Comprehensive Physiology Wiley Online Library

Electrophysiology of Salivary and Pancreatic Acinar Cells

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 Electrophysiological Methods in Study of Acinar Cells
1.1 Intracellular Microelectrodes
1.2 Patch‐Clamp Methods
2 Cell‐To‐Cell Communication
3 Resting Membrane Properties
3.1 Resting Membrane Potential
3.2 Electrodiffusional Control of Membrane Potential
4 Potassium Channels in Resting Membrane
4.1 Single‐Channel Recording from Excised Membrane Patches
4.2 Single‐Channel Current Recording from Intact Acinar Cells
4.3 Whole‐Cell Current Recording from Single Acinar Cells
5 Electrogenic Pumps
5.1 Sodium‐Potassium Pump
5.2 Sodium–Amino Acid Cotransport
6 Membrane Effects of Stimulants
6.1 Activation of Potassium Channels
6.2 Activation of Chloride Channels
6.3 Activation of Nonselective Cation Channels
6.4 Stimulant‐Evoked Membrane Potential Changes
6.5 Stimulation‐Evoked Changes in Membrane Capacitance
6.6 Source of Calcium Used as Messenger for Stimulant‐Evoked Membrane Changes
7 Importance of Electrogenic Processes For Acinar Cell Function
7.1 Potassium Channels and Fluid Secretion
7.2 Nonselective Cation Channels and Fluid Secretion
7.3 Potassium Channels and Protein Secretion
Figure 1. Figure 1.

Two simple models explaining stimulant‐evoked Cl uptake and membrane hyperpolarization in exocrine acinar cells. Left: model originally proposed by Lundberg 54 for salivary glands. Autonomic neurotransmitters activate directly an electrogenic active Cl transport. Right: model proposed by Petersen and Maruyama 97,119. Stimulant secretagogues [acetylcholine (ACh) and norepinephrine (NA) in salivary glands, other neurotransmitters or hormones in other glands; see ref. 89] evoke an increase in intracellular Ca2+ concentration ([Ca2+]i) that activates K+ channels. Exit of K+ allows K+ reuptake through Na+‐K+ pump and Na+‐K+‐2Cl cotransporter. In steady‐state stimulated condition all these processes constitute operational Cl pump that is electrogenic. i, Inner surface of plasma membrane; o, outer surface of plasma membrane.

From Petersen 93
Figure 2. Figure 2.

Intracellular recording with 2 separate microelectrodes from neighboring acinar cells. A: recording configuration. Both intracellular electrodes can be used for membrane potential recording and for current injection. If electrode is filled with secretagogue (shaded electrodes), current injection of appropriate polarity can be used for local agonist application. B: membrane potential recordings from mouse pancreas. Lower trace, rectangular hyperpolarizing current pulse 10‐9 A) of 100 ms duration. Upper traces, effects on membrane potential in injected cell as well as its neighbor during rest (a) and during acetylcholine application (b). Vertical calibration, 10 mV and 10‐9 A; horizontal calibration, 20 ms.

From Petersen and Ueda 104
Figure 3. Figure 3.

Different patch‐clamp recording configurations and procedures used to establish them. Patch‐clamp experiments start in cell‐attached configuration (top left). Tip of fire‐polished recording pipette is gently brought into contact with cell surface and slight suction is applied to pipette resulting in high‐resistance seal between tip of glass pipette and cell membrane. Detergent saponin (or digitonin) can be introduced briefly into bath to permeabilize cell membrane (outside isolated patch area), allowing equilibration between cell interior (i) and exterior (o) (open cell‐attached configuration). Starting from original cell‐attached situation, patch membrane can be mechanically pulled off leading to formation of closed membrane vesicle in pipette tip. Outer surface of vesicle can be disrupted by passing pipette tip briefly through air‐water interface of bath, leading to excised inside‐out membrane configuration. If short pulse of suction is applied to pipette in cell‐attached mode, membrane patch can be broken and direct continuity between pipette and cell interior established. Currents flowing across entire cell membrane can now be recorded (whole‐cell recording). If pipette is then pulled away from cell, excised membrane fragments will reseal so that excised outside‐out membrane patch is obtained.

From Petersen and Petersen 99
Figure 4. Figure 4.

Measurement of current fluctuation with help of patch‐clamp whole‐cell recording configuration. Recording obtained from single isolated rat parotid acinar cell dialyzed with K+‐rich and Ca2+‐free, ethylene glycol‐bis(β‐aminoethylether)‐N,N'‐tetraacetic acid (EGTA)‐containing solution. Holding potential was −60 mV. A: single trace of whole‐cell current elicited by 80 mV depolarizing voltage step. B: mean current elicited by 80 mV depolarizing voltage steps obtained by averaging 20 single sweeps. C: current fluctuation obtained by subtracting trace B from trace A.

From Maruyama et al. 62
Figure 5. Figure 5.

Single‐channel current recording from excised (inside‐out) pig pancreatic acinar membrane patches. A: single‐channel current traces from inside‐out membrane patch exposed to extracellular solution on outside (pipette) and intracellular solution on inside (bath) (see cartoon), a: Ca2+ concentration in bath ([Ca2+]i) was 10‐8 M; b: in same patch, [Ca2+]i was adjusted to 10‐7 M. Membrane potentials (MP) at which individual traces were obtained are indicated. Horizontal dashed lines, current level when channel is closed. Upward deflections represent outward current. Upper 2 traces, response at membrane potential of 40 mV. Although channel was generally open, there were times when channel went from open state into state of prolonged closure. B: single‐channel current‐voltage (i/V) relationships from excised membrane patches. Circles, experiment shown in A (normal Na+/K+ gradients). Closed circles, [Ca2+]i = 10‐8 M; open circles, [Ca2+]i = 10‐7 M. Triangles, different experiment in which intracellular solution was used on both sides of membrane (K+/K+). Closed triangles, [Ca2+]i = 10‐8 M; open triangles, [Ca2+]i = 10‐7 M. C: relationship between open‐state probability (P) of K+ channel and membrane potential (MP) at 3 different levels of [Ca2+]i: open circles, 10‐8 M; closed circles, 10‐7 M; and open and closed squares, 10‐6 M. Open and closed circles, experiment shown in A. Triangles (dashed line), experiment on in situ (cell‐attached) patch. In intact cell [Ca2+]i was apparently between 10‐7 M and 10‐8 M.

From Maruyama et al. 69. Reprinted by permission from Nature, copyright 1983, Macmillan Journals Limited
Figure 6. Figure 6.

Single‐channel current recording in cell‐attached configuration (isolated pig pancreatic acinar cell). Cartoon shows recording configuration. Recording pipette is filled with intracellular solution (high K+ concentration), whereas bath fluid is extracellular solution (high Na+ concentration). Traces, single‐channel currents through K+ channel in isolated patch area at different pipette potentials (Vp).

From Petersen 93
Figure 7. Figure 7.

Effects of intracellular and extracellular tetraethylammonium (TEA) on whole‐cell currents in pig pancreatic acinar cell evoked by depolarizing or hyperpolarizing voltage steps from holding potential (HP) of −40 mV. Pipette contained intracellular K+‐rich solution to which 2 mM TEA had been added. Bath was filled with extracellular Na+‐rich solution. A: whole‐cell currents before, during presence of TEA 2 mM) in bath, and after return to control situation. Voltage steps of ±20 to 70 mV (in one case 80 mV) were applied. B: whole‐cell current‐voltage relationship in control situations before and after TEA application and in presence of 2 mM TEA in bath.

From Iwatsuki and Petersen 42
Figure 8. Figure 8.

Whole‐cell currents in isolated pig pancreas acinar cell at different intracellular ionized Ca2+ concentrations (strongly buffered Ca2+‐EGTA solution) plotted as function of membrane potential.

From Maruyama and Petersen 67
Figure 9. Figure 9.

Simplified diagram for transport of amino acids and cations across basolateral plasma membrane of acinar cells. Na+‐amino acid cotransport system is coupled to Na+‐K+ pump that is closely linked to the K+ channels, i, Inner surface of plasma membrane; o, outer surface of plasma membrane.

From Singh and Petersen 115
Figure 10. Figure 10.

Whole‐cell recording from isolated mouse pancreatic acinar cell with Na+‐rich solutions on both sides of membrane. Transmembrane current at membrane voltage of 0 mV after external addition and removal of l‐alanine (ala) or N‐methylaminoiso‐butyric acid (MeAIB). Inward currents are plotted downward.

From Jauch et al. 44
Figure 11. Figure 11.

Single‐channel current recording from basolateral membrane patch in pig pancreatic acinar cell. All records were obtained in cell‐attached configuration. Recording pipette was filled with K+‐rich (intracellular) solution containing 2.5 mM Ca2+, whereas bath was filled with Na+‐rich (extracellular) solution. Resting membrane potential was about −60 mV. Upper traces, recording made on slow time base. In control situation there are only a few inward current steps. Application of 10‐6 M cholecystokinin‐5 (CCK‐5) to bath evokes, after latency of ∼18 s, a clear and sustained increase in frequency of channel openings (inward current steps). Single‐channel current amplitude is also enhanced because of membrane hyperpolarization (to about −80 mV). Lower graphs, plots of channel open‐state probability as function of membrane potential in control situation and during stimulation with 10‐6 M and 5 X 10‐6 M CCK‐5. Inset, single‐channel current recorded on fast time base together with idealized current trace obtained from computerized threshold analysis of digitized data (bottom trace). Inset trace was obtained in presence of CCK‐5 10‐6 M) at membrane potential of −50 mV.

From Suzuki et al. 118
Figure 12. Figure 12.

Whole‐cell voltage‐clamp current recordings from single pig pancreatic acinar cell. Bath contained extracellular Na+‐rich solution and pipette was filled with intracellular K+‐rich solution without Ca2+ (containing 0.5 mM EGTA). Currents associated with depolarizing voltage steps are shown as upward deflections (outward current) and with hyperpolarizing steps as downward deflections. Holding potential was −40 mV, and 90 ms voltage steps to −20, 0, +20, and +40, as well as −60, −80, and in one case −100 mV, were applied. Control, currents recorded before stimulation; CCK‐5, 3 min after start of continued exposure to 10‐6 M CCK‐5; CCK‐5 with TEA, 3 min after addition of 5 mM TEA still in presence of CCK‐5; TEA, 3 min after discontinuation of CCK‐5 stimulation but still in presence of TEA. Relationship between steady‐state currents and membrane potential in different experimental situations obtained from displayed current traces is shown in graph.

From Suzuki et al. 118
Figure 13. Figure 13.

Activation of single‐channel current in mouse pancreatic acinar cell by local CCK application. A: schematic of experimental system. a: Recording from in situ membrane patch. Recording pipette was always filled with extracellular bath solution. No agonist was present in pipette solution. CCK pipette contained the octapeptide CCK‐8 in concentration of 5 μM. Spontaneous diffusion of CCK from tip of micropipette was used as means of stimulation. b: After in situ recording patch‐clamp pipette was withdrawn from acinus to obtain excised inside‐out membrane patch. Single‐channel recordings were made in symmetrical saline solutions or bath fluid was changed to one having an intracellular composition. B: a‐h: consecutive traces from same membrane patch. a: Recording situation as described in A but with CCK pipette far away from acinus under investigation; pipette potential +40 mV. b‐d: Three continuous traces obtained 20, 30, and 40 s, respectively, after tip of CCK pipette had been brought close to patch pipette; pipette potential +40 mV. e: 50 s after start of CCK stimulation. Star, potential of patch pipette was changed (in steps of 5 mV) from +40 to +20 mV. f: Pipette potential +20 mV. Plus sign, potential was changed to 0. g: Pipette potential −40 mV. h: currents from same patch after excision so now inside out [recording situation as shown in A (b)]. Symmetrical extracellular saline solutions; pipette potential +50 mV. i,j: Currents recorded from another excised (inside‐out) membrane patch with intracellular solution (high K+, low Na+) in bath and normal extracellular solution (high Na+, low K+) in pipette. Pipette potential +60 mV in i and −60 mV in j. Dashed lines, current level when all channels are closed. Downward deflections represent current from extracellular to intracellular side of membrane.

From Maruyama and Petersen 65. Reprinted by permission from Nature, copyright 1982, Macmillan Journals Limited
Figure 14. Figure 14.

Conductance change during response to nerve stimulation using two intracellular micro‐electrodes in neighboring cells of cockroach salivary gland acinus (see Fig. 2. A: upper trace, current pulses; lower trace, electrotonic potentials. Period of stimulation indicated by thickening of traces due to stimulus artifacts. Extracellular K+ concentration = 10 mM. B: upper and lower envelopes of voltage trace in A.

From Ginsborg et al. 25
Figure 15. Figure 15.

Intracellular microelectrode recording from pig pancreatic acinar cell. Effects of external Ca2+ removal on ACh‐evoked membrane potential changes. Upper trace, recording during control conditions with normal extracellular Ca2+ levels in superfusion fluid. Lower 2 traces, part of continuous recording made 20 min after switch to superfusion with Ca2+‐free solution containing Ca2+‐chelating agent EGTA 10‐4 M). Horizontal bars, period of 5 X 10‐7 M ACh superfusion.

From Pearson et al. 81
Figure 16. Figure 16.

Effects of acetylcholine (ACh) and epinephrine (Epi), applied by microionophoresis, on membrane potential and resistance of mouse parotid acinar cells. Resting potential of each cell is written to left of its potential recording. Pulses on potential record are produced by passage of rectangular current pulses 2 nA, 100 ms, 1 s‐1) through recording microelectrode. Shorter intervals on time marker trace (top) represent 1 s.

From Roberts and Petersen 109
Figure 17. Figure 17.

Comparison between effect of nerve stimulation (FS) and local acetylcholine (ACh) application on rat pancreatic acinar cell membrane potential and resistance. Upper traces, pen recordings; lower traces, oscilloscope photographs taken at times indicated in pen recording. Calibration in oscilloscope photographs: vertical, 10 mV and 1 nA; horizontal, 20 ms.

From Petersen 90
Figure 18. Figure 18.

Patch‐clamp whole‐cell recording of Ca2+‐evoked capacitance changes in isolated rat pancreatic acinar cells. Cell was dialyzed with K+‐glutamate solution containing Mg2+ and ATP, and Ca2+‐EGTA buffered solutions were used to attain specified free Ca2+ concentrations inside cell. In each of 3 records (A, B, and C), trace labeled C represents capacitance, whereas trace labeled G represents conductance. A: Ca2+ concentration in cell <10‐9 M (nominally Ca2+‐free solution, 1 mM EGTA); B and C: Ca2+ concentration in cell = 5 X 10‐7 M 1 mM Ca2+, 1.2 mM EGTA).

From Maruyama 60
Figure 19. Figure 19.

Dog submaxillary gland in vivo. Submaxillary gland circulation was isolated so that venous blood draining from gland could be collected. Following rest period, chorda tympani (parasympathetic nerves) was stimulated at 20 Hz. Kv, venous plasma K+ concentration (mM); Ks salivary K+ concentration (mM); V, saliva flow rate in mg·g‐1·min‐1; ABF, arterial blood flow through gland in ml·g‐1‐min‐1. At beginning of stimulation K+ concentration in saliva was >40 mM, whereas in steady state it was only 6 mM. At beginning of stimulation concentration of K+ in venous plasma reached >11 mM and later settled down to 2.6 mM.

From Burgen 5
Figure 20. Figure 20.

Semiquantitative model of transport events in exocrine acini. Three different aspects are, for clarity, represented by three separate cells. Upper cell, intracellular Ca2+ release by transmitter (e.g., acetylcholine) or hormone (e.g., epinephrine) and Ca2+ activation of K+ and Cl conductance pathways. Middle cell, relation between transport rates via different routes in steady‐state stimulation situation. Lower cell, overall electrical circuit.

From Suzuki and Petersen 119


Figure 1.

Two simple models explaining stimulant‐evoked Cl uptake and membrane hyperpolarization in exocrine acinar cells. Left: model originally proposed by Lundberg 54 for salivary glands. Autonomic neurotransmitters activate directly an electrogenic active Cl transport. Right: model proposed by Petersen and Maruyama 97,119. Stimulant secretagogues [acetylcholine (ACh) and norepinephrine (NA) in salivary glands, other neurotransmitters or hormones in other glands; see ref. 89] evoke an increase in intracellular Ca2+ concentration ([Ca2+]i) that activates K+ channels. Exit of K+ allows K+ reuptake through Na+‐K+ pump and Na+‐K+‐2Cl cotransporter. In steady‐state stimulated condition all these processes constitute operational Cl pump that is electrogenic. i, Inner surface of plasma membrane; o, outer surface of plasma membrane.

From Petersen 93


Figure 2.

Intracellular recording with 2 separate microelectrodes from neighboring acinar cells. A: recording configuration. Both intracellular electrodes can be used for membrane potential recording and for current injection. If electrode is filled with secretagogue (shaded electrodes), current injection of appropriate polarity can be used for local agonist application. B: membrane potential recordings from mouse pancreas. Lower trace, rectangular hyperpolarizing current pulse 10‐9 A) of 100 ms duration. Upper traces, effects on membrane potential in injected cell as well as its neighbor during rest (a) and during acetylcholine application (b). Vertical calibration, 10 mV and 10‐9 A; horizontal calibration, 20 ms.

From Petersen and Ueda 104


Figure 3.

Different patch‐clamp recording configurations and procedures used to establish them. Patch‐clamp experiments start in cell‐attached configuration (top left). Tip of fire‐polished recording pipette is gently brought into contact with cell surface and slight suction is applied to pipette resulting in high‐resistance seal between tip of glass pipette and cell membrane. Detergent saponin (or digitonin) can be introduced briefly into bath to permeabilize cell membrane (outside isolated patch area), allowing equilibration between cell interior (i) and exterior (o) (open cell‐attached configuration). Starting from original cell‐attached situation, patch membrane can be mechanically pulled off leading to formation of closed membrane vesicle in pipette tip. Outer surface of vesicle can be disrupted by passing pipette tip briefly through air‐water interface of bath, leading to excised inside‐out membrane configuration. If short pulse of suction is applied to pipette in cell‐attached mode, membrane patch can be broken and direct continuity between pipette and cell interior established. Currents flowing across entire cell membrane can now be recorded (whole‐cell recording). If pipette is then pulled away from cell, excised membrane fragments will reseal so that excised outside‐out membrane patch is obtained.

From Petersen and Petersen 99


Figure 4.

Measurement of current fluctuation with help of patch‐clamp whole‐cell recording configuration. Recording obtained from single isolated rat parotid acinar cell dialyzed with K+‐rich and Ca2+‐free, ethylene glycol‐bis(β‐aminoethylether)‐N,N'‐tetraacetic acid (EGTA)‐containing solution. Holding potential was −60 mV. A: single trace of whole‐cell current elicited by 80 mV depolarizing voltage step. B: mean current elicited by 80 mV depolarizing voltage steps obtained by averaging 20 single sweeps. C: current fluctuation obtained by subtracting trace B from trace A.

From Maruyama et al. 62


Figure 5.

Single‐channel current recording from excised (inside‐out) pig pancreatic acinar membrane patches. A: single‐channel current traces from inside‐out membrane patch exposed to extracellular solution on outside (pipette) and intracellular solution on inside (bath) (see cartoon), a: Ca2+ concentration in bath ([Ca2+]i) was 10‐8 M; b: in same patch, [Ca2+]i was adjusted to 10‐7 M. Membrane potentials (MP) at which individual traces were obtained are indicated. Horizontal dashed lines, current level when channel is closed. Upward deflections represent outward current. Upper 2 traces, response at membrane potential of 40 mV. Although channel was generally open, there were times when channel went from open state into state of prolonged closure. B: single‐channel current‐voltage (i/V) relationships from excised membrane patches. Circles, experiment shown in A (normal Na+/K+ gradients). Closed circles, [Ca2+]i = 10‐8 M; open circles, [Ca2+]i = 10‐7 M. Triangles, different experiment in which intracellular solution was used on both sides of membrane (K+/K+). Closed triangles, [Ca2+]i = 10‐8 M; open triangles, [Ca2+]i = 10‐7 M. C: relationship between open‐state probability (P) of K+ channel and membrane potential (MP) at 3 different levels of [Ca2+]i: open circles, 10‐8 M; closed circles, 10‐7 M; and open and closed squares, 10‐6 M. Open and closed circles, experiment shown in A. Triangles (dashed line), experiment on in situ (cell‐attached) patch. In intact cell [Ca2+]i was apparently between 10‐7 M and 10‐8 M.

From Maruyama et al. 69. Reprinted by permission from Nature, copyright 1983, Macmillan Journals Limited


Figure 6.

Single‐channel current recording in cell‐attached configuration (isolated pig pancreatic acinar cell). Cartoon shows recording configuration. Recording pipette is filled with intracellular solution (high K+ concentration), whereas bath fluid is extracellular solution (high Na+ concentration). Traces, single‐channel currents through K+ channel in isolated patch area at different pipette potentials (Vp).

From Petersen 93


Figure 7.

Effects of intracellular and extracellular tetraethylammonium (TEA) on whole‐cell currents in pig pancreatic acinar cell evoked by depolarizing or hyperpolarizing voltage steps from holding potential (HP) of −40 mV. Pipette contained intracellular K+‐rich solution to which 2 mM TEA had been added. Bath was filled with extracellular Na+‐rich solution. A: whole‐cell currents before, during presence of TEA 2 mM) in bath, and after return to control situation. Voltage steps of ±20 to 70 mV (in one case 80 mV) were applied. B: whole‐cell current‐voltage relationship in control situations before and after TEA application and in presence of 2 mM TEA in bath.

From Iwatsuki and Petersen 42


Figure 8.

Whole‐cell currents in isolated pig pancreas acinar cell at different intracellular ionized Ca2+ concentrations (strongly buffered Ca2+‐EGTA solution) plotted as function of membrane potential.

From Maruyama and Petersen 67


Figure 9.

Simplified diagram for transport of amino acids and cations across basolateral plasma membrane of acinar cells. Na+‐amino acid cotransport system is coupled to Na+‐K+ pump that is closely linked to the K+ channels, i, Inner surface of plasma membrane; o, outer surface of plasma membrane.

From Singh and Petersen 115


Figure 10.

Whole‐cell recording from isolated mouse pancreatic acinar cell with Na+‐rich solutions on both sides of membrane. Transmembrane current at membrane voltage of 0 mV after external addition and removal of l‐alanine (ala) or N‐methylaminoiso‐butyric acid (MeAIB). Inward currents are plotted downward.

From Jauch et al. 44


Figure 11.

Single‐channel current recording from basolateral membrane patch in pig pancreatic acinar cell. All records were obtained in cell‐attached configuration. Recording pipette was filled with K+‐rich (intracellular) solution containing 2.5 mM Ca2+, whereas bath was filled with Na+‐rich (extracellular) solution. Resting membrane potential was about −60 mV. Upper traces, recording made on slow time base. In control situation there are only a few inward current steps. Application of 10‐6 M cholecystokinin‐5 (CCK‐5) to bath evokes, after latency of ∼18 s, a clear and sustained increase in frequency of channel openings (inward current steps). Single‐channel current amplitude is also enhanced because of membrane hyperpolarization (to about −80 mV). Lower graphs, plots of channel open‐state probability as function of membrane potential in control situation and during stimulation with 10‐6 M and 5 X 10‐6 M CCK‐5. Inset, single‐channel current recorded on fast time base together with idealized current trace obtained from computerized threshold analysis of digitized data (bottom trace). Inset trace was obtained in presence of CCK‐5 10‐6 M) at membrane potential of −50 mV.

From Suzuki et al. 118


Figure 12.

Whole‐cell voltage‐clamp current recordings from single pig pancreatic acinar cell. Bath contained extracellular Na+‐rich solution and pipette was filled with intracellular K+‐rich solution without Ca2+ (containing 0.5 mM EGTA). Currents associated with depolarizing voltage steps are shown as upward deflections (outward current) and with hyperpolarizing steps as downward deflections. Holding potential was −40 mV, and 90 ms voltage steps to −20, 0, +20, and +40, as well as −60, −80, and in one case −100 mV, were applied. Control, currents recorded before stimulation; CCK‐5, 3 min after start of continued exposure to 10‐6 M CCK‐5; CCK‐5 with TEA, 3 min after addition of 5 mM TEA still in presence of CCK‐5; TEA, 3 min after discontinuation of CCK‐5 stimulation but still in presence of TEA. Relationship between steady‐state currents and membrane potential in different experimental situations obtained from displayed current traces is shown in graph.

From Suzuki et al. 118


Figure 13.

Activation of single‐channel current in mouse pancreatic acinar cell by local CCK application. A: schematic of experimental system. a: Recording from in situ membrane patch. Recording pipette was always filled with extracellular bath solution. No agonist was present in pipette solution. CCK pipette contained the octapeptide CCK‐8 in concentration of 5 μM. Spontaneous diffusion of CCK from tip of micropipette was used as means of stimulation. b: After in situ recording patch‐clamp pipette was withdrawn from acinus to obtain excised inside‐out membrane patch. Single‐channel recordings were made in symmetrical saline solutions or bath fluid was changed to one having an intracellular composition. B: a‐h: consecutive traces from same membrane patch. a: Recording situation as described in A but with CCK pipette far away from acinus under investigation; pipette potential +40 mV. b‐d: Three continuous traces obtained 20, 30, and 40 s, respectively, after tip of CCK pipette had been brought close to patch pipette; pipette potential +40 mV. e: 50 s after start of CCK stimulation. Star, potential of patch pipette was changed (in steps of 5 mV) from +40 to +20 mV. f: Pipette potential +20 mV. Plus sign, potential was changed to 0. g: Pipette potential −40 mV. h: currents from same patch after excision so now inside out [recording situation as shown in A (b)]. Symmetrical extracellular saline solutions; pipette potential +50 mV. i,j: Currents recorded from another excised (inside‐out) membrane patch with intracellular solution (high K+, low Na+) in bath and normal extracellular solution (high Na+, low K+) in pipette. Pipette potential +60 mV in i and −60 mV in j. Dashed lines, current level when all channels are closed. Downward deflections represent current from extracellular to intracellular side of membrane.

From Maruyama and Petersen 65. Reprinted by permission from Nature, copyright 1982, Macmillan Journals Limited


Figure 14.

Conductance change during response to nerve stimulation using two intracellular micro‐electrodes in neighboring cells of cockroach salivary gland acinus (see Fig. 2. A: upper trace, current pulses; lower trace, electrotonic potentials. Period of stimulation indicated by thickening of traces due to stimulus artifacts. Extracellular K+ concentration = 10 mM. B: upper and lower envelopes of voltage trace in A.

From Ginsborg et al. 25


Figure 15.

Intracellular microelectrode recording from pig pancreatic acinar cell. Effects of external Ca2+ removal on ACh‐evoked membrane potential changes. Upper trace, recording during control conditions with normal extracellular Ca2+ levels in superfusion fluid. Lower 2 traces, part of continuous recording made 20 min after switch to superfusion with Ca2+‐free solution containing Ca2+‐chelating agent EGTA 10‐4 M). Horizontal bars, period of 5 X 10‐7 M ACh superfusion.

From Pearson et al. 81


Figure 16.

Effects of acetylcholine (ACh) and epinephrine (Epi), applied by microionophoresis, on membrane potential and resistance of mouse parotid acinar cells. Resting potential of each cell is written to left of its potential recording. Pulses on potential record are produced by passage of rectangular current pulses 2 nA, 100 ms, 1 s‐1) through recording microelectrode. Shorter intervals on time marker trace (top) represent 1 s.

From Roberts and Petersen 109


Figure 17.

Comparison between effect of nerve stimulation (FS) and local acetylcholine (ACh) application on rat pancreatic acinar cell membrane potential and resistance. Upper traces, pen recordings; lower traces, oscilloscope photographs taken at times indicated in pen recording. Calibration in oscilloscope photographs: vertical, 10 mV and 1 nA; horizontal, 20 ms.

From Petersen 90


Figure 18.

Patch‐clamp whole‐cell recording of Ca2+‐evoked capacitance changes in isolated rat pancreatic acinar cells. Cell was dialyzed with K+‐glutamate solution containing Mg2+ and ATP, and Ca2+‐EGTA buffered solutions were used to attain specified free Ca2+ concentrations inside cell. In each of 3 records (A, B, and C), trace labeled C represents capacitance, whereas trace labeled G represents conductance. A: Ca2+ concentration in cell <10‐9 M (nominally Ca2+‐free solution, 1 mM EGTA); B and C: Ca2+ concentration in cell = 5 X 10‐7 M 1 mM Ca2+, 1.2 mM EGTA).

From Maruyama 60


Figure 19.

Dog submaxillary gland in vivo. Submaxillary gland circulation was isolated so that venous blood draining from gland could be collected. Following rest period, chorda tympani (parasympathetic nerves) was stimulated at 20 Hz. Kv, venous plasma K+ concentration (mM); Ks salivary K+ concentration (mM); V, saliva flow rate in mg·g‐1·min‐1; ABF, arterial blood flow through gland in ml·g‐1‐min‐1. At beginning of stimulation K+ concentration in saliva was >40 mM, whereas in steady state it was only 6 mM. At beginning of stimulation concentration of K+ in venous plasma reached >11 mM and later settled down to 2.6 mM.

From Burgen 5


Figure 20.

Semiquantitative model of transport events in exocrine acini. Three different aspects are, for clarity, represented by three separate cells. Upper cell, intracellular Ca2+ release by transmitter (e.g., acetylcholine) or hormone (e.g., epinephrine) and Ca2+ activation of K+ and Cl conductance pathways. Middle cell, relation between transport rates via different routes in steady‐state stimulation situation. Lower cell, overall electrical circuit.

From Suzuki and Petersen 119
References
 1. Anderson, C. R., and C. F. Stevens. Voltage‐clamp analysis of acetylcholine produced end‐plate current fluctuations of frog neuromuscular junction. J. Physiol. Lond. 235: 655–691, 1973.
 2. Baker, P. F. The regulation of intracellular calcium in giant axons of Loligo.and Myxicola. Ann. NY Acad. Sci. 307: 250–268, 1978.
 3. Berridge, M. J., and R. F. Irvine. Inositol trisphosphate, a novel second messenger in cellular signal transduction. Nature Lond. 312: 315–321, 1984.
 4. Bundgaard, M., M. Møller, and J. H. Poulsen. Localization of sodium pump sites in cat salivary glands. J. Physiol. Lond. 273: 339–353, 1977.
 5. Burgen, A. S. V. The secretion of potassium in saliva. J. Physiol. Lond. 132: 20–39, 1956.
 6. Burgen, A. S. V., and N. G. Emmelin. Physiology of the Salivary Glands. London: Arnold, 1961.
 7. Chandler, D. E., and J. A. Williams. Intracellular divalent cation release in pancreatic acinar cells during stimulus‐secretion coupling. I. Use of chlorotetracycline as a fluorescent probe. J. Cell Biol. 76: 371–385, 1978.
 8. Chandler, D. E., and J. A. Williams. Intracellular divalent cation release in pancreatic acinar cells during stimulus‐secretion coupling. II. Subcellular localization of the fluorescent probe chlorotetracycline. J. Cell Biol. 76: 386–399, 1978.
 9. Christensen, H. N. Organic ion transport during seven decades. The amino acids. Biochim. Biophys. Acta 779: 255–269, 1984.
 10. Evans, L. A. R., D. Pirani, D. I. Cook, and J. A. Young. Intraepithelial current flow in rat pancreatic secretory epithelia. Pfluegers Arch. 407, Suppl. 2: S107–S111, 1986.
 11. Evans, M. G., and A. Marty. Calcium‐dependent chloride currents in isolated cells from rat lacrimal glands. J. Physiol. Lond. 378: 437–460, 1986.
 12. Evans, M. G., A. Marty, and Y. Tan. Blockage of a Ca‐activated Cl conductance by furosemide in rat lacrimal glands. Pfluegers Arch. 406: 65–68, 1986.
 13. Fernandez, J. H., E. Neher, and B. D. Gomperts. Capacitance measurements reveal stepwise fusion events in degranulating mast cells. Nature Lond. 312: 453–455, 1984.
 14. Findlay, I. A patch‐clamp study of potassium channels and whole‐cell currents in acinar cells of the mouse lacrimal gland. J. Physiol. Lond. 350: 179–195, 1984.
 15. Findlay, I., M. J. Dunne, and O. H. Petersen. ATP‐sensitive inward rectifier and voltage‐ and calcium‐activated K+ channels in cultured pancreatic islet cells. J. Membr. Biol. 88: 165–172, 1985.
 16. Findlay, I., M. J. Dunne, S. Ullrich, C. B. Wollheim, and O. H. Petersen. Quinine inhibits Ca2+‐independent K+ channels whereas tetraethylammonium inhibits Ca2+‐activated K+ channels in insulin‐secreting cells. FEBS Lett. 185: 4–8, 1985.
 17. Findlay, I., and O. H. Petersen. Acetylcholine‐evoked uncoupling restricts the passage of Lucifer Yellow between pancreatic acinar cells. Cell Tissue Res. 225: 633–638, 1982.
 18. Findlay, I., and O. H. Petersen. The extent of dye‐coupling between exocrine acinar cells of the mouse pancreas. Cell Tissue Res. 232: 121–127, 1983.
 19. Findlay, I., and O. H. Petersen. Acetylcholine stimulates a Ca2+‐dependent Cl‐ conductance in mouse lacrimal acinar cells. Pfluegers Arch. 403: 328–330, 1985.
 20. Frömter, E., and J. Diamond. Route of passive ion permeation in epithelia. Nat. New Biol. 235: 9–13, 1972.
 21. Gallacher, D. V., Y. Maruyama, and O. H. Petersen. Patch‐clamp study of rubidium and potassium conductances in single cation channels from mammalian exocrine acini. Pfluegers Arch. 401: 361–367, 1984.
 22. Gallacher, D. V., and A. P. Morris. A patch‐clamp study of K+ currents in resting and acetylcholine stimulated mouse submandibular acinar cells. J. Physiol. Lond. 373: 379–395, 1986.
 23. Gallacher, D. V., and O. H. Petersen. Electrophysiology of parotid acini: effects of electrical field stimulation and ionophoresis of neurotransmitters. J. Physiol. Lond. 305: 43–57, 1980.
 24. Ginsborg, B. L., and C. R. House. Stimulus‐response coupling in gland cells. Annu. Rev. Biophys. Bioeng. 9: 55–80, 1980.
 25. Ginsborg, B. L., C. R. House, and E. M. Silinsky. Conductance changes associated with the secretory potential in the cockroach salivary gland. J. Physiol. Lond. 236: 723–731, 1974.
 26. Graf, J., and O. H. Petersen. Cell membrane potential and resistance in liver. J. Physiol. Lond. 284: 105–126, 1978.
 27. Hamill, O. P., A. Marty, E. Neher, B. Sakmann, and F. J. Sigworth. Improved patch‐clamp technique for high‐resolution current recording from cells and cell‐free membrane patches. Pfluegers Arch. 391: 85–100, 1981.
 28. Hodgkin, A. L. The Conduction of the Nervous Impulse. Liverpool, UK: Liverpool Univ. Press, 1964.
 29. Hokin, L. E. Receptors and phosphoinositide‐generated second messengers. Annu. Rev. Biochem. 54: 205–235, 1985.
 30. Hootman, S. R., D. L. Ochs, and J. A. Williams. Intracellular mediators of Na+‐K+ pump activity in guinea pig pancreatic acinar cells. Am. J. Physiol. 249 (Gastrointest. Liver Physiol.12): G470–G478, 1985.
 31. Horn, R., and J. B. Patlak. Single‐channel currents from excised patches of muscle membrane. Proc. Natl. Acad. Sci. USA 77: 6930–6934, 1980.
 32. Iwatsuki, N., and O. H. Petersen. Acetylcholine‐like effects of intracellular calcium application in pancreatic acinar cells. Nature Lond. 268: 147–149, 1977.
 33. Iwatsuki, N., and O. H. Petersen. Pancreatic acinar cells: localization of acetylcholine receptors and the importance of chloride and calcium for acetylcholine‐evoked depolarization. J. Physiol. Lond. 269: 723–733, 1977.
 34. Iwatsuki, N., and O. H. Petersen. Pancreatic acinar cells: the acetylcholine equilibrium potential and its ionic dependency. J. Physiol. Lond. 269: 735–751, 1977.
 35. Iwatsuki, N., and O. H. Petersen. Electrical coupling and uncoupling of exocrine acinar cells. J. Cell Biol. 79: 533–545, 1978.
 36. Iwatsuki, N., and O. H. Petersen. Intracellular Ca2+ injection causes membrane hyperpolarization and conductance increase in lacrimal acinar cells. Pfluegers Arch. 377: 185–187, 1978.
 37. Iwatsuki, N., and O. H. Petersen. In vitro action of bombesin on amylase secretion, membrane potential and membrane resistance in rat and mouse pancreatic acinar cells. A comparison with other secretagogues. J. Clin. Invest. 61: 41–46, 1978.
 38. Iwatsuki, N., and O. H. Petersen. Direct visualization of cell to cell coupling: transfer of fluorescent probes in living mammalian pancreatic acini. Pfluegers Arch. 380: 277–281, 1979.
 39. Iwatsuki, N., and O. H. Petersen. Pancreatic acinar cells: the effect of CO2, NH4Cl and acetylcholine on intercellular communication. J. Physiol. Lond. 291: 317–326, 1979.
 40. Iwatsuki, N., and O. H. Petersen. Amino acid‐evoked membrane potential and resistance changes in pancreatic acinar cells. Pfluegers Arch. 386: 153–159, 1980.
 41. Iwatsuki, N., and O. H. Petersen. Amino acids evoke short‐latency membrane conductance increase in pancreatic acinar cells. Nature Lond. 283: 492–494, 1980.
 42. Iwatsuki, N., and O. H. Petersen. Action of tetraethylammonium on calcium‐activated potassium channels in pig pancreatic acinar cells studied by patch‐clamp single‐channel and whole‐cell current recording. J. Membr. Biol. 86: 139–144, 1985.
 43. Iwatsuki, N., and O. H. Petersen. Inhibition of Ca2+‐activated K+ channels in pig pancreatic acinar cells by Ba2+, Ca2+, quinine and quinidine. Biochim. Biophys. Acta 819: 249–257, 1985.
 44. Jauch, P., Y. Maruyama, O. H. Petersen, H. A. Kolb, and P. Läuger. Electrophysiological study of the alanine‐sodium cotransporter in pancreatic acinar cells. In: Ion Gradient‐Coupled Transport, edited by F. Alvardo and C. H. Van Os. Amsterdam: Elsevier, 1986, p. 241–244. (INSERM Symp. Ser., vol. 26.).
 45. Jauch, P., O. H. Petersen, and P. Läuger. Electrogenic properties of the sodium‐alanine cotransporter in pancreatic acinar cells: I. Tight‐seal whole‐cell recordings. J. Membr. Biol. 94: 99–115, 1986.
 46. Katz, B., and R. Miledi. The statistical nature of the acetylcholine potential and its molecular components. J. Physiol. Lond. 224: 665–699, 1972.
 47. Kondo, S., and I. Schulz. Calcium ion uptake in isolated pancreas cells induced by secretagogues. Biochim. Biophys. Acta 419: 76–92, 1976.
 48. Kurtzer, R. J., and M. L. Roberts. Calcium‐dependent K+ efflux from rat submandibular gland. The effects of trifluoroperazine and quinidine. Biochim. Biophys. Acta 693: 479–484, 1982.
 49. Laugier, R., and O. H. Petersen. Effects of intracellular EGTA injection on stimulant‐evoked membrane potential and resistance changes in pancreatic acinar cells. Pfluegers Arch. 386: 147–152, 1980.
 50. Laugier, R., and O. H. Petersen. Pancreatic acinar cells: electrophysiological evidence for stimulant‐evoked increase in membrane calcium permeability in the mouse. J. Physiol. Lond. 303: 61–72, 1980.
 51. Laugier, R., and O. H. Petersen. Two different types of electrogenic amino acid action on pancreatic acinar cells. Biochim. Biophys. Acta 641: 216–221, 1981.
 52. Loewenstein, W. R. Junctional intercellular communication: the cell‐to‐cell membrane channel. Physiol. Rev. 61: 829–913, 1981.
 53. Lundberg, A. The electrophysiology of the submaxillary gland of the cat. Acta Physiol. Scand. 35: 1–25, 1955.
 54. Lundberg, A. Secretory potentials and secretion in the sublingual gland of the cat. Nature Lond. 177: 1080–1081, 1956.
 55. Lundberg, A. Secretory potentials in the sublingual gland of the cat. Acta Physiol. Scand. 40: 21–34, 1957.
 56. Lundberg, A. The mechanism of establishment of secretory potentials in sublingual gland cells. Acta Physiol. Scand. 40: 35–58, 1957.
 57. Lundberg, A. Anionic dependence of secretion and secretory potentials in sublingual gland cells. Acta Physiol. Scand. 40: 101–112, 1957.
 58. Lundberg, A. Electrophysiology of salivary glands. Physiol. Rev. 38: 21–40, 1958.
 59. Marty, A., Y. P. Tan, and A. Trautmann. Three types of calcium‐dependent channel in rat lacrimal glands. J. Physiol. Lond. 357: 293–325, 1984.
 60. Maruyama, Y. Ca2+‐induced excess capacitance fluctuation studied by phase‐sensitive detection method in exocrine pancreatic acinar cells. Pfluegers Arch. 407: 561–563, 1986.
 61. Maruyama, Y., D. V. Gallacher, and O. H. Petersen. Voltage and Ca2+‐activated K+ channel in baso‐lateral acinar cell membranes of mammalian salivary glands. Nature Lond. 302: 827–829, 1983.
 62. Maruyama, Y., A. Nishiyama, T. Izumi, N. Hoshimiya, and O. H. Petersen. Ensemble noise and current relaxation analysis of K+ current in single isolated salivary acinar cells from rat. Pfluegers Arch. 406: 69–72, 1986.
 63. Maruyama, Y., A. Nishiyama, and T. Teshima. Two types of cation channels in the basolateral cell membrane of human salivary gland acinar cells. Jpn. J. Physiol. 36: 219–223, 1986.
 64. Maruyama, Y., and O. H. Petersen. Single‐channel currents in isolated patches of plasma membrane from basal surface of pancreatic acini. Nature Lond. 299: 159–161, 1982.
 65. Maruyama, Y., and O. H. Petersen. Cholecystokinin activation of single‐channel currents is mediated by internal messenger in pancreatic acinar cells. Nature Lond. 300: 61–63, 1982.
 66. Maruyama, Y., and O. H. Petersen. Voltage clamp study of stimulant‐evoked currents in mouse pancreatic acinar cells. Pfluegers Arch. 399: 54–62, 1983.
 67. Maruyama, Y., and O. H. Petersen. Control of K+ conductance by cholecystokinin and Ca2+ in single pancreatic acinar cells studied by the patch‐clamp technique. J. Membr. Biol. 79: 293–300, 1984.
 68. Maruyama, Y., and O. H. Petersen. Single calcium‐dependent cation channels in mouse pancreatic acinar cells. J. Membr. Biol. 81: 83–87, 1984.
 69. Maruyama, Y., O. H. Petersen, P. Flanagan, and G. T. Pearson. Quantification of Ca2+‐activated K+ channels under hormonal control in pig pancreas acinar cells. Nature Lond. 305: 228–232, 1983.
 70. Matthews, E. K., O. H. Petersen, and J. A. Williams. Pancreatic acinar cells: acetylcholine‐induced membrane depolarization, calcium efflux and amylase release. J. Physiol. Load. 234: 689–701, 1973.
 71. Meda, P., I. Findlay, E. Kolod, L. Orci, and O. H. Petersen. Short and reversible uncoupling evokes little change in the gap junctions of pancreatic acinar cells. J. Ultrastruct. Res. 83: 69–84, 1983.
 72. Muallem, S., M. Schoeffield, S. Pandol, and G. Sachs. Inositol trisphosphate modification of ion transport in rough endoplasmic reticulum. Proc. Natl. Acad. Sci. USA 82: 4433–4437, 1985.
 73. Nagel, W. Inhibition of potassium conductance by barium in frog skin epithelium. Biochim. Biophys. Acta 552: 346–357, 1979.
 74. Nauntofte, B., and J. H. Poulsen. Effects of Ca2+ and furosemide on Cl‐ transport and O2 uptake in rat parotid acini. Am. J. Physiol. 251 (Cell Physiol. 20): C175–C185, 1986.
 75. Neher, E., and A. Marty. Discrete changes of cell membrane capacitance observed under conditions of enhanced secretion in bovine adrenal chromaffin cells. Proc. Natl. Acad. Sci. USA 79: 6712–6716, 1982.
 76. Neher, E., and B. Sakmann. Single‐channel currents recorded from membrane of denervated frog muscle fibres. Nature Lond. 260: 799–802, 1976.
 77. Neyton, J., and A. Trautmann. Single‐channel currents of an intercellular junction. Nature Lond. 317: 331–335, 1985.
 78. Nishiyama, A., and O. H. Petersen. Membrane potential and resistance measurement in acinar cells from salivary glands in vitro: effect of acetylcholine. J. Physiol. Lond. 242: 173–188, 1974.
 79. Ochs, D. L., J. I. Korenbrot, and J. A. Williams. Relationship between free cytosolic calcium and amylase release by pancreatic acini. Am. J. Physiol. 249 (Gastrointest. Liver Physiol.12): G389–G398, 1985.
 80. Palade, G. Intracellular aspects of the process of protein synthesis. Science Wash. DC 189: 347–358, 1975.
 81. Pearson, G. T., P. M. Flanagan, and O. H. Petersen. Neural and hormonal control of membrane conductance in the pig pancreatic acinar cell. Am. J. Physiol. 247 (Gastrointest. Liver Physiol.10): G520–G526, 1984.
 82. Pearson, G. T., and O. H. Petersen. Nervous control of membrane conductance in mouse lacrimal acinar cells. Pfluegers Arch. 400: 51–59, 1984.
 83. Pedersen, G. L., and O. H. Petersen. Membrane potential measurement in parotid acinar cells. J. Physiol. Lond. 234: 217–227, 1973.
 84. Petersen, O. H. Some factors influencing stimulation‐induced release of potassium from the cat submandibular gland to fluid perfused through the gland. J. Physiol. Lond. 208: 431–447, 1970.
 85. Petersen, O. H. Formation of saliva and potassium transport in the perfused cat submandibular gland. J. Physiol. Lond. 216: 129–142, 1971.
 86. Petersen, O. H. Electrogenic sodium pump in pancreatic acinar cells. Proc. R. Soc. Lond. B Biol. Sci. 184: 115–119, 1973.
 87. Petersen, O. H. Electrical coupling between pancreatic acinar cells. J. Physiol. Lond. 250: 2P–4P, 1975.
 88. Petersen, O. H. Electrophysiology of mammalian gland cells. Physiol. Rev. 56: 535–577, 1976.
 89. Petersen, O. H. Electrophysiology of Gland Cells. New York: Academic, 1981.
 90. Petersen, O. H. Stimulus‐excitation coupling in plasma membranes of pancreatic acinar cells. Biochim. Biophys. Acta 694: 163–184, 1982.
 91. Petersen, O. H. Importance of electrical cell‐cell communication in secretory epithelia. In: Gap Junctions, edited by M. V. Bennett and D. C. Spray. Cold Spring Harbor, NY: Cold Spring Harbor, 1985, p. 315–324.
 92. Petersen, O. H. Potassium channels and fluid secretion. News Physiol. Sci. 1: 92–95, 1986.
 93. Petersen, O. H. Calcium‐activated potassium channels and fluid secretion by exocrine glands. Am. J. Physiol. 251 (Gastrointest. Liver Physiol.14): G1–G13, 1986.
 94. Petersen, O. H., I. Findlay, N. Iwatsuki, J. Singh, D. V. Gallacher, C. M. Fuller, G. T. Pearson, M. J. Dunne, and A. P. Morris. Human pancreatic acinar cells: studies of stimulus‐secretion coupling. Gastroenterology 89: 109–117, 1985.
 95. Petersen, O. H., and Y. Maruyama. Cholecystokinin and acetylcholine activation of single‐channel currents via second messenger in pancreatic acinar cells. In: Single‐Channel Recording, edited by B. Sakmann and E. Neher. New York: Plenum, 1983, p. 425–435.
 96. Petersen, O. H., and Y. Maruyama. What is the mechanism of the calcium influx to pancreatic acinar cells evoked by secretagogues? Pfluegers Arch. 396: 82–84, 1983.
 97. Petersen, O. H., and Y. Maruyama. Calcium‐activated potassium channels and their role in secretion. Nature Lond. 307: 693–696, 1984.
 98. Petersen, O. H., Y. Maruyama, J. Graf, R. Laugier, A. Nishiyama, and G. T. Pearson. Ionic currents across pancreatic acinar cell membranes and their role in fluid secretion. Philos. Trans. R. Soc. Lond. B Biol. Sci. 296: 151–166, 1981.
 99. Petersen, O. H., and C. C. H. Petersen. The patch‐clamp technique: recording ionic currents through single pores in the cell membrane. News Physiol. Sci. 1: 5–8, 1986.
 100. Petersen, O. H., and H. G. Philpott. Pancreatic acinar cells: the anion selectivity of the acetylcholine‐opened chloride pathway. J. Physiol. Lond. 306: 481–492, 1980.
 101. Petersen, O. H., and J. H. Poulsen. Secretory potentials, potassium transport and secretion in the cat submandibular gland during perfusion with sulphate Locke's solution. Experientia 24: 919–920, 1968.
 102. Petersen, O. H., and J. Singh. Amino acid‐evoked membrane current in voltage‐clamped mouse pancreatic acini. J. Physiol. Lond. 319: 99P–100P, 1981.
 103. Petersen, O. H., and J. Singh. Acetylcholine‐evoked potassium release in the mouse pancreas. J. Physiol. Lond. 365: 319–329, 1985.
 104. Petersen, O. H., and N. Ueda. Pancreatic acinar cells: the role of calcium in stimulus‐secretion coupling. J. Physiol. Lond. 254: 583–606, 1976.
 105. Petersen, O. H., and N. Ueda. Secretion of fluid and amylase in the perfused rat pancreas. J. Physiol. Lond. 264: 819–835, 1977.
 106. Poulsen, J. H., and S. W. Bledsoe. Salivary gland K+ transport: in vivo studies with K+‐specific microelectrodes. Am. J. Physiol. 234 (Endocrinol. Metab. Gastrointest. Physiol.3): E79–E83, 1978.
 107. Poulsen, J. H., and B. Oakley II. Intracellular potassium ion activity in resting and stimulated mouse pancreas and submandibular gland. Proc. R. Soc. Lond. B Biol. Sci. 204: 99–104, 1979.
 108. Roberts, M. L., N. Iwatsuki, and O. H. Petersen. Parotid acinar cells: ionic dependence of acetylcholine‐evoked membrane potential changes. Pfluegers Arch. 376: 159–167, 1978.
 109. Roberts, M. L., and O. H. Petersen. Membrane potential and resistance changes induced in salivary gland acinar cells by microiontophoretic application of acetylcholine and adrenergic agonists. J. Membr. Biol. 30: 297–312, 1978.
 110. Saito, Y., T. Ozawa, H. Hayashi, and A. Nishiyama. Acetylcholine‐induced change in intracellular Cl‐ activity of the mouse lacrimal acinar cells. Pfluegers Arch. 405: 108–111, 1985.
 111. Schulz, I. Electrolyte and fluid secretion in the exocrine pancreas. In: Physiology of the Gastrointestinal Tract. (2nd ed.), edited by L. R. Johnson. New York: Raven, 1987, p. 1147–1171.
 112. Schulz, I., and H. H. Stolze. The exocrine pancreas: the role of secretagogues, cyclic nucleotides and calcium in enzyme secretion. Annu. Rev. Physiol. 42: 127–156, 1980.
 113. Sigworth, F. J. The variance of sodium current fluctuations at the node of Ranvier. J. Physiol. Lond. 307: 97–129, 1980.
 114. Silva, P., J. Stoff, M. Field, L. Fine, J. N. Forrest, and F. H. Epstein. Mechanism of active chloride secretion by shark rectal gland: role of Na‐K‐ATPase in chloride transport. Am. J. Physiol. 233 (Renal Fluid Electrolyte Physiol.2): F298–F306, 1977.
 115. Singh, J., and O. H. Petersen. The effects of l‐alanine and acetylcholine on membrane potential, 45Ca2+ and 86Rb+ efflux and amylase secretion in the isolated mouse pancreas. Q. J. Exp. Physiol. Cogn. Med. Sci. 69: 531–540, 1984.
 116. Stanley, E. F., and G. Ehrenstein. A model for exocytosis based on the opening of calcium‐activated potassium channels in vesicles. Life Sci. 37: 1985–1995, 1985.
 117. Streb, H., E. Bayerdorffer, W. Haase, R. F. Irvine, and I. Schulz. Effect of inositol‐1,4,5‐trisphosphate on isolated subcellular fractions of rat pancreas. J. Membr. Biol. 81: 241–253, 1984.
 118. Suzuki, K., C. C. H. Petersen, and O. H. Petersen. Hormonal activation of single K+ channels via internal messenger in isolated pancreatic acinar cells. FEBS Lett. 192: 307–312, 1985.
 119. Suzuki, K., and O. H. Petersen. The effects of Na+ and Cl‐ removal and of loop diuretics on acetylcholine‐evoked membrane potential changes in mouse lacrimal acinar cells. Q. J. Exp. Physiol. Cogn. Med. Sci. 70: 437–445, 1985.
 120. Suzuki, K., and O. H. Petersen. Patch‐clamp studies of K+ channels in guinea‐pig pancreatic acinar cells (Abstract). J. Physiol. Lond. 378: 62P, 1986.
 121. Trautmann, A., and A. Marty. Activation of Ca‐dependent K channels by carbamoylcholine in rat lacrimal glands. Proc. Natl. Acad. Sci. USA 81: 611–615, 1984.
 122. Tyrakowski, T., S. Milutinovic, and I. Schulz. Studies on isolated subcellular components of cat pancreas. III. Alanine‐sodium co‐transport in isolated plasma membrane vesicles. J. Membr. Biol. 38: 333–346, 1978.
 123. Ueda, N., and O. H. Petersen. The dependence of caerulein‐evoked pancreatic fluid secretion on the extracellular calcium concentration. Pfluegers Arch. 370: 179–183, 1977.
 124. Welsh, M. J. Barium inhibition of basolateral membrane potassium conductance in tracheal epithelium. Am. J. Physiol. 244 (Renal Fluid Electrolyte Physiol.13): F639–F645, 1983.
 125. Young, J. A., and E. W. Van Lennep. Transport in salivary and salt glands. In: Membrane Transport in Biology. Transport Organs, edited by G. Giebisch. Berlin: Springer‐Verlag, 1979, p. 563–692.

Contact Editor

Submit a note to the editor about this article by filling in the form below.

* Required Field

How to Cite

O. H. Petersen, Y. Maruyama. Electrophysiology of Salivary and Pancreatic Acinar Cells. Compr Physiol 2011, Supplement 18: Handbook of Physiology, The Gastrointestinal System, Salivary, Gastric, Pancreatic, and Hepatobiliary Secretion: 25-50. First published in print 1989. doi: 10.1002/cphy.cp060302