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.
The use of ruthenium complexes in antitumor therapy was launched two decades ago. In view of their low toxicity and good selectivity for solid tumor metastasis, ruthenium complexes have great potential as alternative drugs to cisplatin in cancer chemotherapy. A series of monochloro ruthenium complexes, Ru(terpy) (NN)Cl+ (NN, bidentate nitrogen ligand), containing different electron‐donating groups were prepared. The reactivity towards the formation of Ru‐DNA adduct were revealed by gel mobility shift assay. Their DNA binding sites of Ru(terpy)(tmephen)Cl+ were located predominantly at the purine residues i.e., guanine and adenine, by terminating DNA elongation in vitro using PCR and primer extension techniques. Surprisingly, the ability of Ru(terpy)(tmephen)Cl+ to inhibit cell growth was found to be approximately two times better than that of a known cross‐linking agent, Ru(bpy)2Cl2. Therefore, the increase in liability of the chloro ligand was demonstrated to improve the reactivity of these ruthenium complexes towards the covalent bond formation in Ru‐DNA adducts and result also in a significant inhibition of cell growth. Based on our results, these ruthenium complexes modified with electron‐rich groups provide new consideration in the tune of ruthenium‐based drugs in cancer chemotherapy.  相似文献   

2.
A series of novel ruthenium(II)–cymene complexes ( 1 – 8 ) containing substituted pyridyl–thiazole ligands, [Ru(η6p‐cymene)(L)Cl]Cl (L = N,N‐chelating derivatives), have been synthesized and characterized using elemental analysis, infrared, 1H NMR and 13C NMR spectroscopies and mass spectrometry. All these complexes not only display marked cytotoxicity in vitro against three different human cancer cell lines (HeLa, A549 and MDA‐MB‐231), but also exhibit promising anti‐metastatic activity at sub‐cytotoxic concentrations. Cell cycle analysis shows that the ruthenium(II) complex‐induced growth inhibition was mainly caused by S‐phase cell cycle arrest. Further protein level analysis suggests that compound 5 may exert antitumor activity via a p53‐independent mechanism.  相似文献   

3.
Ruthenium polypyridyl complexes are widely used as light harvesters in dye‐sensitized solar cells. Since one of the potential applications of single‐wall carbon nanotubes (SWCNTs) and their derived materials is their use as active components in organic and hybrid solar cells, the study of the photochemistry of SWCNTs with tethered ruthenium polypyridyl complexes is important. A water‐soluble ruthenium tris(bipyridyl) complex linked through peptidic bonds to SWCNTs (Ru‐SWCNTs) was prepared by radical addition of thiol‐terminated SWCNT to a terminal C?C double bond of a bipyridyl ligand of the ruthenium tris(bipyridyl) complex. The resulting macromolecular Ru‐SWCNT (≈500 nm, 15.6 % ruthenium complex content) was water‐soluble and was characterized by using TEM, thermogravimetric analysis, chemical analysis, and optical spectroscopy. The emission of Ru‐SWCNT is 1.6 times weaker than that of a mixture of [Ru(bpy)3]2+ and SWCNT of similar concentration. Time‐resolved absorption optical spectroscopy allows the detection of the [Ru(bpy)3]2+‐excited triplet and [Ru(bpy)3]+. The laser flash studies reveal that Ru‐SWCNT exhibits an unprecedented two‐photon process that is enabled by the semiconducting properties of the SWCNT. Thus, the effect of the excitation wavelength and laser power on the transient spectra indicate that upon excitation of two [Ru(bpy)3]2+ complexes of Ru‐SWCNT, a disproportionation process occurs leading to delayed formation of [Ru(bpy)3]+ and the performance of the SWCNT as a semiconductor. This two‐photon delayed [Ru(bpy)3]+ generation is not observed in the photolysis of [Ru(bpy)3]3+; SWCNT acts as an electron wire or electron relay in the disproportionation of two [Ru(bpy)3]2+ triplets in a process that illustrates that the SWCNT plays a key role in the process. We propose a mechanism for this two‐photon disproportionation compatible with i) the need for high laser flux, ii) the long lifetime of the [Ru(bpy)3]2+ triplets, iii) the semiconducting properties of the SWNT, and iv) the energy of the HOMO/LUMO levels involved.  相似文献   

4.
Light‐activated ruthenium polypyridyl anticancer prodrugs often suffer from poor water solubility, poor selectivity, and/or ill‐defined intracellular targets. Coordination of the d ‐ or l ‐glucose thioether ligand 3 (2‐(2‐(2‐(methylthio)ethoxy)ethoxy)ethyl‐β‐glucopyranoside) to the highly lipophilic ruthenium complex [Ru(tpy)(dppn)(H2O)]2+ ([ 1 ]2+; dppn=benzo[i]dipyrido‐[3,2‐a:2′,3′‐c]phenazine, tpy=2,2′:6′,2′′‐terpyridine) solved all these problems at once. The two enantiomers of [Ru(tpy)(dppn)( 3) ][PF6]2, [d ‐ 2 ][PF6]2 and [l ‐ 2 ][PF6]2, were soluble in water, which allowed the influence of the chirality of the glucose moiety on uptake, toxicity, and intracellular localization of the prodrug to be probed without changing any other physicochemical properties. Both compounds showed mild, but different, cytotoxicity in A549 (human lung carcinoma) and MCF‐7 (human breast adenocarcinoma) cancer cells in the dark, whereas following low doses of visible light irradiation (3.1 J cm?2 at λ = 454 nm), a similar, but high cytotoxicity (EC50 < 1 μm ), was observed. Irrespective of the chirality, both slightly emissive Ru complexes were found in the mitochondria, and two modes of action may contribute to light‐induced cell death: 1) the glucose thioether ligand is photosubstituted by water, thus [ 1 ]2+, which interacts with DNA at an exceptionally high 400:1 base pair/Ru ratio, is released; 2) both [ 1 ]2+ and [ 2 ]2+ produce massive amounts of singlet oxygen, which leads to very efficient photodynamic DNA cleavage.  相似文献   

5.
An organometallic salt composed of a new cationic p‐cymene ruthenium chloro complex containing a chelating benzaldehyde semicarbazone ligand and of the known anionic p‐cymene ruthenium trichloro complex, [(η6p‐cymene)Ru(bzsc)Cl]+[(η6p‐cymene)RuCl3] ( 1 ) (bzsc = benzaldehyde semicarbazone) was synthesized and further characterized by IR, 1H NMR, and UV/Vis spectroscopy HR‐ESI mass spectrometry, and elemental analysis. The single‐crystal structure of 1 was also determined. The in vitro anticancer activities of the complex was evaluated against three human cancer cell lines (SGC‐7901, BEL‐7404 and CNE‐1), and the IC50 values were 20.7, 71.1 and 42.6 μM, respectively.  相似文献   

6.
Half‐sandwich ruthenium, rhodium and iridium complexes ( 1 – 12 ) were synthesized with aldoxime ( L1 ), ketoxime ( L2 ) and amidoxime ( L3 ) ligands. Ligands have the general formula [PyC(R)NOH], where R = H ( L1 ), R = CH3 ( L2 ) and R = NH2 ( L3 ). Reaction of [{(arene)MCl2}2] (arene = p ‐cymene, benzene, Cp*; M = Ru, Rh, Ir) with ligands L1 – L3 in 1:2 metal precursor‐to‐ligand ratio yielded complexes such as [{(arene)MLκ2(N∩N)Cl}]PF6. All the ligands act as bidentate chelating nitrogen donors in κ2(N∩N) fashion while forming complexes. In vitro anti‐tumour activity of complexes 2 and 10 against HT‐29 (human colorectal cancer), BE (human colorectal cancer) and MIA PaCa‐2 (human pancreatic cancer) cell lines and non‐cancer cell line ARPE‐19 (human retinal epithelial cells) revealed a comparable activity although complex 2 demonstrated greater selectivity for MIA PaCa‐2 cells than cisplatin. Further studies demonstrated that complexes 3 , 6 , 9 and 12 induced significant apoptosis in Dalton's ascites lymphoma (DL) cells. In vivo anti‐tumour activity of complex 2 on DL‐bearing mice revealed a statistically significant anti‐tumour activity (P  = 0.0052). Complexes 1 – 12 exhibit HOMO–LUMO energy gaps from 3.31 to 3.68 eV. Time‐dependent density functional theory calculations explain the nature of electronic transitions and were in good agreement with experiments.  相似文献   

7.
A novel fluorescent porous thin film based on the precipitation of the [(2,2′‐bipyridyl)(4‐(2‐pyrrol‐1‐ylethyl)‐4′‐methyl‐2,2′‐bipyridyl)2]ruthenium(II) (BF4)2 complex (pyr‐Ru) was fabricated by easily spreading 2 µL of pyr‐Ru (1 mM in acetonitrile solution) onto the surface of a platinum electrode and drying it in ambient conditions. The morphology of the resulting pyr‐Ru thin film was characterized by scanning electron microscopy (SEM) and fluorescence microscopy. The coating exhibits fluorescent properties of the ruthenium complex and a porous structure with pore diameters of micrometers. The solid‐state electrochemistry and electrochemiluminescence behaviors of the porous pyr‐Ru thin film were investigated in aqueous solution by cyclic voltammetry and step potential.  相似文献   

8.
Selenadiazole derivatives (SeDs) have been found to show promise in chemo‐/radiotherapy applications by activating various downstream signaling pathways. However, the functional role of SeDs on angiogenesis, which is pivotal for tumor progression and metastasis, has not yet been elucidated. In the present study, we have examined the antiangiogenic activities of SeDs and elucidated their underlying mechanisms. The results showed that the as‐synthesized SeDs not only enhanced their anticancer activities against several human cancer cells but also showed more potent inhibition on human umbilical vein endothelial cells (HUVECs). The in vitro results suggested that SeDs, especially 1 a , dose‐dependently inhibited the vascular endothelial growth factor (VEGF)‐induced cell migration, invasion, and capillary‐like structure formation of HUVECs. Compound 1 a also significantly suppressed VEGF‐induced angiogenesis in a Matrigel plug assay as part of a C57/BL6 mice assay by means of down regulation of VEGF. Furthermore, we found that 1 a significantly inhibited MCF‐7 human breast tumor growth in nude mice without severe systematic cytotoxicity. Compound 1 a was more effective in inhibiting cell proliferation and induced a much more pronounced apoptosis effect in endothelial cells than MCF‐7 cells, which implies that endothelial cells might be the primary target of 1 a . Further mechanistic studies on tumor growth inhibition effects and neovessel formation suppression demonstrated that 1 a inhibited cell viability of MCF‐7 and HUVECs by induction of cell apoptosis, accompanied by poly(adenosine diphosphate ribose)polymerase (PARP) cleavage and caspase activation. Additionally, the 1 a ‐induced antiangiogenesis effect was achieved by abolishing the VEGF‐VEGFR2‐ERK/AKT (ERK=extracellular signal–regulated kinases; AKT=protein kinease B) signal axis and enhanced the apoptosis effect by triggering reactive oxygen species (ROS)‐mediated DNA damage. Taken together, these results clearly demonstrate the antiangiogenic potency of SeDs and the underlying molecular mechanisms.  相似文献   

9.
Histone deacetylases inhibitors (HDACis) have gained much attention as a new class of anticancer agents in recent years. Herein, we report a series of fluorescent ruthenium(II) complexes containing N1‐hydroxy‐N8‐(1,10‐phenanthrolin‐5‐yl)octanediamide ( L ), a suberoylanilide hydroxamic acid (SAHA) derivative, as a ligand. As expected, these complexes show interesting chemiphysical properties, including relatively high quantum yields, large Stokes shifts, and long emission lifetimes. The in vitro inhibitory effect of the most effective drug, [Ru(DIP)2 L ](PF6)2 ( 3 ; DIP: 4,7‐diphenyl‐1,10‐phenanthroline), on histone deacetylases (HDACs) is approximately equivalent in activity to that of SAHA, and treatment with complex 3 results in increased levels of the acetylated histone H3. Complex 3 is highly active against a panel of human cancer cell lines, whereas it shows relatively much lower toxicity to normal cells. Further mechanism studies show that complex 3 can elicit cell cycle arrest and induce apoptosis through mitochondria‐related pathways and the production of reactive oxygen species. These data suggest that these fluorescent ruthenium(II)–HDACi conjugates may represent a promising class of anticancer agents for potential dual imaging and therapeutic applications targeting HDACs.  相似文献   

10.
Coordination of a redox‐active pyridine aminophenol ligand to RuII followed by aerobic oxidation generates two diamagnetic RuIII species [ 1 a (cis) and 1 b (trans)] with ligand‐centered radicals. The reaction of 1 a / 1 b with excess NaN3 under inert atmosphere resulted in the formation of a rare bis(nitrido)‐bridged trinuclear ruthenium complex with two nonlinear asymmetrical Ru‐N‐Ru fragments. The spontaneous reduction of the ligand centered radical in the parent 1 a / 1 b supports the oxidation of a nitride (N3?) to half an equivalent of N2. The trinuclear omplex is reactive toward TEMPO‐H, tin hydrides, thiols, and dihydrogen.  相似文献   

11.
The title complex, di‐μ‐chloro‐bis­[chloro­(η6p‐cymene)ruthenium(II)]–9H‐carbazole (1/2), [Ru2Cl4(C10H14)2]·2C12H9N, is composed of one [RuCl26p‐cymene)]2 and two 9H‐carbazole mol­ecules. There are one‐half of a dinuclear complex and one 9H‐carbazole mol­ecule per asymmetric unit. In the dinuclear complex, each of the two crystallographically equivalent Ru atoms is in a pseudo‐tetra­hedral environment, coordinated by a terminal Cl atom, two bridging Cl atoms and the aromatic hydro­carbon, which is linked in a η6 manner; the Ru⋯Ru separation is 3.688 (3) Å. The title complex has a crystallographic centre of symmetry located at the mid‐point of the Ru⋯Ru line. Inter­molecular N—H⋯Cl and π–π stacking inter­actions are observed. These inter­actions form a four‐pointed star‐shaped ring and one‐dimensional linear chains of edge‐fused rings running parallel to the [100] direction, which stabilize the crystal packing.  相似文献   

12.
《化学:亚洲杂志》2018,13(19):2923-2933
A family of novel imine‐N‐heterocyclic carbene ruthenium(II) complexes of the general formula [(η6p‐cymene)Ru(C^N)Cl]PF6 (where C^N is an imine‐N‐heterocyclic carbene chelating ligand with varying substituents) have been prepared and characterized. In this imine‐N‐heterocyclic carbene chelating ligand framework, there are three potential sites that can be modified, which distinguishes this class of ligand and provides a body of flexibilities and opportunities to tune the cytotoxicity of these ruthenium(II) complexes. The influence of substituent effects of three tunable domains on the anticancer activity and catalytic ability in converting coenzyme NADH to NAD+ is investigated. This family of complexes displays an exceedingly distinct anticancer activity against A549 cancer cells, despite their close structural similarity. Complex 9 shows the highest anticancer activity in this series against A549 cancer cells (IC50=14.36 μm ), with an approximately 1.5‐fold better activity than the clinical platinum drug cisplatin (IC50=21.30 μm ) in A549 cancer cells. Mechanistic studies reveal that complex 9 mediates cell death mainly through cell stress, including cell cycle arrest, inducing apoptosis, increasing intracellular reactive oxygen species (ROS) levels, and depolarization of the mitochondrial membrane potential (MMP). Furthermore, lysosomal damage is also detected by confocal microscopy.  相似文献   

13.
A new bridging ligand, 2,3‐di(2‐pyridyl)‐5‐phenylpyrazine (dpppzH), has been synthesized. This ligand was designed so that it could bind two metals through a NN‐CNN‐type coordination mode. The reaction of dpppzH with cis‐[(bpy)2RuCl2] (bpy=2,2′‐bipyridine) affords monoruthenium complex [(bpy)2Ru(dpppzH)]2+ ( 12+ ) in 64 % yield, in which dpppzH behaves as a NN bidentate ligand. The asymmetric biruthenium complex [(bpy)2Ru(dpppz)Ru(Mebip)]3+ ( 23+ ) was prepared from complex 12+ and [(Mebip)RuCl3] (Mebip=bis(N‐methylbenzimidazolyl)pyridine), in which one hydrogen atom on the phenyl ring of dpppzH is lost and the bridging ligand binds to the second ruthenium atom in a CNN tridentate fashion. In addition, the RuPt heterobimetallic complex [(bpy)2Ru(dpppz)Pt(C?CPh)]2+ ( 42+ ) has been prepared from complex 12+ , in which the bridging ligand binds to the platinum atom through a CNN binding mode. The electronic properties of these complexes have been probed by using electrochemical and spectroscopic techniques and studied by theoretical calculations. Complex 12+ is emissive at room temperature, with an emission λmax=695 nm. No emission was detected for complex 23+ at room temperature in MeCN, whereas complex 42+ displayed an emission at about 750 nm. The emission properties of these complexes are compared to those of previously reported Ru and RuPt bimetallic complexes with a related ligand, 2,3‐di(2‐pyridyl)‐5,6‐diphenylpyrazine.  相似文献   

14.
Supported ruthenium hydroxide catalysts (Ru(OH)x/support) were prepared with three different TiO2 supports (anatase TiO2 (TiO2(A), BET surface area: 316 m2 g?1), anatase TiO2 (TiO2(B), 73 m2 g?1), and rutile TiO2 (TiO2(C), 3.2 m2 g?1)), as well as an Al2O3 support (160 m2 g?1). Characterizations with X‐ray diffraction (XRD), X‐ray photoelectron spectroscopy (XPS), electron spin resonance (ESR), and X‐ray absorption fine structure (XAFS) showed the presence of monomeric ruthenium(III) hydroxide and polymeric ruthenium(III) hydroxide species. Judging from the coordination numbers of the nearest‐neighbor Ru atoms and the intensities of the ESR signals, the amount of monomeric hydroxide species increased in the order of Ru(OH)x<Ru(OH)x/TiO2(C)<Ru(OH)x/Al2O3<Ru(OH)x/TiO2(B)<Ru(OH)x/TiO2(A). These supported ruthenium hydroxide catalysts, especially Ru(OH)x/TiO2(A), showed high catalytic activities and selectivities for liquid‐phase hydrogen‐transfer reactions, such as racemization of chiral secondary alcohols and the reduction of carbonyl compounds and allylic alcohols. The catalytic activities of Ru(OH)x/TiO2(A) for these hydrogen‐transfer reactions were at least one order of magnitude higher than those of previously reported heterogeneous catalysts, such as Ru(OH)x/Al2O3. These catalyses were truly heterogeneous, and the catalysts recovered after the reactions could be reused several times without loss of catalytic performance. The reaction rates monotonically increased with an increase in the amount of monomeric ruthenium hydroxide species, which suggests that the monomeric species are effective for these hydrogen‐transfer reactions.  相似文献   

15.
An essentially molecular ruthenium–benzene complex anchored at the aluminum sites of dealuminated zeolite Y was formed by treating a zeolite‐supported mononuclear ruthenium complex, [Ru(acac)(η2‐C2H4)2]+ (acac=acetylacetonate, C5H7O2?), with 13C6H6 at 413 K. IR, 13C NMR, and extended X‐ray absorption fine structure (EXAFS) spectra of the sample reveal the replacement of two ethene ligands and one acac ligand in the original complex with one 13C6H6 ligand and the formation of adsorbed protonated acac (Hacac). The EXAFS results indicate that the supported [Ru(η6‐C6H6)]2+ incorporates an oxygen atom of the support to balance the charge, being bonded to the zeolite through three Ru? O bonds. The supported ruthenium–benzene complex is analogous to complexes with polyoxometalate ligands, consistent with the high structural uniformity of the zeolite‐supported species, which led to good agreement between the spectra and calculations at the density functional theory level. The calculations show that the interaction of the zeolite with the Hacac formed on treatment of the original complex with 13C6H6 drives the reaction to form the ruthenium–benzene complex.  相似文献   

16.
A novel polypyridine ligand, dipyrido[3,2‐a:2′,3′‐c]phenazine‐11‐carboxylic acid methyl ester (=dppz‐11‐CO2Me), and its ruthenium(II) complex, [Ru(bpy)2(dppz‐11‐CO2Me)]2+ ( 1 ), were synthesized and characterized. The binding properties of this complex to calf‐thymus DNA (CT‐DNA) were investigated by different spectrophotometric methods and viscosity measurements. The results suggest that the complex binds to DNA in an intercalative mode and serves as a molecular ‘light switch’ for DNA. When irradiated at 365 nm, the complex 1 promoted the photocleavage of plasmid pBR‐322 DNA.  相似文献   

17.
Palladium(II) complexes are generally reactive toward substitution/reduction, and their biological applications are seldom explored. A new series of palladium(II) N‐heterocyclic carbene (NHC) complexes that are stable in the presence of biological thiols are reported. A representative complex, [Pd(C^N^N)(N,N′‐nBu2NHC)](CF3SO3) ( Pd1 d , HC^N^N=6‐phenyl‐2,2′‐bipyridine, N,N′‐nBu2NHC=N,N′‐di‐n‐butylimidazolylidene), displays potent killing activity toward cancer cell lines (IC50=0.09–0.5 μm ) but is less cytotoxic toward a normal human fibroblast cell line (CCD‐19Lu, IC50=11.8 μm ). In vivo anticancer studies revealed that Pd1 d significantly inhibited tumor growth in a nude mice model. Proteomics data and in vitro biochemical assays reveal that Pd1 d exerts anticancer effects, including inhibition of an epidermal growth factor receptor pathway, induction of mitochondrial dysfunction, and antiangiogenic activity to endothelial cells.  相似文献   

18.
Acrolein, methacrolein, methyl vinyl ketone, ethyl vinyl ketone, 3‐methyl‐3‐en‐2‐one, and divinyl ketone were coordinated to a cationic cyclopentadienyl ruthenium(II) Lewis acid incorporating the electron‐poor bidentate BIPHOP–F ligand. Analysis by NOESY and ROESY NMR techniques allowed the determination of conformations of enals and enones present in solution in CD2Cl2. The results were compared to solid‐state structures and to the facial selectivities of catalytic asymmetric Diels–Alder reactions with cyclopentadiene. X‐Ray structures of four Ru‐enal and Ru‐enone complexes show the α,β‐unsaturated C=O compounds to adopt an anti‐s‐trans conformation. In solution, enals assume both anti‐s‐trans and anti‐s‐cis conformations. An additional conformation, syn‐s‐trans, is present in enone complexes. Enantioface selectivity in the cycloaddition reactions differs for enals and enones. Reaction products indicate enals to react exclusively in the anti‐s‐trans conformation, whereas with enones, the major product results from the syn‐s‐trans conformation. The alkene in s‐cis conformations, while present in solution, is shielded and cannot undergo cycloaddition. A syn‐s‐trans conformation is found in the solid state of the bulky 6,6‐dimethyl cyclohexanone‐Ru(II) complex. The X‐ray structure of divinyl ketone is unique in that the Ru(II) center binds the enone via a η2 bond to one of the alkene moieties. In solution, coordination to Ru–C=O oxygen is adopted. A comparison of facial preference is also made to the corresponding indenyl Lewis acids.  相似文献   

19.
The crystal structure of the title compound, chloro(η5‐cyclopenta­dienyl){(1R,2S)‐2‐[(di­phenyl­phosphino)­methyl­amino]‐1‐phenyl­propyl di­phenyl­phosphinite‐κ2P,P′}ruthenium(II), [Ru(C5H5)Cl(C34H33NOP2)], is reported. The pseudo‐octa­hedral complex is chiral and the configuration at the Ru atom is S. The seven‐membered metallacycle adopts a boat‐like conformation.  相似文献   

20.
The objective of this work was to combine the ABC‐transporter inhibition characteristic of esters of the Schisandra sphenantera metabolite schisandrol A with the growth inhibitory and anti‐migratory effects typical of titanium complexes. To this end dimethyltitanocene, (h5‐C5H5)2Ti(CH3)2, was reacted with the vicinal diol schisandrol A to afford the schisandroxy(methyl)titanocene ( 1 ) as a stable water‐soluble solid. In MTT assays against seven tumor cell lines it proved distinctly more cytotoxic than schisandrol A or dimethyltitanocene or combinations of these. It reached single‐digit micromolar IC50(72 h) values against cells of leukemia HL‐60, melanoma 518A2 and also resistant cervix carcinoma KB‐V1/vbl and resistant breast carcinoma MCF‐7/topo. Non‐malignant fibroblasts were virtually insensitive to it [IC50 (72 h) > 50 µM ]. In addition, the new complex inhibited the p‐gp drug transporters of KB‐V1/vbl cells and prevented a regrowth and closure of scratch wounds in cancer cell cultures (‘wound‐healing’ assay) when applied in concentrations < 1 µM , which is an indication of a potential anti‐migratory and anti‐invasive activity in solid tumors. Copyright © 2010 John Wiley & Sons, Ltd.  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