首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Dermatophytes are fungi that cause infections of keratinized tissues. We have recently demonstrated the susceptibility of the dermatophyte Trichophyton rubrum to photodynamic treatment (PDT) with 5,10,15-Tris(4-methylpyridinium)-20-phenyl-[21 H ,23 H ]-porphine trichloride (Sylsens B) in 5 m m citric acid/sodium citrate buffer (pH 5.2, formulation I). In this work, we examined the penetration of Sylsens B in healthy and with T. rubrum infected skin and we investigated the susceptibility of T. rubrum to PDT using formulation I and UVA-1 radiation (340–550 nm). Skin penetration studies were performed with formulations I and II (Sylsens B in PBS, pH 7.4) applied on dermatomed skin, human stratum corneum (SC), disrupted SC by T. rubrum growth and SC pretreated with a detergent. No penetration was observed in healthy skin. Disruption of SC by preceding fungal growth caused Sylsens B penetration at pH 7.4, but not at pH 5.2. However, chemically damaged SC allowed Sylsens B to penetrate also at pH 5.2. UVA-1 PDT was applied ex vivo during two fungal growth stages of two T. rubrum strains (CBS 304.60 and a clinical isolate). Both strains could be killed by UVA-1 alone (40 J/cm2). Combined with formulation I (1 and 10 μ m Sylsens B for, respectively, CBS 304.60 and the clinical isolate), only 18 J/cm2 UVA-1 was required for fungal kill. Therefore, PDT with 10 μ m Sylsens B (formulation I) and 18 J/cm2 UVA-1 could be considered as effective and safe. This offers the possibility to perform clinical studies in future.  相似文献   

2.
Owing to the accessibility of skin to light, many applications of photodynamic treatment (PDT) have been developed within dermatology. The recent increase of dermatological antimicrobial PDT investigations is related to the growing problem of bacterial and fungal resistance to antibiotics. This review focuses on the susceptibility of dermatophytic fungi, in particular Trichophyton rubrum, to PDT and shows its potential usefulness in treatment of clinical dermatophytoses. There are no data indicating significant differences in PDT susceptibility between various dermatophytes and it is unlikely that treatment problems of especially T. rubrum with current antimycotics would occur in case of PDT. Red light 5-aminolevulinic acid-mediated PDT is after repeated sessions successful in in vivo treatment of onychomycosis (fungal nail infection) caused by various dermatophytes. Regarding skin dermatophytoses, UVA-1 PDT with cationic porphyrins appears to be safe and efficient. Most effective toward T. rubrum ex vivo is 5,10,15-tris(4-methylpyridinium)-20-phenyl-[21H,23H]-porphine trichloride (Sylsens B) when combined with UVA-1 radiation or red light; this creates the possibility of efficiently treating nail infections and remaining spores in hair follicles. If the promising in vitro and ex vivo results could be transferred to clinical practice, then PDT has a good prospect to become a worthy alternative to established antifungal drugs.  相似文献   

3.
Singlet oxygen (1O2) is the primary oxidant generated in photodynamic therapy (PDT) protocols involving sensitizers resulting in type II reactions. 1O2 can give rise to additional reactive oxygen species (ROS) such as the hydroxyl radical (?OH). The current study was designed to assess 3′‐p‐(aminophenyl) fluorescein (APF) and 3′‐p‐(hydroxyphenyl) fluorescein (HPF) as probes for the detection of 1O2 and ?OH under conditions relevant to PDT. Cell‐free studies indicated that both APF and HPF were converted to fluorescent products following exposure to 1O2 generated by irradiation of a water‐soluble photosensitizing agent (TPPS) and that APF was 35‐fold more sensitive than HPF. Using the 1O2 probe singlet oxygen sensor green (SOSG) we confirmed that 1 mm NaN3 quenched 1O2‐induced APF/HPF fluorescence, while 1% DMSO had no effect. APF and HPF also yielded a fluorescent product upon interacting with ?OH generated from H2O2 via the Fenton reaction in a cell‐free system. DMSO quenched the fluorogenic interaction between APF/HPF and ?OH at doses as low as 0.02%. Although NaN3 was expected to quench ?OH‐induced APF/HPF fluorescence, co‐incubating NaN3 with APF or HPF in the presence of ?OH markedly enhanced fluorescence. Cultured L1210 cells that had been photosensitized with benzoporphyhrin derivative exhibited APF fluorescence immediately following irradiation. Approximately 50% of the cellular fluorescence could be suppressed by inclusion of either DMSO or the iron‐chelator desferroxamine. Combining the latter two agents did not enhance suppression. We conclude that APF can be used to monitor the formation of both 1O2 and ?OH in cells subjected to PDT if studies are performed in the presence and absence of DMSO, respectively. That portion of the fluorescence quenched by DMSO will represent the contribution of ?OH. This procedure could represent a useful means for evaluating formation of both ROS in the context of PDT.  相似文献   

4.
Photodynamic therapy (PDT) can treat superficial, early‐stage disease with minimal damage to underlying tissues and without cumulative dose‐limiting toxicity. Treatment efficacy is affected by disease physiologic properties, but these properties are not routinely measured. We assessed diffuse reflectance spectroscopy (DRS) for the noninvasive, contact measurement of tissue hemoglobin oxygen saturation (StO2) and total hemoglobin concentration ([tHb]) in the premalignant or superficial microinvasive oral lesions of patients treated with 5‐aminolevulinic acid (ALA)‐PDT. Patients were enrolled on a Phase 1 study of ALA‐PDT that evaluated fluences of 50, 100, 150 or 200 J cm?2 delivered at 100 mW cm?2. To test the feasibility of incorporating DRS measurements within the illumination period, studies were performed in patients who received fractionated (two‐part) illumination that included a dark interval of 90–180 s. Using DRS, tissue oxygenation at different depths within the lesion could also be assessed. DRS could be performed concurrently with contact measurements of photosensitizer levels by fluorescence spectroscopy, but a separate noncontact fluorescence spectroscopy system provided continuous assessment of photobleaching during illumination to greater tissue depths. Results establish that the integration of DRS into PDT of early‐stage oral disease is feasible, and motivates further studies to evaluate its predictive and dosimetric value.  相似文献   

5.
The reactions of 5,10,15,20-tetraphenylporpine, 5,10,15-triphenyl-20-(4-hexadecanoxyphenyl)porphine, 5,10,15,20-tetra-(4-butoxyphenyl)porphine, and 2,3,7,8,12,13,17,18-ocaethylporphine with H2PtCl6 in boiling phenol and with PdCl2 in boiling dimethylformamide are studied by spectrophotometry. Due to a strong electronic effect of the substituents, the reactivity of the tetrapyrrole macrocycle during porphyrinate formation changes by more than two orders of magnitude. Platinum(II) 5,10,15-triphenyl-20-(4-hexadecanoxyphenyl)porphinate and palladium(II) 5,10,15,20-tetra-(4-butoxyphenyl)porphinate have been synthesized and identified for the first time.  相似文献   

6.
The reactions of p‐nitrophenyl acetate (PNPA) with a series of monopyridinium oximes, viz. 2‐PAM (2‐hydroxyiminomethyl‐1‐methylpyridinium iodide), 3‐PAM (3‐hydroxyiminomethyl‐1‐methylpyridinium iodide), and 4‐PAM (4‐hydroxyiminomethyl‐1‐methylpyridinium iodide) have been studied in the presence of cationic surfactants of same hydrophobic chain length (C16) within the concentration range of 0.5–6.0 mM at pH 8.0 under the pseudo‐first‐order condition. The observed rate constant (kobs) increases with increasing surfactant concentration culminating into a maximum, and this has been analyzed in detail following the concepts of micellar catalysis. The structure–activity relationship of the investigated oximes has been discussed, and 2‐PAM was found to be the most reactive among all the three investigated oximes for the cleavage of PNPA. Esterolytic decomposition of p‐nitrophenyldiphenyl phosphate with oximate ions (? CH?NO?) was followed in cetyltrimethylammonium bromide micelles at pH 9.0, and 4‐PAM was the most reactive oxime for the micellar hydrolysis of phosphate ester. The apparent acid dissociation constants (pKa) of the investigated oximes have been determined spectrophotometrically. © 2011 Wiley Periodicals, Inc. Int J Chem Kinet 43: 569–578, 2011  相似文献   

7.
Dermatophytes are fungi that can cause infections (known as tinea) of the skin, hair and nails because of their ability to use keratin. Superficial mycoses are probably the most prevalent of infectious diseases worldwide. One of the most distinct limitations of the current therapeutic options is the recurrence of the infection and duration of treatment. The present study shows that Trichophyton rubrum in suspension culture is susceptible to photodynamic treatment (PDT), a completely new application in this area. T. rubrum could be effectively killed with the use of the light-activated porphyrins deuteroporphyrin monomethylester (DP mme) and 5,10,15-tris(4-methylpyridinium)-20-phenyl-[21H,23H]-porphine trichloride (Sylsens B). The photodynamic efficacy was compared with that of some other photosensitizers that are well known in the field of PDT: the porphyrins deuteroporphyrin and hematoporphyrin, the drug Photofrin and several phthalocyanines. It was demonstrated that with the use of broadband white light, the phthalocyanines and Photofrin displayed a fungistatic effect for about 1 week, whereas all the porphyrins caused photodynamic killing of the dermatophyte. Sylsens B was the most effective sensitizer and showed no dark toxicity; therefore, in an appropriate formulation, it could be a promising candidate for the treatment of various forms of tinea. For Sylsens B and DP mme, which displayed the best results, a concentration-dependent uptake by T. rubrum was established.  相似文献   

8.
The contributions of the amino and imino resonance forms to the ground‐state structures of 2‐amino‐4‐methylpyridinium nitrate, C6H9N2+·NO3, and the previously reported 2‐amino‐5‐methylpyridinium nitrate [Yan, Fan, Bi, Zuo & Zhang (2012). Acta Cryst. E 68 , o2084], were studied using a combination of IR spectroscopy, X‐ray crystallography and density functional theory (DFT). The results show that the structures of 2‐amino‐4‐methylpyridine and 2‐amino‐5‐methylpyridine obtained upon protonation are best described as existing largely in the imino resonance forms.  相似文献   

9.
Three novel porphyrins, including two Schiff‐bases porphyrins, 5,10,15‐triphenyl‐20‐[4‐(2‐(4‐formyl)phenoxy)ethoxy]phenyl porphyrin ( H2Pp ( 1 )), 5,10,15‐triphenyl‐20‐[4‐(2‐(4‐hydroxyimino)phenoxy)ethoxy]phenyl porphyrin ( H2Pp ( 2 )) and 5,10,15‐triphenyl‐20‐[4‐(2‐(4‐m‐hydroxyanilinodeneformyl)phenoxy)ethoxy]phenyl porphyrin ( H2Pp ( 3 )), as well as three metalloporphyrins ( CuPp ( 1a ), ZnPp ( 1b ), and CoPp ( 1c )) of porphyrin H2Pp ( 1 ) were synthesized. Their molecular structures were characterized by 1H‐NMR, MS, UV/VIS, and FT‐IR spectra. Furthermore, they were evaluated by their cytotoxicities against human epidermal squamous cell carcinoma cell (A431) and normal human horn cells (HaCaT) in vitro with MTT assay. Interestingly, these porphyrins and metalloporphyrins, which had a negligible cytotoxicity to HaCaT cells, showed highly cytotoxicity against A431 cells with IC50 values in the range of 6.6–9.8 μM , and metalloporphyrins exhibited higher cytotoxicity than that of metal‐free porphyrins.  相似文献   

10.
2‐Nitro‐5,10,15‐tri(4‐tert‐butylphenyl)subporphyrin 2 was prepared by the nitration of 5,10,15‐tri(4‐tert‐ butylphenyl)subporphyrin 1a with five equivalents of Cu(NO3)2 ? 5 H2O in a mixed EtOAc/Ac2O solution and was reduced into 2‐amino‐5,10,15‐tri(4‐tert‐butylphenyl)subporphyrin 3 . Bromination of 5,10,15‐triphenylsubporphyrin 1b with 1.5 equivalents of N‐bromosuccinimide (NBS) gave 2‐bromo‐5,10,15‐triphenylsubporphyrin, which was converted into various 2‐arylamino‐5,10,15‐triphenylsubporphyrins ( 4a , 4b , 4c , 4d ) and 2‐benzamido‐5,10,15‐triphenylsubporphyrin 5 through Pd‐catalyzed cross‐coupling reactions. These molecules constitute the first examples of mono‐β‐substituted subporphyrins. These subporphyrins exhibit significantly perturbed optical and electrochemical properties, which reflect a large influence of the peripherally attached substituents on the electronic networks of subporphyrins.  相似文献   

11.
When irradiated, fullerene efficiently generates reactive oxygen species (ROS) and is an attractive photosensitizer for photodynamic therapy (PDT). Ideally, photosensitizers for PDT should be water-soluble and tumor-specific. Because cancer cells endocytose glucose more effectively than normal cells, the characteristics of fullerene as a photosensitizer were improved by combining it with glucose. The cytotoxicity of PDT was studied in several cancer cell lines cultured with C60-(Glc)1 (d -glucose residue pendant fullerene) and C60-(6Glc)1 (a maltohexaose residue pendant fullerene) subsequently irradiated with UVA1. PDT alone induced significant cytotoxicity. In contrast, PDT with the glycoconjugated fullerene exhibited no significant cytotoxicity against normal fibroblasts, indicating that PDT with these compounds targeted cancer cells. To investigate whether the effects of PDT with glycoconjugated fullerene were because of the generation of singlet oxygen (1O2), NaN3 was added to cancer cells during irradiation. NaN3 extensively blocked PDT-induced apoptosis, suggesting that PDT-induced cell death was a result of the generation of 1O2. Finally, to investigate the effect of PDT in vivo, melanoma-bearing mice were injected intratumorally with C60-(Glc)1 and irradiated with UVA1. PDT with C60-(Glc)1 suppressed tumor growth. These findings indicate that PDT with glycoconjugated fullerene exhibits tumor-specific cytotoxicity both in vivo and in vitro via the induction of 1O2.  相似文献   

12.
The structures of seven A2Cu4X10 compounds containing quasi‐planar oligomers are reported: bis(1,2,4‐trimethylpyridinium) hexa‐μ‐chlorido‐tetrachloridotetracuprate(II), (C8H12N)2[Cu4Cl10], (I), and the hexa‐μ‐bromido‐tetrabromidotetracuprate(II) salts of 1,2,4‐trimethylpyridinium, (C8H12N)2[Cu4Br10], (II), 3,4‐dimethylpyridinium, (C7H10N)2[Cu4Br10], (III), 2,3‐dimethylpyridinium, (C7H10N)2[Cu4Br10], (IV), 1‐methylpyridinium, (C6H8N)2[Cu4Br10], (V), trimethylphenylammonium, (C9H14N)2[Cu4Br10], (VI), and 2,4‐dimethylpyridinium, (C7H10N)2[Cu4Br10], (VII). The first four are isomorphous and contain stacks of tetracopper oligomers aggregated through semicoordinate Cu...X bond formation in a 4(,) stacking pattern. The 1‐methylpyridinium salt also contains oligomers stacked in a 4(,) pattern, but is isomorphous with the known chloride analog instead. The trimethylphenylammonium salt contains stacks of oligomers arranged in a 4(,) stacking pattern similar to the tetramethylphosphonium analog. These six structures feature inversion‐related organic cation pairs and hybrid oligomer/organic cation layers derived from the parent CuX2 structure. The 2,4‐dimethylpyridinium salt is isomorphous with the known (2‐amino‐4‐methylpyridinium)2Cu4Cl10 structure, in which isolated stacks of organic cations and of oligomers in a 4(,) pattern are found. In bis(3‐chloro‐1‐methylpyridinium) octa‐μ‐bromido‐tetrabromidopentacuprate(II), (C6H7ClN)[Cu5Br12], (VIII), containing the first reported fully halogenated quasi‐planar pentacopper oligomer, the oligomers stack in a 5(,) stacking pattern as the highest nuclearity [CunX2n+2]2− oligomer compound known with isolated stacking. Bis(2‐chloro‐1‐methylpyridinium) dodeca‐μ‐bromido‐tetrabromidoheptacuprate(II), (C6H7ClN)2[Cu7Br16], (IX), contains the second heptacopper oligomer reported and consists of layers of interleaved oligomer stacks with a 7[(,)][(−,−)] pattern isomorphous with that of the known 1,2‐dimethylpyridinium analog. All the oligomers reported here are inversion symmetric.  相似文献   

13.
Os(II) hydridocarbonyl complexes of coumarinyl azoimidazoles, [Osh(CO)(PPh3)2(CZ‐4R‐R′)]0/+ ( 3 , 4 ) (CZ‐R‐H = 2‐(coumarinyl‐6‐azo)‐4‐substituted imidazole or 1‐alkyl‐2‐(coumarinyl‐6‐azo)‐4‐substituted imidazole), were characterized from spectroscopic data and the single‐crystal X‐ray data for one of the complexes, [Osh(CO)(PPh3)2(CZ‐4‐Ph)] ( 3c ) (CZ‐4‐Ph = 2‐(coumarinyl‐6‐azo)‐4‐phenylimidazolate), confirmed the structure. The complexes show higher emission (quantum yield ? = 0.0163–0.16) and longer lifetime (τ = 1.4–10.3 ns) than free ligands (? = 0.0012–0.0185 and τ = 0.685–1.306 ns). Cyclic voltammetry shows quasi‐reversible metal oxidation at 0.67–0.94 V for [Os(III)/Os(II)] and 1.21–1.36 V for [Os(IV)/Os(III)] and subsequent azo reductions (?0.68 to ?0.95 V for [? N?N? ]/[? N N? ]? and irreversible < ?1.2 V for [? N N? ]?/[? N? N? ]2?) of the chelated coumarinyl azoimidazole. The complexes are photostable and show better photovoltaic power conversion efficiency than free ligands. Also, the complexes were used as catalysts for the oxidation of primary/secondary alcohols to aldehydes/ketones using oxidizing agents like N‐methylmorpholine N‐oxide, t‐BuOOH and H2O2. Density functional theory computation was carried out from the optimized structures and the data obtained were used to interpret the electronic and photovoltaic properties. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

14.
Crystals of bis(2‐ethyl‐3‐hydroxy‐6‐methylpyridinium) succinate–succinic acid (1/1), C8H12NO+·0.5C4H4O42−·0.5C4H6O4, (I), and 2‐ethyl‐3‐hydroxy‐6‐methylpyridinium hydrogen succinate, C8H12NO+·C4H5O4, (II), were obtained by reaction of 2‐ethyl‐6‐methylpyridin‐3‐ol with succinic acid. The succinate anion and succinic acid molecule in (I) are located about centres of inversion. Intermolecular O—H...O, N—H...O and C—H...O hydrogen bonds are responsible for the formation of a three‐dimensional network in the crystal structure of (I) and a two‐dimensional network in the crystal structure of (II). Both structures are additionally stabilized by π–π interactions between symmetry‐related pyridine rings, forming a rod‐like cationic arrangement for (I) and cationic dimers for (II).  相似文献   

15.
A technique is introduced that monitors the depletion of intracellular ground state oxygen concentration ([3O2]) during photodynamic therapy of Mat‐LyLu cell monolayers and cell suspensions. The photosensitizer Pd(II) meso‐tetra(4‐carboxyphenyl)porphine (PdT790) is used to manipulate and indicate intracellular [3O2] in both of the in vitro models. The Stern–Volmer relationship for PdT790 phosphorescence was characterized in suspensions by flowing nitrogen over the suspension while short pulses of 405 nm light were used to excite the sensitizer. The bleaching of sensitizer and the oxygen consumption rate were also measured during continuous exposure of the cell suspension to the 405 nm laser. Photodynamic therapy (PDT) was conducted in both cell suspensions and in cell monolayers under different treatment conditions while the phosphorescence signal was acquired. The intracellular [3O2] during PDT was calculated by using the measured Stern–Volmer relationship and correcting for sensitizer photobleaching. In addition, the amount of oxygen that was consumed during the treatments was calculated. It was found that even at large oxygen consumption rates, cells remain well oxygenated during PDT of cell suspensions. For monolayer treatments, it was found that intracellular [3O2] is rapidly depleted over the course of PDT.  相似文献   

16.
A new class of near‐infrared (NIR)‐absorptive (>900 nm) photosensitizer based on a phenothiazinium scaffold is reported. The stable solid compound, o‐DAP, the oxidative form of 3,7‐bis(4‐methylaminophenyl)‐10H‐phenothiazine, can generate reactive oxygen species (ROS, singlet oxygen and superoxide) under appropriate irradiation conditions. After biologically evaluating the intracellular uptake, localization, and phototoxicity of this compound, it was concluded that o‐DAP is photostable and a potential selective photodynamic therapy (PDT) agent under either NIR or white light irradiation because its photodamage is more efficient in cancer cells than in normal cells and is without significant dark toxicity. This is very rare for photosensitizers in PDT applications.  相似文献   

17.
Four complexes containing the [UO2(oda)2]2− anion (oda is oxydiacetate) are reported, namely dipyridinium dioxidobis(oxydiacetato)uranate(VI), (C5H6N)2[U(C4H4O5)2O2], (I), bis(2‐methylpyridinium) dioxidobis(oxydiacetato)uranate(VI), (C8H8N)2[U(C4H4O5)2O2], (II), bis(3‐methylpyridinium) dioxidobis(oxydiacetato)uranate(VI), (C8H8N)2[U(C4H4O5)2O2], (III), and bis(4‐methylpyridinium) dioxidobis(oxydiacetato)uranate(VI), (C8H8N)2[U(C4H4O5)2O2], (IV). The anions are achiral and are located on a mirror plane in (I) and on inversion centres in (II)–(IV). The four complexes are assembled into three‐dimensional structures via N—H...O and C—H...O interactions. Compounds (III) and (IV) are isomorphous; the [UO2(oda)2]2− anions form a porous matrix which is nearly identical in the two structures, and the cations are located in channels formed in this matrix. Compounds (I) and (II) are very different from (III) and (IV): (I) forms a layered structure, while (II) forms ribbons.  相似文献   

18.
The photophysical parameters for the photosensitizer Pd(II) meso‐Tetra(4‐carboxyphenyl) porphine (PdT790) acquired in a previous study were incorporated into the PDT oxygen diffusion models for cell suspensions and cell monolayers. The time‐dependent phosphorescence signals generated by the diffusion models are shown to match signals previously measured by M.A.W. and M.S.P. when reasonable physical and photophysical parameters are used. Simulations were performed to investigate the effects of metabolic and photodynamic oxygen consumption rates on the PDT dose in each of the treatment geometries. It was found that in cell suspensions of <1 million cells per mL, PDT should not be inhibited by hypoxia if the photodynamic consumption rate is <1 mm  s?1. For cell monolayers the optimal photodynamic oxygen consumption rate was found to depend on the metabolic rate of oxygen consumption. If cells remained well oxygenated in the absence of PDT, then maximum PDT dose was delivered with the lowest practical photodynamic oxygen consumption rate. Simulations of PDT treatments for multicell tumor spheroids showed that large anoxic cores develop within the spheroids and, as a consequence, less PDT dose is delivered in comparison with similar treatments in cell suspensions and cell monolayers.  相似文献   

19.
Four new coordination compounds, [Cd(L1)2]n ( 1 ), [Mn(L1)2]n ( 2 ), [Zn(L1)(NA)] ( 3 ), and [Pb(L1)2(H2O)] ( 4 ) were obtained on the basis of the in‐situ ligand reactions of quinoline‐2‐carbonitrile (QCN) and NaN3 under solvothermal conditions. 1 and 2 are 1D isostructural chains, where the central metal atoms are six‐coordinate by six nitrogen atoms in a distorted octahedron. The cycloaddition reaction of QCN and NaN3 in the presence of hydrated ZnCl2 occur with the aid of the ancillary ligand nicotinic acid (NA), where NA not only provides an acidic environment but also serves as an ancillary ligand. The extended structure of 3 is a 1D ladder‐like polymer. Compound 4 is a mononuclear Pb2+ compound. All the compounds were structurally characterized by X‐ray crystallography and their luminescent properties were investigated in detail.  相似文献   

20.
This research studied the effectiveness of the photoactive compound methylene blue (MB) activated with red LED light (576–672 nm) compared to that of caspofungin (CAS) on 1 Candida albicans and 3 Candida parapsilosis strains. Results were evaluated in terms of SMIC50 for CAS or in PDI (photodynamic inactivation)‐SMIC50 for MB (minimal inhibitory concentration inhibiting sessile biofilm to 50% in comparison to the control without CAS or after irradiation in comparison to the control without MB). While all strains were susceptible to CAS in planktonic form, the SMIC50 was determined to be >16 μg mL?1 when CAS was added to a 24 h biofilm. However, PDI‐MIC50s (1.67 mW cm?2, fluence 15 J cm?2) were 0.0075–0.03 mmol L?1. For biofilm, PDI‐SMIC50s were in the range from 0.7 to 1.35 mmol L?1. MB concentration of 1 mmol L?1 prevented a biofilm being formed ex vivo on mouse tongues after irradiation regardless of the application time, in contrast to CAS, which was only effective at a concentration of 16 μg mL?1 when it was added at the beginning of biofilm formation. PDI seems to be a promising method for the prevention of microbial biofilms that do not respond significantly to conventional drugs.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号