• Non ci sono risultati.

Calcium SignalingLawrence D. Gaspers, Nicola Pierobon, Andrew P. Thomas

N/A
N/A
Protected

Academic year: 2022

Condividi "Calcium SignalingLawrence D. Gaspers, Nicola Pierobon, Andrew P. Thomas"

Copied!
11
0
0

Testo completo

(1)

Calcium Signaling

Lawrence D. Gaspers, Nicola Pierobon, Andrew P. Thomas

18

18.1

Ca

2+

Signaling in Liver

Ca

2+

is a ubiquitous and versatile signaling molecule controlling the activity of a broad spectrum of bio- logical events [2–4, 13, 74]. Agonist-evoked changes in cytosolic free Ca

2+

([Ca

2+

]

i

) can be spatially local- ized or global in nature, with information encoded in the frequency of Ca

2+

spikes or the magnitude of the Ca

2+

increase. The ability to control Ca

2+

events in both space and time permits the transmission of a diverse set of intracellular messages to downstream Ca

2+

-sensitive targets. Calcium ions can exert con- trol by either binding directly to the target protein, as is the case for intramitochondrial dehydrogenas- es [41], or indirectly through Ca

2+

-binding proteins, such as calmodulin or calcineurin [40, 42] and even at the genetic level through transcriptional repres- sors (i.e. DREAM) [9].

Ca

2+

signaling is likely to be of critical impor- tance for hepatocyte function within the intact liver.

Hepatocytes are large polarized epithelia cells with different Ca

2+

-sensitive metabolic activities occur- ring within the basolateral and apical membrane domains. One pertinent example is the actin- and myosin-containing microfilaments surrounding the apical membrane, which has been implicated in Ca

2+

-dependent contraction of the bile canaliculi [7, 33, 78, 79, 83]. Moreover, the functional capac- ity of the hepatic lobule displays an asymmetrical distribution between hepatocytes; ATP-utilizing pathways, such as gluconeogenesis and urea syn- thesis, predominate in the periportal zone, whereas glutamine production and glycolytic activity are higher in pericentral hepatocytes [31, 32]. Each dis- tinct metabolic zone presumably contains it own unique repertoire of hormone receptors, intracellu- lar Ca

2+

requirements and Ca

2+

sensors. Thus, inter- cellular Ca

2+

movements between lobular zones may be required for the liver to function in a coordinated fashion under high metabolic demand.

18.2

Phosphoinositide Signaling Pathway

Glycogenolytic hormones, such as α

1

-adrenergic ag- onists, vasopressin and ATP, evoke [Ca

2+

]

i

increases by stimulating the production of the Ca

2+

-mobiliz- ing second messenger, inositol-1,4,5-trisphosphate (IP

3

). This is achieved through receptor-associated heterotrimeric G proteins, which activate phos- phohypherationinositide-specific phospholipase C (PLC) [2, 13, 53]. Under basal conditions, heterot- rimeric G proteins consisting of α, β, and γ subunits are associated together in an inactive GDP-bound state. Agonist binding to cell surface receptors stim- ulates the exchange of GDP for GTP on the G α subu- nit, promoting its dissociation from the βγ dimer.

The G γ subunits are modified by protein prenyla- tion, which promotes the association of the βγ dimer with membranes. The activated G α subunits also re- main attached to the plasma membrane through a covalently attached palmitate moiety in the COOH terminal. All α subunits of the G

q

subfamily interact with the C2 domain of PLC- β mediating the trans- location of the enzyme to the plasma membrane.

The G βγ dimer has also been proposed to enhance

membrane localization of PLC- β by binding to the

pleckstrin homology (PH) domain in the NH

2

ter-

minal [53]. Membrane localization increases the hy-

drolysis of a minor membrane phospholipid, phos-

phatidylinositol-4,5-bisphosphate (PIP

2

) resulting

in the co-release of two second messengers: IP

3

and

diacylglycerol (DAG). The lipophilic DAG recruits

protein kinase C (PKC) to the plasma membrane

activating one branch of the phosphoinositide-de-

pendent signaling pathway. At the same time, IP

3

is

released into the cytosol where it mobilizes inter-

nal Ca

2+

stores in the endoplasmic reticulum upon

binding to the IP

3

receptors (IP

3

R) [2, 48]. The de-

pletion of luminal Ca

2+

stores stimulates the influx

of external Ca

2+

to sustain the agonist signal and

refill the internal stores [51]. The signaling pathway

is turned off by intrinsic GTPase activity of the G α

subunit slowly hydrolyzing GTP back to GDP, which

(2)

allows reassembly with the G βγ dimer into an inac- tive complex.

Epidermal growth factor and hepatocyte growth factor are important agonists regulating liver growth and regeneration. These growth factors also acti- vate bifurcated signaling pathways, including Ca

2+

mobilization in isolated hepatocytes. In this case, growth factor challenge evokes receptor dimeriza- tion resulting in autophosphorylation of the receptor on specific tyrosine residues. These phosphorylated protein domains are high affinity docking sites for SH2-containing proteins, like PLC- γ. After binding to the SH2-domain, PLC- γ is activated by receptor- mediated phosphorylation. Moreover, binding to growth factor receptors tethers the enzyme close to its substrate, PIP

2

, in the plasma membrane [2, 13, 53]. The end result of these events is an increase in IP

3

and DAG formation.

18.3

IP3-Dependent [Ca

2+

]

i

Oscillations in Liver

At the single cell level, receptor-mediated activa- tion of the phosphoinositide signaling pathway of- ten generates complex spatiotemporal intracellular [Ca

2+

]

i

responses. Peter Cobbold and coworkers were the first group to demonstrate that hormones coupled to PLC activation elicit a series of discrete, baseline-separated [Ca

2+

]

i

spikes or oscillations in aequorin-injected hepatocytes [81, 82]. We have used Ca

2+

-sensitive fluorescent indicators in com- bination with digital fluorescence imaging mi- croscopy and laser scanning confocal microscopy techniques to confirm these initial observations in isolated hepatocytes [59, 60, 71] and have adapted these approaches to extend the study of Ca

2+

signal- ing into the intact liver [21, 49, 55, 56, 72–74].

In agreement with our previous studies con- ducted in primary cultured hepatocytes, challeng- ing the liver with a continuous perfusion of sub- maximal agonist concentrations evokes periodic [Ca

2+

]

i

spikes in the individual hepatocytes of the perfused liver (Fig. 18.1). In both cultured cells and hepatocytes within the intact liver, the agonist con- centration determines the frequency of the [Ca

2+

]

i

oscillations, whereas the amplitude and kinetics of the individual [Ca

2+

]

i

oscillation are independent of agonist concentration [55, 59, 72]. This phenome- non has been termed frequency-modulation [1] and is thought to regulate Ca

2+

-sensitive targets with greater fidelity than other types of [Ca

2+

]

i

signals (e.g. amplitude-modulation), especially at low levels of agonist stimulation [71, 74]. Frequency encoded Ca

2+

signals have been shown to be more effec- tive in controlling mitochondrial metabolism [24]

initiating gene expression [39] and capable of dif- ferentially activating transcription factors [17]. To use the information encoded in the incoming Ca

2+

signals necessitates that downstream Ca

2+

-sensitive proteins can decode or count the [Ca

2+

]

i

spikes. The rates of activation and inactivation of PKC [46] or Ca

2+

/calmodulin-dependent protein kinase II [16]

in response to oscillatory Ca

2+

signals suggest that these enzymes are capable of discriminating be- tween Ca

2+

signals that differ in spike frequency, converting the digitized Ca

2+

response into discrete levels of kinase activity [16]. Moreover, we have shown that mitochondria can decode the frequency of Ca

2+

signals into different time-averaged levels of NAD(P)H production; the result of a complex in- terplay between Ca

2+

-activated intramitochondrial dehydrogenases and Ca

2+

-dependent stimulation of the respiratory chain [24, 57].

The traces in Fig. 18.1 (a, b) illustrate two exam- ples of the agonist concentration exerting control over [Ca

2+

]

i

signals in hepatocytes of the perfused liver. At low vasopressin concentrations, individual

Fig. 18.1 Vasopressin-evoked [Ca2+]i signals in the intact per- fused liver. a, b Fura-2/AM-loaded livers were challenged with the indicated hormone concentrations. c [Ca2+]i oscillations in

two hepatocytes displaced 70 µm along the hepatic plate. All traces are typical, single-cell [Ca2+]i responses.

(3)

[Ca

2+

]

i

spikes are separated by relatively long peri- ods during which [Ca

2+

]

i

is maintained close to ba- sal values. Increasing the hormone concentration from 50 to 200 pM vasopressin evoked either an increase in the number of Ca

2+

spikes per minute (Fig. 18.1a) or resulted in a sustained increase in [Ca

2+

]

i

(Fig. 18.1b). Note that there is a difference in hormone sensitivity between the two hepatocytes in response to the initial agonist challenge.

The [Ca

2+

]

i

spikes in individual hepatocytes of the perfused liver are also spatially organized as regenerative intracellular Ca

2+

waves, with similar properties to those described for isolated hepato- cytes. In both cases, the rate of [Ca

2+

]

i

rise was simi- lar throughout the cell, but the Ca

2+

responses were offset in direct proportion to the distance from the site of Ca

2+

wave initiation. Intracellular Ca

2+

waves commenced predominately close to the basolateral membrane or sometimes occurred more diffusely throughout the cell (Fig. 18.4). The latter observa- tion may reflect Ca

2+

waves propagating in the Z-di- rection or from out of the focal plane. Once initiated, intracellular Ca

2+

waves propagate throughout the cell at a constant velocity and are unaffected by ago- nist type or concentration. The calculated rates of Ca

2+

wave propagation were 15–25 µm/s in both iso- lated hepatocytes [60] and individual hepatocytes of the perfused liver [55]. Depleting extracellular Ca

2+

for short periods prior to agonist challenge did not alter the rate of intracellular Ca

2+

waves (unpub- lished observations) suggesting that Ca

2+

influx is not mandatory. However, it should be noted the ex- tracellular Ca

2+

omission did slow the rate of inter- cellular Ca

2+

waves (see below).

18.4

Intercellular Ca

2+

Waves

Ca

2+

signals are not limited to the boundaries of the one cell, but rather the [Ca

2+

]

i

increase can be com- municated to neighboring cells giving rise to inter- cellular Ca

2+

waves. Ca

2+

waves passing through multiple cells and up to 1 mm in distance have been reported in a variety of cell preparations, which preserved the functional integrity of the intact tis- sue [64] and in perfused organs [43, 49, 55, 84, 85].

Intercellular Ca

2+

waves have been evoked by local or global application of Ca

2+

-mobilizing hormones [55], mechanical stimulation of a single cell [63] or microinjection/photo release of Ca

2+

-mobilizing second messengers [10, 63]. Intercellular Ca

2+

waves can display remarkable multicellular organiza- tion, propagating as radial [55], circuitous or even spiral Ca

2+

waves [26] at speeds up to 100 µm/s. A

potentially interesting property of some Ca

2+

waves is their ability to pass between different cell types.

Such “heterotypic” calcium signaling has been dem- onstrated to occur between photocytes and their supporting cells, oocytes and follicular cells, mixed glia and in astrocyte/endothelial cell co-cultures [19, 38, 44, 62]. Intercellular Ca

2+

waves in the as- trocyte syncitium have been reported to pass into neighboring neurons and modulate their excitabil- ity [45]. Conversely, electrical stimulation of neural circuits can trigger Ca

2+

waves in glial cell popula- tions in both hippocampal slices [15] and the neu- romuscular junction [30, 52].

The mechanism whereby Ca

2+

waves pass be- tween cells is the subject of some debate: both intra- cellular and extracellular pathways have been pro- posed. Calcium waves may propagate by diffusion of a second messenger through gap junctions or by the secretion of a Ca

2+

-mobilizing agonist, poten- tially ATP. Evidence for a role of gap junctions has been provided by studies demonstrating blockage of intercellular Ca

2+

waves by gap junction inhibitors [6, 75] or electroporation of antibodies to connexins [5], a gap junction constituent. Indeed, gap junctions are permeable to both Ca

2+

and the second messen- ger, IP

3

[61]. Moreover, the extent of intercellular Ca

2+

waves has been correlated with the expression levels of connexin proteins. Thus, intercellular Ca

2+

waves are not observed in cell lines expressing low levels of connexins but occur in cultures transfected with connexins [12]. Heterogeneity in gap junction permeability between cells may provide an elegant means for “routing” Ca

2+

waves through the tissue, while the level of connexin expression or subtype composition may regulate the rate of intercellular Ca

2+

waves.

In other cell preparations, Ca

2+

waves propagate

across cell-free boundaries and the path of these

Ca

2+

waves can be influenced by the direction of

extracellular perfusate flow [28, 47]. Such studies

suggest that upon stimulation certain cell types se-

crete an extracellular message; most likely ATP that

induces Ca

2+

increases in neighboring cells [23, 28,

47]. The field is further complicated by the observa-

tion that overexpressing connexins in gap-junction-

deficient cell lines enhances ATP release during

mechanical stimulation [14]. This discrepancy may

be explained by the fact that unpaired connexins or

hemichannels in the plasma membrane appear to

form Ca

2+

-regulated ATP-secreting channels [22,

69]. Indeed, photorelease of caged IP

3

and the con-

sequential rise in Ca

2+

has been shown to stimulate

ATP release, presumably mediated through con-

nexin hemichannels [8]. Moreover, focal mechani-

cal stimulation of single cultured hepatocytes has

been reported to release ATP evoking a [Ca

2+

]

i

rise

(4)

in neighboring hepatocytes and bile ductal cells through the activation of metabotrophic purinergic receptors [65]. Finally, there is evidence that hor- monal stimulation can increase cell volume in the perfused liver [29, 76], which is a potent stimulus for ATP efflux in rat and human hepatocytes [18, 20]. Thus, it is possible that both gap junctions and ATP release are involved in propagating the agonist- evoked intercellular Ca

2+

wave. The relative contri- butions of these different pathways in mediating the propagation of intercellular Ca

2+

waves will most likely vary considerably depending upon the tissue and agonist type.

18.5

Intercellular Ca

2+

Waves in the Perfused Liver

Remarkably, IP

3

-dependent Ca

2+

signals are or- ganized at the multicellular level in the perfused

liver and do not occur asynchronously. The traces in Fig. 18.1c depict vasopressin-evoked [Ca

2+

]

i

re- sponses in two hepatocytes separated by 70 µm along the same hepatic plate. The [Ca

2+

]

i

spikes in the two cells have the same periodicity, but are phase-shifted in time. The mechanism underlying this observation is the coordinated propagation of intercellular Ca

2+

waves from cell to cell along the hepatic plate (Figs. 18.2, 18.4). Thus, the rising phase of each [Ca

2+

]

i

spike passes sequentially into neighboring hepatocytes.

Fig. 18.2 shows a series of confocal images of fluo-3 fluorescence in two adjacent hepatic lobules.

The periportal (PP) and pericentral (PC) zones were identified by infusing fluorescein-labeled bovine serum albumin (F-BSA) into the portal circulation and recording the flow pattern of the dye. F-BSA does not readily cross the hepatocyte plasma mem- brane or access the bile canaliculi, and thus is an excellent marker for the sinusoidal space. The PP zones are defined as those regions displaying the first increase in fluorescence during F-BSA infusion

Fig. 18.2 Fluo-3/AM-loaded liver was sequentially challenged with 1–10 nM glucagon (top) followed by 80 pM vasopressin (bottom). The images (597×406 µm) were acquired with laser

scanning confocal microscopy. Ca2+ increases are depicted in white and time(s) after hormone infusion is shown bottom right.

(5)

(Fig. 18.3). The white overlay superimposed onto the grayscale images in Figs. 18.2 and 18.4 depicts the sites of fluorescence increase at the indicated time points during hormone challenge.

Each vasopressin-evoked [Ca

2+

]

i

spike initiates from a small number of hepatocytes located just outside the portal tract, and then propagates in a ra- dial fashion into the portal vein and outward to the pericentral region (Fig. 18.2, bottom panels). The initiator cells act like pacemakers entraining the fre- quency and spatial pattern of [Ca

2+

]

i

oscillations for the entire lobule. These intercellular [Ca

2+

]

i

waves are not self-propagating and require the continuous presence of the hormone to pass through the entire lobule. The direction of the translobular Ca

2+

wave is periportal to pericentral during low hormone challenge. This orientation is also observed during retrograde perfusion via the hepatic vein [55], but is reversed at supramaximal vasopressin doses that generate sustained [Ca

2+

]

i

increases. Thus, the path of Ca

2+

wave propagation is independent of the di- rection of perfusion flow.

Coordinated intralobular [Ca

2+

]

i

waves were routinely observed when livers were stimulated with vasopressin or α

1

-adrenergic agonists, but not for all agonists reported to elevate [Ca

2+

]

i

in isolated hepa- tocytes. Glucagon is a key gluconeogenic and glycog- enolytic hormone that regulates glucose output from the liver. Several groups have previously shown that physiological levels of glucagon can evoke [Ca

2+

]

i

increases in hepatocyte suspensions [11, 68], which may be mediated by a rise in either IP

3

or cAMP [68, 77]. Challenging the perfused liver with 1–10 nM glucagon elicited asynchronous [Ca

2+

]

i

oscilla- tions in a limited number of hepatocytes (Fig. 18.2, top panels) that were randomly distributed in both periportal and pericentral regions. These [Ca

2+

]

i

signals propagated along the hepatic plate for rela- tively short distances (several cells) before dissipat- ing. Moreover, these asynchronous Ca

2+

responses were similar to the Ca

2+

signals obtained with sub- threshold vasopressin doses (>10 pM). Supraphysi- ological levels of glucagon (100 nM) did evoke a sustained increase in [Ca

2+

]

i

in all of the cells in the hepatic lobule. However, these were non-oscillatory Ca

2+

signals and intercellular Ca

2+

waves followed the progression of the hormone infusion. These data suggest that physiological doses of glucagon do not evoke an increase in IP

3

production sufficient to fully mobilize internal Ca

2+

stores.

Stimulating the nerve tracts surrounding the portal vein enhances glucose and lactate release from the perfused liver, while inducing an overflow of norepinephrine into the hepatic vein [27, 67].

These data are consistent with hepatic sympathetic nerves controlling glycogen breakdown in the liver.

Most postganglionic sympathetic neurons store and co-release ATP along with norepinephrine [80].

Norepinephrine binds to α

1

-adrenergic receptors on the hepatocytes stimulating IP

3

-dependent [Ca

2+

]

i

signals with a consequent increase in phosphorylase a activity. On the other hand, little is known about the role of ATP in controlling glucose metabolism.

In our hands, low levels of ATP (1–10 µM) evoked [Ca

2+

]

i

oscillations in hepatocytes mainly near the portal tract and these Ca

2+

waves propagated across a third to a half of the lobule before termi- nating (unpublished observations). The pattern of Ca

2+

responses was not due to an uneven distribu- tion of purinergic receptors; maximal ATP (100 µM) challenge evoked a sustained [Ca

2+

]

i

increase in all hepatocytes in the lobule [49]. It is possible that during low ATP challenge the ligand is degraded by ecto-ATPase before reaching the pericentral hepa-

Fig. 18.3 Periportal (PP) and pericentral (PC) zones of the he- patic lobule are determined by infusing fluorescein-conjugated bovine serum albumin (F-BSA) into the portal vein. Time(s) rela-

tive to F-BSA infusion is shown top right. The sites of initial fluo- rescence increase are indicated (arrows).

(6)

tocytes. Since intercellular Ca

2+

waves require the continuous presence of agonist to propagate across the entire lobule, ATP hydrolysis could explain the early termination of the Ca

2+

wave. The functional consequences of these “truncated” Ca

2+

waves are currently unknown, but may be important in main- taining some of the distinct metabolic zones in the hepatic lobule.

When examined with sufficient spatial and temporal resolution, intracellular and intercellu- lar [Ca

2+

]

i

waves were easily observed propagating from cell to cell along the hepatic plates (Fig. 18.4).

Intercellular [Ca

2+

]

i

waves show a clear delay at each cell–cell boundary, which presumably reflects the period required to regenerate a propagating Ca

2+

- mobilizing second messenger in the adjacent hepa- tocyte. The propagation of intercellular Ca

2+

waves, in both isolated hepatocyte triplets [75] and the perfused liver (unpublished observations), are sen- sitive to gap junction inhibitors and experimental maneuvers designed to disrupt cell-to cell-contacts.

Moreover, vasopressin-evoked intercellular Ca

2+

waves are observed after prior desensitization of purinergic receptors with supramaximal ATP con- centrations [49]. Taken together, these data suggest that intercellular Ca

2+

waves in liver are predomi- nately mediated through gap junctions. The second messenger involved in propagating the Ca

2+

signal between hepatocytes has yet to be determined; how- ever, hepatic gap junctions are permeable to both Ca

2+

and IP

3

[61]. Although the rate of propaga- tion of intracellular Ca

2+

waves was independent of hormone concentration, the time delay for the Ca

2+

signal to pass between cells was inversely related to the agonist dose. Thus, the strength of the extracel-

lular stimuli, in the intact liver, can be encoded in both the frequency of [Ca

2+

]

i

spiking and the rate at which these Ca

2+

signals propagate across the lob- ule. Frequency-modulated Ca

2+

signals will most likely be the determining factor regulating the in- tensity of the cellular response in individual hepa- tocytes, whereas the rate of intercellular Ca

2+

wave propagation may determine the degree of coordina- tion between the lobular zones.

As mentioned previously, [Ca

2+

]

i

oscillations oc- cur asynchronously throughout the lobule at sub- threshold vasopressin concentrations, the result of limited intercellular Ca

2+

waves. Challenging the liver with bradykinin, to evoke Ca

2+

mobilization in hepatic sinusoidal endothelial cells, markedly increased the frequency of vasopressin-induced [Ca

2+

]

i

oscillations plus the rate and extent to which the intercellular Ca

2+

waves propagated across the lobule. The effects of bradykinin were mimicked by nitric oxide donors and blocked by inhibitors of nitric oxide formation [49]. These data suggest that crosstalk between nitric oxide and Ca

2+

signaling pathways may be another important mechanism in fine-tuning liver function.

18.6

Ca

2+

and Mitochondrial Metabolism

A rise in [Ca

2+

]

i

stimulates contractile, secretory or metabolic pathways augmenting ATP demand in the cytosol. To maintain cellular energy homeos- tasis and cell viability, the rate of ATP production must match utilization. This requires coordinating

Fig. 18.4 Confocal images (143×105 µm) illustrating the cell-to-cell propagation of [Ca2+]i oscillations. Dots show the initiating hepatocytes and arrows depict the ap- parent direction of the Ca2+

waves. Ca2+ increases are depicted in white and time(s) after vasopressin (100 nM) infusion is shown top left.

(7)

the activation of mitochondrial oxidative phosphor- ylation with the flux rate through cytosolic ATP-re- quiring reactions. In many cell types, mitochondria are strategically localized close to Ca

2+

release sites, such that [Ca

2+

]

i

increases from either internal Ca

2+

stores or Ca

2+

influx across the plasma membrane can be rapidly transported into the mitochondrial matrix [24, 37, 41, 54, 57, 58]. The consequent eleva- tion in mitochondrial Ca

2+

([Ca

2+

]

m

) stimulates the Ca

2+

-sensitive intramitochondrial dehydrogenases, resulting in elevation of NAD(P)H [24, 41, 57, 58]. The preferential coupling between increases of [Ca

2+

]

i

and [Ca

2+

]

m

is one proposed mechanism to coor- dinate mitochondrial ATP production with cellular energy demand [41, 58]. However, efficient transfer of Ca

2+

from the cytosol to the mitochondria only occurs during the rising phase of a Ca

2+

spike, when a gradient of high [Ca

2+

]

i

that is sufficient to activate mitochondrial Ca

2+

uptake develops close to the re- lease channel [24, 37, 54, 57, 58]. Consequently, mi- tochondrial metabolism is maximally stimulated by periodic [Ca

2+

]

i

oscillations at frequencies above 0.5 spikes per minute. In contrast, sustained increases in [Ca

2+

]

i

evoked by maximal hormone doses can only transiently elevate [Ca

2+

]

m

, and consequently do not sustain the activity of Ca

2+

-sensitive intrami- tochondrial dehydrogenases for a prolonged period [24, 57, 58].

Agonist-evoked [Ca

2+

]

i

responses can also regu- late the activity of the aspartate/glutamate carrier in the mitochondrial inner membrane [36]. This transporter is an integral component of the malate/

aspartate shuttle, which transfers cytosolic NADH into the mitochondrial matrix for oxidation. Thus, Ca

2+

-dependent stimulation of intramitochondrial dehydrogenases and the transport of cytosolic re- ducing equivalents work in a concerted fashion to increase supply of substrates for oxidative me- tabolism. In isolated hepatocytes, Ca

2+

-mobilizing hormones stimulate an increase in mitochondrial membrane potential ( ∆Ψ

m

), which is the potential energy used to drive ATP production [57, 58]. This appears to occur through a Ca

2+

-dependent inhibi- tion of mitochondrial pyrophosphatases leading to K

+

entry into the mitochondrial matrix, which stim- ulates volume increases and electron flux through the respiratory chain [25]. Finally, Ca

2+

can directly stimulate F

1

F

O

-ATPase synthetic activity and thus, a higher rate of ATP production [70]. Taken together, these data suggest that Ca

2+

ions can coordinate the entire ATP synthesis pathway from substrate supply to the finished product.

18.7

Hormone-Evoked Redox Changes in the Perfused Liver

We have previously shown that the activation of Ca

2+

-sensitive intramitochondrial dehydrogenases can be monitored in real time by changes in the relative redox state of mitochondrial pyridine and flavin nucleotides [21, 24, 57]. Ca

2+

-mobilizing hor- mones stimulate mitochondrial NADH production resulting in an increase in cellular autofluorescence at 360 nm. The reduced flavin has a lower fluores- cent signal at 470 nm than the oxidized form, thus an opposite signal change occurs for flavoproteins during hormone challenge (Fig. 18.5).

Cellular flavoprotein fluorescence originates predominately from the mitochondrial matrix, since the flavin fluorescence is quenched in most cytosolic proteins [66]. It has been estimated that lipoamide dehydrogenases and electron transfer fla- voproteins constitute 75% of total flavoprotein fluo- rescence in isolated mitochondria [34, 35]. Lipoam- ide dehydrogenase is an integral component of the Ca

2+

-sensitive intramitochondrial pyruvate and 2-oxoglutarate dehydrogenases [50] and their fla- vin cofactors are in direct equilibrium with the mi- tochondrial NAD

+

/NADH pool. On the other hand, NADH fluorescence can derive from both cytosolic and mitochondrial compartments. We have utilized 2-photon confocal microscopy to evaluate the rela- tive contribution of the cytosolic and mitochondrial pools to the overall NADH fluorescence signal. The image in Fig. 18.5a was acquired with multi-photon excitation and shows the endogenous pyridine nu- cleotide fluorescence in a fluo4/AM-loaded intact liver. The NAD(P)H signal originates primarily from the mitochondria, the bright punctuate structures within the two hepatocytes. The infrared absorp- tion for fluo4 and pyridine nucleotide are spectrally distinct, therefore fluo4 does not interfere with py- ridine nucleotide measurements. Hence, it should be possible to monitor hormone-evoked NADH increases with multi-photon excitation simultane- ously with Ca

2+

responses using fluo4 and a 488 nm argon laser line.

Challenging perfused livers with vasopressin

concentrations that evoke oscillatory Ca

2+

respons-

es resulted in a sustained increase in NAD(P)H pro-

duction, which was paralleled by a reduction in mi-

tochondrial flavoproteins (Fig. 18.5b). At the end of

the experiment, ketone bodies were infused into the

portal vein to evaluate the relative change in mito-

chondrial redox couples (Fig. 18.5b). We have show

previously that the mitochondrial NAD

+

/NADH

(8)

ratio can be manipulated experimentally with ac- etoacetate and β-hydroxybutyrate [24, 57]. Ketone bodies are interconverted by the mitochondrial ma- trix enzyme, β-hydroxybutyrate dehydrogenase, which utilizes mitochondrial NADH and NAD

+

, re- spectively. In this experiment, vasopressin evoked an approximately 50% reduction in mitochondrial NADH and FAD redox couples in equilibrium with β-hydroxybutyrate dehydrogenase.

The vasopressin-evoked rise in NAD(P)H fluo- rescence propagated across the lobule as a redox wave. In agreement with our isolated hepatocyte studies, hormone-dependent increases in NAD(P)H decayed more slowly than the [Ca

2+

]

i

spike, result- ing in a sustained mitochondrial response at low frequency of Ca

2+

spiking (not shown). Stimula- tion with sub-threshold agonist doses generates asynchronous [Ca

2+

]

i

oscillations, which do not propagate as coordinated intercellular Ca

2+

waves across the lobule. Under these conditions, the mi- tochondrial NAD(P)H responses are not sustained and limited to a few cells surrounding the portal tract. Increasing the agonist dose above the thresh- old (50–150 pM vasopressin) initiated a coordinated intercellular Ca

2+

wave and a concomitant NAD(P)H wave across the lobule. These studies strongly sug- gest that intercellular Ca

2+

waves allow the entire lobule to respond to agonist stimulation in a coor- dinated fashion. Moreover, these studies raise the possibility that other Ca

2+

-dependent processes in the liver are also regulated in a similar manner.

18.8 Summary

The strength of the extracellular agonist signal is thought to be encoded primarily by the frequency of [Ca

2+

]

i

spikes (temporal aspect), whereas the in- tracellular [Ca

2+

]

i

wave serves to propagate the Ca

2+

signal, thus ensuring that the entire cell is exposed to the full strength of the Ca

2+

signal (spatial as- pect). In the intact liver, agonist strength may also be conveyed by the rate of intercellular Ca

2+

wave propagation. Moreover, the translobular movement of Ca

2+

provides a mechanism to coordinate the dif- ferent metabolic zones in the liver. The ensemble of Ca

2+

signals is recognized by downstream Ca

2+

-sen- sitive pathways, such as mitochondrial NADH pro- duction, and converted into steady-state changes in metabolic output.

Selected Reading

Rhee SG. Regulation of phosphoinositide-specific phospholi- pase C. Annu Rev Biochem 2001;70:281–312. (A concise and readable review highlighting the enzymes involved in phos- phoinositide signaling pathways.)

Berridge MJ. Inositol trisphosphate and calcium signalling. Na- ture 1993;361:315–325. (A short, but detailed primer devot- ed to IP3-dependent Ca2+ signaling.)

Duchen MR. Contributions of mitochondria to animal physiol- ogy: from homeostatic sensor to calcium signalling and cell death. J Physiol (Lond) 1999;516:1–17. (A general review de- scribing mitochondrial Ca2+ homeostasis.)

Hajnóczky G, Robb-Gaspers LD, Seitz MB, Thomas AP. Decoding of cytosolic calcium oscillations in the mitochondria. Cell 1995;82:415–424.

Fig. 18.5 a Two-photon image of pyridine nucleotide fluores- cence in the intact liver. b Simultaneous measurement of flavin (blue) and pyridine (gray) nucleotide fluorescence changes dur-

ing vasopressin challenge. Acetoacetate (AcAc) and β-hydroxy- butyrate (β-HB) are added to assess the size of the mitochondrial redox pool.

(9)

References

1. Berridge MJ, Galione A. Cytosolic calcium oscillators. FASEB J 1988;2:3074–3082.

2. Berridge MJ. Inositol trisphosphate and calcium signalling.

Nature 1993;361:315–325.

3. Berridge MJ, Bootman MD, Lipp P. Calcium – a life and death signal [news]. Nature 1998;395:645–648.

4. Berridge MJ, Lipp P, Bootman MD. The versatility and univer- sality of calcium signalling. Nat Rev Mol Cell Biol 2000;1:11–

21.

5. Boitano S, Dirksen ER, Evans WH. Sequence-specific anti- bodies to connexins block intercellular calcium signaling through gap junctions. Cell Calcium 1998;23:1–9.

6. Boitano S, Evans WH. Connexin mimetic peptides revers- ibly inhibit Ca(2+) signaling through gap junctions in airway cells. Am J Physiol Lung Cell Mol Physiol 2000;279:L623–

L630.

7. Bouscarel B, Kroll SD, Fromm H. Signal transduction and hepatocellular bile acid transport: cross talk between bile acids and second messengers. Gastroenterology 1999;117:433–452.

8. Braet K, Vandamme W, Martin PE et al. Photoliberating inosi- tol-1,4,5-trisphosphate triggers ATP release that is blocked by the connexin mimetic peptide gap 26. Cell Calcium 2003;33:37–48.

9. Carrion AM, Link WA, Ledo F et al. DREAM is a Ca2+-regulated transcriptional repressor. Nature 1999;398:80–84.

10. Carter TD, Chen XY, Carlile G et al. Porcine aortic endothe- lial gap junctions: identification and permeation by caged InsP3. J Cell Sci 1996;109:1765–1773.

11. Charest R, Blackmore PF, Berthon B, Exton JH. Changes in free cytosolic Ca2+ in hepatocytes following alpha 1-adren- ergic stimulation. Studies on Quin-2-loaded hepatocytes. J Biol Chem 1983;258:8769–8773.

12. Charles AC, Naus CC, Zhu D et al. Intercellular calcium signal- ing via gap junctions in glioma cells. J Cell Biol 1992;118:195–

201.

13. Clapham DE. Calcium signaling. Cell 1995;80:259–268.

14. Cotrina ML, Lin JH, Alves-Rodrigues A et al. Connexins regu- late calcium signaling by controlling ATP release. Proc Natl Acad Sci USA 1998;95:15735–15740.

15. Dani JW, Chernjavsky A, Smith SJ. Neuronal activity triggers calcium waves in hippocampal astrocyte networks. Neuron 1992;8:429–440.

16. De Koninck P, Schulman H. Sensitivity of CaM kinase II to the frequency of Ca2+ oscillations [see comments]. Science 1998;279:227–230.

17. Dolmetsch RE, Xu K, Lewis RS. Calcium oscillations increase the efficiency and specificity of gene expression [see com- ments]. Nature 1998;392:933–936.

18. Dunkelberg JC, Feranchak AP, Fitz JG. Liver cell volume regu- lation: size matters. Hepatology 2001;33:1349–1352.

19. Dunlap K, Takeda K, Brehm P. Activation of a calcium-de- pendent photoprotein by chemical signalling through gap junctions. Nature 1987;325:60–62.

20. Feranchak AP, Fitz JG. Adenosine triphosphate release and purinergic regulation of cholangiocyte transport. Semin Liver Dis 2002;22:251–262.

21. Gaspers LD, Patel S, Morgan AJ et al. Monitoring generation and propagation of calcium signals in multicellular prepa- rations. In: Tepikin A, ed. Calcium signalling: a practical ap- proach. Oxford, UK: Oxford University Press, 2001:155–175.

22. Goodenough DA, Paul DL. Beyond the gap: functions of un- paired connexon channels. Nat Rev Mol Cell Biol 2003;4:285–

294.

23. Guthrie PB, Knappenberger J, Segal M et al. ATP released from astrocytes mediates glial calcium waves. J Neurosci 1999;19:520–528.

24. Hajnóczky G, Robb-Gaspers LD, Seitz MB, Thomas AP. De- coding of cytosolic calcium oscillations in the mitochondria.

Cell 1995;82:415–424.

25. Halestrap AP. The regulation of the matrix volume of mam- malian mitochondria in vivo and in vitro and its role in the control of mitochondrial metabolism. Biochim Biophys Acta 1989;973:355–382.

26. Harris-White ME, Zanotti SA, Frautschy SA, Charles AC. Spiral intercellular calcium waves in hippocampal slice cultures. J Neurophysiol 1998;79:1045–1052.

27. Hartmann H, Beckh K, Jungermann K. Direct control of gly- cogen metabolism in the perfused rat liver by the sympa- thetic innervation. Eur J Biochem 1982;123:521–526.

28. Hassinger TD, Guthrie PB, Atkinson PB et al. An extracellular signaling component in propagation of astrocytic calcium waves. Proc Natl Acad Sci USA 1996;93:13268–13273.

29. Haussinger D, Lang F. Cell volume and hormone action.

Trends Pharmacol Sci 1992;13:371–373.

30. Jahromi BS, Robitaille R, Charlton MP. Transmitter release in- creases intracellular calcium in perisynaptic Schwann cells in situ. Neuron 1992;8:1069–1077.

31. Jungermann K, Kietzmann T. Zonation of parenchymal and nonparenchymal metabolism in liver. Annu Rev Nutr 1996;16:179–203.

32. Jungermann K, Kietzmann T. Role of oxygen in the zonation of carbohydrate metabolism and gene expression in liver.

Kidney Int 1997;51:402–412.

33. Kitamura T, Brauneis U, Gatmaitan Z, Arias IM. Extracellular ATP, intracellular calcium and canalicular contraction in rat hepatocyte doublets. Hepatology 1991;14:640–647.

34. Kunz WS, Kunz W. Contribution of different enzymes to fla- voprotein fluorescence of isolated rat liver mitochondria.

Biochim Biophys Acta 1985;841:237–246.

35. Kunz WS. Evaluation of electron-transfer flavoprotein and alpha-lipoamide dehydrogenase redox states by two-chan- nel fluorimetry and its application to the investigation of beta-oxidation. Biochim Biophys Acta 1988;932:8–16.

36. Lasorsa FM, Pinton P, Palmieri L et al. Recombinant expres- sion of the Ca(2+)-sensitive aspartate/glutamate carrier in- creases mitochondrial ATP production in agonist-stimulated

(10)

Chinese hamster ovary cells. J Biol Chem 2003;278:38686–

38692.

37. Lawrie AM, Rizzuto R, Pozzan T, Simpson AW. A role for cal- cium influx in the regulation of mitochondrial calcium in en- dothelial cells. J Biol Chem 1996;271:10753–10759.

38. Leybaert L, Paemeleire K, Strahonja A, Sanderson MJ. Inosi- tol-trisphosphate-dependent intercellular calcium signal- ing in and between astrocytes and endothelial cells. Glia 1998;24:398–407.

39. Li W, Llopis J, Whitney M et al. Cell-permeant caged InsP3 ester shows that Ca2+ spike frequency can optimize gene expression [see comments]. Nature 1998;392:936–941.

40. Liu J, Albers MW, Wandless TJ et al. Inhibition of T cell sig- naling by immunophilin-ligand complexes correlates with loss of calcineurin phosphatase activity. Biochemistry 1992;31:3896–3901.

41. McCormack JG, Halestrap AP, Denton RM. Role of calcium ions in regulation of mammalian intramitochondrial me- tabolism. Physiol Rev 1990;70:391–425.

42. Means AR, VanBerkum MF, Bagchi I et al. Regulatory func- tions of calmodulin. Pharmacol Ther 1991;50:255–270.

43. Nathanson MH, Burgstahler AD, Mennone A et al. Ca2+ waves are organized among hepatocytes in the intact organ. Am J Physiol 1995;269:G167–G171.

44. Newman EA, Zahs KR. Calcium waves in retinal glial cells. Sci- ence 1997;275:844–847.

45. Newman EA, Zahs KR. Modulation of neuronal activity by glial cells in the retina. J Neurosci 1998;18:4022–4028.

46. Oancea E, Meyer T. Protein kinase C as a molecular ma- chine for decoding calcium and diacylglycerol signals. Cell 1998;95:307–318.

47. Osipchuk Y, Cahalan M. Cell-to-cell spread of calcium signals mediated by ATP receptors in mast cells. Nature 1992;359:241–244.

48. Patel S, Joseph SK, Thomas AP. Molecular properties of inosi- tol 1,4,5-trisphosphate receptors. Cell Calcium 1999;25:247–

264.

49. Patel S, Robb-Gaspers LD, Stellato KA et al. Coordination of calcium signalling by endothelial-derived nitric oxide in the intact liver. Nat Cell Biol 1999;1:467–471.

50. Perham RN, Jones DD, Chauhan HJ, Howard MJ. Substrate channelling in 2-oxo acid dehydrogenase multienzyme complexes. Biochem Soc Trans 2002;30:47–51.

51. Putney JW Jr, Bird GS. The inositol phosphate-calcium sign- aling system in nonexcitable cells. Endocr Rev 1993;14:610–

631.

52. Reist NE, Smith SJ. Neurally evoked calcium transients in terminal Schwann cells at the neuromuscular junction. Proc Natl Acad Sci USA 1992;89:7625–7629.

53. Rhee SG. Regulation of phosphoinositide-specific phos- pholipase C. Annu Rev Biochem 2001;70:281–312.

54. Rizzuto R, Brini M, Murgia M, Pozzan T. Microdomains with high Ca2+ close to IP3-sensitive channels that are sensed by neighboring mitochondria. Science 1993;262:744–747.

55. Robb-Gaspers LD, Thomas AP. Coordination of Ca2+ signal- ing by intercellular propagation of Ca2+ waves in the intact liver. J Biol Chem 1995;270:8102–8107.

56. Robb-Gaspers LD, Anderson PA, Thomas AP. Imaging whole organs. In: Rizzuto R, Fasolato C, eds. Imaging living cells.

New York: Springer-Verlag Inc., 1998:115–139.

57. Robb-Gaspers LD, Burnett P, Rutter GA et al. Integrating cytosolic calcium signals into mitochondrial metabolic re- sponses. EMBO J 1998;17:4987–5000.

58. Robb-Gaspers LD, Rutter GA, Burnett P et al. Coupling be- tween cytosolic and mitochondrial calcium oscillations: role in the regulation of hepatic metabolism. Biochim Biophys Acta 1998;1366:17–32.

59. Rooney TA, Sass EJ, Thomas AP. Characterization of cytosolic calcium oscillations induced by phenylephrine and vaso- pressin in single fura-2-loaded hepatocytes. J Biol Chem 1989;264:17131–17141.

60. Rooney TA, Sass EJ, Thomas AP. Agonist-induced cytosolic calcium oscillations originate from a specific locus in single hepatocytes. J Biol Chem 1990;265:10792–10796.

61. Saez JC, Connor JA, Spray DC, Bennett MV. Hepatocyte gap junctions are permeable to the second messenger, inositol 1,4,5-trisphosphate, and to calcium ions. Proc Natl Acad Sci USA 1989;86:2708–2712.

62. Sandberg K, Bor M, Ji H et al. Angiotensin II-induced calcium mobilization in oocytes by signal transfer through gap junc- tions. Science 1990;249:298–301.

63. Sanderson MJ, Charles AC, Dirksen ER. Mechanical stimula- tion and intercellular communication increases intracellular Ca2+ in epithelial cells. Cell Regul 1990;1:585–596.

64. Sanderson MJ, Charles AC, Boitano S, Dirksen ER. Mecha- nisms and function of intercellular calcium signaling. Mol Cell Endocrinol 1994;98:173–187.

65. Schlosser SF, Burgstahler AD, Nathanson MH. Isolated rat hepatocytes can signal to other hepatocytes and bile duct cells by release of nucleotides. Proc Natl Acad Sci USA 1996;93:9948–9953.

66. Scholz R, Thurman RG, Williamson JR et al. Flavin and pyri- dine nucleotide oxidation-reduction changes in perfused rat liver. I. Anoxia and subcellular localization of fluorescent flavoproteins. J Biol Chem 1969;244:2317–2324.

67. Seseke FG, Gardemann A, Jungermann K. Signal propaga- tion via gap junctions, a key step in the regulation of liver metabolism by the sympathetic hepatic nerves. FEBS Lett 1992;301:265–270.

68. Staddon JM, Hansford RG. Evidence indicating that the glu- cagon-induced increase in cytoplasmic free Ca2+ concentra- tion in hepatocytes is mediated by an increase in cyclic AMP concentration. Eur J Biochem 1989;179:47–52.

69. Stout CE, Costantin JL, Naus CC, Charles AC. Intercellular calcium signaling in astrocytes via ATP release through con- nexin hemichannels. J Biol Chem 2002;277:10482–10488.

70. Territo PR, Mootha VK, French SA, Balaban RS. Ca(2+) activa- tion of heart mitochondrial oxidative phosphorylation: role of the F(0)/F(1)-ATPase. Am J Physiol Cell Physiol 2000;278:

C423–C435.

(11)

71. Thomas AP, Renard DC, Rooney TA. Spatial and temporal organization of calcium signalling in hepatocytes. Cell Cal- cium 1991;12:111–126.

72. Thomas AP, Renard-Rooney DC, Hajnóczky G et al. Subcel- lular organization of calcium signalling in hepatocytes and the intact liver. Ciba Found Symp 1995;188:18–35.

73. Thomas AP, Robb-Gaspers LD, Rooney TA et al. Spatial or- ganization of oscillating calcium signals in liver. Biochem Soc Trans 1995;23:642–648.

74. Thomas AP, Bird GS, Hajnóczky G et al. Spatial and temporal aspects of cellular calcium signaling. FASEB J 1996;10:1505–

1517.

75. Tordjmann T, Berthon B, Claret M, Combettes L. Coordinated intercellular calcium waves induced by noradrenaline in rat hepatocytes: dual control by gap junction permeability and agonist. EMBO J 1997;16:5398–5407.

76. vom Dahl S, Hallbrucker C, Lang F, Haussinger D. Regulation of cell volume in the perfused rat liver by hormones. Bio- chem J 1991;280(Pt 1):105–109.

77. Wakelam MJ, Murphy GJ, Hruby VJ, Houslay MD. Activation of two signal-transduction systems in hepatocytes by gluca- gon. Nature 1986;323:68–71.

78. Watanabe S, Phillips MJ. Ca2+ causes active contraction of bile canaliculi: direct evidence from microinjection studies.

Proc Natl Acad Sci USA 1984;81:6164–6168.

79. Watanabe S, Miyazaki A, Hirose M et al. Myosin in hepa- tocytes is essential for bile canalicular contraction. Liver 1991;11:185–189.

80. Westfall DP, Todorov LD, Mihaylova-Todorova ST. ATP as a cotransmitter in sympathetic nerves and its inactivation by releasable enzymes. J Pharmacol Exp Ther 2002;303:439–

444.

81. Woods NM, Cuthbertson KS, Cobbold PH. Repetitive tran- sient rises in cytoplasmic free calcium in hormone-stimu- lated hepatocytes. Nature 1986;319:600–602.

82. Woods NM, Cuthbertson KS, Cobbold PH. Agonist-induced oscillations in cytoplasmic free calcium concentration in sin- gle rat hepatocytes. Cell Calcium 1987;8:79–100.

83. Wurzinger R, Englisch R, Roka S et al. Intracellular calcium in the isolated rat liver: correlation to glucose release, K(+) balance and bile flow. Cell Calcium 2001;30:403–412.

84. Ying X, Minamiya Y, Fu C, Bhattacharya J. Ca2+ waves in lung capillary endothelium. Circ Res 1996;79:898–908.

85. Zimmermann B, Walz B. Serotonin-induced intercellular calcium waves in salivary glands of the blowfly Calliphora erythrocephala. J Physiol 1997;500:17–28.

Riferimenti

Documenti correlati

Methodologies, including quality assurance such as data harmonization, completeness and plausibility tests have been applied according to the ICP Forests Manual, parts X and XV

Da qui l’interesse rivolto alla narrazione come forma più empirica, agile, flessibile, di accedere alla conoscenza, come elemento organizzatore che consente di selezionare i dati e

Let us start by very briefly reviewing the derivation of the acoustic metric for a BEC system, and show that the equations for the phonons of the condensate closely mimic the dynamics

Meta-analysis with Duval and Tweedie adjustment for publication bias demonstrated that SGA resulted in significant improvements of (hypo-)manic symptoms of bipolar mixed depression

Table 3: Overview of the variable ratings of the slightly emotional words (with ratings between 3.0 - 6.5 on the general emotionality scale) and of the highly emotional words

Da una parte, il progetto rivoluzionario di af- fermazione del monopolio statuale sul diritto e di semplificazione delle complessità giuridiche dell’Ancien Régime, è passato anche

This last case study, instead, shows how the proposed controller is able to track the desired Cartesian pose, while preserving the possibility to act in the null-space with a

Both of these have properties of fundamental measurement, but the latter (which is none other than a Rasch model) offers better statistical properties for the estimation of ability