Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 15 de 15
Filter
Add more filters










Publication year range
1.
J Chromatogr Sci ; 50(3): 162-6, 2012 Mar.
Article in English | MEDLINE | ID: mdl-22337791

ABSTRACT

High-performance liquid chromatography method for the determination of propofol has been developed and validated. Following a liquid extraction using ethyl acetate and hexane, samples were separated by reverse-phase high-performance liquid chromatography on an XBridge C(18) column and quantified using fluorescence detection at an excitation of 276 nm and an emission of 310 nm. The mobile phase was a mixture of water (pH 4.0) and acetonitrile, with a flow rate of 1.5 mL/min. The standard curve ranged from 5-2000 ng/mL. Intra- and inter-assay variability for propofol was less than 10%, and the average recovery was greater than 95%. This assay is suitable for use in pharmacokinetic studies.


Subject(s)
Chromatography, High Pressure Liquid/methods , Propofol/blood , Acetonitriles/chemistry , Animals , Dogs , Hydrogen-Ion Concentration , Limit of Detection , Linear Models , Methanol/chemistry , Reproducibility of Results , Spectrometry, Fluorescence/methods
2.
J Zoo Wildl Med ; 42(2): 241-6, 2011 Jun.
Article in English | MEDLINE | ID: mdl-22946401

ABSTRACT

A description of bioencapsulation of metronidazole in adult brine shrimp (Artemia) for 2.5 g/L, 5 g/L, and 10 g/L treatment baths is presented. Metronidazole was detected in adult brine shrimp tissue after enrichment periods of 15 min, 30 min, 1 hr, 2 hr, 4 hr, 8 hr, 12 hr, and 24 hr. The assays were performed using high performance liquid chromatography. There was a positive relationship in both dose and time. When evaluating percent uptake, all three baths demonstrated a similar pattern. All three bath concentrations had a high initial concentration that fell at 30 min and slowly began to increase through the end of the study. Survival of shrimp was not affected by bath concentration but decreased over time in all treatment baths comparatively. It can be concluded that metronidazole can be successfully bioencapsulated in adult Artemia.


Subject(s)
Anti-Infective Agents/chemistry , Anti-Infective Agents/metabolism , Artemia/metabolism , Metronidazole/chemistry , Metronidazole/metabolism , Animals , Time Factors
3.
Biomed Chromatogr ; 24(8): 908-13, 2010 Aug.
Article in English | MEDLINE | ID: mdl-20041399

ABSTRACT

Carboplatin is an antineoplastic drug administered to treat different tumoral conditions in canine oncology. The objective of this study was to validate a high-performance chromatographic (HPLC) method which could be applied in canine pharmacokinetic studies. Following ultrafiltration using a Centrifree device, standards, quality controls and plasma samples were separated by isocratic reversed-phase HPLC on an Inertsil ODS-2 (250 x 4.6 mm i.d.) analytical column and quantified using UV detection at 220 nm. The mobile phase was potassium phosphate (pH 4.5), with a flow-rate of 1.0 mL/min. The procedure produced a linear curve (r(2) > 0.999) over the concentration range 1-200 microg/mL. The lower limit of quantification was 1 microg/mL. The intra-assay and inter-assay precision was approximately 90%. The overall recovery was approximately 90%. The method was illustrated with a preliminary pharmacokinetic analysis on nine dogs treated with carboplatin at our hospital. Carboplatin disposition followed a monocompartmental structure in dogs and was characterized by a short half-life (50 min).


Subject(s)
Antineoplastic Agents/blood , Carboplatin/blood , Chromatography, High Pressure Liquid/methods , Animals , Antineoplastic Agents/pharmacokinetics , Carboplatin/pharmacokinetics , Chromatography, High Pressure Liquid/veterinary , Dogs , Veterinary Drugs/blood , Veterinary Drugs/pharmacokinetics
4.
Toxicol Sci ; 107(1): 122-34, 2009 Jan.
Article in English | MEDLINE | ID: mdl-18996888

ABSTRACT

Saccharomyces cerevisiae bioluminescent bioreporter assays were developed previously to assess a chemical's estrogenic or androgenic disrupting potential. S. cerevisiae BLYES, S. cerevisiae BLYAS, S. cerevisiae BLYR, were used to assess their reproducibility and utility in screening 68, 69, and 71 chemicals for estrogenic, androgenic, and toxic effects, respectively. EC(50) values were 6.3 +/- 2.4 x 10(-10)M (n = 18) and 1.1 +/- 0.5 x 10(-8)M (n = 13) for BLYES and BLYAS, using 17beta-estradiol and 5alpha-dihydrotestosterone over concentration ranges of 2.5 x 10(-12) through 1.0 x 10(-6)M, respectively. Based on analysis of replicate standard curves and comparison to background controls, a set of quantitative rules have been formulated to interpret data and determine if a chemical is potentially hormonally active, toxic, both, or neither. The results demonstrated that these assays are applicable for Tier I chemical screening in Environmental Protection Agency's Endocrine Disruptor Screening and Testing Program as well as for monitoring endocrine-disrupting activity of unknown chemicals in water.


Subject(s)
Biosensing Techniques/methods , Dihydrotestosterone/agonists , Estradiol/agonists , Saccharomyces cerevisiae/metabolism , Decision Trees , Dihydrotestosterone/analysis , Dihydrotestosterone/metabolism , Dose-Response Relationship, Drug , Estradiol/analysis , Estradiol/metabolism , Genes, Reporter , Luminescent Measurements , Luminescent Proteins/metabolism , Reproducibility of Results , Saccharomyces cerevisiae/genetics , Sensitivity and Specificity
5.
Chem Res Toxicol ; 20(9): 1359-63, 2007 Sep.
Article in English | MEDLINE | ID: mdl-17672510

ABSTRACT

A diverse series of polarized alpha,beta-unsaturated and related compounds were evaluated for reactivity with a spectrophotometric assay using the sulfhydryl group in the form of the cysteine residue of the tripeptide GSH as a model nucleophile. The reactive end point (RC 50) calculations were compared to previously described structural alerts based on conventional organic chemistry. This comparison focused on polarized alpha,beta-unsaturates, including ones containing an aldehyde, ketone, ester, sulfoxide, sulfone, sulfonate, nitro, or cyano moiety as well as ortho- and para-pyridino compounds and ortho- and para-quinones. The alerts were coded by substructure and are available in open-source software ( http://sourceforge.net/projects/chemeval). Comparisons of reactivity between selected analogues revealed that only the polarized alpha,beta-unsaturates were reactive. These results verified the coded structural alerts that define the applicability domain for Michael acceptor electrophiles.


Subject(s)
Alkenes/chemistry , Alkynes/chemistry , Cysteine/chemistry , Models, Chemical , Computer Simulation , Molecular Structure , Spectrophotometry
6.
Chem Res Toxicol ; 20(3): 558-62, 2007 Mar.
Article in English | MEDLINE | ID: mdl-17319700

ABSTRACT

A diverse series of aliphatic alpha,beta-unsaturated esters, ketones, and aldehydes were evaluated for reactivity with the model nucleophile sulfhydryl group in the form of the cysteine residue of the tripeptide glutathione; the reactive end point (RC50) was then related to aquatic toxicity (IGC50) assessed in the Tetrahymena pyriformis population growth impairment assay. The substructure specific to all tested reactive substances, an olefin conjugated to a carbonyl group, is inherently electrophilic and conveys the potential to act by way of Michael-type nucleophilic addition. All such unsaturated compounds are inherently acutely toxic. However, their toxicity is difficult to model with conventional descriptors since toxicity is independent of both hydrophobicity and molecular orbital electrophilicity but dependent on the specific molecular structure. While methacrylates typically did not attain an RC50 value at saturation, a linear relationship [log (IGC50(-1)) = 0.936[log (RC50(-1))] + 0.508, where n = 41, r2 = 0.846, q2 = 0.832, s = 0.35, F = 214, and Pr > F = 0.0001] was observed between aquatic toxicity and reactivity for the other carbonyl-containing alpha,beta-unsaturated chemicals.


Subject(s)
Organic Chemicals/chemistry , Organic Chemicals/toxicity , Sulfhydryl Compounds/chemistry , Acrolein/chemistry , Aldehydes/chemistry , Aldehydes/toxicity , Algorithms , Animals , Data Interpretation, Statistical , Esters/chemistry , Esters/toxicity , Glutathione/chemistry , Glutathione/toxicity , Ketones/chemistry , Ketones/toxicity , Kinetics , Methylmethacrylate/chemistry , Tetrahymena pyriformis/chemistry , Tetrahymena pyriformis/drug effects , Tetrahymena pyriformis/metabolism
7.
Environ Toxicol Pharmacol ; 23(1): 10-7, 2007 Jan.
Article in English | MEDLINE | ID: mdl-21783731

ABSTRACT

Aquatic toxicity data in the TETRATOX assay and reactivity data in an abiotic thiol assay were collected for a series of aliphatic isothiocyanates. These compounds can act as Michael-type acceptors with N-hydro-C-mercapto-addition to cellular thiols as a molecular mechanism of action. Comparison of both toxicity and reactivity among the analogues revealed that derivatives with a branch hydrocarbon moiety, especially branched in the ß-position were less toxic and less reactive. In contrast, the di-isothiocyanate and the allyl and propargyl derivatives are more toxic than their 1-alkyl homologues. The toxicity and reactivity differences are consistent but except for the tert-butyl-derivative not remarkable. The differences are due to variations in steric hindrance at the reaction center. For the mono-isothiocyanates compounds toxicity (IGC(50)) is linearly related to thiol reactivity (EC(50)): log(1/IGC(50))=1.33(log(1/EC(50)))-0.41; n=23, s=0.24, r(2)=0.911, q(2)=0.907, F=215.

8.
Inorg Chem ; 43(15): 4702-7, 2004 Jul 26.
Article in English | MEDLINE | ID: mdl-15257599

ABSTRACT

The cis-dithiolate N2S2Ni complex bismercaptoethanediazacycloheptanenickel(II), (bme-dach)Ni or Ni-1', takes up two equivalents of sulfur dioxide in which thiolate-sulfur to SO2-sulfur interactions are well-defined by X-ray crystallography. Ni-1' x 2SO2, C9H18N2NiO4S4, yields monoclinic crystals belonging to the P2(1)/c space group: a = 10.308(4) angstroms, b = 13.334(5) angstroms, c = 10.842(4) angstroms, alpha = 90 degrees, beta = 91.963(6) degrees, gamma = 90 degrees, and Z = 4. Further characterization by nu(SO) IR spectroscopy, thermal gravimetric analysis, electronic spectroscopy, and visual color changes upon reversible SO2 adduct formation establish Ni-1' and the analogous bismercaptoethanediazacyclooctane derivative, (bme-daco)Ni, Ni-1, to be viable candidates for technical development as chemical sensors of this noxious gas. Visual SO2 detection limits of Ni-1 and Ni-1' are established at 25 and 100 ppm, respectively. Both the Ni-1' x 2SO2 adduct and the Ni-1' reactant are air stable. In addition, the stability of Ni-1' x SO2 to vacuum and removal of SO2 by heating make Ni-1' a possible storage/controlled release complex for SO2 gas.

9.
Inorg Chem ; 42(24): 7809-18, 2003 Dec 01.
Article in English | MEDLINE | ID: mdl-14632496

ABSTRACT

Several synthetic approaches for the preparation of double metal cyanide (DMC) derivatives of iron(II) and zinc(II) are described. These include (1) metathesis reactions of ZnCl(2) or ZnI(2) with KCpFe(CN)(2)CO in aqueous solution, (2) reactions of KCpFe(CN)(2)CO and its phosphine-substituted analogues with Zn(CH(3)CN)(4)(BF(4))(2) and subsequent displacement of acetonitrile at the zinc centers by the addition of a neutral (phosphine) or anionic (phenoxide) ligand, and (3) reactions of the protonated HCpFe(CN)(2)(phosphine) complexes with Zn(N(SiMe(3))(2))(2), followed by the addition of phenols. All structures are based on a diamond-shaped planar arrangement of the Fe(2)(CN)(4)Zn(2) core with various appended ligands at the metal sites. Although attempts to replace the iodide ligands in [CpFe(mu-CN)(2)PPh(3)ZnI(THF)](2) with acetate using silver acetate failed, two novel cationic mixed-metal cyanide salts based on the [CpFe(PPh(3))(mu-CN)(2)Zn(NC(5)H(5))](2)(2+) framework were isolated from pyridine solution and their structures were defined by X-ray crystallography. The anionic ligand bound to zinc in these derivatives, which serve as an anionic polymerization initiator, was shown to be central to the catalytic copolymerization reaction of CO(2)/epoxide to provide polycarbonates and cyclic carbonates. The structurally stabilized phosphine-strapped complexes [CpFe(mu-CN)(2)Zn(X)THF](2)(mu-dppp), where X = I or phenolate, were shown to be thermally stable under the conditions (80 degrees C) of the copolymerization reaction by in situ infrared spectroscopy. Both of these derivatives were proposed to serve as mimics for the heterogeneous DMC catalysts in the patent literature, with the derivative where the initiator is a phenolate being more active for the production of polycarbonates.

10.
Inorg Chem ; 42(21): 6915-22, 2003 Oct 20.
Article in English | MEDLINE | ID: mdl-14552643

ABSTRACT

The synthesis of nickel(II) and palladium(II) salicylaldiminato complexes incorporating the water-soluble phosphine 1,3,5-triaza-7-phosphaadamantane(PTA) has been achieved employing two preparative routes. Reaction of the original ethylene polymerization catalyst developed by Grubbs and co-workers (Organometallics 1998, 17, 3149), (salicylaldiminato)Ni(Ph)PPh(3), with PTA using a homogeneous methanol/toluene solvent system resulted in the formation of the PTA analogues in good yields. Alternatively, complexes of this type may be synthesized via a direct approach utilizing (tmeda)M(CH(3))(2) (M = Ni, Pd), the corresponding salicylaldimine, and PTA. Yields by this method were generally near quantitative. The complexes were characterized in solution by (1)H/(13)C/(31)P NMR spectroscopy and in the solid-state by X-ray crystallography. All derivatives exhibited square-planar geometry with the bulky isopropyl groups on the aniline being perpendicular to the plane formed by the metal center and its four ligands. Such orientation of these sterically encumbering groups is responsible for polymer chain growth during olefin polymerization in favor of chain termination via beta-hydride elimination. Polymerization reactions were attempted using the nickel-PTA complexes in a biphasic toluene/water mixture in an effort to initiate ethylene polymerization by trapping the dissociated phosphine ligand in the water layer, thereby eliminating the need for a phosphine scavenger. Unfortunately, because of the strong binding ability of the small, donating phosphine(PTA) as compared to PPh(3), phosphine dissociation did not occur at a temperature where the complexes are thermally stable.

11.
J Am Chem Soc ; 125(25): 7586-91, 2003 Jun 25.
Article in English | MEDLINE | ID: mdl-12812499

ABSTRACT

The catalysis of the reaction of carbon dioxide with epoxides (cyclohexene oxide or propylene oxide) using the (salen)Cr(III)Cl complex as catalyst, where H(2)salen = N,N'-bis(3,5-di-tert-butylsalicylidene)-1,2-cyclohexenediimine (1), to provide copolymer and cyclic carbonate has been investigated by in situ infrared spectroscopy. As previously demonstrated for the cyclohexene oxide/CO(2) reaction in the presence of complex 1, coupling of propylene oxide and carbon dioxide was found to occur by way of a pathway first-order in catalyst concentration. Unlike the cyclohexene oxide/carbon dioxide reaction catalyzed by complex 1, which affords completely alternating copolymer and only small quantities of trans-cyclic cyclohexyl carbonate, under similar conditions propylene oxide/carbon dioxide produces mostly cyclic propylene carbonate. Comparative kinetic measurements were performed as a function of reaction temperature to assess the activation barrier for production of cyclic carbonates and polycarbonates for the two different classes of epoxides, i.e., alicyclic (cyclohexene oxide) and aliphatic (propylene oxide). As anticipated in both instances the unimolecular pathway for cyclic carbonate formation has a larger energy of activation than the bimolecular enchainment pathway. That is, the energies of activation determined for cyclic propylene carbonate and poly(propylene carbonate) formation were 100.5 and 67.6 kJ.mol(-1), respectively, compared to the corresponding values for cyclic cyclohexyl carbonate and poly(cyclohexylene carbonate) production of 133 and 46.9 kJ.mol(-1). The small energy difference in the two concurrent reactions for the propylene oxide/CO(2) process (33 kJ.mol(-1)) accounts for the large quantity of cyclic carbonate produced at elevated temperatures in this instance.

12.
Inorg Chem ; 42(2): 581-9, 2003 Jan 27.
Article in English | MEDLINE | ID: mdl-12693242

ABSTRACT

The reactions of zinc halides with 2,6-di-methoxypyridine or 3-trifluoromethylpyridine in dichloromethane have led to the formation of quite different complexes. Specifically, reactions involving pyridine containing electron donating methoxy substitutents have provided salts of the type [Zn(2,6-dimethoxypyridine)4][Zn2X6], as revealed by elemental analysis and X-ray crystallography. On the other hand, simple bis-pyridine adducts of zinc halides were isolated from the reactions involving the pyridine ligand with electron withdrawing substituents and characterized by X-ray crystallography, for example, Zn(3-trifluoromethylpyridine)2Br2. These zinc complexes were shown to be catalytically active for the coupling of carbon dioxide and epoxides to provide high molecular weight polycarbonates and cyclic carbonates, with the order of reactivity being Cl > or = Br > I, and 2,6-di-methoxypyridine > 3-trifluoromethylpyridine. Polycarbonate production from carbon dioxide and cyclohexene oxide was shown to be first-order in both metal precursor complex and cyclohexene oxide, as monitored by in situ infrared spectroscopy at 80 degrees C and 55 bar pressure. For reactions carried out in CO2 swollen epoxide solutions in the absence of added quantities of pyridine, the copolymer produced contained significant polyether linkages. Alternatively, reactions performed in the presence of excess pyridine or in hydrocarbon solvent, although slower in rate, afforded completely alternating copolymers. For comparative purposes, zinc chloride was a very effective homopolymerization catalyst for polyethers. Additionally, zinc chloride afforded copolymers with 60% carbonate linkages in the presence of high carbon dioxide pressures. In the case of cyclohexene oxide, the copolymer back-biting reaction led exclusively to the production of the trans cyclic carbonate as shown by infrared spectroscopy in v(C=O) region and X-ray crystallography. The unique feature of these catalyst systems is their simplicity.

13.
J Am Chem Soc ; 124(24): 7075-83, 2002 Jun 19.
Article in English | MEDLINE | ID: mdl-12059232

ABSTRACT

The reaction of Cd[N(SiMe(3))(2)](2) with 2 equiv of the corresponding phenol in toluene has led to the isolation of [Cd(O-2,6-R(2)C(6)H(3))(2)](2) derivatives, where R represents the sterically bulky (t)Bu and Ph substituents. The dimeric nature of these complexes in the solid state has been established via X-ray crystallography, i.e., trigonal geometry around cadmium is observed in 1 (R = (t)Bu) where the two cadmium centers are bridged by two phenoxides with each metal containing a terminal phenoxide. Complex 2 (R = Ph) contains an additional interaction of the metal centers with carbon atoms of the aromatic substituents on the phenoxide ligands. These dimeric structures are maintained in weakly coordinating solvents as revealed by (113)Cd NMR in d(2)-methylene chloride, which displays (111)Cd-(113)Cd coupling. Nevertheless, because of the excessive steric requirements of these phenoxide ligands, these dimers are easily disrupted in solution by weak donor ligands such as epoxides. Three bisepoxide adducts have been isolated as crystalline solids and characterized by X-ray crystallography. As previously observed in other Cd(O-2,6-(t)Bu(2)C(6)H(3))(2) x L(2) complexes, these epoxide adducts adopt a crystallographically imposed square-planar geometry about the cadmium centers, with the exception of the exo-2,3-epoxynorbornane derivative, which displays a distorted tetrahedral geometry. Temperature-dependent (113)Cd NMR studies have established that there is little difference in the binding abilities of these epoxides with either complex 1 or complex 2. Importantly, it is concluded from these studies that the lack of reactivity of alpha-pinene oxide and exo-2,3-epoxynorbornane toward copolymerization reactions with carbon dioxide, in the presence of zinc bisphenoxide catalysts, is not due to differences in epoxide metal binding. This is further affirmed by the isolation and crystallographic characterization of the very stable Zn(O-2,6-(t)Bu(2)C(6)H(3))(2) x (exo-2,3-epoxynorbornane)(2) derivative.

14.
J Am Chem Soc ; 124(22): 6335-42, 2002 Jun 05.
Article in English | MEDLINE | ID: mdl-12033862

ABSTRACT

The air-stable, chiral (salen)Cr(III)Cl complex (3), where H(2)salen = N,N'-bis(3,5-di-tert-butyl-salicylidene)-1,2-cyclohexene diamine, has been shown to be an effective catalyst for the coupling of cyclohexene oxide and carbon dioxide to afford poly(cyclohexenylene carbonate), along with a small quantity of its trans-cyclic carbonate. The thus produced polycarbonate contained >99% carbonate linkages and had a M(n) value of 8900 g/mol with a polydispersity index of 1.2 as determined by gel permeation chromatography. The turnover number (TON) and turnover frequency (TOF) values of 683 g of polym/g of Cr and 28.5 g of polym/g of Cr/h, respectively for reactions carried out at 80 degrees C and 58.5 bar pressure increased by over 3-fold upon addition of 5 equiv of the Lewis base cocatalyst, N-methyl imidazole. Although this chiral catalyst is well documented for the asymmetric ring-opening (ARO) of epoxides, in this instance the copolymer produced was completely atactic as illustrated by (13)C NMR spectroscopy. Whereas the mechanism for the (salen)Cr(III)-catalyzed ARO of epoxides displays a squared dependence on [catalyst], which presumably is true for the initiation step of the copolymerization reaction, the rate of carbonate chain growth leading to copolymer or cyclic carbonate formation is linearly dependent on [catalyst]. This was demonstrated herein by way of in situ measurements at 80 degrees C and 58.5 bar pressure. Hence, an alternative mechanism for copolymer production is operative, which is suggested to involve a concerted attack of epoxide at the axial site of the chromium(III) complex where the growing polymer chain for epoxide ring-opening resides. Preliminary investigations of this (salen)Cr(III)-catalyzed system for the coupling of propylene oxide and carbon dioxide reveal that although cyclic carbonate is the main product provided at elevated temperatures, at ambient temperature polycarbonate formation is dominant. A common reaction pathway for alicyclic (cyclohexene oxide) and aliphatic (propylene oxide) carbon dioxide coupling is thought to be in effect, where in the latter instance cyclic carbonate production has a greater temperature dependence compared to copolymer formation.

15.
Inorg Chem ; 41(4): 973-80, 2002 Feb 25.
Article in English | MEDLINE | ID: mdl-11849101

ABSTRACT

Zinc complexes derived from benzoic acids containing electron-withdrawing substituents have been synthesized from Zn(II)(bis-trimethylsilyl amide)(2) and the corresponding carboxylic acid (2,6-X(2)C(6)H(3)COOH, where X = F, Cl, or OMe) in THF and structurally characterized via X-ray crystallography. The 2,6-difluorobenzoate complex crystallizes from THF or CH(3)CN as a seven membered zinc aggregate, where the metal atoms are interconnected by a combination of 10 mu-benzoates and mu(4)-oxo ligands, that is, [(2,6-difluorobenzoate)(10)O(2)Zn(7)](solvent)(2), solvent = THF (1) and CH(3)CN (1a). On the other hand, the 2,6-dichlorobenzoate zinc derivative crystallizes from THF as a dimer, [(2,6-dichlorobenzoate)(4)Zn(2)](THF)(3) (2), where the two zinc centers are bridged by three benzoate ligand. One of the zinc centers possesses a tetrahedral ligand environment where the fourth ligand is a unidentate benzoate, and the other zinc center has an octahedral arrangement of ligands which is accomplished by the additional binding of three THF molecules. Upon dissolution of complex 1 or 2 in the strongly binding pyridine solvent, disruption of these zinc carboxylates occurs with concomitant formation of mononuclear zinc bis-benzoates with three pyridine ligands in the metal coordination sphere. Complexes 1 and 2 were found to be effective catalysts for the copolymerization of cyclohexene oxide and carbon dioxide to afford polycarbonates devoid of polyether linkages, that is, completely alternating copolymers. Although these catalysts or catalyst precursors in the presence of CO(2)/propylene oxide afforded mostly propylene carbonate, they did serve as efficient catalysts for the terpolymerization of carbon dioxide/cyclohexene oxide/propylene oxide. The reactivities of these zinc carboxylates were very similar to those previously reported analogous complexes which have not been structurally characterized. Hence, it is suggested here that all of these zinc carboxylates provide similar catalytic sites for CO(2)/epoxide coupling processes.

SELECTION OF CITATIONS
SEARCH DETAIL
...