To understand the complex time course of cytosolic Ca2+ signalling evoked by hormones and neurotransmitters, it is necessary to know the kinetics of steps in the second-messenger cascade, particularly cooperative and inhibitory interactions between components that might give rise to periodic fluctuations. In the case of inositol trisphosphate (InsP3)-evoked Ca2+ release, fast perfusion studies with subcellular fractions or permeabilised cells can be made if sufficient homogeneous tissue is available. Single-cell studies can be made by combining whole-cell patch-clamp techniques and microspectrofluorimetry with flash photolytic release of InsP3 to give quantitative, time-resolved data of Ca2+ release from stores. A technical description is given here of flash photolysis of caged InsP3, and the results of fast perfusion and flash photolytic experiments are reviewed. Studies of kinetics of Ca2+ release have shown that the InsP3 receptor/channel is regulated first by positive and then by negative feedback by free cytosolic Ca2+ concentration, producing a pulse of Ca2+ release having properties that may be important in the spatial propagation of Ca2+ signals within and between cells. The properties of InsP3-evoked Ca2+ release in single cells differ between peripheral tissues, such as the liver, and Purkinje neurones of the cerebellum. Purkinje neurones need 20–50 times higher InsP3 concentrations and release Ca2+ to change the free cytosolic concentration 30 times faster and to higher peak concentrations than in liver. The Ins P3 receptors in the two cell types appear to differ in apparent affinity, and the greater Ca2+ efflux from stores in Purkinje cells is probably due to a high receptor density.

There are many tissues in which hormone or neurotransmitter action on certain G-protein-coupled surface receptors results in the release of Ca2+ from intracellular stores by InsP3. The rise in free [Ca2+] activates physiological processes, such as ion channel opening, contraction or secretion, or metabolic processes, such as glycogenolysis. This process has been best described in non-excitable tissues, such as liver or exocrine cells. In general, it has a characteristic time course, showing delays that can be many seconds after receptor activation and periodic fluctuations of the intracellular Ca2+ concentration when hormonal stimulation is maintained. The complex kinetics suggests the presence of both regenerative and inhibitory interactions in the second-messenger reactions that regulate the rise and fall of free Ca2+ concentration. Kinetic methods, including flash photolysis and fast mixing, have been used to identify important interactions in the sequence of reactions. These will be reviewed here.

The existence of a similar system in neuronal tissue is supported by the presence of G-protein-coupled ‘metabotropic’ receptors, the presence of high densities of InsP3 receptors and the demonstration of polyphosphoinositide metabolism and Ca2+ mobilisation in tissue derived from central nervous system (CNS). However, kinetic experiments on uniform neuronal cell populations or single cells are more difficult. As a result, the physiological role, means of activation and properties of InsP3-evoked Ca2+ signalling in neurones are unclear. A comparison will be made here between data from neuronal and non-neuronal tissue.

Flash photolysis of photolabile ‘caged’ precursors is a technique now quite widely used in cell physiology to make rapid, precise changes of substrate or ligand concentration at sites within the cytosol or at cell surface receptors. The information obtained from experiments of this kind can be simply qualitative, for instance gaining evidence of the participation of a ligand such as calcium ions in a cellular process, or can be more quantitative, providing data on mechanism such as equilibrium constants and reaction rates. In the latter case, the time course of the re-equilibration following a concentration jump can give the rates of biochemical steps – ligand binding, conformation changes or ion translocation rates – that constitute the physiological response and can show the site of regenerative or inhibitory steps in the overall process. The technique is most usefully applied at sites within unstirred regions of the extracellular or intracellular space only slowly accessible by diffusion, where perfusion methods would be too slow to study kinetics.

Diffusional delays distort both the time course and steady-state concentration of physiological ligands applied by conventional perfusion methods, because of metabolism, uptake and binding. With flash photolysis, a physiologically inert, photolabile ‘caged’ derivative is allowed to reach diffusional equilibrium before photolysis by a brief light pulse releases a known concentration of free ligand at the site of action. Diffusional delays and metabolic breakdown are minimised and the rate of application and concentration of the ligand are determined by the speed and efficiency of photolysis.

The problem can be illustrated by comparing the action of intracellular InsP3 when perfused from a whole-cell patch pipette with that seen with flash photolysis of caged InsP3. Diffusional equilibration of small ions from the pipette is known to take tens of seconds (Fenwick et al. 1982) and for larger molecules may require many minutes (Pusch and Neher, 1988). The time course of loading the Ca2+ dye Fura-2 into a parotid gland acinar cell by diffusion from a patch pipette is shown in Fig. 1A by the rise of the fluorescence traces at 350 and 390nm excitation. Even in these small cells, several minutes were required to reach a steady level of fluorescence, indicating equilibration of dye concentration between pipette and cytosol, after establishing access to the cytosol (at the beginning of the trace). Applying 10 μmol l-1 InsP3 simultaneously with Fura-2 produced an erratic and small increase of free [Ca2+], as monitored by the ratio of Fura-2 fluorescence signals and shown in the top trace. Loading 1 μmol l-1 ‘caged’ InsP3 with the Fura-2 for 3min (Fig. 1B) followed by release of 0.5 μmol l-1 InsP3 in the cytosol by a 1ms flash, given at the time indicated by the arrow, produced an immediate fast rise of free [Ca2+], followed by a slower decline to baseline levels. The time course of this response is more readily interpreted in terms of the kinetics of InsP3-evoked Ca2+ release from stores and subsequent Ca2+ metabolism than that of Fig. 1A. In this system, InsP3 released by photolysis produced Ca2+ release at concentrations as low as 200nmol l-1, compared with concentrations of 10 μmol l-1 required with perfusion from the pipette (Evans and Marty, 1986; Capiod et al. 1987; Wakui et al. 1989). This suggests that metabolism of InsP3 reduces considerably the concentration reaching the stores from the patch pipette. Similar considerations apply also to the second messengers inositol tetrakisphosphate (InsP4), cyclic AMP and cyclic GMP.

Fig. 1.

InsP3-evoked Ca2+ release in single parotid acinar cell – a comparison of pipette loading InsP3 and flash photolysis of caged InsP3 (see Gray et al. 1988). (A) The fluorescence signals (long-pass filter 490nm) produced by excitation at 350nm and 390nm in the lower two traces and the free Ca2+ concentration calculated from the ratio of fluorescence in the upper trace. On rupturing the membrane under a patch pipette containing 200 μmol l-1 Fura-2 and 10 μmol l-1 InsP3 at the time indicated by the arrow, the rise of fluorescence signals excited at 350 and 390nm shows the time course of dye diffusion into the cell from the pipette. The free [Ca2+] calculated from the ratio of fluorescence is independent of dye concentration. In this cell, a fluctuating small rise of [Ca2+] was evoked by the presence of InsP3, although in most cells tested in this way no obvious change of [Ca2+] occurred that could be attributed to InsP3 in the pipette. (B) Recordings of the same variables from a cell loaded for several minutes with 200 μmol l-1 Fura-2 and 1 μmol l-1 caged InsP3. At the time indicated by the arrow, a 1ms pulse of near-ultraviolet light was applied to release 0.5 μmol l-1 InsP3, resulting in a fast Ca2+ transient.

Fig. 1.

InsP3-evoked Ca2+ release in single parotid acinar cell – a comparison of pipette loading InsP3 and flash photolysis of caged InsP3 (see Gray et al. 1988). (A) The fluorescence signals (long-pass filter 490nm) produced by excitation at 350nm and 390nm in the lower two traces and the free Ca2+ concentration calculated from the ratio of fluorescence in the upper trace. On rupturing the membrane under a patch pipette containing 200 μmol l-1 Fura-2 and 10 μmol l-1 InsP3 at the time indicated by the arrow, the rise of fluorescence signals excited at 350 and 390nm shows the time course of dye diffusion into the cell from the pipette. The free [Ca2+] calculated from the ratio of fluorescence is independent of dye concentration. In this cell, a fluctuating small rise of [Ca2+] was evoked by the presence of InsP3, although in most cells tested in this way no obvious change of [Ca2+] occurred that could be attributed to InsP3 in the pipette. (B) Recordings of the same variables from a cell loaded for several minutes with 200 μmol l-1 Fura-2 and 1 μmol l-1 caged InsP3. At the time indicated by the arrow, a 1ms pulse of near-ultraviolet light was applied to release 0.5 μmol l-1 InsP3, resulting in a fast Ca2+ transient.

Although the principle of flash photolysis is straightforward, there are a number of technical problems which will be reviewed briefly with reference to caged InsP3.

Chemistry of photoreactive groups and photolysis mechanisms

Given that the photolytic yield and rate of release of the ligand are good enough to be useful, it is essential to know the extent to which the caged compound itself and by-products formed during photolytic decay interfere with the physiological system under investigation.

Caged compounds developed as physiological probes since 1978 have 2-nitrobenzyl or 2-nitrophenyl moieties (or close derivatives) as photolabile protecting groups attached covalently to the ligand (see Kaplan et al. 1978; McCray and Trentham, 1989). The reduction in activity of the caged ligand is presumably due to lowered binding affinity, produced either by changing the charge near an important binding group or by some steric effect impeding access to the binding site. Because of the remaining structural similarity, the caged ligand may act as a weak agonist or antagonist at the receptors for the free ligand.

The photochemical properties of caged compounds, particularly the mechanism of photolysis as exemplified by caged ATP, have been reviewed in detail recently by McCray and Trentham (1989). The photolysis reaction proceeds essentially in two steps, absorption of a photon of sufficient energy to generate an activated intermediate (or intermediates), followed by one or more slower dark reactions in which the intermediates decay to release products. On absorbing a photon, the activated intermediate is formed very quickly, on a nanosecond time scale. However, the decay to release products during the dark part of the reaction is relatively slow, in some cases in microseconds, but with some reactions very much slower, for example in 5–10ms with caged InsP3. Whether these compounds are useful depends on the rates of the physiological process being investigated. The efficiency and rate of photolysis depend on the nature of the bond cleaved and the substituent groups on adjacent atoms. Changing these substituents can be used to eliminate residual biological activity and to modify the rate and efficiency of photolysis. Descriptions of the effects of chemical substitutions can be found in reviews by Wootton and Trentham (1989), McCray and Trentham (1989), Nerbonne (1986) and Lester and Nerbonne (1982). The importance of the experimental conditions in photolysis is discussed by McCray and Trentham (1989).

Synthesis and properties of caged InsP3

The synthesis and purification of 2-nitrobenzyl derivatives, the effects of substituents on their properties and problems such as the presence of stereoisomers with differing biological activity are discussed in a methodological review by Walker (1991). In particular, he gives recipes for the synthesis and purification of several compounds, including caged InsP3. The approach used to ‘cage’ ATP, with a 2-nitrophenylethyl ester to protect the terminal phosphate (Walker et al. 1988), was applied by Walker et al. (1989) to generate the three photolabile isomeric 2-nitrophenylethyl phosphate esters of InsP3, substituted on the 1, 4 or 5 phosphates. The isomers were separated by anion-exchange HPLC, characterised with nuclear magnetic resonance and their photochemical and physiological properties evaluated. The rate of the dark reaction was determined from the decay of aci-nitro and the formation of nitrosoacetophenone. The scheme of the photolysis reaction is shown in Fig. 2. At physiological ionic strength and pH (0.2mol l-1 and pH7.1), the P4 and P5 isomers photolysed with rates of 225 s-1 and 280 s-1, respectively (half-times close to 3ms), and with quantum yields Qp of 0.65, similar to the value of 0.63 determined for caged ATP in the same conditions. The interactions of the isomers with the InsP3-evoked Ca2+ release process in smooth muscle and with enzymes involved with InsP3 metabolism were tested. At 10 μmol l-1, the P1 isomer produced maximal tension in permeabilised smooth muscle, indicating Ca2+ release from stores. The P4 and P5 caged isomers were inactive at 50 μmol l-1 in releasing Ca2+ and were ineffective in blocking Ca2+ release evoked by InsP3. Similar results have been obtained for Ca2+ release from stores in isolated guinea pig hepatocytes, in which the caged InsP3 was applied from whole-cell patch-clamp pipettes. The potency of InsP3 released photochemically with low levels of photolysis, so that a high caged InsP3 concentration remained, is similar to that obtained with high levels of photolysis and a low cage concentration. The potency in single hepatocytes with whole-cell perfusion and flash photolysis is similar to that obtained by applying InsP3 to permeabilised hepatocyte suspensions (Burgess et al. 1984). A similar concentration range, 0.2–10 μmoll-1, is required for activation of Ca2+ release in both cases, suggesting that inhibition by the unphotolysed cage was small.

Fig. 2.

Schematic photolysis reaction for caged InsP3. The reaction releases a proton and a by-product, 2-nitrosoacetophenone.

Fig. 2.

Schematic photolysis reaction for caged InsP3. The reaction releases a proton and a by-product, 2-nitrosoacetophenone.

The three isomers were tested by Walker et al. (1989) as substrates and inhibitors of 5-phosphatase and 3-kinase. The P1 isomer was hydrolysed to caged Ins(1,4)P2 by phosphatase and the P4 and P5 isomers were neither substrates nor inhibitors of this enzyme. The P5 isomer inhibited 3-kinase, which phosphorylates InsP3 to InsP4, by about twofold at 10 μmol l-1, and the P4 isomer was inactive on this enzyme. Neither was phosphorylated to InsP4. Thus, the commercially available mixture of P4 and P5 isomers may inhibit 3-kinase. Further, the P1 caged Ins(1,4)P2 has been used as a control for the effects of photolysis, release of the inactive bisphosphate in smooth muscle and liver showing no effect on the hormonal response of the system (Walker et al. 1987; Ogden et al. 1990).

Toxicity of photoproducts

The aromatic nitroso ketones released as by-products are reactive particularly with –SH groups of proteins and are potentially toxic. It is necessary to make control experiments in which the by-products of photolysis are released into the system without the ligand of primary interest. In some instances, custom-made analogues such as caged InsP2 may be available, but compounds such as caged ATP or caged phosphate may be suitable. As a guide, it has been found that 50 μmol l-1 or less of nitrosoacetophenone released into cells appears to have no effect, but higher concentrations approaching 1 mmol l-1 are toxic. The toxicity can be alleviated by including thiols – mercaptoethanol, dithiothreitol (DTT) or glutathione – in the solution to react with nitrosoketones released.

Light sources

The optimal wavelengths for photolysis of nitrophenyl derivatives are in the near ultraviolet, 300–360nm, longer than those that damage proteins or nucleic acids and in a region where cells are relatively transparent. Light flashes in the near ultraviolet with sufficient intensity for flash photolysis can be produced by lasers, flashlamps or shuttered arc lamps. The energy density required is about 50–500mJcm-2 at 300–360nm. The pulse should be short enough, 1ms or less, not to limit the rate of photolysis. Details of light sources and their utilisation have been reviewed by McCray and Trentham (1989).

Xenon arc flashlamps produce pulses of 0.5–1ms duration and energies up to 150mJ at 300–350nm (Schott UG11 or Hoya U 350 filter). Details of the design and performance of a lamp available commercially are given by Rapp and Guth (1988). The lamp arc can be focused directly to an image of about 4mm in diameter at a focal length of 3cm.

Calibration of photolysis

The quantum yield, Qp, of a caged molecule is the proportion of molecules which, having absorbed a photon of light, then break down to release products. To determine Qp requires measurement of the light absorbed as well as the extent of photolysis, and is useful for photochemical comparison. For routine use of flash photolysis, it is important to know simply the percentage conversion of cage to product under experimental conditions. How easy this is to determine depends on the complexity of the experimental arrangement. Two approaches have been used. In the most straightforward case, where the light is focused to a large spot from above, a 5–10 μl drop of solution can be exposed to a flash and the changes estimated chemically, for instance by HPLC for caged phosphates (Ogden et al. 1990; Walker et al. 1988). If photolysis is carried out in combination with microspectrofluorimetry, protons released stoichiometrically during photolysis can be measured directly with a pH-sensitive dye (Walker et al. 1988; Khodakhah and Ogden, 1993). As a guide to the extent of photolysis, caged ATP and caged InsP3 (the 2-nitrophenylethyl esters) are photolysed by about 50% with a 500 m Jcm-2 pulse of 300–360 nm light from a xenon arc flashlamp applied to a 10 μl drop.

Simultaneous flash photolysis and Ca2+ measurement

Experiments in which free [Ca2+] is changed by flash photolysis are designed either to study the effects of [Ca2+] on a physiological variable, such as contractile tension, membrane conductance or secretion, or to study the mobilisation and sequestration of free Ca2+ by the cell. In some instances, the Ca2+-sensitivity of the tension or conductance may be known and provides a measure of the free [Ca2+]. However, it is often necessary to make independent measurements of free [Ca2+] with an optical technique, such as Ca2+-binding dyes, and the use of both simultaneously may give greatly enhanced information about the system. The light pulse can produce direct optical interference in two ways, from saturation of the photomultipliers (or other photosensitive element) during the flash and by a slowly decaying visible phosphorescence induced by the flash in glass components, such as coverslips and objective lenses. The direct effect of the flash can be removed by a fast shutter or by electronically gating the dynodes of the photomultiplier during the flash. The phosphorescence is less easily dealt with and requires a near-ultraviolet filter between the preparation and the microscope optics and the use of quartz components for the chamber (Ogden et al. 1991).

A second form of interference arises from continuous photolysis of caged compounds by excitation light during [Ca2+] measurement with fluorescent dyes. The most widely used Ca2+ probes are the fluorescent dyes Indo-1 and Fura-2, which are excited at near-ultraviolet wavelengths and can be used without appreciable photolysis only with weak excitation intensity. Indo-1 and Fura-2 have the advantage of spectral shifts on binding Ca2+, permitting [Ca2+] measurements independently of dye concentration and geometric factors by the use of the ratio of fluorescence at two wavelengths. However, a weak excitation intensity means, in practice, that time resolution is limited by the need to integrate the emitted light over 100ms or more to achieve an acceptable signal/noise ratio (as in Fig. 1), and the good time resolution possible with flash photolysis is lost. There are three commonly used dyes excited at longer wavelengths, Fluo-3 or Calcium Green at fluoroscein (480nm) and Rhod-2 at rhodamine (520nm) wavelengths, which can be used at high excitation intensity.

Time-resolved measurements of free [Ca2+] require a detection system that is linear over the whole [Ca2+] range of even local transients, which may reach 100 μmoll-1 adjacent to Ca2+ channels for short times (e.g. Chad and Eckert, 1984). Distortions arise when the dye is locally approaching saturation and does not contribute proportionately to the spatial average. The fluorescent dyes commonly used (Fura-2, Indo-1, Fluo-3) have equilibrium dissociation constants near the resting [Ca2+] and therefore give a good signal at low [Ca2+] levels. However, their response characteristics approach saturation at relatively low [Ca2+] and, as a result, transient detection and time resolution is poor. A second problem is that they are extensively bound to Ca2+ even at resting levels, buffering Ca2+ to a greater extent than the native buffers. As a result, they distort spatial aspects of Ca2+ signalling that arise from Ca2+ binding to fixed cytosolic buffers by acting as a mobile Ca2+ carrier and speeding Ca2+ diffusion. A fluorescent dye for which the problems of distortion and buffering are minimal is MagFura-2 (Furaptra) which, when excited at 420nm, is quenched by Ca2+ to almost zero fluorescence, like Fura-2, and so is easily calibrated, but has the advantage of a relatively low Ca2+ affinity, around 50 μmol l-1 (see Konishi et al. 1991).

Origin of the delay in hormone responses in guinea pig hepatocytes and other peripheral tissues

Calcium-mobilising hormones bind to cell surface receptors and set in motion a sequence of reactions comprising activation of a pertussis-insensitive G-protein, phospholipase C, and generation of InsP3, which releases Ca2+ from a component of the endoplasmic reticulum. Hepatocytes are a good ‘model’ system for single-cell studies, partly because they have provided much information about the biochemistry of second-messenger action and also because Ca2+ release from stores is the only means of Ca2+ signalling. In hepatocytes, the Ca2+ has a primary function of activating phosphorylase kinase via calmodulin, producing glycogenolysis and glucose release. A secondary action (in guinea pig and rabbit) is to activate K+-permeable and Cl--permeable ion channels in the surface membrane, resulting in rapid KCl loss from the cell (Burgess et al. 1981). The potassium ion conductance is an apamin-sensitive Ca2+-activated channel that can serve as a monitor for intracellular free Ca2+ concentration in the range 0.3–2.0 μmol l-1 (Capiod and Ogden, 1989b). The whole-cell patch-clamp technique permits both the perfusion of the cell with caged InsP3 and the measurement of the K+ conductance to monitor Ca2+ release. Single-cell microelectrode or patch-clamp recordings have shown that, when noradrenaline acting through a-adrenoceptors is applied rapidly at high concentration to single hepatocytes, the K+ conductance increase follows with a variable delay of minimum duration 2s and up to 90s or more (Field and Jenkinson, 1987; Capiod and Ogden, 1989a; Ogden et al. 1990). Flash photolysis of caged InsP3 has been used in a qualitative fashion to indicate the point in the sequence of reactions where the delay may occur. A number of studies have compared the time courses of responses to Ca2+-mobilising hormones and to InsP3 released by photolysis from caged InsP3. Fig. 3 shows, for comparison, in the same hepatocyte the Ca2+-activated K+ conductance following with 8s delay after application of noradrenaline (upper trace) and flash photolytic release of 0.5 μmol l-1 InsP3, with a delay of 0.3s, in the lower trace. A pulse of InsP3 at moderate concentration, 0.5–1 μmoll-1, initiates a rise of [Ca2+]in 100–200ms in most tissues studied, compared with delays in the onset of the effect of Ca2+-mobilising hormones of 1s or longer (smooth muscle: Walker et al. 1987; hepatocytes: Capiod et al. 1988; Xenopus oocytes: Miledi and Parker, 1989; vascular endothelia: Carter and Ogden, 1992). The delays in Ca2+ mobilisation by hormones are therefore at a step prior to, or during, the generation of InsP3. The delays with hormone vary from one cell type to another, and in some cases have been shown to be shorter at high than at low hormone concentration, even though the ensuing response has a rise time and amplitude that are much the same. It should also be mentioned that responses of this kind have, in some cases, been shown to propagate across and between cells, and, in the case of primary cultures of hepatocytes, that the response originates at the same point each time in a particular cell (Rooney et al. 1990). In exocrine acinar cells, there is evidence that the site of InsP3 production is at the opposite pole of the cell from the region of Ca2+ release (Kasai and Augustine, 1990; see Marty, 1991). Thus, the spatial arrangement of the different elements of the signalling pathway and diffusion of intermediates may contribute to the delay.

Fig. 3.

Comparison of the time course of activation of outward K+ current in a single guinea pig hepatocyte under whole-cell patch-clamp with 10 μmol l-1 external noradenaline (upper trace, delay 8s) and later with photolytic release of 0.5 μmol l-1 InsP3 from 1 μmol l-1 caged InsP3 (lower trace, delay 300ms). Membrane potential 0mV. Cl--free solutions.

Fig. 3.

Comparison of the time course of activation of outward K+ current in a single guinea pig hepatocyte under whole-cell patch-clamp with 10 μmol l-1 external noradenaline (upper trace, delay 8s) and later with photolytic release of 0.5 μmol l-1 InsP3 from 1 μmol l-1 caged InsP3 (lower trace, delay 300ms). Membrane potential 0mV. Cl--free solutions.

It is possible that the delay is due to the time taken to accumulate sufficient local concentration of a mediator, InsP3 or Ca2+ or both, to set off a regenerative reaction sequence (Marty et al. 1989; Parker and Miledi, 1989; Crossley et al. 1991). This idea could be tested by release of InsP3 by photolysis during the delay in hormone action to see whether the response can be triggered. Attempts to do this have shown no ‘triggering’ action of InsP3 in hepatocytes (T. Capiod and D. Ogden, unpublished observations) but it must be remembered that InsP3 is released by this means throughout the cell and, if a spatial element is important, experiments with photolysis restricted to small areas of the cell would be needed.

The GDP/GTP exchange reaction and dissociation of subunits of the G-proteins coupling receptors to phospholipase C (PLC) are also thought to be slow (rate 0.03min-1, Pang and Sternweiss, 1990) but catalysed by the activation of hormone receptors. Caged GTP-γ-S is potentially a useful agent to activate G-protein independently of hormonal stimulation, as has been done with G-protein interactions with Ca2+ channels (Dolphin et al. 1988). Experiments with a caged 2-nitrophenylethyl ester of GTP-γ-S show first a delay of several seconds in activation of Ca2+ release and, second, that this delay is not reduced by application of noradrenaline, which would accelerate GTP/GDP exchange. On the basis of this observation, the delay would appear to be after G-protein activation and before InsP3 action on stores, i.e. during production of Ins P3 by PLC.

Mechanism of InsP3-evoked Ca2+ release

Flash photolysis of caged InsP3 introduced into single hepatocytes or other cells during whole-cell patch-clamp can give information on the concentration-dependence, kinetics and regulatory interactions during InsP3 action that would otherwise only be obtained with fast mixing experiments in permeabilised cells or subcellular fractions. Microinjection of InsP3 itself or caged InsP3 followed by photolysis does not give information on the concentration applied because of the difficulty of estimating the quantities injected and the cell volume.

The simplest hypothesis of how the Ca2+ release might occur is gating by cytosolic InsP3 of an ion channel in the store-cytoplasm membrane. In this case, the rate of Ca2+ flux into the cytosol will be proportional to the open probability of the channels. The rate of change of Ca 2+ concentration in the cytosol, d[Ca2+]/dt, is proportional to the net flux of Ca2+ and provides the best measure of open probability of the InsP3 channels within each cell. However, it may be expected to vary from cell to cell, not least because of geometric factors such as the ratio of InsP3 receptor density to cytosolic volume.

Initial delays

Experiments with caged InsP3 and the Ca2+ dye Fluo-3 in permeabilised smooth muscle (Somlyo et al. 1992), in hepatocytes (Ogden et al. 1990, 1991) and in vascular endothelial cells (Carter and Ogden, 1992; see Fig. 4) have shown a delay in the activation of Ca2+ efflux of several hundred milliseconds at low InsP3 concentration (0.1–0.5 μmol l-1), which becomes much shorter, less than 20ms, at high concentration (10–100 μmol l-1). The short delay at high concentration is not much longer than the time required for photolysis and suggests direct activation of the Ca2+ efflux channel by InsP3. The concentration range of InsP3 activation of receptors is given by the range over which the delays were observed to become shorter, from about 0.2 to 5 μmol l-1. These data from photolytic release of InsP3 in single hepatocytes and other tissues agree with data from rapid mixing experiments with permeabilised hepatocytes (Champeil et al. 1989). In permeabilised rat basophilic leukaemia cells, a similar reduction of the delay at high concentration was found with rapid mixing experiments, but in this case over a lower concentration range and with a longer delay (65ms) at high concentration (Meyer et al. 1990). In all tissues tested with both photolytic and fast mixing experiments, the speed of activation at high concentration is consistent with a directly gated channel, being much faster, for instance, than rates of activation of G-protein-coupled channels in the surface membrane (Hartzell et al. 1991).

Fig. 4.

Fluo-3 fluorescence detection of Ca2+ release from stores in porcine cultured aortic endothelial cells with whole-cell patch-clamp to show the delay after photolysis of caged InsP3 and the rate of rise of fluorescence. (A,B) Release of 0.2 and 0.35μmol l-1 InsP3 at the time indicated by the arrow; sequential records from same cell. (C,D) Release of 0.8 and 1.4 μmol l-1 InsP3 in the same cell. Membrane potential -44mV. The right-hand traces are expanded regions of the left-hand traces. From Carter and Ogden (1992).

Fig. 4.

Fluo-3 fluorescence detection of Ca2+ release from stores in porcine cultured aortic endothelial cells with whole-cell patch-clamp to show the delay after photolysis of caged InsP3 and the rate of rise of fluorescence. (A,B) Release of 0.2 and 0.35μmol l-1 InsP3 at the time indicated by the arrow; sequential records from same cell. (C,D) Release of 0.8 and 1.4 μmol l-1 InsP3 in the same cell. Membrane potential -44mV. The right-hand traces are expanded regions of the left-hand traces. From Carter and Ogden (1992).

The delay at low InsP3 concentrations is not predicted for simple models of ligand gating, and two general explanations have been proposed. The first is a high (n=4) cooperativity in InsP3 binding seen in permeabilised rat basophilic leukaemia cells (Meyer et al. 1990). In permeabilised smooth muscle, the delay is consistent with n=1 or n=2 (Somlyo et al. 1992). The second explanation is a cooperative or co-agonist effect between InsP3 and Ca2+ at low free Ca2+ concentrations. This idea is supported by observations of the Ca2+-dependence of InsP3 action in smooth muscle (Iino, 1990; Iino and Endo, 1992), brain microsomes (Finch et al. 1991) and reconstituted InsP3 receptor (Bezprozvanny et al. 1991). Although a co-agonist action of Ca2+ and InsP3 can explain the delay quite well, direct experimental tests are difficult. The experiments of Iino and Endo (1992) show that, at strongly buffered [Ca2+] of 300nmol l-1, the delay to photoreleased InsP3 is removed and the sigmoid initial rise of [Ca2+] is absent and is replaced by a single exponential rise, consistent with loss of the cooperative action of Ca2+.

To summarise, the presence of a delay suggests cooperativity in InsP3 action, as multiple InsP3 binding and/or a co-agonist action of Ca2+. The very short delays at high InsP3 concentration are consistent with a direct ligand-gated mode of channel activation.

The concentration range of InsP3 action, from longest to shortest delay, is 0.1–5 μmol l-1 in the peripheral tissues studied.

Rate of rise of [Ca2+]

To study channel activation as a function of InsP3 concentration or on changing other variables, the maximal rates of rise of free [Ca2+] should be measured in conditions where other Ca2+ fluxes into or out from the cytosol are small. In whole-cell recording from guinea pig liver cells, the Ca2+-activated K+ conductance of the surface membrane provides a good measure of the submembrane free [Ca2+], with a steep Ca2+-dependence in the range 0.3–2 μmol l-1 Ca2+. The rate of rise of the conductance increased with InsP3 concentration between 0.5 and 10 μmol l-1 when compared within the same cell (Ogden et al. 1990). These results have been confirmed with the Ca2+ indicator dye Fluo-3 (Ogden et al. 1991). In endothelial cells, an increase in the rate of rise of the Fluo-3 signal was found in the range 0.2–5μmol l-1 InsP3 (Carter and Ogden, 1992). A dependence of the rate of change of the Fluo-3 fluorescence on InsP3 concentration over a similar range has been found in permeabilised smooth muscle (Somlyo et al. 1992).

It is of interest to know the slope of the relationship between d[Ca2+]/dt and log [InsP3]. The Hill coefficient, n, in principle, gives a lower-limit estimate of the number of InsP3 molecules binding to produce activation. In single intact cells, data with Fluo-3 have proved quantitatively unreliable because of dye binding to proteins and because the high affinity for Ca2+ will underestimate the rate of change of free [Ca2+], especially if there is spatial localisation of the [Ca2+] rise to regions near the release sites. Data on the efflux/InsP3 concentration relationship have come mainly from fast mixing experiments. A high potency of InsP3, active at 4nmol l-1, and a large Hill slope of 3–4 at low concentrations, suggesting high cooperativity, were found in permeabilised rat basophilic leukaemia cells (Meyer et al. 1988). In permeabilised liver cells (Champeil et al. 1989), the maximum rates of rise of [Ca2+] measured with Quin-2 increased with concentration in the range 0.1–10 μmol l-1 with a Hill coefficient of 1.6, indicating a small degree of cooperativity. The cooperativity seen in hepatocyte suspensions was reduced at high (10 μmol l-1) [Ca2+] (Combettes et al. 1993). In an InsP3-sensitive microsomal preparation from the CNS, the concentration range measured by 45Ca2+ efflux was 30nmol l-1 to 10 μmol l-1 with a Hill coefficient of 1.0, under conditions of constant buffered free Ca2+ concentration (Finch et al. 1991). Similar results were found with single-channel recording from InsP3 receptors isolated from cerebellum and incorporated into bilayers (Bezprozvanny et al. 1991). The maximum InsP3-evoked open probability, of up to 15% at 2 μmol l-1, occurred at a free [Ca2+] of 0.3 μmoll-1 and declined at higher [Ca2+]. In smooth muscle, a Hill coefficient of 1–2 was reported by Somlyo et al. (1992) and of 2 by Iino and Endo (1992), both studies performed using flash photolysis of caged InsP3 and with Fluo-3 as the Ca2+ indicator in permeabilised strips.

The strong dependence of InsP3 potency on the free [Ca2+] discussed above suggests that the discrepancies between different studies, particularly in cell suspensions, may result from local Ca2+ accumulation. The data of Iino and Endo (1992), which gave n=2, were carried out with strongly buffered Ca2+ in order to avoid the influence of variable

[Ca2+] on the InsP3 cooperativity estimated, and were measured at the foot of the InsP3 concentration range. It would appear that the value of the Hill coefficient is between 1 and 2, suggesting that at least two InsP3 molecules need to bind to produce channel opening, but some degree of uncertainty remains about the value and significance of this number. An analogy may be made with studies of the nicotinic receptor, where good measurements of the Hill coefficient were difficult to make, even in this more accessible preparation, and in fact the firmest evidence for two bindings in activation was from the biochemical demonstration of two binding sites in each receptor oligomer.

Threshold and quantal effects

There are a number of reports indicating that regenerative phenomena occur during InsP3 action. Parker and Ivorra (1990a, 1993) found that increasing InsP3 concentration by incremental photolysis of caged InsP3 showed a ‘threshold’ concentration for activation of Ca2+-dependent Cl- conductance and for the peak [Ca 2+] estimated by dye in Xenopus oocytes. In these experiments, [Ca2+] was monitored with Rhod-2 fluorescence confocally within the oocyte from a spot of a few micrometres in diameter, to avoid spatial averaging of the signal from a large number of Ca2+ storage elements that may have had different sensitivities. InsP3 concentrations were generated by increasing the duration or transmission of a steady ultraviolet irradiation of caged InsP3 microinjected into the oocyte, but the absolute concentrations were not known. Above threshold, the amplitude of the Ca2+ fluorescence was constant with increasing InsP3 concentration, indicating a constant ‘quantal’ Ca2+ release. The rate of rise of the Ca2+ fluorescence signal was greater as InsP3 concentration was increased, suggesting that the rate of Ca2+ efflux from stores, and therefore receptor activation, increased with concentration even though the peak did not. The data therefore suggest that some process quickly terminated Ca2+ efflux so that the final concentration of Ca2+ did not increase much at high concentration (see discussion below). The advantage of Parker and Ivorra’s approach is the confocal localisation of the observations to a small region of the oocyte’s Ca2+ store, but observations of conserved ‘quantal’ Ca2+ release have been reported in other cells in response to thio-InsP3 or hormonal stimulation (Muallem et al. 1989; Taylor and Potter, 1990). The reasons for the threshold and ‘quantal’ behaviour are not known, but they may be linked with Ca2+ regulation of InsP3 sensitivity.

Cooperative and inhibitory interactions of Ca2+ and InsP3

The complex kinetic behaviour of the response of many cell types to Ca2+-mobilising hormones indicates the presence in the second-messenger system of steps that are regenerative or highly cooperative, and also of inhibitory mechanisms that restore the system. Some mechanisms of this kind have been shown to occur in the InsP3-evoked Ca2+ release step. First, it is well documented that, immediately following InsP3-evoked Ca2+ release, the release mechanism becomes refractory to further InsP3 action and that recovery of sensitivity has a time course of tens of seconds. Fig. 5 shows data from experiments in which consecutive twin pulses of InsP3 were released photolytically in a hepatocyte. The sensitivity of the Ca2+ release mechanism to InsP3, measured by the rate of rise of the Ca2+-activated conductance, is suppressed to zero immediately after the first pulse and recovers with a half-time of 10–20s as the pulse interval is increased. It occurs in Limulus photoreceptors (Payne et al. 1988, 1990), guinea pig hepatocytes (Ogden et al. 1990), Xenopus oocytes (Parker and Ivorra, 1990b), brain synaptosomes (Finch et al. 1991), rat hepatocytes (Combettes et al. 1992, 1993) and in single-channel recordings from cerebellar InsP3 receptors in bilayers (Bezprozvanny et al. 1991). It has been shown to be due to raised cytosolic [Ca2+] and is attributed to an action directly on the channel or with a closely associated protein. This mechanism terminates InsP3-evoked Ca2+ release in guinea pig hepatocytes (and most probably in other tissues). It has a rapid onset (Finch et al. 1991; Iino and Endo, 1992; Levitan et al. 1993) and may be important in the declining phase of hormonally induced Ca2+ spikes and in conserving stored Ca2+. Estimates of the Ca2+ concentration needed to produce half-inhibition in the steady state are 0.5–1 μmol l-1 in brain microsomes (Finch et al. 1991) and 3–15 μmol l-1 in permeabilised hepatocytes (Combettes et al. 1993). The onset of inhibition has been reported to have a time constant of 580ms (brain microsomes, Finch et al. 1991). An immediate inhibition was reported by flash-released Ca2+ in smooth muscle (Iino and Endo, 1992) and complete inhibition within 200ms in Limulus photoreceptors (Levitan et al. 1993). The time course of recovery from inhibition is 10–50s. It is likely that the inhibition characterised by Ca2+ release measurements in cells has a counterpart in a Ca2+-induced increase of InsP3 binding affinity, coupled with a Ca2+-induced desensitisation of the InsP3 receptor, described in hepatocyte microsomes by Pietri et al. (1990).

Fig. 5.

Recovery of InsP3 sensitivity in twin-pulse experiments in guinea pig hepatocytes. (A) Pulses of 1 μmol l-1 InsP3 released at the arrows with an interval of 10s, a recovery period of 100s, then an interval of 5s, same cell. Top traces in each panel are Ca2+-activated K+ current; lower traces show the rate of rise of current, which measures d[Ca2+]/dt. (B) Plot of the rate of rise of the second response relative to that of the first to show the time course of recovery of InsP3 sensitivity. Different symbols show results from different hepatocytes. From Ogden et al. (1990).

Fig. 5.

Recovery of InsP3 sensitivity in twin-pulse experiments in guinea pig hepatocytes. (A) Pulses of 1 μmol l-1 InsP3 released at the arrows with an interval of 10s, a recovery period of 100s, then an interval of 5s, same cell. Top traces in each panel are Ca2+-activated K+ current; lower traces show the rate of rise of current, which measures d[Ca2+]/dt. (B) Plot of the rate of rise of the second response relative to that of the first to show the time course of recovery of InsP3 sensitivity. Different symbols show results from different hepatocytes. From Ogden et al. (1990).

An inhibitory effect of the decline in Ca2+ concentration within the store has been proposed as a regulatory mechanism (Irvine, 1990) but experimental tests do not support an important role for this mechanism (Combettes et al. 1992; Shuttleworth, 1992).

A positive interaction has been demonstrated between InsP3 and cytosolic Ca2+ in preparations from CNS and smooth muscle. Finch et al. (1991) described a strong co-agonist action of Ca2+ and InsP3 with an apparent equilibrium constant for Ca2+ of 0.66 μmol l-1 and a Hill coefficient for InsP3 action of 1 at constant [Ca2+].

Bezprovzanny et al. (1991) described an increase in open probability of cerebellar InsP3 channels at Ca2+ concentrations up to 0.5 μmol l-1. In smooth muscle, Iino (1990) first described an increased rate of InsP3-induced Ca2+ release at a free [Ca2+] up to 0.3 μmoll-1 and showed subsequently that, if [Ca2+] was suddenly increased to about 0.3 μmoll-1, the rate of InsP3-induced efflux was immediately increased (Iino and Endo, 1992; see also Finch et al. 1991; Combette et al. 1993). Thus, InsP3-evoked Ca2+ release has the elements – positive feedback by Ca2+ to increase Ca2+ efflux – that could produce a regenerative component to the response. There is, however, no direct evidence that this occurs in intact cells, particularly that Ca2+ applied in the presence of InsP3 can initiate Ca2+ efflux from stores. This is a point that needs to be tested.

Role of Ca2+ feedback

The calcium ions which accumulate in the cytosol adjacent to sites of Ca2+ efflux would initially increase InsP3-evoked release and, at later times, inhibit release by the mechanism described above, which comes into play at higher Ca2+ concentrations. The result would be a pulse of Ca2+ released into the cytosol. Given that Ca2+ is heavily buffered by immobile binding sites to an extent of 50–100 bound/free (see, for example, Neher and Augustine, 1992), Ca2+ diffusion will be considerably slowed from that in free solution by about the same proportion. A pulse of Ca2+ released at the store will diffuse slowly, on a time scale of 50–100ms over 1 μm. Therefore, conditions exist in which the Ca2+ concentration adjacent to the receptor could rise to high levels as Ca2+ permeates through the channel. Conditions like this are thought to exist in the cytoplasm adjacent to Ca2+ channels in the surface membrane. Calculations based on the dimensions of a channel with adjacent buffering of Ca2+ indicate that a moderate Ca2+ flux, corresponding to a channel current of about 0.5pA, will quickly, in 1–2ms, produce a large Ca2+ concentration of 10 μmol l-1 or higher within 0.1 μm of the channel (see, for example, Chad and Eckert, 1984), sufficient in this case for inhibition of the InsP3-evoked efflux (50% at 3–15 μmol l-1, Combettes et al. 1993). The brief, local high pulse of Ca2+ concentration will diffuse into the adjacent cytosol, becoming diluted over distances of about 1 μm in 100ms. On reaching adjacent receptor channels, the Ca2+ will contribute to their activation by the InsP3 present. It is easy to envisage this cycle occurring at each receptor/channel and producing a spatially propagated wave, possibly locally restricted if channel open time were relatively brief and if the time course were dominated by the diffusion of Ca2+. Control of the cycle could be, in some circumstances, by local Ca2+ buffering and by the channel density.

Mechanism of periodic fluctuations of [Ca2+]

Hormonal stimulation of non-excitable cells often results in repetitive fluctuations of [Ca2+], which may be periodic spikes every 20s to 4–5min, as shown for a single guinea pig hepatocyte with 10 μmol l-1 noradrenaline in Fig. 6B, or disorganised fluctuations about a plateau level, as shown for a different hepatocyte in Fig. 6A (for reviews, see Berridge, 1990; Jacob, 1990; Petersen and Wakui, 1990; Cobbold et al. 1991; Marty, 1991). Models for systems with periodic fluctuations in the concentration of reactants require one or more autocatalytic or cooperative steps, to provide the rise in concentration, and an inhibitory step, or exhaustion of a reactant, to account for the fall. Two approaches can be made experimentally with flash photolysis, one to isolate regenerative and inhibitory steps under controlled, non-fluctuating conditions, the other to probe the sensitivity to intermediates such as InsP3 during different parts of the cycle. The fluctuations of [Ca2+] seen in the continuous presence of InsP3 or its stable thio derivative in several cell types (Evans and Marty, 1986; Capiod et al. 1987; Wakui et al. 1989; Ogden et al. 1990; Payne and Potter, 1991) indicate that regenerative and inhibitory elements are present even in the simplest case of InsP3 interaction with the Ca2+ store. Taking the first approach, the cooperative and inhibitory interactions of Ca2+ with InsP3, described above for measurements under stable conditions, could form the basis of the simple type of InsP3-evoked Ca2+ fluctuations with periods of 10–30s. However, it is clear that there are three or more different patterns of Ca2+ fluctuations or spiking that occur with hormonal stimulation and may involve additional steps, such as interactions between the receptor/G-protein/PLC and the diacylglycerol/protein kinase C/Ca2+ pathways (Cobbold et al. 1991; Llano and Marty, 1987; Sanchez-Bueno et al. 1990). In slowly spiking systems, the interval between spikes, of 1min or more, is too long to be accounted for by the Ca2+ inhibition of InsP3 action discussed above, which has a half-time for recovery of 10–20s. Taking the second approach, photolytic release of InsP3 during the interspike interval in slowly spiking REF52 fibroblastic cells produced a sharp rise in [Ca2+], confirming in this case that the release process was not inhibited and that cytosolic InsP3 was below the concentration required to initiate Ca2+ release (Harootunian et al. 1988). Although this would appear to be a good experimental approach, systematic studies with caged InsP3 or with caged Ca2+ chelators to test the sensitivity to InsP3 and [Ca2+] during different parts of the cycle have yet to be made.

Fig. 6.

Fluctuations of intracellular Ca2+ concentration in a guinea pig hepatocyte evoked by 10 μmol l-1 noradrenaline acting via a-adrenoceptors. Ca2+ concentration was monitored by Ca2+-dependent K+ conductance with whole-cell patch-clamp at 0mV, Cl--free solutions. (A) Disorganised fluctuations of Ca2+ concentration; (B) regular spiking of Ca2+ concentration in a different cell.

Fig. 6.

Fluctuations of intracellular Ca2+ concentration in a guinea pig hepatocyte evoked by 10 μmol l-1 noradrenaline acting via a-adrenoceptors. Ca2+ concentration was monitored by Ca2+-dependent K+ conductance with whole-cell patch-clamp at 0mV, Cl--free solutions. (A) Disorganised fluctuations of Ca2+ concentration; (B) regular spiking of Ca2+ concentration in a different cell.

InsP3-evoked Ca2+ release in neurones and glia

Glial cells in culture, both astrocytes and Schwann cells, respond to neurotransmitters with Ca2+ mobilisation from intracellular stores. The response may be evoked with excitatory amino acids, monoamines, peptides, ATP and cholinergic ligands. In more physiological conditions, propagated Ca2+ signalling has been observed among glia in organotypic cultures of hippocampal slices following stimulation of input pathways (Dani et al. 1992). Experiments with caged InsP3 in cultured cerebellar astrocytes show Ca2+ mobilisation with characteristics similar to those seen in nonexcitable cells; the concentration range is 0.2–5 μmol l-1 and responses are transient, lasting 5–20s, and show delays of up to 200ms at low concentrations (Khodakhah and Ogden, 1993). The role of Ca2+ mobilisation in glial physiology has not been demonstrated but, by analogy with other cells, it may play a part in the regulation of ion fluxes and volume, via Ca2+-activated channels, and in the regulation of metabolic processes such as glycogenolysis. It is difficult to record from glia in situ, and the significance of data in culture is unclear since it is likely that glial characteristics, such as the sensitivity to neurotransmitters and the role of second-messenger systems, change in culture.

In neurones there is evidence of Ca2+ mobilisation by neurotransmitters that is independent of extracellular Ca2+ (dorsal root ganglion cells, Murphy and Miller, 1988; cerebellar Purkinje cells, Llano et al. 1991). The kinetics of InsP3-evoked Ca2+ release has been studied with experiments in Purkinje neurones in thin cerebellar slices. A large, rapid Ca2+ mobilisation from stores was found with photolytic release of InsP3. The properties of Ca2+ release in Purkinje cells differ from those in peripheral tissues, described above. The InsP3 concentrations needed are in the region of 10 μmol l-1 or higher, 20 times higher than those needed in astrocytes and other nonexcitable cells, with maximal efflux still not attained at 80 μmoll-1 InsP3 (Khodakhah and Ogden, 1993). The overall characteristics are similar, but with shorter delays of up to 70ms at a low concentration, 10 μmol l-1, of InsP3, this being reduced to less than 20ms at concentrations above 33 μmol l-1. The fluorescence increase was very large compared with the Ca2+-induced fluorescence produced in the same cell either by depolarisation in voltage-clamp or with action potential discharge. The rate of rise of the Ca2+ concentration is much greater, up to 1mmol l-1 s-1, more than 30 times the maximum seen in hepatocytes, and the peak levels of Ca2+ concentration are also much higher. The decline of the Ca2+ transient is quite fast in most cells, the response to a pulse of InsP3 lasting a few seconds as in other tissues. Examples of the Ca2+ transients evoked by three concentrations of InsP3 in a Purkinje cell are shown in Fig. 7. It was essential to monitor Ca2+ concentration with the low-affinity dye Furaptra (Magfura-2) in whole-cell patch-clamp. The basis for the difference in properties between liver and Purkinje cell InsP3 receptors may be in the isoform of receptor present, suggesting that the type 1 isoform present in Purkinje cells may have a lower InsP3 affinity. The density of receptors per unit cytosolic volume, which is very high in Purkinje cells, would result in a high d[Ca2+]/dt compared with peripheral tissues.

Fig. 7.

Free Ca2+ concentration in a Purkinje cell soma in a rat cerebellar slice. Whole-cell patch-clamp, membrane potential -70mV, 500 μmol l-1 Furaptra. Flash photolysis of caged InsP3 to release 11 μmol l-1, 19 μmol l-1 and 38 μmol l-1 InsP3 into the soma. From Khodakhah and Ogden (1993).

Fig. 7.

Free Ca2+ concentration in a Purkinje cell soma in a rat cerebellar slice. Whole-cell patch-clamp, membrane potential -70mV, 500 μmol l-1 Furaptra. Flash photolysis of caged InsP3 to release 11 μmol l-1, 19 μmol l-1 and 38 μmol l-1 InsP3 into the soma. From Khodakhah and Ogden (1993).

The physiological role of ‘metabotropic’ Ca2+ release from stores in neurones is difficult to demonstrate, but Ca2+ release by InsP3 will result in short-term changes in excitability due to activation of Ca2+-gated K+ channels (Khodakhah and Ogden, 1993), and possibly in slower changes, such as long-term depression in cerebellar Purkinje cells. The reason for the lower apparent affinity of InsP3 receptors and the much faster and larger Ca2+ mobilisation by InsP3 in Purkinje cell may be related to a synaptic rather than hormonal function of the phosphoinositide system.

Ca2+ influx during Ca2+ mobilisation

The rise of cytosolic [Ca2+] during hormonal Ca2+ mobilisation has been shown, in many cell types, to have an initial phase independent of the presence of external Ca2+. This phase is therefore thought to be due entirely to release from Ca2+ stores. A later phase of the response, in which the [Ca2+] often shows periodic fluctuations dependent on the presence of external Ca2+, is thought to require Ca2+ influx (see reviews by Jacob, 1990; Marty, 1991). Current hypotheses concerning the regulation of this influx invoke InsP3, InsP4 or a combination of the two, possibly acting via Ca2+ stores either by means of a ligand-regulated connection of the store to the external solution or by depletion of the store itself, which then activates refilling from the external solution. Evidence has been obtained for InsP3-induced Ca2+ influx, by both 1,4,5-and 2,4,5-isomers alone (e.g. Kuno and Gardner, 1987; Parker and Miledi, 1987; Matthews et al. 1989), for influx requiring both InsP3 and InsP4 (e.g. Morris et al. 1987) and for influx requiring InsP4 alone (Luckhoff and Clapham, 1992). The problems of access and metabolism encountered with InsP3 applied from the patch pipette also occur with InsP4, but a ‘caged’ InsP4 is not available to obtain evidence of its role or of the time course and concentrations required.

Ca2+ influx current via store depletion (Hoth and Penner, 1992) has been demonstrated in several tissues and appears to require stringent buffering of intracellular [Ca2+] and drastic depletion of stores compared with that expected with physiological stimulation. Photolysis of caged InsP3 has been used to show a Ca2+ influx in T-cells with milder conditions (McDonald et al. 1993). The time course of Ca2+ influx suggested that it may be via the stores, but similar experiments with caged InsP3 in other systems have not been reported. The question of whether Ca2+ influx is via the stores or directly through the surface membrane in physiological conditions has not been satisfactorily resolved.

Caged cyclic AMP and cyclic GMP

The role of cyclic AMP as a second messenger in cytosolic Ca2+ regulation has been investigated with flash photolysis in two systems: enhanced Ca2+ influx through non-inactivating high-threshold Ca2+ channels in cardiac muscle; and Ca2+ release from stores in hepatocytes. Two derivatives of cyclic AMP have been used as caged precursors, both as esters of the phosphate: the 4,5-dimethoxy-2-nitrobenzyl ester and the 2-nitrophenylethyl ester. The photochemistry of the caged cyclic nucleotides is discussed by Wootton and Trentham (1989). The 2-nitrophenylethyl esters photolyse slowly, with rates of about 5 s-1 at pH7 and 100mmol l-1 salt, and fairly efficiently, with a Qp about

0.4. The 4,5-dimethoxy-2-nitrobenzyl esters photolyse more rapidly but with lower efficiencies. Caged cyclic nucleotides are lipophilic and may partition within cell membranes, where photolysis rates and efficiencies may differ from those in the cytosol. Photolytic release of cyclic AMP increased the amplitude and the rate of decay of Ca2+ current in heart, probably through phosphorylation via protein kinase A (Nerbonne et al. 1984; Charnet et al. 1991). It may be noted that cyclic AMP released in the cytosol acts with little delay, unlike a-adrenoceptor stimulation by noradrenaline, which augments Ca2+ current after a delay of several seconds.

The Ca2+-mobilising actions via adenylyl cyclase of β-receptor stimulation in hepatocytes are less well documented than those via InsP3. This pathway can produce Ca2+ mobilisation alone in a smaller proportion of cells, often after very long delays of 1 min or more (Capiod et al. 1991). Cyclic AMP released from either cage mobilises Ca2+ with an average delay of 3s. The cyclic AMP mechanism is insensitive to heparin, providing evidence that InsP3 is not involved by an indirect route (Noel and Capiod, 1991).

The authors’ work was supported by the MRC, INSERM, The British Heart Foundation and Glaxo PLC.

Berridge
,
M. J.
(
1990
).
Calcium oscillations
.
J. biol. Chem
.
265
,
9583
9586
.
Bezprozvanny
,
I.
,
Watras
,
J.
and
Ehrlich
,
B. E.
(
1991
).
Bell-shaped calcium-response curves of Ins(1,4,5)P3-gated and calcium-gated channels from endoplasmic reticulum of cerebellum
.
Nature
351
,
751
754
.
Burgess
,
G. M.
,
Claret
,
M.
and
Jenkinson
,
D. H.
(
1981
).
Effects of quinine and apamin on the calcium dependent potassium permeability of mammalian hepatocytes and red cells
.
J. Physiol., Lond.
317
,
67
90
.
Burgess
,
G. M.
,
Irvine
,
R. F.
,
Berridge
,
M. J.
,
McKinney
,
J. S.
and
Putney
,
J. W.
(
1984
).
Actions of inositol phosphates on Ca2+ pools in guinea-pig hepatocytes
.
Biochem. J.
224
,
741
746
.
Capiod
,
T.
,
Field
,
A. C.
,
Ogden
,
D. C.
and
Sandford
,
C. A.
(
1987
).
Internal perfusion of guinea-pig hepatocytes with buffered calcium ion or inositol 1,4,5-trisphosphate activates potassium and chloride conductances
.
FEBS Lett.
217
,
247
252
.
Capiod
,
T.
,
Noel
,
J.
,
Combettes
,
L.
and
Claret
,
M.
(
1991
).
Cyclic AMP-evoked oscillations of intracellular [Ca2+] in guinea-pig hepatocytes
.
Biochem. J.
275
,
277
280
.
Capiod
,
T.
and
Ogden
,
D. C.
(
1989a
).
Properties of membrane ion conductances evoked by hormonal stimulation of guinea pig and rabbit isolated hepatocytes
.
Proc. R. Soc. Lond. B
236
,
187
201
Capiod
,
T.
and
Ogden
,
D. C.
(
1989b
).
The properties of calcium activated potassium ion channels in guinea pig isolated hepatocytes
.
J. Physiol., Lond.
409
,
285
295
.
Capiod
,
T.
,
Ogden
,
D. C.
,
Trentham
,
D. R.
and
Walker
,
J. W.
(
1988
).
Activation of membrane conductance in isolated guinea pig hepatocytes by photochemically generated intracellular InsP3 of free calcium ions
.
J. Physiol., Lond.
406
,
124P
.
Carter
,
T. D.
and
Ogden
,
D. C.
(
1992
).
Kinetics of intracellular calcium release by InsP3 and extracellular ATP in porcine cultured endothelial cells
.
Proc. R. Soc. Lond. B
250
,
235
241
.
Chad
,
J. E.
and
Eckert
,
R. O.
(
1984
).
Calcium domains associated with individual channels can account for anomalous voltage relations of Ca-dependent responses
.
Biophys. J.
45
,
993
996
.
Champeil
,
P.
,
Combettes
,
L.
,
Berthon
,
B.
,
Doucet
,
E.
,
Orlowski
,
S.
and
Claret
,
M.
(
1989
).
Fast kinetics of calcium release induced by myo-inositol trisphosphate in permeabilized rat hepatocytes
.
J. biol. Chem
.
264
,
17665
17673
.
Charnet
,
P.
,
Richard
,
S.
,
Gurney
,
A. M.
,
Ouadid
,
H.
,
Tiaho
,
F.
and
Nargeot
,
J.
(
1991
).
Modulation of Ca2+ currents in isolated frog atrial cells studied with photosensitive probes. Regulation by cyclic AMP and Ca2+: a common pathway?
J. molec. cell. Cardiol
.
23
,
343
356
.
Cobbold
,
P. H.
,
Sanchez-Bueno
,
A.
and
Dixon
,
C. J.
(
1991
).
The hepatocyte calcium oscillator
.
Cell Calcium
12
,
87
95
.
Combettes
,
L.
,
Claret
,
M.
and
Champeil
,
P.
(
1992
).
Do submaximal InsP3 concentrations only induce the partial discharge of permeabilized hepatocyte calcium pools because of the concomitant reduction of intraluminal concentration?
Febs Lett.
301
,
287
290
.
Combettes
,
L.
,
Claret
,
M.
and
Champeil
,
P.
(
1993
).
Calcium control on InsP3-induced discharge of calcium from permeabilised hepatocyte pools
.
Cell Calcium
14
,
279
292
.
Crossley
,
I.
,
Whalley
,
T.
and
Whitaker
,
M.
(
1991
).
Guanosine 5-thiotriphosphate may stimulate phosphoinositide messenger production in sea urchin eggs by a different route than the fertilizing sperm
.
Cell Regulation
2
,
121
133
.
Dani
,
J. W.
,
Chernjavsky
,
A.
and
Smith
,
S. J.
(
1992
).
Neuronal activity triggers calcium waves in hippocampal astrocyte networks
.
Neuron
8
,
429
440
.
Dolphin
,
A. C.
,
Wootton
,
J. F.
,
Scott
,
R. H.
and
Trentham
,
D. R.
(
1988
).
Photoactivation of intracellular guanosine triphosphate analogues reduces the amplitude and slows the kinetics of voltage-activated calcium channel currents in sensory neurones
.
Pflügers Arch.
411
,
628
636
.
Evans
,
M. G.
and
Marty
,
A.
(
1986
).
Calcium-dependent chloride currents in isolated cells from rat lacrimal glands
.
J. Physiol., Lond.
378
,
437
460
.
Fenwick
,
E. M.
,
Marty
,
A.
and
Neher
,
E.
(
1982
).
A patch clamp study of bovine chromaffin cells and of their sensitivity to acetylcholine
.
J. Physiol., Lond.
331
,
577
597
.
Field
,
A. C.
and
Jenkinson
,
D. H.
(
1987
).
The effect of noradrenaline on the ion permeability of isolated mammalian hepatocytes, studied by intracellular recording
.
J. Physiol., Lond.
392
,
493
512
.
Finch
,
E. A.
,
Turner
,
T. J.
and
Goldin
,
S. M.
(
1991
).
Calcium as a coagonist of inositol 1,4,5-trisphosphate induced calcium release
.
Science
252
,
443
446
.
Gray
,
P. T. A.
,
Ogden
,
D. C.
,
Trentham
,
D. R.
and
Walker
,
J. W.
(
1988
).
Elevation of cytosolic free calcium by rapid photolysis of caged InsP3 in single rat parotid acinar cells
.
J. Physiol., Lond.
410
,
90P
.
Harootunian
,
A. T.
,
Kao
,
J. P. Y.
and
Tsien
,
R. Y.
(
1988
).
Agonist-induced calcium oscillations in depolarized fibroblasts and their manipulation by photoreleased Ins(1,4,5)P3, Ca2+ and Ca2+ buffer
.
Cold Spring Harbor Symp. quant. Biol
.
53
,
935
943
.
Hartzell
,
H. C.
,
Mery
,
P. F.
,
Fischmeister
,
R.
and
Szabo
,
G.
(
1991
).
Sympathetic regulation of cardiac calcium current is due exclusively to cyclic AMP-dependent phosphorylation
.
Nature
351
,
573
576
.
Hoth
,
M.
and
Penner
,
R.
(
1992
).
Depletion of intracellular calcium stores activates a calcium current in mast cells
.
Nature
355
,
353
356
.
Iino
,
M.
(
1990
).
Biphasic Ca2+ dependence of inositol 1,4,5-trisphosphate induced Ca2+ release in smooth muscle cells of the guinea pig taenia caeci
.
J. gen. Physiol
.
95
,
1103
1122
.
Iino
,
M.
and
Endo
,
M.
(
1992
).
Calcium dependent immediate feedback control of InsP3 induced Ca2+ release
.
Nature
360
,
76
78
.
Irvine
,
R. F.
(
1990
).
Quantal Ca2+ release and the control of Ca2+ entry by inositol phosphates – a possible mechanism
.
Febs Lett.
263
,
5
9
.
Jacob
,
R.
(
1990
).
Calcium oscillations in electrically non-excitable cells
.
Biochim. biophys. Acta
1052
,
427
438
.
Kaplan
,
J. H.
,
Forbrush
,
B.
and
Hoffman
,
J. F.
(
1978
).
Rapid photolytic release of adenosine 5-trisphosphate from a protected analogue: utilization by the Na:K pump of human red cell ghosts
.
Biochemistry, N.Y.
17
,
1929
1935
.
Kasai
,
H.
and
Augustine
,
G. J.
(
1990
).
Cytosolic Ca2+ gradients triggering unidirectional fluid secretion from exocrine pancreas
.
Nature
348
,
735
738
.
Khodakhah
,
K.
and
Ogden
,
D. C.
(
1993
).
Functional heterogeneity of calcium release by inositol 1,4,5-trisphosphate in single Purkinje neurones, cultured cerebellar astrocytes and peripheral tissues
.
Proc. natn. Acad. Sci. U.S.A.
90
,
4976
4980
.
Konishi
,
M.
,
Hollingworth
,
S.
,
Hawkins
,
A. B.
and
Bayor
,
S. M.
(
1991
).
Myoplasmic Ca2+ transients in intact frog skeletal muscle fibres monitored with the fluorescent indicator furaptra
.
J. gen. Physiol.
97
,
271
302
.
Kuno
,
M.
and
Gardner
,
P.
(
1987
).
Ion channels activated by inositol 1,4,5-trisphosphate in plasme membrane of human T-lymphocytes
.
Nature
326
,
301
304
.
Lester
,
H. A.
and
Nerbonne
,
J. M.
(
1982
).
Physiological and pharmacological manipulations with light flashes
.
A. Rev. Biophys. Bioenerg.
11
,
151
175
.
Levitan
,
I.
,
Hillman
,
P.
and
Payne
,
R.
(
1993
).
Fast desensitisation of the response to Ins P3 in Limulus ventral photoreceptors
.
Biophys. J.
64
,
1354
1360
.
Llano
,
I.
,
Dreessen
,
J.
,
Kano
,
M.
and
Konnerth
,
A.
(
1991
).
Intradendritic release of calcium induced by glutamate in cerebellar Purkinje cells
.
Neuron
7
,
577
583
.
Llano
,
I.
and
Marty
,
A.
(
1987
).
Protein kinase C activators inhibit the inositol trisphosphate-mediated muscarinic current responses in rat lacrimal cells
.
J. Physiol., Lond
.
394
,
239
248
.
Luckhoff
,
A.
and
Clapham
,
D. E.
(
1992
).
Inositol 1,3,4,5-tetrakisphosphate activates an endothelial Ca2+-permeable channel
.
Nature
355
,
356
358
.
Marty
,
A.
(
1991
).
Calcium release and internal calcium regulation in acinar cells of exocrine glands
.
J. Membr. Biol
.
124
,
189
197
.
Marty
,
A.
,
Horn
,
R.
,
Tan
,
Y. P.
and
Zimmerberg
,
J.
(
1989
).
Delay of the Ca2+ mobilization response to muscarinic stimulation
. In
Secretion and its Control, 42nd Annual Symposium of the Society of General Physiologists
(ed.
C. M.
Armstrong
and
G. S.
Oxford
), pp.
97
110
.
New York
:
Rockfeller University Press
.
Matthews
,
G.
,
Neher
,
E.
and
Penner
,
R.
(
1989
).
Second messenger-activated calcium influx in rat peritoneal mast cells
.
J. Physiol., Lond
.
418
,
105
130
.
McCray
,
J. A.
and
Trentham
,
D. R.
(
1989
).
Properties and uses of photoreactive caged compounds
.
A. Rev. Biophys. biophys. Chem
.
18
,
239
270
.
McDonald
,
T. V.
,
Premack
,
B. A.
and
Gardner
,
P.
(
1993
).
Flash photolysis of caged InsP3 activates plasma membrane calcium current in human T cells
.
J. biol. Chem
.
268
,
3889
3896
.
Meyer
,
T.
,
Holowka
,
D.
and
Stryer
,
L.
(
1988
).
Highly cooperative opening of calcium channels by inositol 1,4,5-trisphosphate
.
Science
240
,
653
656
.
Meyer
,
T.
,
Wensel
,
T.
and
Stryer
,
L.
(
1990
).
Kinetics of calcium channel opening by inositol 1,4,5-trisphosphate
.
Biochemistry, N.Y.
29
,
32
37
.
Miledi
,
R.
and
Parker
,
I.
(
1989
).
Latencies of membrane currents evoked in Xenopus oocytes by receptor activation, inositol trisphosphate and calcium
.
J. Physiol., Lond
.
415
,
189
210
.
Morris
,
A. P.
,
Gallacher
,
D. V.
,
Irvine
,
R. F.
and
Petersen
,
O. H.
(
1987
).
Synergism of inositol trisphosphate and tetrakisphosphate in activating Ca2+ dependent K channels
.
Nature
330
,
653
655
.
Muallem
,
S.
,
Pandol
,
S. J.
and
Beeker
,
T. G.
(
1989
).
Hormone evoked calcium release from intracellular stores is a quantal process
.
J. biol. Chem
.
264
,
205
212
.
Murphy
,
S. N.
and
Miller
,
R.
(
1988
).
A glutamate receptor regulates Ca2+ mobilisation in hippocampal neurons
.
Proc. natn. Acad. Sci. U.S.A
.
85
,
8737
8741
.
Neher
,
E.
and
Augustine
,
G. J.
(
1992
).
Calcium gradients and buffers in bovine chromaffin cells
.
J. Physiol., Lond
450
,
273
301
.
Nerbonne
,
J. M.
(
1986
).
Design and application of photolabile intracellular probes
.
Soc. gen. Physiol. Ser.
40
,
417
445
.
Nerbonne
,
J. M.
,
Richard
,
S.
,
Nargeot
,
J.
and
Lester
,
H. A.
(
1984
).
New photoactivable cyclic nucleotides produce intracellular jumps in cyclic AMP and cyclic GMP concentrations
.
Nature
310
,
74
76
.
Noel
,
J.
and
Capiod
,
T.
(
1991
).
Photolytic release of cyclic AMP activates Ca2+-dependent K+ permeability in guinea-pig liver cells
.
Pflügers Arch.
417
,
546
548
.
Ogden
,
D. C.
,
Capiod
,
T.
and
Carter
,
T.
(
1991
).
The delay in activation of plasmalemmal K conductance by Ca2+ ions released by inositol trisphosphate in guinea-pig hepatocytes
.
J. Physiol., Lond
434
,
39P
.
Ogden
,
D. C.
,
Capiod
,
T.
,
Walker
,
J. W.
and
Trentham
,
D. R.
(
1990
).
Kinetics of the conductance evoked by noradrenaline, inositol trisphosphate or Ca2+ in guinea-pig isolated hepatocytes
.
J. Physiol., Lond.
422
,
585
602
.
Pang
,
I. H.
and
Sternweis
,
P. C.
(
1990
).
Purification of unique a subunits of GTP-binding regulatory proteins (G proteins) by affinity chromatography with immobilized-subunits
.
J. biol. Chem.
265
,
18707
18712
.
Parker
,
I.
and
Ivorra
,
I.
(
1990a
).
Localized all-or-none calcium liberation by inositol trisphosphate
.
Science
250
,
977
979
.
Parker
,
I.
and
Ivorra
,
I.
(
1990b
).
Inhibition by Ca2+ of inositol trisphosphate-mediated Ca2+ liberation: a possible mechanism for oscillatory release of Ca2+
.
Proc. natn. Acad. Sci. U.S.A
.
87
,
260
264
.
Parker
,
I.
and
Ivorra
,
I.
(
1993
).
Confocal microfluorimetry of Ca signals evoked in Xenopus oocytes by photoreleased inositol trisphosphate
.
J. Physiol., Lond
.
461
,
133
165
.
Parker
,
I.
and
Miledi
,
R.
(
1987
).
InsP3 activates a voltage dependent Ca2+ influx in Xenopus oocytes
.
Proc. R. Soc. Lond B
231
,
27
36
Parker
,
I.
and
Miledi
,
R.
(
1989
).
Nonlinearity and facilitation in phosphoinositide signalling studied by the use of caged inositol trisphosphate in Xenopus oocytes
.
J. Neurosci.
9
,
4068
4077
.
Payne
,
R.
,
Flores
,
T. M.
and
Fein
,
A.
(
1990
).
Feedback inhibition by calcium limits the release of calcium by inositol trisphosphate in Limulus ventral photoreceptors
.
Neuron
4
,
547
555
.
Payne
,
R.
and
Potter
,
B. V. L.
(
1991
).
Injection of inositol trisphosphorothioate into Limulus ventral photoreceptors causes oscillations of free cytosolic calcium
.
J. gen. Physiol
.
97
,
1165
1186
.
Payne
,
R.
,
Walz
,
B.
,
Levy
,
S.
and
Fein
,
A.
(
1988
).
The localization of Ca2+ release in Limulus photoreceptors and its control by negative feedback
.
Phil. Trans. R. Soc. B
320
,
359
370
.
Petersen
,
O. H.
and
Wakui
,
M.
(
1990
).
Oscillating intracellular Ca2+ signals evoked by activation of receptors linked to inositol lipid hydrolysis – mechanism of generation
.
J. Membr. Biol
.
118
,
93
105
.
Pietri
,
F.
,
Hilly
,
M.
and
Mauger
,
J. P.
(
1990
).
Calcium mediates the interconversion of two states of the liver inositol 1,4,5-trisphosphate receptor
.
J. biol. Chem
.
265
,
17478
17485
.
Pusch
,
M.
and
Neher
,
E.
(
1988
).
Rates of diffusional exchange between small cells and a measuring patch pipette
.
Pflügers Arch.
411
,
204
211
.
Rapp
,
G.
and
Guth
,
K.
(
1988
).
A low cost high intensity flash device for photolysis experiments
.
Pflügers Arch.
411
,
200
203
.
Rooney
,
T.
,
Sass
,
E. J.
and
Thomas
,
A. P.
(
1990
).
Agonist induced cytosolic calcium oscillations originate from a specific locus in single hepatocytes
.
J. biol. Chem
.
265
,
10792
10796
.
Sanchez-Bueno
,
A.
,
Dixon
,
C. J.
,
Woods
,
N. M.
,
Cuthbertson
,
K. S. R.
and
Cobbold
,
P. H.
(
1990
).
Inhibitors of protein kinase C prolong the falling phase of each free Ca2+ transient in a hormone-stimulated hepatocyte
.
Biochem. J.
268
,
627
632
.
Shuttleworth
,
T.
(
1992
).
Ca2+ release from InsP3 sensitive stores is not modulated by intraluminal Ca
.
J. biol. Chem.
267
,
3573
3576
.
Somlyo
,
A. V.
,
Horiuti
,
K.
,
Trentham
,
D. R.
,
Kitazawa
,
T.
and
Somlyo
,
A. P.
(
1992
).
Kinetics of InsP3-induced Ca2+-release and contraction induced by photolysis of caged D-myo inositol 1,4,5-trisphosphate in smooth muscle – the effects of heparin, procaine and adenine nucleotides
.
J. biol. Chem.
267
,
22316
22322
.
Taylor
,
C. W.
and
Potter
,
B. V. L.
(
1990
).
The size of inositol 1,4,5-trisphosphate-sensitive Ca2+ stores depends on inositol 1,4,5-trisphosphate concentration
.
Biochem. J.
266
,
189
194
.
Wakui
,
M.
,
Potter
,
B. V. L.
and
Petersen
,
O. H.
(
1989
).
Pulsatile intracellular calcium release does not depend on fluctuations in inositol trisphosphate concentration
.
Nature
339
,
317
320
.
Walker
,
J. W.
(
1991
).
Caged molecules activated by light
. In
Cellular Neurobiology: A Practical Approach
(ed.
J.
Chad
and
H.
Wheal
), pp.
179
204
. Oxford: IRL Press.
Walker
,
J. W.
,
Feeney
,
J.
and
Trentham
,
D. R.
(
1989
).
Photolabile precursors of inositol phosphates. Preparation and properties of 1-(2-nitrophenyl)ethyl esters of myo-inositol 1,4,5-trisphosphate
.
Biochemistry, N.Y.
28
,
3272
3280
.
Walker
,
J. W.
,
Reid
,
G. P.
,
McCray
,
J. A.
and
Trentham
,
D. R.
(
1988
).
Photolabile 1-(2-nitrophenyl)ethyl phosphate esters of adenine nucleotide analogues. Synthesis and mechanism of photolysis
.
J. Am. chem. Soc.
110
,
7170
7177
.
Walker
,
J. W.
,
Somlyo
,
A. V.
,
Goldman
,
Y. E.
,
Somlyo
,
A. P.
and
Trentham
,
D. R.
(
1987
).
Kinetics of smooth and skeletal muscle activation by laser pulse photolysis of caged inositol 1,4,5-trisphosphate
.
Nature
327
,
249
252
.
Wootton
,
J. F.
and
Trentham
,
D. R.
(
1989
).
‘Caged’ compounds to probe the dynamics of cellular processes: synthesis and properties of some novel photosensitive P-2-nitrobenzylesters of nucleotides
. In
Photochemical Probes in Biochemistry
(ed.
P. E.
Nielsen
), pp.
277
296
. NATO ASI Series C, vol. 272. Dordrecht: Kluwer.