Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 16 de 16
Filter
Add more filters










Publication year range
1.
J Tissue Viability ; 29(4): 291-296, 2020 Nov.
Article in English | MEDLINE | ID: mdl-32653294

ABSTRACT

AIM: The purpose of this study was to investigate the degradation of desmocollin-1 (DSC1), a member of the desmosomal cadherin family in patients with diabetes, as well as the factors associated with the suppression of DSC1 degradation. METHODS: This cross-sectional study included 60 cases of foot callus involving 30 patients with diabetes (DM) and 30 matched volunteers without diabetes (non-DM). DSC1 degradation in samples from debrided calluses was analysed using western blotting. Skin hydration, a factor reported to suppress DSC1 degradation, was measured using a mobile moisture device. RESULTS: Full-length DSC1 (approximately 100 kDa) was detected in six participants only in the DM group, and no relationship was found between the suppression of DSC1 degradation and decreased skin hydration in the DM group. There was no significant difference in skin hydration values between the DM and non-DM groups. CONCLUSION: DSC1 degradation was suppressed in the DM group. There was no relationship between the suppression of DSC1 degradation and decreased skin hydration in the DM group. Current external force callus care would not be sufficient. This study highlights the need to develop novel callus care to enhance the degradation of DSC1.


Subject(s)
Bony Callus/physiopathology , Desmocollins/analysis , Skin/physiopathology , Aged , Aged, 80 and over , Ankle Brachial Index , Blotting, Western/methods , Body Mass Index , Cross-Sectional Studies , Diabetes Complications , Female , Foot/physiopathology , Humans , Male , Middle Aged , Statistics, Nonparametric
2.
Wound Repair Regen ; 27(5): 540-547, 2019 09.
Article in English | MEDLINE | ID: mdl-31145519

ABSTRACT

A biofilm plays a crucial role in delaying wound healing. Sharp debridement, a possible effective method for eliminating biofilms, can only be applied to the wound with visible necrotic tissue; thus, no option has been available for eliminating biofilms that are not accompanied by necrotic tissue. Wound blotting was recently developed to visualize biofilm noninvasively and quickly, and ultrasonic debridement is available for biofilm removal. Therefore, the purpose of this study was to investigate the efficacy of "biofilm-based wound care system (BWCS)," a combination of wound blotting as a point-of-care testing and ultrasonic debridement, for promoting wound healing. Firstly, the cross-sectional study was conducted to examine the proportion of biofilm removal by ultrasonic debridement in pressure ulcers [Study 1]. Subsequently, the retrospective cohort study was conducted to examine the effectiveness of BWCS for healing of chronic wounds [Study 2]. The proportions of wound healing between wounds treated with BWCS and those with standard care in the home-visiting clinic were compared by Kaplan-Meier curve, and the Cox proportional hazard modeling was used to assess the effect of BWCS on wound healing. In Study 1, the median of biofilm removal proportion was 38.9% (interquartile range, 12.9-68.0%) for pressure ulcers treated with standard care and 65.2% (41.1-78.8%) for those treated with ultrasonic debridement (p = 0.009). In Study 2, the proportion of wound healing within 90 days was significantly higher in wounds treated with BWCS than in those treated with standard care (p = 0.001). The adjusted hazard ratio of BWCS for wound healing was 4.5 (95% confidence interval, 1.3-15.0; p = 0.015). In conclusion, we demonstrated that our novel approach, BWCS, can be a promising therapeutic strategy for visualizing biofilms that are not accompanied by necrotic tissue and promoting healing in chronic wounds.


Subject(s)
Debridement/methods , Necrosis/therapy , Pressure Ulcer/therapy , Wound Healing/physiology , Wound Infection/therapy , Aged , Aged, 80 and over , Anti-Infective Agents, Local , Biofilms , Cross-Sectional Studies , Debridement/instrumentation , Female , Humans , Male , Middle Aged , Necrosis/microbiology , Point-of-Care Systems , Pressure Ulcer/microbiology , Treatment Outcome , Wound Infection/microbiology
3.
J Phys Chem A ; 118(25): 4415-24, 2014 Jun 26.
Article in English | MEDLINE | ID: mdl-24866729

ABSTRACT

Hemiporphyrazines are a large family of phthalocyanine analogues in which two isoindoline units are replaced by other rings. Here we report unambiguous identification of 20π-electron structure of triazolehemiporphyrazines (1, 2) and thiazolehemiporphyrazine (3) by means of X-ray analysis, various spectroscopic methods, and density functional theory (DFT) calculations. The hemiporphyrazines were compared in detail with dibenzotetraazaporphyrin (4), a structurally related 18π-electron molecule. X-ray analysis revealed that tetrakis(2,6-dimethylphenyloxy)triazolehemiporphyrazine (1b) adopted planar geometry in the solid state. A weak absorption band with a pronounced vibronic progression, observed for all the hemiporphyrazines, was attributed to the lowest π-π* transition with the electric-dipole-forbidden nature. In the case of intrinsically chiral vanadyl triazolehemiporphyrazine (2), a large dissymmetry (g) factor was detected for the CD signal corresponding to the lowest π-π* transition with the magnetic-dipole-allowed nature. Molecular orbital analysis and NICS calculations showed that the azolehemiporphyrazines have a 20π-electron system with a weak paratropic ring current.


Subject(s)
Azo Compounds/chemistry , Azoles/chemistry , Electrons , Magnetic Resonance Spectroscopy , Porphyrins/chemistry , Quantum Theory , Molecular Structure
4.
J Phys Chem A ; 117(41): 10576-87, 2013 Oct 17.
Article in English | MEDLINE | ID: mdl-24053738

ABSTRACT

The proton location and proton transfer (PT) dynamics of a hydrogen bond are under the influence of the static and dynamical properties of the solvent and counterions. In the present study, the N-H distances were determined for salts of 1,8-bis(dimethylamino)naphthalene, DMANH(+)X(-) (X(-) = BPh4(-), ClO4(-), and Cl(-)), in acetonitrile (AN) solution, and DMANH(+)Br(-) in water by observing the (15)N spin-lattice relaxation caused by the (15)N-(1)H magnetic dipolar coupling under assumption that the PT time was shorter than the NH reorientation time (∼10(-11) s). The obtained N-H distances decreased in the following order: DMANH(+)BPh4(-) > DMANH(+)ClO4(-) > DMANH(+)Br(-)/H2O > DMANH(+)Cl(-), indicating that interactions with the environment affect the PT potentials. To understand the results at the molecular level, Car-Parrinello molecular dynamics simulations were performed for DMANH(+), DMANH(+) in water, and DMANH(+)-Cl(-) ion-pair in AN. The results of simulation suggest that (1) the N-H distance decreases in the presence of a solvent and counterion; (2) the PT time is probably ∼10(-12) s, which confirms the above assumption used for the NMR relaxation data analyses; and (3) fluctuation of the interactions with the solvent or counterion has a significant role in PT. Quantum nuclear effects on the hydrogen bond were also examined.


Subject(s)
1-Naphthylamine/analogs & derivatives , Hydrogen/chemistry , Nitrogen/chemistry , Protons , Solvents/chemistry , 1-Naphthylamine/chemistry , Hydrogen Bonding , Ions , Magnetic Resonance Spectroscopy , Molecular Structure
5.
Dalton Trans ; 42(19): 6873-8, 2013 May 21.
Article in English | MEDLINE | ID: mdl-23503591

ABSTRACT

Vibrational circular dichroism (VCD) spectroscopy was applied to study the structural change of a Ni(II) complex (denoted by [Ni(II)L]) with a chiral Schiff base ligand, (R)- or (S)-2,2'-bis(salicylideneamino)-1,1'-binaphthyl (denoted by H2L), in solution. The major signals in the mid-IR region were assigned on the basis of comparison with the DFT-calculated spectra. The complex transformed reversibly between the square-planar, tetrahedral and octahedral configurations, depending on solvents and temperature. The observed changes in the VCD peaks accompanying the transformation were analyzed in terms of the conformational change of the chiral ligand with a focus on the twisting angle in the Schiff base backbone and the dihedral angle of the binaphthyl group.


Subject(s)
Coordination Complexes/chemistry , Naphthalenes/chemistry , Nickel/chemistry , Schiff Bases/chemistry , Circular Dichroism , Coordination Complexes/chemical synthesis , Quantum Theory , Solvents/chemistry , Temperature
6.
Dalton Trans ; 42(21): 7579-85, 2013 Jun 07.
Article in English | MEDLINE | ID: mdl-23536142

ABSTRACT

Mononuclear and dinuclear square planar palladium(II) complexes (denoted by [(hfac)Pd(II)(L-LH)] and [(hfac)Pd(II)(L-L)Pd(II)(hfac)], respectively) were synthesized. Here hfac(-), HL-L(-) and L-L(2-) denote hexafluoroacetylacetonato, monoprotonated and non-protonated bis-ß-diketonato ligands, respectively. Three bis-ß-diketones were used as HL-LH: 1,2-diacetyl-1,2-dibenzoylethane (denoted by dabeH2), 1,2-diacetyl-1,2-bis(3-methylbutanoyl)ethane (baetH2) and 1,2-diacetyl-1,2-propanoylethane (dpeH2). Both the monomeric and dimeric Pd(II) complexes were chiral due to the orthogonal twisting of the two non-symmetric diketonato moieties in HL-L(-) and L-L(2-), respectively. Optical resolution of [(hfac)Pd(II)(dabe)Pd(II)(hfac)] was achieved chromatographically on a chiral column to obtain a pair of optical antipodes which were stable against racemization. As for the other complexes, resolution was possible only after replacing hfac(-) with a bulky ligand such as dibenzoylmethanato (dbm(-)). Although a dinuclear complex with a symmetric bis-ß-diketonato ligand, [(hfac)Pd(II)(taet)Pd(II)(hfac)] (taet(2-) = 1,1,2,2-tetraacetylethanato), was achiral, the replacement of the terminal ligands with non-symmetric ß-diketonates yielded an axially chiral complex such as [(phacac)Pd(II)(taet)Pd(II)(phacac)], wherein phacac(-) denotes 1-phenyl-1,3-butanedionato. The UV and CD spectra of the Pd(II) complexes were analyzed with the help of the TDDFT calculations. The chiral monomeric species, [(dbm)Pd(II)(R- or S-baetH)], formed a heterometallic tetranuclear complex, [Fe(III){(dbm)Pd(II)(R- or S-baet)}3], in methanol solution.

7.
Dalton Trans ; 42(1): 232-7, 2013 Jan 07.
Article in English | MEDLINE | ID: mdl-23037769

ABSTRACT

Mononuclear and dinuclear ruthenium(III) complexes with bis-ß-diketonato ligands (denoted by [Ru(acac)(2)(L-LH)] and [Ru(acac)(2)(L-L)Ru(acac)(2)], respectively) were synthesized, where acac, L-LH(-) and L-L(2-) denote acetylacetonato, monoprotonated and unprotonated bis-ß-diketonato ligands, respectively. The following three ligands were used as the bis-ß-diketonato ligand (L-L(2-)): 1,2-diacetyl-1,2-dibenzoylethanato (denoted by dabe(2-)), 1,2-diacetyl-1,2-bis(3-methylbutanoyl)ethanato (baet(2-)) and 1,2-diacetyl-1,2-dipropanoylethanato (dpe(2-)). For the mononuclear and the meso-type dinuclear complexes, a pair of diastereomeric species were identified as Δ- (or Λ-) [Ru(acac)(2)(R- or S-L-LH)] and [Δ-Ru(acac)(2)(R- or S-L-L)Λ-Ru(acac)(2)], respectively. The possibility of thermal inversion in coordinated L-LH(-) (mononuclear) or L-L(2-) (dinuclear) was pursued by monitoring the changes in the electronic circular dichroism or the (1)H NMR spectra. No inversion occurred for the dinuclear complexes, when their chloroform solutions were kept at 50 °C for ca. 100 h. In contrast, some of the mononuclear complexes underwent the inversion of axial chirality to give an equilibrium mixture under the same conditions. The reaction followed the first-order rate law and the overall first-order rate constants (k) of [Ru(acac)(2)(L-LH)] were determined to be k = 0.13, 0.0048 and less than 0.001 h(-1) for L-LH(-) = dabeH(-), baetH(-) and dpeH(-), respectively. The results suggest that the main factor determining the barrier height of the internal rotation is not the steric but the electronic properties of the carbon-carbon bond connecting the two ß-diketonato moieties.

8.
Phys Chem Chem Phys ; 14(31): 11090-8, 2012 Aug 21.
Article in English | MEDLINE | ID: mdl-22767155

ABSTRACT

Two-dimensional potential energy surfaces (PESs) were calculated for the degenerate intramolecular proton transfer (PT) in two N-H···N hydrogen-bonded systems, (Z)-2-(2-pyridylmethylidene)-1,2-dihydropyridine (1) and monoprotonated di(2-pyridyl) ether (2), at the MP2/cc-pVDZ level of theory. The calculated PES had two minima in both cases. The energy barrier in 1 was higher than the zero-point energy (ZPE) level, while that in 2 was close to the ZPE. Vibrational wavefunctions were obtained by solving time-independent Schrödinger equations with the calculated PESs. The maximum points of the probability density were shifted from the energy minima towards the region where the covalent N-H bond was elongated and the N···N distance shortened. The effects of a polar solvent on the PES were investigated with the continuum or cluster models in such a way that the solute-solvent electrostatic interactions could be taken into account under non-equilibrated conditions. A solvated contact ion-pair was modelled by a cluster consisting of one cation 2, one chloride ion and 26 molecules of acetonitrile. The calculation with this model suggested that the bridging proton is localised in the deeper well due to the significant asymmetry of the PES and the high potential barrier.


Subject(s)
Models, Theoretical , Pyridines/chemistry , Pyridinium Compounds/chemistry , Hydrogen/chemistry , Hydrogen Bonding , Nitrogen/chemistry , Protons , Solvents/chemistry
9.
Molecules ; 17(2): 2161-8, 2012 Feb 21.
Article in English | MEDLINE | ID: mdl-22354192

ABSTRACT

Ab initio calculations were carried out for a benzyl-substituted iminium cation derived from (E)-crotonaldehyde and a chiral imidazolidinone that was developed as an organocatalyst by MacMillan et al. At the MP2 level of theory it is predicted that the phenyl group is close to the iminium moiety in the most stable conformer, suggesting that the cation-π interaction contributes to the stabilization of this conformer. Energy decomposition analyses on model systems indicate that the electrostatic and polarization terms make significant contribution to the attractive interactions between the benzene ring and the iminium cation.


Subject(s)
Aldehydes/chemistry , Imidazolidines/chemistry , Imines/chemistry , Benzene/chemistry , Catalysis , Cations/chemistry , Models, Chemical , Static Electricity
10.
Dalton Trans ; 41(3): 747-51, 2012 Jan 21.
Article in English | MEDLINE | ID: mdl-21897931

ABSTRACT

Dinuclear Ru(III) complexes, [Ru(III)(acac)(2)(dabe)Ru(III)(acac)(2)] (acacH = acetylacetone; dabeH(2) = 1, 2-diacetyl-1,2-dibenzoylethane) and [Ru(III)(acac)(2)(tbet)Ru(III)(acac)(2)] (tbetH(2) = 1,1,2,2-tetrabenzoylethane) were synthesized by reacting [Ru(acac)(2)(CH(3)CN)(2)]PF(6) with dabeH(2) and tbetH(2) respectively, in toluene. The X-ray structural analysis of a meso-type dinuclear Ru(III) complex, ΔΛ-[Ru(III)(acac)(2)(dabe)Ru(III)(acac)(2)], showed that the bridging part became chiral due to the orthogonal twisting of two non-symmetrical ß-diketonato moieties. To confirm this conclusion, the complex was resolved chromatographically to provide a pair of optical antipodes. Such chirality in the bridging part was not generated for [Ru(III)(acac)(2)(tbet)Ru(III)(acac)(2)], because the ß-diketonato moieties in tbet(2-) are symmetrical.

11.
Chem Commun (Camb) ; 46(30): 5473-5, 2010 Aug 14.
Article in English | MEDLINE | ID: mdl-20596573

ABSTRACT

Star-burst tetranuclear Ru(iii) complexes, Delta-(or Lambda-)[{Delta-(or Lambda-)-Ru(iii)(acac)(2)(taet)}(3)Ru(iii)] (acac = acetylacetonato; taet = tetraacetylethanato), were prepared and resolved into eight enantiomers on a column packed with an ion-exchange adduct of a clay and Delta-[Ru(ii)(phen)(3)](2+) (phen = 1,10-phenanthroline), demonstrating that the dual chiralities of the central and peripheral regions were discriminated on the modified clay surface.

12.
FEBS J ; 276(17): 4763-79, 2009 Sep.
Article in English | MEDLINE | ID: mdl-19656186

ABSTRACT

The ATP-pyrophosphate exchange reaction catalyzed by Arg-tRNA, Gln-tRNA and Glu-tRNA synthetases requires the assistance of the cognate tRNA. tRNA also assists Arg-tRNA synthetase in catalyzing the pyrophosphorolysis of synthetic Arg-AMP at low pH. The mechanism by which the 3'-end A76, and in particular its hydroxyl group, of the cognate tRNA is involved with the exchange reaction catalyzed by those enzymes has yet to be established. We determined a crystal structure of a complex of Arg-tRNA synthetase from Pyrococcus horikoshii, tRNA(Arg)(CCU) and an ATP analog with Rfactor = 0.213 (Rfree = 0.253) at 2.0 A resolution. On the basis of newly obtained structural information about the position of ATP bound on the enzyme, we constructed a structural model for a mechanism in which the formation of a hydrogen bond between the 2'-OH group of A76 of tRNA and the carboxyl group of Arg induces both formation of Arg-AMP (Arg + ATP --> Arg-AMP + pyrophosphate) and pyrophosphorolysis of Arg-AMP (Arg-AMP + pyrophosphate --> Arg + ATP) at low pH. Furthermore, we obtained a structural model of the molecular mechanism for the Arg-tRNA synthetase-catalyzed deacylation of Arg-tRNA (Arg-tRNA + AMP --> Arg-AMP + tRNA at high pH), in which the deacylation of aminoacyl-tRNA bound on Arg-tRNA synthetase and Glu-tRNA synthetase is catalyzed by a quite similar mechanism, whereby the proton-donating group (-NH-C+(NH2)2 or -COOH) of Arg and Glu assists the aminoacyl transfer from the 2'-OH group of tRNA to the phosphate group of AMP at high pH.


Subject(s)
Adenylyl Imidodiphosphate/chemistry , Arginine-tRNA Ligase/chemistry , Arginine/chemistry , RNA, Transfer, Amino Acyl/chemistry , Amino Acid Sequence , Binding Sites , Crystallography, X-Ray , Hydrogen Bonding , Models, Molecular , Molecular Sequence Data , Organophosphates/chemistry , Protein Binding , Pyrococcus horikoshii/enzymology , Saccharomyces cerevisiae/enzymology , Thermus thermophilus/enzymology
13.
Inorg Chem ; 48(10): 4354-61, 2009 May 18.
Article in English | MEDLINE | ID: mdl-19391627

ABSTRACT

Twenty four kinds of Ru(III) complexes expressed by the formula of Delta- or Lambda-[Ru((-)- or (+)-tfac)(n)(acac)(3-n)] (n = 1, 2, and 3, (-)- or (+)-tfac = (-)- or (+)-3-trifluoroacetylcamphorato and acac = acetylacetonato) were prepared in a pure diastereomeric form. The separation of these diastereomers was accomplished chromatographically by rational use of two antipodal chiral columns. The separated complexes were identified by means of mass spectra, (1)H NMR, electronic circular dichroism, and partly X-ray diffraction analyses. When a pair of Delta- and Lambda-[Ru(acac)(3)] (n = 0) was added to this group, they constituted the complete series of mixed ligand complexes including ligand chirality under the condition of no mixing of chiral ligands. The vibrational circular dichroism (VCD) spectra of their CDCl(3) solutions were recorded in the wavenumber region of 1000 approximately 1800 cm(-1). The results provided a benchmark for systematically examining the effects of the stereochemical properties on VCD spectra such as the degree of ligand substitution, DeltaLambda configurations, geometrical isomerism, and ligand chirality. As a result, the geometrical isomers of trans- or cis-[Ru((-)- or (+)-tfac)(2)(acac)] and mer- or fac-[Ru((-)- or (+)-tfac)(3)] were clearly differentiated by their VCD signals, which was hardly possible from their IR spectra alone.

14.
J Phys Chem A ; 111(24): 5136-43, 2007 Jun 21.
Article in English | MEDLINE | ID: mdl-17530835

ABSTRACT

The ultrafast electron transfer occurring upon Soret excitation of three new porphyrin-ferrocene (XP-Fc) dyads has been studied by femtosecond up-conversion and pump-probe techniques. In the XP-Fc dyads (XP-Fcs) designed in this study, the ferrocene moiety is covalently bonded to the meso positions of 3,5-di-tert-butylphenyl zinc porphyrin (BPZnP-Fc), pentafluorophenyl zinc porphyrin (FPZnP-Fc), and 3,5-di-tert-butylphenyl free-base porphyrin (BPH2P-Fc). Charge separation and recombination in the XP-Fcs were confirmed by transient absorption spectra, and the lifetimes of the charge-separated states were estimated from the decay rate of the porphyrin radical anion band to be approximately 20 ps. The charge-separation rates of the XP-Fcs were found to be >10(13) s-1 from the S2 state and 6.3x10(12) s-1 from the S1 state. Charge separation from the S2 state was particularly efficient for BPZnP-Fc, whereas the main reaction pathway was from the S1 state for BPH2P-Fc. Charge separation from the S2 and S1 states occurred at virtually the same rate in benzene and tetrahydrofuran and was much faster than their solvation times. Analysis of these results using semiquantum Marcus theory indicates that the magnitude of the electronic-tunneling matrix element is rather large and far outside the range of nonadiabatic approximation. The pump-probe data show the presence of vibrational coherence during the reactions, suggesting that wavepacket dynamics on the adiabatic potential energy surface might regulate the ultrafast reactions.


Subject(s)
Ferrous Compounds/chemistry , Porphyrins/chemistry , Electrons , Ferrous Compounds/metabolism , Metallocenes , Molecular Structure , Photochemistry , Porphyrins/metabolism , Thermodynamics , Time Factors
15.
Chem Commun (Camb) ; (18): 2066-7, 2004 Sep 21.
Article in English | MEDLINE | ID: mdl-15367977

ABSTRACT

A hydrogen bonding effect on photocurrent generation has been evaluated successfully in a mixed film of porphyrin and fullerene with hydrogen bonding on an ITO electrode, which exhibits efficient cathodic photocurrent generation as compared to the reference system without hydrogen bonding.

16.
Chemistry ; 10(2): 474-83, 2004 Jan 23.
Article in English | MEDLINE | ID: mdl-14735516

ABSTRACT

Zinc porphyrin-naphthalenediimide (ZnP-NIm) dyads and zinc porphyrin-pyromellitdiimide-naphthalenediimide (ZnP-Im-NIm) triad have been employed to examine the effects of metal ions on photoinduced charge-separation (CS) and charge-recombination (CR) processes in the presence of metal ions (scandium triflate (Sc(OTf)(3)) or lutetium triflate (Lu(OTf)(3)), both of which can bind with the radical anion of NIm). Formation of the charge-separated states in the absence and in the presence of Sc(3+) was confirmed by the appearance of absorption bands due to ZnP(.) (+) and NIm(.) (-) in the absence of metal ions and of those due to ZnP(.) (+) and the NIm(.) (-)/Sc(3+) complex in the presence of Sc(3+) in the time-resolved transient absorption spectra of dyads and triad. The lifetimes of the charge-separated states in the presence of 1.0 x 10(-3) M Sc(3+) (14 micros for ZnP-NIm, 8.3 micros for ZnP-Im-NIm) are more than ten times longer than those in the absence of metal ions (1.3 micros for ZnP-NIm, 0.33 micros for ZnP-Im-NIm). In contrast, the rate constants of the CS step determined by the fluorescence lifetime measurements are the same, irrespective of the presence or absence of metal ions. This indicates that photoinduced electron transfer from (1)ZnP(*) to NIm in the presence of Sc(3+) occurs without involvement of the metal ion to produce ZnP(.) (+)-NIm(.) (-), followed by complexation with Sc(3+) to afford the ZnP(.) (+)-NIm(.) (-)/Sc(3+) complex. The one-electron reduction potential (E(red)) of the NIm moiety in the presence of a metal ion is shifted in a positive direction with increasing metal ion concentration, obeying the Nernst equation, whereas the one-electron oxidation potential of the ZnP moiety remains the same. The driving force dependence of the observed rate constants (k(ET)) of CS and CR processes in the absence and in the presence of metal ions is well evaluated in terms of the Marcus theory of electron transfer. In the presence of metal ions, the driving force of the CS process is the same as that in the absence of metal ions, whereas the driving force of the CR process decreases with increasing metal ion concentration. The reorganization energy of the CR process also decreases with increasing metal ion concentration, when the CR rate constant becomes independent of the metal ion concentration.

SELECTION OF CITATIONS
SEARCH DETAIL
...