 Volume 2(Issue 8) pp:1007-1019
Publication Date(Web):28 JUN 2007
DOI:10.1002/asia.200600411

He(I) photoelectron spectroscopy was used to examine the valence-shell electronic structure of three new and seven previously known bicyclo[1.1.1]pentane derivatives, 1,3-Y2-C5X6 (for X=H, Y=H, Cl, Br, I, CN; for X=F, Y=H, Br, I, CN). A larger series (X=H or F, Y=H, F, Cl, Br, I, At, CN) has been studied computationally with the SAC-CI (symmetry-adapted cluster configuration interaction) method. The outer-valence ionization spectra calculated by the SAC-CI method, including spin–orbit interaction, reproduced the experimental photoelectron spectra well, and quantitative assignments are given. When the extent of effective through-cage interaction between the bridgehead halogen lone-pair orbitals was defined in the usual way by orbital-energy splitting, it was found to be larger than that mediated by other cages such as cubane, and was further enhanced by hexafluorination. The origin of the orbital-energy splitting is analyzed in terms of cage structure, and it is pointed out that its relation to the degree of interaction between the bridgehead substituents is not as simple as is often assumed.

Co-reporter:Raül Crespo;Mari Carmen Piqueras
Theoretical Chemistry Accounts 2007 Volume 118( Issue 1) pp:81-87
Publication Date(Web):2007 July
DOI:10.1007/s00214-007-0246-1
We present a multistate complete active space second-order perturbation theory (MS-CASPT2) study of the low-lying valence excited states of a peralkylated tetrasilacyclopentane, c-(CH2Si4Me8) (1). The lowest-lying calculated valence excited states are located in the 37,400–50,500 cm−1 region, in perfect agreement with experimental observations, 38,600–50,000 cm−1. The MS-CASPT2 results provide a very good description of the nature, intensity, and position of the observed absorption bands, including the recently detected fourth electronic transition. The computational results have been used as a benchmark for evaluating a time-dependent density functional theory (TD-DFT) procedure suitable for calculations on longer oligosilanes.
Co-reporter:Zbyněk Janoušek;Kamesh Vyakaranam;Ludvig Eriksson
Heteroatom Chemistry 2006 Volume 17(Issue 3) pp:217-223
Publication Date(Web):10 APR 2006
DOI:10.1002/hc.20224

We report an improved synthesis of 1-halocarba-closo-dodecaborate anions 1-Hal–CB11H and their efficient conversion to the undecamethylated anions 1-Hal–CB11Me (Hal = Cl, Br, I) and the hexamethylated anions 1-Hal-(7–12)-(CH3)6–CB11H (Hal = F, Cl) by treatment with methyl triflate in sulfolane in the presence of calcium hydride to remove the triflic acid byproduct. © 2006 Wiley Periodicals, Inc. Heteroatom Chem 17:217–223, 2006; Published online in Wiley InterScience (www.interscience.wiley.com). DOI 10.1002/hc.20224

Co-reporter:Dominik Horinek;
Proceedings of the National Academy of Sciences 2005 102(40) pp:14175-14180
Publication Date(Web):September 26, 2005
DOI:10.1073/pnas.0506183102
Molecular dynamics simulations of the response to oscillating electric field elicited from an altitudinal dipolar molecular rotor mounted on the Au(111) surface and previously studied experimentally in static fields show unidirectional rotation in one of the three pairs of conformational enantiomers. The simulations are based on the universal force field and take into account electronic friction in the metal through its effect on the image charges. The rotor consists of two cobalt sandwich posts whose upper decks carry a biphenyl-like rotator with a dipole moment perpendicular to the rotation axle, mounted parallel to the surface. A phase diagram of rotor performance at 10 K as a function of field frequency and amplitude contains five unidirectional rotation regions: synchronous, half-synchronous (every other cycle skipped), quarter-synchronous (only indistinctly), asynchronous, and essentially no response. The nature of the subharmonic “single-molecule parametric oscillator” behavior is understood in mechanistic detail. Simulations at higher temperatures distinguish the thermal (“Brownian”) and driven regimes of rotation, elucidated in terms of time-dependent potential energy surfaces for the rotation.
Co-reporter:Heather A. Fogarty;Roman Imhof
PNAS 2004 101 (29 ) pp:10517-10522
Publication Date(Web):2004-07-20
DOI:10.1073/pnas.0403209101
Magnetic circular dichroism (MCD) of five peralkylated tetrasilanes (1–5) conformationally constrained to angles ranging from nearly 0° to 180° and of the open chain tetrasilane Si4Me10 (6) shows a clear conformational dependence and permits the detection of previously hidden transitions. In the tetrasilane CH2Si4Me8 (1), with the smallest dihedral angle, comparison of MCD with absorption spectra reveals four low-energy electronic transitions. In the tetrasilanes 2–4, three distinct transitions are apparent. In tetrasilanes 5 and 6, MCD reveals the very weak transition that has been predicted to be buried under the first intense peak and to which the anomalous thermochromism of 6 and other short-chain oligosilanes has been attributed.
Co-reporter:Natalia Varaksa;Lubomír Pospíšil;Thomas F. Magnera
PNAS 2002 Volume 99 (Issue 8 ) pp:5012-5017
Publication Date(Web):2002-04-16
DOI:10.1073/pnas.082098299
The adsorption of the trigonal connector, 1,3,5-tris[10-(3-ethylthiopropyl)dimethylsilyl-1,10-dicarba-closo-decaboran-1-yl]benzene (1), from acetonitrile/0.1 M LiClO4 on the surface of mercury at potentials ranging from +0.3 to −1.4 V (vs. aqueous Ag|AgCl|1 M LiCl) was examined by voltammetry, Langmuir isotherms at controlled potentials, and impedance measurements. No adsorption is observed at potentials more negative than ∼ −0.85 V. Physisorption is seen between ∼ −0.85 and 0 V. At positive potentials, adsorbate-assisted anodic dissolution of mercury occurs and an organized surface layer is formed. Although the mercury cations are reduced at −0.10 V, the surface layer remains metastable to potentials as negative as −0.85 V. Its surface areas per molecule and per redox center are compatible with a regular structure with the connectors 1 woven into a hexagonal network by RR′S→Hg←SRR′ or RR′S→Hg2+←SRR′ bridges. The structure is simulated closely by geometry optimization in the semiempirical AM1 approximation.
Co-reporter:Thomas F. Magnera
PNAS 2002 Volume 99 (Issue 8 ) pp:4788-4792
Publication Date(Web):2002-04-16
DOI:10.1073/pnas.052016299
Co-reporter:Donald E. David, V. Balaji, Josef Michl, Herbert M. Urbassek
International Journal of Mass Spectrometry 2001 Volume 212(1–3) pp:477-489
Publication Date(Web):15 November 2001
DOI:10.1016/S1387-3806(01)00528-0
We have measured the initial and fluence-dependent sputtering yields of condensed neat CO2, NO2, N2O, CH4, C2H4, and NH3 for kilo-electron-volt-energy rare-gas ion bombardment. The results of the measurements are rationalized in terms of the gas-flow model of condensed-gas sputtering, modified by the chemical reactions expected to occur during and after gas flow. The three triatomic target molecules studied have similar physical properties. The different initial sputtering yields observed are rationalized in terms of chemical effects occurring during bombardment. Only moderate fluence dependencies are observed. Sputtering of CH4 and C2H4 targets occurs with only moderate initial sputtering yields, compatible with no gas-flow participation in the sputtering mechanism. With increasing incident ion fluence, the sputtering yields rise due to the build up of highly volatile products. For higher bombardment fluences two different results are observed. In the case of lighter bombarding ions (He+, Ne+, and, for the C2H4 target, also Ar+) the yields eventually decrease, probably due to polymerization reaction products coating the target. For heavier projectile ions the yields plateau at a high level. We attribute this latter behavior to a strong gas-flow contribution which entraps even the heavier carbonaceous products in the gas flow and prevents their accumulation on the surface. This behavior has not been observed before. The sputtering behavior of NH3 follows closely that of CH4, except that yield decreases due to polymerization are absent.
Co-reporter:Jaroslav Vacek
PNAS 2001 Volume 98 (Issue 10 ) pp:5481-5486
Publication Date(Web):2001-05-08
DOI:10.1073/pnas.091100598
Classical molecular dynamics is applied to the rotation of a dipolar molecular rotor mounted on a square grid and driven by rotating electric field E(ν) at T ≃ 150 K. The rotor is a complex of Re with two substituted o-phenanthrolines, one positively and one negatively charged, attached to an axial position of Rh in a [2]staffanedicarboxylate grid through 2-(3-cyanobicyclo[1.1.1]pent-1-yl)malonic dialdehyde. Four regimes are characterized by a, the average lag per turn: (i) synchronous (a < 1/e) at E(ν) = |E(ν)| > Ec(ν) [Ec(ν) is the critical field strength], (ii) asynchronous (1/e < a < 1) at Ec(ν) > E(ν) > Ebo(ν) > kT/μ, [Ebo(ν) is the break-off field strength], (iii) random driven (a ≃ 1) at Ebo(ν) > E(ν) > kT/μ, and (iv) random thermal (a ≃ 1) at kT/μ > E(ν). A fifth regime, (v) strongly hindered, W > kT, Eμ, (W is the rotational barrier), has not been examined. We find Ebo(ν)/kVcm−1 ≃ (kT/μ)/kVcm−1 + 0.13(ν/GHz)1.9 and Ec(ν)/kVcm−1 ≃ (2.3kT/μ)/kVcm−1 + 0.87(ν/GHz)1.6. For ν > 40 GHz, the rotor behaves as a macroscopic body with a friction constant proportional to frequency, η/eVps ≃ 1.14 ν/THz, and for ν < 20 GHz, it exhibits a uniquely molecular behavior.
Co-reporter:Dr. Bo Albinsson;Dr. Hiroyuki Teramae;Dr. John W. Downing; Josef Michl
Chemistry - A European Journal 1996 Volume 2(Issue 5) pp:
Publication Date(Web):21 JAN 2006
DOI:10.1002/chem.19960020512

Infrared and ultraviolet spectra of the gauche and anti conformers of matrix-isolated permethyl-n-tetrasilane have been obtained separately by taking advantage of thermally induced gauche-to-anti conversion and of wavelength-selective photochemical destruction of either conformer. The resolved UV spectrum of the gauche conformer provides the first piece of experimental evidence in favor of the recently proposed reinterpretation of conformational effects on tetrasilane electronic states. According to this, it is not the energy but the intensity of the lowest singlet excitation that changes dramatically as the SiSiSiSi dihedral angle is varied, as a result of an avoided crossing between s̀s̀* and s̀π* states. Implications for the general understanding of sigma conjugation in simple terms are discussed. Unconstrained MP2/6-31 G* optimization predicts the existence of a third backbone conformer (ortho), with a dihedral angle of about 90°. Its predicted (HF/3-21 G*) mid-IR spectrum is indistinguishable from that of the gauche conformer, and the matrix-isolation spectra thus provide no evidence for or against its presence.

Pyridine, 4,4',4'',4'''-(1,2,4,5-benzenetetrayltetra-2,1-ethynediyl)tetrakis-
Bicyclo[1.1.1]pentane, 1-iodo-3-methyl-
Pyridazine, 3,6-bis(1,1-dimethylethyl)-
3,4'-dibromobiphenyl
3-(4-bromophenyl)aniline
Nitropropane
1,4-Naphthalenedione,2-chloro-3-(methylamino)-
1,4-Naphthalenedione, 2-chloro-3-(methylnitrosoamino)-
1,4-Naphthalenedione, 2,3-diamino-