Registered users receive a variety of benefits including the ability to customize email alerts, create favorite journals list, and save searches.
Please note that a BioOne web account does not automatically grant access to full-text content. An institutional or society member subscription is required to view non-Open Access content.
Contact helpdesk@bioone.org with any questions.
The effects of gymnochrome A were tested on the electrical activity of the frog atrial heart muscle. Gymnochrome A (1–5 μM) did not alter the resting potential. Gymnochrome A (5 μM) slowed the initial depolarizing phase of the spontaneously beating action potential. Under voltage-clamp conditions gymnochrome A (5 μM) did not affect the electrical constant of the membrane and the kinetic parameters of the peak Na current (INa) recorded in the Ringer solution containing tetraethylammonium (2 mM) and Cd2 (1 mM) but shifted the membrane potential at which the current both activated and reached its maximal value toward more negative membrane potentials. It did not alter the reversal potential for INa, indicating that the selectivity of the Na channels had not changed. These observations suggest that gymnochrome A binds to the membrane and shifts the activation of INa on the voltage axis by modifying the free negative fixed charges present at the membrane surface rather than by occupying a specific site on the Na channel. Photoexcited gymnochrome A transiently triggered an early outward current which lengthened the time-to-peak of INa and decreased its amplitude. In addition, photoexcited gymnochrome A blocked the background K current. This is, to our knowledge, the first time that such effects are reported on the cardiac muscle. These observations suggest that the photoexcitation of gymnochrome produces physico-chemical effects which lead to intracellular changes. Further experiments are required to determine their nature.
The potent photodynamic properties of hypericin (HY) elicit a range of light-dependent virucidal and tumoricidal activities. Yet, a relatively low reduction/oxidation potential endows HY with electron accepting and donating properties enabling it to act as both an oxidizing and a reducing agent. HY can thus compete as an electron acceptor from bioenergized reduction/oxidation reactions generating its excitation energy for biological activities from physiological reduction/oxidation reactions in the absence of light. Our studies show that HY can inhibit the growth of highly metastatic murine breast adenocarcinoma and squamous cell carcinoma tumors in culture. Furthermore, we show that HY can interfere with the growth of these tumors in mice reducing tumor size and prolonging animal survival in complete absence of light. While there is no evidence that HY induces apoptosis in these cells in the dark, 3H-thymidine incorporation into DNA was significantly reduced indicating effects that are apparently cytostatic in nature compared to the cytocidal effects of HY with light.
In a recent clinical study we showed that hypericin accumulates selectively in urothelial lesions following intravesical administration of the compound to patients. In the present study the efficacy of hypericin as a photochemotherapeutic tool against urinary bladder carcinoma was investigated using the AY-27 cells (chemically induced rat bladder carcinoma cells). The uptake of hypericin by the cells increased by prolonging the incubation time and increasing the extracellular hypericin concentration. Photodynamic treatment of the cells incubated with 0.8 and 1.6 μM hypericin concentrations resulted in remarkable cytotoxic effects the extent of which depended on the fluence rates. Photoactivation of 1.6 μM hypericin by 0.5, 1.0 or 2.0 mW/cm2 for 15 min resulted in 3, 30 and 95% of the antiproliferative effect, respectively. Increasing the photoactivating light dose from 0.45 to 3.6 J/cm2 resulted in a five-fold increase in hypericin photodynamic activity. Irrespective of the fluence rates and irradiation times incubation of the cells with 10 μM hypericin induced rapid and extensive cell death in all conditions. The type of cell death (apoptosis or necrosis) induced by photoactivated hypericin depended largely on the hypericin concentration and the postirradiation time. At lower hypericin concentrations and shorter postirradiation times apoptosis was the prominent mode of cell death; increasing the hypericin concentration and/or prolonging the postirradiation time resulted in increased necrotic cell death. Cell pretreatment with the singlet oxygen quencher histidine, but not with the free-radical quenchers, significantly protected the cells from photoactivated hypericin–induced apoptosis, at least when a relatively low concentration (1.25 μM) was used. This result suggests the involvement of a Type-II photosensitization process. However, cells treated with higher hypericin concentrations (2.5–5 μM) were inadequately protected by histidine. Since hypericin is thus shown to be a potent and efficient photosensitizer, and since the conditions used were the same as when hypericin is used clinically to locate early-stage urothelial carcinoma lesions, hypericin may well become very important for the photodynamic treatment of superficial bladder carcinoma.
Annelies Vantieghem, Yan Xu, Wim Declercq, Peter Vandenabeele, Geertrui Denecker, Jackie R. Vandenheede, Wilfried Merlevede, Peter A. de Witte, Patrizia Agostinis
In this study we show that overexpression of Bcl-2 in PC60R1R2 cells reveals a caspase-dependent mechanism of cytochrome c release following photodynamic therapy (PDT) with hypericin. Bcl-2 overexpression remarkably delayed cytochrome c release, procaspase-3 activation and poly(adenosine diphosphate-ribose)polymerase cleavage during PDT-induced apoptosis while it did not protect against PDT-induced necrosis. PDT-treated cells showed a reduction in the mitochondrial membrane potential which occurred with similar kinetics in PC60R1R2 and PC60R1R2/Bcl-2 cells, and was affected neither by the permeability transition pore inhibitor cyclosporin A nor by the caspase inhibitor N-benzyloxycarbonyl-Val-Ala-Asp-fluoromethylketone (zVAD-fmk). Hypericin-induced mitochondrial depolarization coincided with cytochrome c release in PC60R1R2 cells while it precedes massive cytochrome c efflux in PC60R1R2/Bcl-2 cells. Preincubation of PC60R1R2 cells with zVAD-fmk or cyclosporin A did not prevent the mitochondrial efflux of cytochrome c, and caspase inhibition only partially protected the cells from PDT-induced apoptosis. In contrast, in PC60R1R2/Bcl-2 cells cytochrome c release and apoptosis were suppressed by addition of zVAD-fmk or cyclosporin A. These observations suggest that the progression of the PDT-induced apoptotic process in Bcl-2–overexpressing cells involves a caspase-dependent feed-forward amplification loop for the release of cytochrome c.
Aluminum ion complexed 5,8-di-Br-hypocrellin B is a new water-soluble perylenequinonoid derivative with enhanced absorption over hypocrellin B (HB) in the phototherapeutic window (600–900 nm). Electron paramagnetic resonance and 9,10-diphenyl-anthracene bleaching methods were used to investigate the photosensitizing activity of [Al2(5,8-di-Br-HB)Cl4]n in the presence of oxygen. Singlet oxygen, superoxide anion radical and hydroxyl radical can be generated by [Al2(5,8-di-Br-HB)Cl4]n photosensitization. Singlet oxygen (1O2) is formed via energy transfer from triplet-state [Al2(5,8-di-Br-HB)Cl4]n to ground-state molecular oxygen. 1O2 participates in the generation of a portion of superoxide anion radical (O2·−). Besides superoxide anion radical (O2·−) may originate from the electron transfer between the triplet-state [Al2(5,8-di-Br-HB)Cl4]n and the ground-state molecular oxygen. ·OH is formed through the Fenton–Haber–Weiss reaction and the decomposition of DMPO–1O2 adduct. Compared with HB [Al2(5,8-di-Br-HB)Cl4]n primarily remains and enhances the generation efficiency of superoxide anion radical and hydroxyl radical but that of singlet oxygen decreases.
The electron-transfer properties of the hypericin derivatives, dibromo-, hexaacetyl-, hexamethyl- and desmethylhypericin, were studied. Cyclovoltammetric measurements revealed that dibromo- and desmethylhypericin have almost the same redox potentials as the parent hypericin. Substitution of the hydroxyl groups by acetoxy leads to less negative E½ values, whereas methoxy substitution induces more negative values. Electron paramagnetic resonance (EPR)/electron nuclear double resonance/general TRIPLE spectroscopy and quantum mechanical calculations were used to establish the structure of the one-electron reduced stages of hypericin derivatives. Proton loss in the bay region, already demonstrated for hypericin, was also found for dibromo- and desmethylhypericin. The spin and charge of the radical ions are predominately confined to the central biphenoquinone moiety of the hypericin skeleton. Generation of the radical ions by in situ electrolysis indicates that the redox potentials of hypericin, dibromo- and desmethylhypericin, containing hydroxyls at the 1, 3, 4, 6, 8 and 13 positions, largely depend on the solvent. With phosphate-buffered saline (pH 7.4)/dimethylsulfoxide (DMSO) as the solvent the EPR spectra of the corresponding radical ions appear at markedly lower potentials than in pure DMSO and N,N′-dimethylformamide. However, this effect is not observable for hexaacetyl- and hexamethylhypericin-lacking hydroxyl groups. In all cases the EPR data and calculations revealed the presence of 7,14 tautomers.
B. M. Showalter, A. Datta, P. K. Chowdhury, J. Park, P. Bandyopadhyay, P. K. Choudhury, S. Kesavan, Y. Zeng, G. A. Kraus, M. S. Gordon, J. P. Toscano, J. W. Petrich
Using time-resolved infrared spectroscopy, ab initio quantum mechanical calculations and synthetic organic chemistry a region in the infrared spectrum of triplet hypericin has been found between 1400 and 1500 cm−1 corresponding to the translocation of the hydrogen atom between the enol and the keto oxygens, O···H···O. This result is discussed in the context of the photophysics of hypericin and of eventual measurements to observe directly the excited-state H-atom transfer.
In the present study, hypericin analogs with an increased hydrophilic character were synthesized. As chemical modifications alter the lipophilicity/hydrophilicity balance together with the photophysical/chemical background of the molecule the influence of these structural changes on the cellular uptake, retention and subcellular localization in HeLa cells was investigated. Besides, their photocytotoxic effects using three cell lines (HeLa, MCF-7, A431), as well as their plasma protein binding were also assessed. To assess the relative hydrophilic/lipophilic character of hypericin and analogs their retention times were determined on a reversed phase high performance liquid chromatography (C-18) column. The retention time of all the hypericin analogs was <46 min, except for dibenzyltetramethylhypericin (118 min), while the retention time of hypericin was >200 min (solvent system: methanol/citrate buffer 30 mM pH 7; 70/30). Hypericin, hexa-, penta- and dibenzyltetramethylhypericin displayed a potent antiproliferative effect at the nanomolar range after photosensitization (3.6 J/cm2). On the contrary, photoactivated tetrasulfonhypericin and fringelite D had no antiproliferative effect on the three cell lines, whereas hypericin polyethylene glycol showed only an intermediate cytotoxic effect on A431 cells. In dark conditions no antiproliferative effect was observed for any photosensitizer. The antiproliferative photoeffect correlated well with the intracellular accumulation as measured using HeLa cells. In general, the photocytotoxic hypericin analogs concentrated to a large extent, while the noncytotoxic compounds were not taken up by the HeLa cells. Furthermore, confocal laser microscopy revealed that all photosensitizers mainly concentrated in the perinuclear region, probably corresponding with Golgi apparatus and the endoplasmic reticulum, except for tetrasulfonhypericin which located at the plasma membrane. In addition, the plasma protein binding studies illustrated that hypericin bind extensively to the low-density lipoproteins, while the other hypericin analogs were mainly bound to heavy proteins (mostly albumin) and to a small extent to low-density lipoproteins.
Surface-enhanced Raman spectroscopy, resonance Raman spectroscopy and molecular modeling were employed to study the interaction of hypericin (Hyp) with human (HSA), rat (RSA) and bovine (BSA) serum albumins. The identification of the binding site of Hyp in serum albumins as well as the structural model for Hyp/HSA complex are presented. The interactions mainly reflect: (1) a change of the strength of H bonding at the N1–H site of Trp; (2) a change of the Trp side-chain conformation; (3) a change of the hydrophobicity of the Trp environment; and (4) a formation of an H-bond between the carbonyl group of Hyp and a proton donor in HSA and RSA which leads to a protonated-like carbonyl in Hyp. Our results indicate that Hyp is rigidly bound in IIA subdomain of HSA close to Trp214 (distance 5.12 Å between the centers of masses). In the model presented the carbonyl group of Hyp is hydrogen bonded to Asn458. Two other candidates for hydrogen bonds have been identified between the bay-region hydroxyl group of Hyp and the carbonyl group of the Trp214 peptidic link and between the peri-region hydroxyl group of Hyp and the Asn458 carbonyl group. It is shown that the structures of the Hyp/HSA and Hyp/RSA complexes are similar to, and in some aspects different from, those found for the Hyp/BSA complex. The role of aminoacid sequence in the IIA subdomains of HSA, RSA and BSA is discussed to explain the observed differences.
The photobleaching of hypocrellin B (HB) and its derivative butylamino-substituted hypocrellin B (BAHB), both of which are potent sensitizers for photodynamic therapy (PDT), were investigated by studies of absorption spectra and quenching experiments and by the determination of photoproducts. Control experiments indicated that the sensitizer, oxygen and light were essential for the photobleaching of HB and BAHB, which suggested that it was a photodynamic process, e.g. the photobleaching processes of both HB and BAHB were mainly self-sensitized photooxidations. The illumination of HB with visible light in oxygenated nonpolar solvents generated singlet oxygen efficiently [Φ(1O2) = 0.76] which in turn attacked the sensitizer HB with the subsequent formation of an endoperoxide product. The endoperoxide of HB is unstable at room temperature and undergoes loss of singlet oxygen with regeneration of the parent HB. The singlet oxygen released from the endoperoxide of HB was detected with chemical trapping experiments. The quenching experiments indicated that in increasingly polar solvents the superoxide anion mechanism (type I) as well as the singlet oxygen mechanism (type II) contributed to the photobleaching of HB. The introduction of the electron-donating butylamino group not only enhanced the yield of the superoxide anion generation but also altered the position of attack in the BAHB molecule by the activated oxygen species. No endoperoxide product was observed, and no singlet oxygen released from the photobleaching process of BAHB was detected. The photobleaching process of BAHB was more complex. Both the singlet oxygen and superoxide anion mechanism played important roles in the photobleaching of BAHB in all organic solvent used here, even in aerobic nonpolar solvents such as CHCl3.
Electron spin resonance technique and spin-trapping methods were used to determine the photoproduction of 1O2 and O2·− by hypomycin B (HMB), a novel perylenequinonoid pigment (PQP) possessing only one hydroxyl group. It was found that the yields of 1O2 and O2·− for HMB were comparable to those for hypocrellin A, a typical natural PQP with good photosensitivity. In addition, the absorption and fluorescence spectra for HMB were investigated. The pKa values in the ground and excited states of HMB were determined to be 8.94 and 5.54, respectively. Thus, the photodynamic mechanisms of HMB may involve not only the photogeneration of 1O2 and O2·− but also the light-induced acidification. Consequently, HMB is proposed to be a good photodynamic therapeutic agent.
Molecular interactions between hypericin and α-, β- and γ-crystallin proteins have been studied by means of absorption and steady-state fluorescence spectroscopy, aiming to clarify if and how the pigment binds to the proteins and to investigate the effects of visible-light irradiation on these molecular systems. Such a study is a prerequisite for assessing the possibility of using hypericin as a mild antidepressant and/or as a photodynamic agent for the treatment of eye tumors and eye viral and bacterial diseases without side injuries to the lens. We have shown that in dark-kept samples, with increasing α-crystallin concentration, both the fluorescence emission intensity and the ratio of the absorption maxima around 590 and 550 nm of hypericin increase. These effects have been attributed to the monomerization of nonfluorescent hypericin aggregates caused by the binding of the pigment to α-crystallin. The binding constant of hypericin has been evaluated to be of the order of 3.0 (mg/mL)−1, corresponding to a dissociation constant of the order of 0.3 mg/mL. Following irradiation with light of wavelengths over 400 nm, at an irradiance of 20 mW/cm2, both tryptophan and hypericin fluorescence emission intensities decrease. These effects are suggested to be the consequence of a spatial rearrangement of the protein framework which takes place following the α-crystallin photopolymerization sensitized by hypericin itself described in the literature. For the sake of comparison hypericin has been studied also in the presence of βH-, βL- and γ-crystallins at the same concentration.
Three new hypocrellin derivatives, amino- or amino acid–substituted on the side ring of hypocrellin B (HB), were synthesized by the reactions of HB with 3-methoxypropylamine, 6-aminohexanoic acid and γ-amino-n-butyric acid, respectively. The structures of these compounds were characterized with proton nuclear magnetic resonance spectra, infrared spectra and mass spectra. The UV–visible absorption spectra, singlet oxygen–generating quantum yield and amphiphilicities of hypocrellin derivatives were measured and compared with HB, the parent compound. These derivatives showed strong absorption in the domain of the phototherapeutic window (600–900 nm) and improved amphiphilicity. HB and the derivatives were preliminarily tested for their photodynamic effects on human oral cavity epithelial carcinoma KB cell line in vitro. Two amino acid–substituted hypocrellins showed phototoxicity to the KB cell line. At an inhibitory dosage of 50% killing only 0.51 μmol L−1 compound 3 (or 0.88 μmol L−1 compound 2) and 0.5 J cm−2 irradiation were required. The hypocrellins exhibited some dark toxicity to the KB cell line. HB and amino acid–substituted hypocrellins showed lower dark toxicity to the KB cell line than amino-substituted hypocrellins in the assessment of cell survival.
Hypericin in organic solvents displays two types of electronic spectra: one type which shows a distinct solvatochromic effect, the stable form, and the other, the unstable form, which lacks this property. The latter type is formed in dry nonprotic solvents (e.g. tetrahydrofuran, EtOAc) and can be converted to the stable form on addition of protic solvents. In order to establish the tautomeric structure of the unstable form we applied conventional nuclear magnetic resonance techniques as well as two-dimensional gradient-enhanced heteronuclear multiple-quantum correlation, gradient-enhanced ROESY and one-dimensional nuclear Overhauser effect difference experiments. All these techniques pointed to the fact that the unstable form has the 7,14-diketo tautomeric structure, like the stable form, and not the 1,6-diketo structure. Electronic spectroscopy indicated that the unstable form has acidic properties and therefore possesses two free OH groups at C3 and C4 at the bay region of the molecule.
To optimize a hypericin derivative as a potential photodynamic therapy agent its light-induced singlet oxygen/superoxide radical formation capability should be enhanced and its long-wavelength absorption band should be bathochromically shifted to better match medicinal lasers. A heavy-atom–substituted derivative was realized by electrophilic iodination of hypericin to yield 2,5-diiodo-hypericin. Using photodestruction of bilirubin IXα this derivative was demonstrated to exhibit an enhanced light-induced singlet oxygen/superoxide radical formation capability as compared to hypericin. With respect to a bathochromically shifted derivative styryl residues were attached to the methyl groups of hypericin by de novo ring synthesis. Although the long-wavelength absorption band of this derivative displayed a bathochromic shift of nearly 40 nm it unfortunately immediately underwent an intramolecular [2 2] cycloaddition to yield the corresponding cyclobutane derivative in which the added conjugation system became interrupted.
A series of hypericin analogs were found to differ in their cytotoxic activity induced by ambient light levels. These analogs vary in their ability to partition into cells, to generate singlet oxygen as well as in other photophysical properties. The data suggest that the biological activity of hypericin is due to a combination of factors whose roles may vary under different circumstances.
Hypericin (HYP) has been reported to have photodependent cytotoxic activity in a variety of cancer cell lines. However, this activity has yet to be rigorously tested in vivo in tumor models. In this study LNCaP, PC-3 and DU-145 cells were used to test the cytotoxic effects of HYP in vitro, precursory to an in vivo study designed to investigate the effects of HYP in an established murine model for prostate cancer. Specifically, the model used employs immunocompromised nude mice bearing the LNCaP solid tumor xenograft. In vitro cytotoxicity experiments indicated that the dose causing 50% lethality for HYP in LNCaP, PC-3 and DU-145 cells were 2.07, 2.15 and 2.23 μM, respectively, following irradiation with red light (590 nm) for 30 min at a fluence rate of 0.1 J/cm2/s. Cells treated with HYP in the absence of photoirradiation showed no signs of cytotoxicity. A tissue distribution study was also carried out using the LNCaP solid tumor model to determine whether or not HYP is distributed to the target tissue. HYP was broadly distributed in tissues studied, including LNCaP tumor xenograft tissue. Furthermore, tumor tissue eliminated HYP at a slower rate than any of the other tissues examined. Interestingly, HYP levels were maintained in serum 24 h after oral administration (5 mg/kg dose). A pilot study designed to examine the efficacy of HYP treatment in nude mice bearing LNCaP tumors conducted over 28 days suggested that HYP, in combination with photoirradiation, inhibits both tumor growth and the elevation of prostate-specific antigen levels. Although the results reported for the current studies are preliminary they do provide evidence for an application of HYP PDT to prostate cancer which warrants further investigation.
In an attempt to develop an efficient chemotherapeutic agent targeted at malignant cells that express receptors to gonadotropin releasing hormone (GnRH) we coupled [d-Lys6]GnRH covalently to an emodin derivative, i.e. emodic acid (Emo) to yield [d-Lys6(Emo)]GnRH. Emodin is a naturally occurring anthraquinone which is widely used as a laxative and has other versatile biological activities. Physico-chemical studies employing electron paramagnetic resonance and electrochemistry of the conjugate as well as the (Emo) moiety showed that these compounds could be easily reduced either chemically, photochemically or enzymatically to their corresponding semiquinones. In the presence of oxygen the semiquinones generated reactive oxygen species (ROS), mainly superoxide and hydroxyl radicals, which were detected by the spin trapping method. Moreover, upon irradiation with visible light these compounds produced ROS and a highly reactive excited triplet state of Emo, which by itself may cause the oxidation of certain electron acceptors such as amino acids and bases of nucleic acids. Thus, [d-Lys6]GnRH-photosensitizer conjugates may be potentially used for targeted photodynamic chemotherapy aimed at treating cancer cells that carry GnRH receptors. These conjugates may also induce cytotoxicity in the dark similar to common conventional chemotherapeutic agents. The peptidic moiety, [d-Lys6]GnRH, was found to be stable toward highly reactive ROS generated either from enzymatic reduction or upon photoirradiation. The physico-chemical properties of Emo were only marginally influenced by the peptidic [d-Lys6]GnRH carrier.
We describe a series of new long-wave absorbing and fluorescing cyanine dyes and labels (based on a general logic for the design of such dyes), their spectra, covalent and noncovalent linkage to proteins, their use in single molecule detection (SMD) and as donors and acceptors, respectively, in fluorescence resonance energy transfer studies. The new labels represent water-soluble and reactive fluorophores whose quantum yields increase substantially if noncovalently or covalently bound to proteins. Due to their strong absorptions between 550 and 700 nm they are excitable by light-emitting diodes or diode lasers. Their high absorbances (ε around 100 000) and adequate fluorescence quantum yields (ϕ up to 0.68 if bound to proteins) along with their availability as reactive NHS esters make them viable labels for proteins and oligomers, e.g. in context with SMD or fluorescence energy transfer immunoassay which is demonstrated for the system HSA/anti-HSA.
Lower extracellular pH in tumors as compared to normal tissues has been proposed to be a factor contributing to the tumor selective uptake of several photosensitizers. Therefore, the pH dependence of absorption and fluorescence spectral properties of four different drugs relevant for photodynamic therapy (hematoporphyrin IX [HpIX], disulfonated meso-tetraphenylporphine [TPPS2a], meso-tetra(3-hydroxyphenyl)porphine [mTHPP] and meso-tetra(3-hydroxyphenyl)chlorin [mTHPC]) has been examined. Spectral analysis of the dyes dissolved in phosphate buffered saline (PBS) indicates pH-dependent modification in the physiologically important region (6.0–8.0) only in the case of HpIX. This modification is probably related to the protonation of carboxylic groups. Spectral changes of HpIX in PBS observed at acidic pH values <5, as well as those of the rest of the drugs (inflection points of titration curves occurred at about 5.1, 3.8 and 2.4 for TPPS2a, mTHPP and mTHPC, respectively), are likely to be due to the protonation of imino nitrogens. The tumor localizing properties of mTHPP and mTHPC reported in the literature appear to be due to factors other than pH-dependent changes in the lipophilicity of the drugs.
Infrared spectra of N4-hydroxycytosine isolated in argon and nitrogen low-temperature matrixes are reported. The compound was found to adopt the syn structure of the imino-oxo tautomeric form exclusively. A photoreaction (λ > 250 nm) converting this form into the anti isomer was observed. The reaction is reversible and leads to a photostationary point. The initial infrared spectrum and the spectrum of the photoproduct were assigned to the syn and anti isomers, respectively. This assignment is based on a close agreement between the experimental spectra and the spectra theoretically simulated at the DFT(B3LYP)/6-31 G** level of theory.
Joseph E. Tietge, Stephen A. Diamond, Gerald T. Ankley, David L. DeFoe, Gary W. Holcombe, Kathleen M. Jensen, Sigmund J. Degitz, Gregory E. Elonen, Edward Hammer
Recent reports concerning the lethal effects of solar ultraviolet-B (UV-B) (290–320 nm) radiation on amphibians suggest that this stressor has the potential to impact some amphibian populations. In this study embryos and larvae of three anuran species, Rana pipiens,Rana clamitans and Rana septentrionalis, were exposed to full-spectrum solar radiation and solar radiation filtered to attenuate UV-B radiation or UV-B and ultraviolet-A (UV-A) (290–380 nm) radiation to determine the effects of each wavelength range on embryo and larval survival. Ambient levels of solar radiation were found to be lethal to all three species under exposure conditions that eliminated shade and refuge. Lethality was ameliorated by filtration of UV-B radiation alone, demonstrating that ambient UV-B radiation is sufficient to cause mortality. Although several studies have qualitatively demonstrated the lethality of UV-B to early life stage amphibians this study demonstrates that the larval life stages of the three species tested are more sensitive than the embryonic stages. This suggests that previous reports that have not included the larval life stage may underestimate the risk posed to some anuran populations by increasing UV-B exposure. Furthermore, this study reports quantitative UV-B dosimetry data, collected in conjunction with the exposures, which can be used to begin the assessment of the impact of environmental changes which increase UV-B exposure of these anurans.
Applying various restriction enzymes on a specially designed 1.5 kb DNA fragment revealed that the inhibitory effects of psoralens UVA irradiation (PUVA) treatment on restriction endonuclease activities are caused by recognition inhibition. In this study restriction enzymes that have a 5′-TpA sequence at the cleaving site (KpnI, XbaI, PmeI and DraI), and the noncleaving site (PacI) in recognition sites, or have two 5′-TpA sequences at the recognition site, and a nonspecific sequence between the recognition and the cleaving sites (BciVI), were inhibited by PUVA treatment. Most of the other restriction enzymes used in this study, which do not have a 5′-TpA sequence at their restriction site, were not inhibited by PUVA treatment, although a 5′-TpA sequence is located adjacent (SmaI) or very close (BamHI, SacI and PstI) to the recognition and cleaving sites for these enzymes. Because SphI, which does not have 5′-TpA at its restriction site, was strongly inhibited by PUVA treatment, the 5′-CpA sequence is suggested to be a new binding site of psoralens after UVA irradiation.
Ultraviolet radiation (UVR) is known to affect a variety of cellular functions, including gene expression. A number of signaling pathways have been suggested to mediate these effects, including the participation of activator protein-1 (AP-1), activator protein-2 (AP-2) and nuclear factor-κB (NF-κB). The divergent results from previous studies could be explained, at least in part, by the source of UVR with different spectral characteristics as well as the type of cells employed as targets. In this study we have utilized UVA-340 as an energy source with output which closely matches the spectrum of natural sunlight over the range of 295–350 nm for irradiation of cultured fibroblasts. Combination of electrophoretic mobility shift assays and Northern analyses revealed activation of AP-1 but not NF-κB or AP-2. Inhibition studies further suggested the participation of protein kinase-C, but not protein kinase-A, and that an inhibitor of mitogen-activated protein kinase (MEK-1/2) did not alter the AP-1 activation. Free radical quenchers, sodium azide and N-acetylcysteine, did not affect the AP-1 binding activity. Finally, UVA-340 was shown to enhance transcriptional expression of the type-VII collagen gene (COL7A1), which is endogenously expressed in dermal fibroblast in an AP-1 dependent manner. Introduction of a mutation into the AP-1 site of the COL7A1 promoter abolished this activation. Thus, our results obtained by utilizing a novel energy source, UVA-340, mimicking natural sunlight at UVB and lower UVA range indicate a role for AP-1 in mediating the enhanced gene expression by UVR. Collectively, these results suggest that AP-1 is an important mediator of UVR action in fibroblasts.
We have previously demonstrated that decreases in skin elasticity, accompanied by increases in the tortuosity of elastic fibers, are important early events in wrinkle formation. In order to study the role of elastases in the degeneration of elastic fibers during wrinkle formation we examined the effects of an inhibitor of skin fibroblast elastase, N-phenethylphosphonyl-l-leucyl-l-tryptophane (NPLT), on wrinkle formation in hairless mice skin following UV irradiation. Dorsal skins of hairless mice were exposed daily to UV light for 18 weeks at doses of 65–95 mJ/cm2 and treated topically with 100 μL of 1 mM NPLT immediately after each UV irradiation. Wrinkles on dorsal skins were evaluated from week 6 through week 18. The daily exposure of mouse skin to UV light with less than 1 minimal erythemal dose significantly enhanced the activity of elastase in the exposed skin by week 4, and the elevated levels of elastase activity were significantly reduced by the in vitro incubation with NPLT in a dose-dependent manner to a level similar to that in unexposed mice skin, indicating that NPLT can efficiently inhibit the UV-inducible elastase activity. Topical application of NPLT significantly suppressed wrinkle formation when compared with vehicle controls by week 15 of treatment (P < 0.05). Histochemistry of elastic fibers with Orcein staining demonstrated that there were no obvious decreases of the fine elastic fibers in UV-exposed NPLT-treated skin in contrast to their marked decreases in the UV-exposed vehicle-treated skin. These findings suggest that skin fibroblast elastase plays a decisive role in wrinkle formation through the degeneration of elastic fiber.
The electronic excited-state behavior of photosystem II (PSII) in Mantoniella squamata, as influenced by the xanthophyll cycle and the transthylakoid pH gradient (ΔpH), was examined in vivo.Mantoniella is distinguished from other photosynthetic organisms by two main features namely (1) a unique light-harvesting complex that serves both photosystems I (PSI) and II (PSII); and (2) a violaxanthin (V) cycle that undergoes only one de-epoxidation step in excess light to accumulate the monoepoxide antheraxanthin (A) as opposed to the epoxide-free zeaxanthin (Z). The cells were treated first with high light to induce the ΔpH and A accumulation, followed by herbicide-induced closure of PSII traps and a chilling treatment, to sustain and stabilize the ΔpH and nigericin-sensitive fluorescence level in the dark. De-epoxidation was controlled with subsaturating concentrations of dithiothreitol (DTT) and was 5–10 times more sensitive to DTT than higher plant thylakoids. The PSII energy dissipation involved two steps: (1) the pH activation of the xanthophyll binding site that was associated with a narrowing and slight attenuation of the main 2 ns (ns = 10−9 s) fluorescence lifetime distribution; and (2) the concentration-dependent binding of A to the activated binding site yielding a second distribution centered around 0.9 ns. Consistent with the model of Gilmore et al. (1998) (Biochemistry 37, 13 582–13 593), the fractional intensity of the 0.9 ns component depended almost entirely on the A concentration and correlated linearly with the decrease of the steady-state chlorophyll a fluorescence intensity.
Photochemical internalization (PCI) is a novel method for the endosomal or lysosomal release of membrane-impermeable molecules into the cytosol of target cells. This novel technology is based on the photodynamically induced rupture of endocytic vesicles preloaded with molecules of therapeutic interest. PCI of the ribosome-inactivating plant toxin gelonin and the immunotoxin monoclonal antibody 31 (MOC31)-gelonin has been performed previously by the use of the endocytic vesicle–localizing photosensitizers TPPS2a and AlPcS2a and light, demonstrating synergistic toxicity against the more than 20 different cell lines tested, most of them of neoplastic origin. In this study we demonstrate that 5-aminolevulinic acid (5-ALA)–induced protoporphyrin IX (PpIX) is also capable of inducing PCI of MOC31-gelonin in the human colon adenocarcinoma cell line WiDr. The cells were incubated with 1 mM 5-ALA for up to 8 h in serum-free medium and from 24 to 96 h in serum-containing medium. Fluorescence microscopical studies indicate a partial plasma membrane localization of PpIX when 5-ALA was applied under serum-free conditions. This plasma membrane localization was not seen when 5-ALA was given in the presence of serum. There was a granular component of the PpIX localization in addition to a diffuse cytoplasmic localization. The granular component resembled the localization of the fluorescent dye conjugate Alexa-gelonin and the lysosomal localizing dye acridine orange. Our present results provide evidence for an endocytic vesicle–associated fraction of PpIX after 5-ALA incubation of the WiDr cells. We demonstrate that PCI, by combining 5-ALA, MOC31-gelonin and light, induces a synergistic cytotoxic effect against the WiDr cells.
The in vivo pharmacokinetics of protoporphyrin IX (PpIX) after administration of 5-aminolevulinic acid (ALA) cannot be described accurately by mathematical models using first-order rate processes. We have replaced first-order reaction rates by dose-dependent (Michaelis–Menten [MM]) reaction rates in a mathematical compartment model. Different combinations of first-order and dose-dependent reaction rates were evaluated to see which one would improve the goodness-of-fit to experimentally determined in vivo PpIX fluorescence kinetics as a function of concentration. The mathematical models that were evaluated are all based on a three-compartment model for drug distribution, conversion to PpIX and subsequent conversion to heme. Implementation of dose-dependent reaction rates improved the goodness-of-fit and enabled interpolation to other drug doses. For most data sets the time constant for delivery to the target cells turned out to be dose dependent. For all data sets the use of MM rates for the conversion of ALA to PpIX yielded better fits. The clearance of PpIX turned out to be a first-order process for all doses and types of administration. Fluorescence curves measured on a specific tissue type but obtained in different studies with different measurement techniques could be described with a single set of parameters.
We originally proposed that the subcellular target for one class of photosensitizing agents was the mitochondrion. This classification was based on effects that occur within minutes of irradiation of photosensitized cells: rapid loss of the mitochondrial membrane potential (ΔΨm), release of cytochrome c into the cytosol and activation of caspase-3. These effects were followed by the appearance of an apoptotic morphology within 30–90 min. Fluorescence localization studies on three sensitizers initially classified as ‘mitochondrial’ revealed that these agents bind to a variety of intracellular membranes. The earliest detectable effect of photodamage is the selective loss of the antiapoptotic protein bcl-2 leaving the proapoptotic protein bax undamaged. Bcl-2 photodamage can be detected directly after irradiation of cells at 10°C. Subsequent warming of cultures to 37°C results in loss of ΔΨm, release of cytochrome c and activation of caspase-3. The latter appears to amplify the other two effects. Based on results reported here we propose that the apoptotic response to these photosensitizers is derived from selective photodamage to the antiapoptotic protein bcl-2 while leaving the proapoptotic protein bax unaffected.
Clinical follow-up studies have revealed that PUVA-treated patients are at increased risk of skin cancer, particularly squamous cell carcinoma (SCC). However, since psoralen and UVA (PUVA) is not only a potent mutagen and carcinogen but also an immunosuppressor, and since other (co)carcinogenic factors often exist in psoriasis patients, the exact causes and mechanisms of PUVA-associated SCC are still not completely understood. In order to fill this gap the tools of molecular epidemiology are being used to study the SCC mutational spectra of p53 and Ha-ras, two of the most commonly mutated genes in human cancers. A previous mutation analysis revealed that SCC in PUVA-treated patients often carried mutated p53 genes and that many of the mutations had the UV fingerprint (i.e. C→T or CC→TT transitions at dipyrimidine sites). In the present study DNA-sequencing analysis revealed a total of 18 Ha-ras missense or nonsense mutations at exons 1–4 in 13 of 17 SCC (76%) from 8 of 11 (73%) PUVA-treated psoriasis patients. Six of the 18 mutations (33%) were of UV-fingerprint type (C→T transitions), five (28%) were at 5′-TpG sites (i.e. potential psoralen-binding sites and thus potentially caused by PUVA) and seven were of other type (39%), including six G:C→T:A transversions at hotspot codon 12. In addition, in the case of 6 of the 11 subjects (55%) both tumor and normal skin samples contained a T:A→C:G base change at codon 27 (a 5′-ATT site), a change previously hypothesized to be a possible silent Ha-ras polymorphism at one allele. When we compared the present Ha-ras mutation spectrum with the p53 mutation spectrum from a previous study of the samples, we found that approximately half of the tumors harbored mutations in both Ha-ras and p53. Together, our results indicate that Ha-ras mutations are present in a large proportion of PUVA-associated SCC and that UVB, PUVA and other agents may induce Ha-ras mutations and act together with p53 in the formation of SCC in psoriasis patients.
We have previously shown that the rhodacyanine dye, Rhodac, exhibits a potent photocytotoxic activity in HeLa cells. In this study several aspects of the photobiological activity of Rhodac were further examined. Rhodac displayed no selective cytotoxicity toward several malignant cell lines after photosensitization (3.6 J/cm2), although HeLa cells were found to be the most sensitive. Interestingly, MCF-7/Adr cells, a multidrug-resistant subline, were less sensitive to the antiproliferative effect of photoactivated Rhodac. The subcellular localization, as revealed by confocal laser microscopy, demonstrated that the dye was mainly concentrated in the cytosolic membranes of the perinuclear region. The Rhodac-induced inhibition of HeLa cell proliferation after light exposure was found to be strictly oxygen dependent. In addition, photoactivated Rhodac induced poly(adenosine 5′ diphosphate-ribose)polymerase cleavage, caspase-3 activation and apoptosis in HeLa cells. In the current work it was further demonstrated that Rhodac binds specifically to high-density lipoproteins and low-density lipoproteins, while no binding was observed to very low-density and heavy proteins. To sum up, our results show that Rhodac is an interesting and potent photosensitizer. Further in vivo experiments are required to elucidate whether the lipoprotein binding leads to a selective uptake of Rhodac in tumor cells and to address its efficacy in photodynamic therapy.
Photodynamic therapy (PDT) with topical aminolevulinic acid (ALA) has been shown in previous studies to improve psoriasis. However, topical ALA-PDT may not be practical for the treatment of extensive disease. In order to overcome this limitation we have explored the potential use of oral ALA administration in psoriatic patients. Twelve patients with plaque psoriasis received a single oral ALA dose of 10, 20 or 30 mg/kg followed by measurement of protoporphyrin IX (PpIX) fluorescence in the skin and circulating blood cells. Skin PpIX levels were determined over time after ALA administration by the quantification of the 635 nm PpIX emission peak with in vivo fluorescence spectroscopy under 442 nm laser excitation. Administration of ALA at 20 and 30 mg/kg induced preferential accumulation of PpIX in psoriatic as opposed to adjacent normal skin. Peak fluorescence intensity in psoriatic and normal skin occurred between 3 and 5 h after the administration of 20 and 30 mg/kg, respectively. Ratios of up to 10 for PpIX fluorescence between psoriatic versus normal skin were obtained at the 30 mg/kg dose of ALA. Visible PpIX fluorescence was also observed on normal facial skin, and nonspecific skin photosensitivity occurred only in patients who received the 20 or 30 mg/kg doses. PpIX fluorescence intensity was measured in circulating blood cells by flow cytometry. PpIX fluorescence was higher in monocytes and neutrophils as compared to CD4 and CD8 T lymphocytes. PpIX levels in these cells were higher in patients who received higher ALA doses and peaked between 4 and 8 h after administration of ALA. There was only a modest increase in PpIX levels in circulating CD4 and CD8 T lymphocytes. In conclusion oral administration of ALA induced preferential accumulation of PpIX in psoriatic plaques as compared to adjacent normal skin suggesting that PDT with oral ALA should be further explored for the treatment of psoriasis.
Several previous studies have suggested that the peripheral benzodiazepine receptor (PBR) on the mitochondrial surface was an important target for photodynamic therapy (PDT). In this study we compared PBR affinity vs photodynamic efficacy of protoporphyrin-IX (PP-IX) and two structural analogs, PP-III and PP-XIII, using murine leukemia L1210 cells in culture. The results indicate that the three agents have approximately equal hydrophobicity, affinity for L1210 cells and ability to initiate photodamage leading to an apoptotic response. But only PP-IX had significant affinity for the PBR. These data indicate that the relationship between PDT efficacy and PBR affinity may hold only for sensitizers with the PP-IX configuration.
The present study investigated the rhythmic changes in glutathione status in midgut gland and hemolymph as well as in glutathione reductase (GR) activity in the crayfish Procambarus clarkii. In order to determine the circadian nature of these rhythms different groups of crayfish were submitted to constant-darkness conditions for 24 or 72 h after they had spent 15 days under light–dark 12:12 cycles. The animals of the different batches were killed at 6 h intervals during a 24 h cycle. Reduced glutathione (GSH) and oxidized glutathione (GSSG) in hemolymph and midgut as well as midgut GR activity were determined in midgut gland and hemolymph by fluorometric and spectrophotometric methods. Data analysis by chronogram and single Cosinor revealed circadian rhythmicity for GSH and GSSG concentration in both tissues as well as midgut GR activity. The rhythm parameters revealed oxidative stress induced by light. The possible correlation between the glutathione rhythm and other metabolic and behavioral rhythms of crayfish as well as the importance of the glutathione circadian temporal order in the adaptation of crayfish are discussed.
New variants of green fluorescent protein (GFP) can be engineered by circular permutation of their amino acid sequence. We characterized a series of permuted enhanced GFP (PEGFP) with new termini introduced at N144-Y145 and linkers of 1, 3, 5 and 6 residues inserted between G232 and M1, as well as a variant with an extended 7-residues linker between K238 and M1. A minimum linker length of 3 residues was necessary for a functional chromophore to be formed, and linkers exceeding 4 residues yielded almost the same fluorescence quantum yield as enhanced GFP (EGFP). PEGFP exhibited dual-wavelength absorption and fluorescence excitation with peaks at 395 and 490 nm but single-wavelength emission at 512 nm. Fluorescence emission increased with increasing pH for all excitation wavelengths with a pKa of 7.7. Between the pH values of 6 and 8 optical absorption showed an isobestic point at 445 nm. PEGFP rapidly denatured in urea between 50 and 60°C. Renaturation proceeded with a short (∼29 s) and a longer (>150 s) time constant. Transient transfection of HEK293 and HeLa cells revealed the expression dynamics of PEGFP to be similar to that of EGFP. Laser-scanning microscopy of HeLa cells demonstrated that the PEGFP are particularly well suited as fluorescent indicators in two-photon imaging.
Nicole M. Haralampus-Grynaviski, Laura E. Lamb, John D. Simon, Jeffrey R. Krogmeier, Robert C. Dunn, Anna Pawlak, Malgorzata Rózanowska, Tadeusz Sarna, Janice M. Burke
The emission spectra of single lipofuscin granules are examined using spectrally resolved confocal microscopy and near-field scanning optical microscopy (NSOM). The emission spectrum varies among the granules examined revealing that individual granules are characterized by different distributions of fluorophores. The range of spectra observed is consistent with in vivo spectra of human retinal pigment epithelium cells. NSOM measurements reveal that the shape of the spectrum does not vary with position within the emissive regions of single lipofuscin granules. These results suggest that the relative distribution of fluorophores within the emissive regions of an individual granule is homogeneous on the spatial scale ∼150 nm.
This article is only available to subscribers. It is not available for individual sale.
Access to the requested content is limited to institutions that have
purchased or subscribe to this BioOne eBook Collection. You are receiving
this notice because your organization may not have this eBook access.*
*Shibboleth/Open Athens users-please
sign in
to access your institution's subscriptions.
Additional information about institution subscriptions can be foundhere