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 15 毫秒
1.
Conventional tools used in prosthetic revision surgery have a limited range of action within the narrow cement mantle. Water jet cutting technology permits tiny and precisely controlled cuts, and may therefore be an alternative method of bone cement removal. Our study compares the cutting performance on bone cement (PMMA) and bone of a pulsed water jet and a continuous water jet. The aim of the study was to establish whether selective removal of PMMA is possible. 55 bone specimens (bovine femora) and 32 specimens of PMMA were cut with a continuous and a pulsed water jet at different pressures (40 MPa, 60 MPa) and pulse frequencies (0Hz, 50Hz, 250Hz). To ensure comparability of the results, the depths of cut were related to the hydraulic power of that part of the jet actually impinging on the material. While for PMMA the power-related depth of cut increased significantly with the pulse frequency, this did not apply to bone. The cuts produced in bone were sharp-edged. Since PMMA is more brittle than bone, the water jet caused cracks that enlarged further until particles of bone broke away. Although selective removal of PMMA without doing damage to the bone was not possible at the investigated settings of the jet parameters, the results do show that a pulsed water jet can cut bone cement much more effectively than bone. This is an important advantage over conventional non-selective tools for the removal of bone cement.  相似文献   

2.
The physical mechanisms that enable short pulses of high-intensity ultraviolet laser radiation to remove tissue, in a process known as laser ablation, remain obscure. The thermodynamic response of biological tissue to pulsed laser irradiation was investigated by measuring and subsequently analyzing the stress transients generated by pulsed argon fluorine (ArF, lambda = 193 nm) and krypton fluorine (KrF, lambda = 248 nm) excimer laser irradiation of porcine dermis using thin-film piezoelectric transducers. For radiant exposures that do not cause material removal, the stress transients are consistent with rapid thermal expansion of the tissue. At the threshold radiant exposure for ablation, the peak stress amplitude generated by 248 nm irradiation is more than an order of magnitude larger than that produced by 193 nm irradiation. For radiant exposures where material removal is achieved, the temporal structure of the stress transient indicates that the onset of material removal occurs during irradiation. In this regime, the variation of the peak compressive stress with radiant exposure is consistent with laser-induced rapid surface vaporization. For 193 nm irradiation, ionization of the ablated material occurs at even greater radiant exposures and is accompanied by a change in the variation of peak stress with radiant exposure consistent with a plasma-mediated ablation process. These results suggest that absorption of ultraviolet laser radiation by the extracellular matrix of tissue leads to decomposition of tissue on the time scale of the laser pulse. The difference in volumetric energy density at ablation threshold between the two wavelengths indicates that the larger stresses generated by 248 nm irradiation may facilitate the onset of material removal. However, once material removal is achieved, the stress measurements demonstrate that energy not directly responsible for target decomposition contributes to increasing the specific energy of the plume (and plasma, when present), which drives the gas dynamic expansion of ablated material. This provides direct evidence that ultraviolet laser ablation of soft biological tissues is a surface-mediated process and not explosive in nature.  相似文献   

3.
Time-resolved imaging was used to examine the use of pulsed laser microbeam irradiation to produce cell lysis. Lysis was accomplished through the delivery of 6 ns, lambda=532 nm laser pulses via a 40x, 0.8 NA objective to a location 10 microm above confluent monolayers of PtK2 cells. The process dynamics were examined at cell surface densities of 600 and 1000 cells/mm2 and pulse energies corresponding to 0.7x, 1x, 2x, and 3x the threshold for plasma formation. The cell lysis process was imaged at times of 0.5 ns to 50 micros after laser pulse delivery and revealed the processes of plasma formation, pressure wave propagation, and cavitation bubble dynamics. Cavitation bubble expansion was the primary agent of cell lysis with the zone of lysed cells fully established within 600 ns of laser pulse delivery. The spatial extent of cell lysis increased with pulse energy but decreased with cell surface density. Hydrodynamic analysis indicated that cells subject to transient shear stresses in excess of a critical value were lysed while cells exposed to lower shear stresses remained adherent and viable. This critical shear stress is independent of laser pulse energy and varied from approximately 60-85 kPa for cell monolayers cultured at a density of 600 cells/mm2 to approximately 180-220 kPa for a surface density of 1000 cells/mm2. The implications for single cell lysis and microsurgery are discussed.  相似文献   

4.
The physical mechanisms that achieve tissue removal through the delivery of short pulses of high-intensity infrared laser radiation, in a process known as laser ablation, remain obscure. The thermodynamic response of biological tissue to pulsed infrared laser irradiation was investigated by measuring and analyzing the stress transients generated by Q-sw Er:YSGG (lambda = 2.79 microns) and TEA CO2 (lambda = 10.6 microns) laser irradiation of porcine dermis using thin-film piezoelectric transducers. For radiant exposures that do not produce material removal, the stress transients are consistent with thermal expansion of the tissue samples. The temporal structure of the stress transients generated at the threshold radiant exposure for ablation indicates that the onset of material removal is delayed with respect to irradiation. Once material removal is achieved, the magnitude of the peak compressive stress and its variation with radiant exposure are consistent with a model that considers this process as an explosive event occurring after the laser pulse. This mechanism is different from ArF- and KrF-excimer laser ablation where absorption of ultraviolet radiation by the collagenous tissue matrix leads to tissue decomposition during irradiation and results in material removal via rapid surface vaporization. It appears that under the conditions examined in this study, explosive boiling of tissue water is the process that mediates the ablation event. This study provides evidence that the dynamics and mechanism of tissue ablation processes can be altered by targeting tissue water rather than the tissue structural matrix.  相似文献   

5.
The enzyme adenosine deaminase (adenosine aminohydrolase EC 3.5.4.4) from calf intestinal mucosa is commercially available at high purity grade yet, at the sensitivity at which fluorescence studies may be undertaken, a nonpeptidic fluorescence is detectable at lambda exmax = 350 nm and lambda emmax = 420 nm. A sevenfold decrease of this nonpeptidic fluorescence was obtained upon irradiation by the third harmonic (355 nm) of a Nd:YAG laser for 16 min, at 5 mJ/pulse, with a pulse width of 6 ns at a repetition rate of 10 Hz. The decline of fluorescence was accompanied by a negligible loss of enzymatic activity. Moreover, the integrity of the protein was ascertained by (i) its fluorescence (lambda exmax = 305 nm, lambda emmax = 335 nm) and lifetime distribution and (ii) its kinetics in the presence of the substrate adenosine and two inhibitors, all of which remained essentially unaltered. Laser photobleaching is a simple way to achieve a fluorescence grade adenosine deaminase.  相似文献   

6.
Using the method of spin trapping and electron paramagnetic resonance, free radicals have been detected accompanying laser ablation of cardiovascular tissue. Radicals were detected using both visible and ultraviolet laser energy from argon-ion and excimer laser sources. The results are discussed in terms of the relative efficiency of the laser wavelengths to produce free radicals and a comparison of the types of radicals produced by the action of pulsed versus cw laser energy.  相似文献   

7.
Absorption changes (deltaA) at 820 nm, following laser flash excitation of spinach chloroplasts and Chlorella cells, were studied in order to obtain information on the reduction time of the photooxidized primary donor of Photosystem II at physiological temperatures. In the microsecond time range the difference spectrum of deltaA between 750 and 900 nm represents a peak at 820 nm, attributable to a radical-cation of chlorophyll a. In untreated dark-adapted material the signal can be attributed solely to P+-700; it decays in a polyphasic manner with half-times of 17 microseconds, 210 microseconds and over 1 ms. The oxidized primary donor of Photosystem II (P+II) is not detected with a time resolution of 3 microseconds. After treatment with 3--10 mM hydroxylamine, which inhibits the donor side of Photosystem II, P+II is observed and decays biphasically (a major phase with t1/2=20--40 microseconds, and a minor phase with t1/2 congruent to 200 microseconds), probably by reduction by an accessory electron donor. In the nanosecond range, which was made accessible by a new fast-response flash photometer operating at 820 nm, it was found the P+II is reduced with a half-time of 25--45 ns in untreated dark-adapted chloroplasts. It is assumed that the normal secondary electron donor is responsible for this fast reduction.  相似文献   

8.
The determination of the rate of release of Ca2+ by pulsed photolysis of the photolabile chelator DM-nitrophen is important for its use in time-resolved physiological studies: the rate of substrate or effector release should be faster than the processes they initiate. Flash photolysis of DM-nitrophen using a 50-ns pulse from a frequency-doubled ruby laser (with emission at 347 nm having energy of ca. 10-20 mJ) yields short-lived photochromic or aci-nitro intermediates. At pH 6.9, double-exponential decay of a photochromic intermediate was observed for DM-nitrophen itself and its Ca2+ complex (tau 1/2 values of 24 and 570 microseconds, and 32 and 220 microseconds respectively), while only monoexponential decay of the DM-nitrophen-Mg2+ complex was detected (tau 1/2 = 31 microseconds). Only the photochemistry of DM-nitrophen-Ca2+ was found to be pH sensitive (monoexponential decay, tau 1/2 approximately 115 microseconds at pH 7.9 and 8.9). Use of the Ca(2+)-sensitive metallochromic dye antipyrylazo III in conjunction with pulsed photolysis of DM-nitrophen-Ca2+ enabled an upper limit of the half-time of release of Ca2+ to be established of ca. 180 microseconds (the rate of association of Ca2+ with the dye was probably rate determining). The rate of Ca2+ photorelease may, however, be faster than this. Thus, the DM-nitrophen-Ca2+ complex releases Ca2+ on photolysis sufficiently rapidly for the study of many Ca(2+)-dependent physiological processes with improved kinetic resolution over conventional mixing methods.  相似文献   

9.
Laser-induced native fluorescence detection with a KrF excimer laser (λ=248 nm) was used to investigate the capillary electrophoretic (CE) profiles of human urine, saliva and serum without the need for sample derivatization. All separations were carried out in sodium phosphate and/or sodium tetraborate buffers at alkaline pH in a 50-μm I.D. capillary. Sodium dodecyl sulfate was added to the buffer for micellar electrokinetic chromatography (MEKC) analysis of human urine. Although inherently a pulsed source, the KrF excimer laser was operated at a high pulse repetition rate of 553, 1001 or 2009 Hz to simulate a continuous wave excitation source. Detection limits were found to vary with pulse rate, as expected, in proportion to average excitation power. The following detection limits (3σ) were determined in free solution CE: tryptophan, 4 nM; conalbumin, 10 nM; α-lactalbumin, 30 nM. Detection limits for indole-based compounds and catecholamine urinary metabolites under MEKC separation conditions were in the range 7–170 nM.  相似文献   

10.
Resonance Raman (RR) spectra are reported for amino acid and amine adducts of pyridoxal 5'-phosphate (PLP) and 5'-deoxypyridoxal (5'-dPL) in aqueous solution. For the valine adducts, a detailed study has been carried out on solutions at pH and pD 5, 9, and 13, values at which the pyridine and imine protons are successively ionized, and on the adducts formed from 15N-valine, alpha-deuterovaline, and N-methyl-PLP. Good quality spectra were obtained, despite the strong fluorescence of pyridoxal Schiff bases, by adding KI as a quencher, and by exciting the molecules on the blue side of their absorption bands: 406.7 nm (cw Kr+ laser) for the pH 5 and 9 species (lambda max = 409 and 414 nm), and 354.7 nm (pulsed YAG laser, third harmonic) for the pH 13 species (lambda max = 360 nm). A prominent band at 1646 cm-1 is assigned to the imine C=N stretch via its 13 cm-1 15N shift. A 12 cm-1 down-shift of the band in D2O confirms that the Schiff base linkage is protonated at pH 9. Deprotonation at pH 13 shifts VC = N from 1646 to 1629 cm-1, values typical of conjugated Schiff bases. The strongest band in the spectrum, at 1338 cm-1, shifts to 1347 cm-1 upon pyridine protonation at pH 5, and is assigned to a ring mode with a large component of phenolate C-O stretch. A shoulder on its low-frequency side is assigned to the C4-C4' stretch. Large enhancements of these modes can be understood qualitatively in terms of the dominant resonance structures contributing to the ground and resonant excited states. A number of weaker bands are observed, and assigned to pyridine ring modes. These modes gain significantly in intensity, while the exocyclic modes diminish, when the spectra are excited at 266 nm (YAG laser, fourth harmonic) in resonance with ring-localized electronic transitions.  相似文献   

11.
The linear dichroism spectrum of rhodopsin in sonicated bovine disk membranes was measured 30, 60, 170, and 600 ns after room temperature photolysis with a linearly polarized, 7-ns laser pulse (lambda = 355 or 477 nm). A global exponential fitting procedure based on singular value decomposition was used to fit the linear dichroism data to two exponential processes which differed spectrally from one another and whose lifetimes were 42 +/- 7 ns and 225 +/- 40 ns. These results are interpreted in terms of a sequential model where bathorhodopsin (BATHO, lambda max = 543 nm) decays toward equilibrium with a blue shifted intermediate (BSI, lambda max = 478 nm). BSI then decays to lumirhodopsin (LUMI, lambda max = 492 nm). It has been suggested that two bathorhodopsins decay in parallel to their products. However, a Monte Carlo simulation of partial photolysis of solid-state visual pigment samples shows that one mechanism which creates populations of BATHO having different photolysis rates at 77 K may not be responsible for the two decay rates reported here at room temperature. The angle between the cis band and 498-nm band transition dipoles of rhodopsin is determined to be 38 degrees. The angles between both these transition dipoles and those of the long-wave-length bands of BATHO, BSI, and LUMI are also determined. It is shown that when BATHO is formed its transition dipole moves away from the original cis band transition dipole direction. The transition dipole then moves roughly twice as much towards the original cis band direction when BSI appears. Production of LUMI is associated with return of the transition dipole almost to the original orientation relative to the cis band, but with some displacement normal to the plane which contains the previous motions. The correlation between the lambda max of an intermediate and its transition dipole direction is discussed.  相似文献   

12.
In order to elucidate the photophysical mechanisms of cellular phototoxicity sensitized by doxycycline, MGH-U1 human bladder carcinoma cells in vitro were treated with 20.7 microM doxycycline and irradiated with either a pulsed (lambda = 355 nm, pulse duration = 24 ps) or a continuous wave (lambda = 351 nm) laser. Cumulative radiant exposure and irradiance were systematically varied in experiments with both lasers. Phototoxicity was assessed by epifluorescence microscopy of unfixed cells using rhodamine 123 labeling of mitochondria. With the continuous wave source, the cumulative radiant exposure required for induction of phototoxic injury was independent of irradiance. With the 24-ps-pulsed source, a significantly lower cumulative radiant exposure was required to induce the phototoxicity when the peak irradiance was 5.8 x 10(7) or 1.3 x 10(8) watts cm-2 compared with when peak irradiance was either lower (6.0 x 10(6) watts cm-2) or higher (7.6 x 10(8) watts cm-2). The measured fluorescence lifetimes of doxycycline in buffered saline solution were longer than the laser pulse duration of 24 ps. The increased efficiency of photosensitization at the optimal peak irradiance in the ps domain appears to result from sequential multiphoton absorption involving higher excited states of the singlet manifold. At the highest irradiance studied, on the other hand, reduced efficiency of photosensitization is attributed to increased photodegradation of doxycycline from higher excited states by processes such as photoionization. A model consistent with these observations is presented along with calculations, based on simple rate equations, that fit the essentials of the proposed model.  相似文献   

13.
We report the results of a study on generation of reactive oxygen species (ROS) and changes in the membrane potential of mitochondria of carcinoma of cervix (HeLa) and Chinese hamster ovary (CHO) cells following exposure to continuous wave (cw) or pulsed Nd: YAG laser (1064 nm). For a given laser irradiation, the generation of ROS and induced changes in the membrane potential of mitochondria were more pronounced for HeLa cells as compared to CHO cells. However, in both the cells the laser dose required to elicit a given change was much lower with pulsed laser exposure compared to that required with a cw laser exposure. This suggests involvement of photothermal effects in the laser irradiation induced changes. Mechanistic studies using quenchers for ROS suggest that laser irradiation leads to generation of hydroxyl radicals.  相似文献   

14.
Raman spectra of oxy- and deoxyhemoglobin obtained with 218 and 200 nm pulsed (7 ns) laser excitation show changes (loss of 880 cm-1 tryptophan band intensity, increase in the 830/850 cm-1 tyrosine doublet intensity ratio) which are attributed to the aromatic contacts (Trp beta 37-Tyr alpha 140 and Tyr alpha 42-Asp beta 99) that are specific to the T quaternary structure. At high concentration (2 mM in heme) HbCO shows the same spectral signatures as HbO2. As the HbCO concentration is decreased, however, the spectra approach those shown by deoxy-Hb. This dilution effect is attributable to photolysis, which increases with decreasing concentration. The results imply that the HbCO photoproduct shows the same aromatic environments as does deoxy-Hb. Thus, T-like contacts are apparently formed at the alpha 1 beta 2 interface within 7 ns of photolysis, a time short compared to the spectral alterations of the heme group (approximately 100 ns, approximately 1 microsecond, and approximately 20 microseconds) which have previously been attributed to tertiary and quaternary relaxations.  相似文献   

15.
A study was made of the lethal effect of pulse laser (second harmonic Nd+3:YAG laser of 532 nm, pulse length 3.3.10(-11) s, peak intensity from 4.10(12) to 1.10(14) W/m2) on HeLa cells at the phases of active and stationary growth, and lethal and mutagenic effects of this radiation on E. coli cells. As was shown, HeLa cells at both growth phases and E. coli cells exhibited low sensitivity to laser radiation at lambda = 532 nm.  相似文献   

16.
Irradiation of amelanotic melanoma B78H1 cells in the presence of liposome-delivered Ni(II)-octabutoxy-naphthalocyanine with a Q-switched Ti:sapphire laser operated in a pulsed mode (850 nm, 30 ns pulses, 10 Hz, 120 mJ pulse -1) promotes a photothermal sensitization process leading to extensive cell inactivation. The photoprocess occurs with identical efficiency in N2-saturated and air-equilibrated media, indicating that this photosensitization modality does not require the presence of oxygen.  相似文献   

17.
Destruction of lambda phage DNA is studied under nanosecond pulse laser irradiation (lambda = 355 nm) of DNA-dye complexes in solution at 77K (dye--acridine orange, 8-methoxypsoralen, ethidium bromide). Free radicals induced by laser radiation are found to participate in DNA sugar-phosphate chain scission. It was observed that the quantity of DNA double-strand breaks correlated with that of the free radicals and that of oxygen influenced DNA laser destruction.  相似文献   

18.
Using fast flash photolysis, we have measured the binding of CO to carboxymethylated cytochrome c and to heme c octapeptide as a function of temperature (5 degrees-350 degreesK) over an extended time range (100 ns(-1) ks). Experiments used a microsecond dye laser (lambda = 540 nm), and a mode-locked frequency-doubled Nd-glass laser (lambda = 530 nm). At low temperatures (5 degrees-120 degreesK) the rebinding exhibits two components. The slower component (I) is nonexponential in time and has an optical spectrum corresponding to rebiding from an S = 2, CO-free deoxy state. The fast component (I*) is exponential in time with a lifetime shorter than 10 mus and an optical spectrum different from the slow component. In myoglobin and the separated alpha and beta chains of hemoglobin, only process I is visible. The optical absorption spectrum of I* and its time dependence suggest that it may correspond to recombination from an excited state in which the iron has not yet moved out of the heme plane. The temperature dependences of both processes have been measured. Both occur via quantum mechanical tunneling at the lowest temperatures and via over-the-barrier motion at higher temperatures.  相似文献   

19.
The carbonate radical anion (CO(3)) is believed to be an important intermediate oxidant derived from the oxidation of bicarbonate anions and nitrosoperoxocarboxylate anions (formed in the reaction of CO(2) with ONOO(-)) in cellular environments. Employing nanosecond laser flash photolysis methods, we show that the CO(3) anion can selectively oxidize guanines in the self-complementary oligonucleotide duplex d(AACGCGAATTCGCGTT) dissolved in air-equilibrated aqueous buffer solution (pH 7.5). In these time-resolved transient absorbance experiments, the CO(3) radicals are generated by one-electron oxidation of the bicarbonate anions (HCO(3)(-)) with sulfate radical anions (SO(4)) that, in turn, are derived from the photodissociation of persulfate anions (S(2)O(8)(2-)) initiated by 308-nm XeCl excimer laser pulse excitation. The kinetics of the CO(3) anion and neutral guanine radicals, G(-H)( small middle dot), arising from the rapid deprotonation of the guanine radical cation, are monitored via their transient absorption spectra (characteristic maxima at 600 and 315 nm, respectively) on time scales of microseconds to seconds. The bimolecular rate constant of oxidation of guanine in this oligonucleotide duplex by CO(3) is (1.9 +/- 0.2) x 10(7) m(-1) s(-1). The decay of the CO(3) anions and the formation of G(-H)( small middle dot) radicals are correlated with one another on the millisecond time scale, whereas the neutral guanine radicals decay on time scales of seconds. Alkali-labile guanine lesions are produced and are revealed by treatment of the irradiated oligonucleotides in hot piperidine solution. The DNA fragments thus formed are identified by a standard polyacrylamide gel electrophoresis assay, showing that strand cleavage occurs at the guanine sites only. The biological implications of these oxidative processes are discussed.  相似文献   

20.
The dynamics of Re(2,2'-bipyridine)(CO)3Cl MLCT state formation and decay were determined after femtosecond UV laser excitation and picosecond pulsed X-ray excitation, in an N,N-dimethylformamide (DMF) solution as well as in its solid form. At room temperature, after UV excitation, this MLCT excited state emits both in DMF solution and in the solid form. Transient absorption spectra were measured in solution at various delay times following excitation by a 160 fs, 390 nm laser pulse. There was a prompt absorption increase at around 460 nm occurring within the pump probe convolution (<1 ps), which was assigned to the formation of the 3MLCT state. This transient absorbance was constant over 100 ps. In contrast to the solution state, in the solid state, the emission maximum slightly red-shifts with increasing time after laser excitation. In both solid and solution the emission rises within the system response time. The solid sample exhibited a 1.4 ns emission decay that was not observed for the solution sample. The emission rise from a solid sample after 20 ps pulsed X-ray excitation was significantly slower than the system's time resolution. It is proposed that kinetically energetic electrons are ejected following X-ray induced ionisation, creating ionised tracks in which energetic cations and electrons take time to recombine yielding delayed 3MLCT states that emit.  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