Comprehensive Physiology Wiley Online Library

Mechanisms of Transmitter Action in the Vertebrate Central Nervous System

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 Historical Considerations
2 Identification of Specific Messengers for Neuronal Pathways
2.1 Criteria for Neurotransmitters
3 Mechanisms of Neuronal Excitability and Synaptic Transmission
3.1 Membrane Ion Channels
3.2 Conventional Synaptic Mechanisms
3.3 Unconventional Messenger Mechanisms
3.4 Neuromodulator Action
3.5 Neuromediator or Second‐Messenger Actions
4 Experimental Strategies
4.1 Measures of Cell Excitability as Indices of Transmitter Action
4.2 Methods of Drug Application
4.3 Comparison of In Vivo and In Vitro Preparations
4.4 Anatomical Description of Selected Brain Regions
5 Effects of Neuromessengers and Comparison with Endogenous Transmitters in Selected Brain Regions
5.1 Amino Acids
5.2 Monoamines
5.3 Nucleosides and Nucleotides
5.4 Peptides
5.5 Other Peptides: Special Problems of Peptide Electrophysiology
5.6 Peptides for Future Study
6 Factors Accounting for Variability in Response to Transmitters
6.1 Pathway Stimulation
6.2 Exogenous Application of Messengers
7 Integration: Synaptic Interactions and Coexistence of Neuromessengers
7.1 Spatial Domain
7.2 Temporal Domain
7.3 Mechanistic Domain
7.4 Cotransmitters
8 Conclusions: Speculations on Need for Diversity of Messages
Figure 1. Figure 1.

Comparison of conventional and newly recognized mechanisms for generation of action potentials and responses to transmitters. A: conventional mechanisms. When excitatory postsynaptic potentials (EPSPs; left) are large enough, threshold for spike initiation is exceeded, triggering the voltage‐dependent conductances that elicit rising and falling phases of action potential (middle). Classic EPSP is due to opening of nonspecific ion channels that allow flow of cations (Na+, K+). Membrane potential is driven (thin arrows) toward equilibrium potential (E) for EPSP (EEPSP), usually around −20−0 mV. Spike results from driving membrane potential toward sum of ENa, EK, and ECa. Afterhyperpolarization (AHP) is produced by opening of voltage‐dependent or Ca2+‐dependent K+ channels, driving membrane potential toward EK. Classic inhibitory postsynaptic potential (IPSP; right) is generated by opening of Cl or K+ channels, driving membrane potential toward ECl or EK (usually −70 to −100 mV). RMP, resting membrane potential; g, conductance. B: newly recognized synaptic mechanisms. Several alternative mechanisms have been proposed as the basis for synaptic events, including those schematized here. Some transmitter actions might be elicited by activation of electrogenic pumps (left); for example, pumping Na+ or Ca2+ out of cell would hyperpolarize membrane, whereas pumping K+ out would depolarize it. However, there is little strong evidence for this type of transmitter action. Generation of synaptic potentials by reducing ion conductance is on more solid experimental ground. Thus the slow EPSP in sympathetic ganglia may be partly due to closing of K+ channels, thus driving membrane potential away from EK (thick arrows) and toward equilibrium potential for next most diffusible ion(s). Slow IPSP might result from closing of Na+ channels (however, see chapter by North in this Handbook), driving membrane potential away from ENa and toward that of next most important ion (K+). In addition at least 1 class of ion channel associated with slow EPSP is a voltage‐dependent K+ channel: the M‐channel. M‐channel is only open over a finite range of potentials (probably −10 to −40 mV). Other voltage‐dependent loci of transmitter action are the conductances responsible for action potential or associated afterpotentials (right). Thus transmitters alter either voltage‐dependent Ca2+ conductance of spike, voltage‐dependent K+ conductance, or Ca2+‐dependent K+ conductance. Such actions would either shorten or lengthen Ca2+ component of spike, or decrease or augment AHP after spike.

Figure 2. Figure 2.

Sample records from cultured CNS neuron illustrating types of events that can be measured with extracellular (top trace), intracellular (middle trace), and single‐channel (bottom trace) recording techniques. For extracellular recordings, relatively large‐tipped microelectrode is placed close to neuronal surface, and neuronal activity is measured from an extracellular vantage point. This technique can detect action potentials but not subthreshold changes in membrane potential. Extra‐cellular recordings provide information about firing rate and pattern and type of event generating pattern. Recorded activity reflects both synaptic and intrinsic influences, single events being generated by integration of excitatory and inhibitory synaptic input, passive and active cable properties of membrane, endogenously generated activity (if present), and any metabolic or second‐messenger influences. For intracellular recording, fine‐tipped microelectrode is inserted through membrane and used to measure potential changes at membrane level. This technique can detect action potentials, synaptic potentials, and subthreshold changes in membrane potential. For single‐channel recording, relatively large‐tipped microelectrode is placed in close contact with neuronal membrane and measures activity of individual ion channels in membrane patch. Channel activity appears as current flow through open channel. Transitions between open and closed states are instantaneous, producing boxlike events. This technique provides information about individual ion channels that underlie chemical and electrical excitability and produce the events recorded with intracellular and extracellular recording techniques.

D. L. Gruol, unpublished observations
Figure 3. Figure 3.

Concentrations of ionic species responsible for electrical excitability, in extracellular and intracellular fluids of CNS. Values are expressed as mM in approximate range most often reported (see refs. 193,417). Width of channel or pore (gap in membrane) reflects relative resting permeability (p) for these ions. Note low permeability and intracellular concentration for Ca2+.

Figure 4. Figure 4.

Simple equivalent electrical circuit analogy of neuronal membrane. Battery (E) generates unequal distribution of electrical charge (positive outside, negative inside), in analogy to action of ion pumps and selective ionic permeability. Resistor (R) symbolizes resistance (limited permeability or conductance) in membrane to flow of ions across it. Capacitor (C) signifies capacity of membrane to store a charge for finite periods of time.

Figure 5. Figure 5.

Ion channels and related processes contributing to electrical activity of excitable cells. Most activity is generated by flow of ions through such conductance channels. Channels that are always open (top) are termed leak channels; they largely generate the resting membrane potential (RMP). Receptor‐mediated channels are depicted on right. These channels are activated or inactivated by chemicals (mostly hormones or neurotransmitters) and are thought to account for conventional non‐voltage‐dependent responses evoked by activation of synaptic pathways. Traditional voltage‐dependent channels (left) open only at certain membrane potentials; they contribute to the generation of an action potential, after the membrane potential is brought to threshold (trigger) level of depolarization by injected current or chemical activation of passive conductances (see Fig. 1A). One result of voltage‐sensitive conductance is shown at bottom left; entry of Ca2+ during action potential triggers efflux of K+, resulting in membrane repolarization (and sometimes hyperpolarizing afterpotential) at conclusion of action potential. Such ion‐sensitive channels are opened only when particular ions (e.g., Ca2+) are present. Existence of voltage‐sensitive, receptor‐mediated ion channels (bottom left) is a relatively new concept; as with conventional synapses, such channels may be opened or closed by neurotransmitters, but only at certain membrane potentials. Because many voltage‐dependent conductances are associated with action‐potential mechanisms, activation of such receptors might be expected to alter properties of the spike or its afterpotentials. Activation or inhibition of electrogenic ion pumps (bottom right) could also contribute to receptor‐ or nonreceptor‐coupled changes in membrane potential. Generation of cyclic nucleotides by nucleotide cyclases, possibly through activation of transmitter receptors, could open or close ion channels directly (or perhaps via protein phosphorylation) or alter voltage‐sensitive conductances or membrane pumps, thus significantly altering neuronal excitability. Equivalent channel effects or mechanisms could apply to presynaptic actions of transmitters (see nerve terminal, top right).

Adapted from Siggins 701
Figure 6. Figure 6.

Possible interactions of cAMP‐mediated synaptic receptor mechanisms and their consequent actions on neuronal physiology. Neuron receives synaptic contact from dopaminergic and noradrenergic terminals. Both types of terminals have been described on separate cell types but not on same cell. Both catecholamines activate adenylate cyclase, dopamine (DA) via DA receptor, and norepinephrine (NE) via β‐receptor. cAMP that is formed and survives phosphodiesterase catabolism can activate cAMP‐dependent protein kinases. Two such kinase actions are illustrated: 1) to phosphorylate or dephosphorylate proteins in synaptic membrane, which could then alter resistive or capacitative properties of membrane (same symbols at bottom left as in Fig. 4); and 2) to phosphorylate nuclear histones, thereby changing extent to which genetic properties are expressed in current metabolic properties of cell. Latter mechanism might also alter electrical or pharmacological properties of cell membrane through classic DNA‐RNA protein route over longer time course of action. Prostaglandins of E series inhibit responses to NE in hippocampal pyramidal neurons 685 and in cerebellar Purkinje neurons 347,711; prostaglandins potentiate DA in caudate nucleus 712,714. Antipsychotic drugs can block both NE and DA, depending on target cell.

Figure 7. Figure 7.

Traditional methods for assessing changes in ionic conductance produced by neuromessengers. A: (left) brief injection (20–200 ms) of rectangular negative current (I) pulse through recording barrel is recorded as hyperpolarizing voltage (V) deflection, which, after nulling of effects of current on electrode resistances with balanced‐bridge circuit, reflects impedance properties of membrane. Slowness of falling and rising phases of this voltage pulse (electro‐tonic potential) represents time constant of cell, which is proportional to product of specific membrane resistance and membrane capacitance. Amplitude of electrotonic potential is proportional to membrane resistance; this value and that of input current allow calculation of input resistance (sum of membrane resistance for whole cell membrane and other, smaller resistances, e.g., that of cytoplasm). If such a current pulse is injected repetitively at regular intervals (1–5 s) and if recording time base is retarded, pulses appear as short downward deflections (A, right). Application of conventional inhibitory transmitter, such as GABA, during sequence reveals expected hyperpolarization, and resultant reduction in size of downward deflections suggests a reduction of input resistance (increase of membrane conductance). However, this simple method may not be adequate for analysis of drug effects on those neurons that rectify in response to injection of some current values or with drugs having voltage‐dependent effects. In both cases, nonlinear I‐ V curves would result. In these cases, construction of I‐V curves before and during messenger administration is preferable and also allows estimation of the reversal potential for messenger effect (see Fig. 8). B: multiple pulses of both polarities and various intensities (usually from 0.05 to 2 nA) are injected into cell at regular intervals, and resultant electrotonic potentials are often displayed as shown on left, using repeated fast sweeps of oscilloscope. “Sag” in the electrotonic potential seen after earlier “hump” is representative of certain central neurons such as hippocampal pyramids (see refs. 601,717,718) and probably reflects time‐dependent activation of some current. Plotting the size of these potentials (either at peak of hump or at steady state) against current used to generate each potential yields IV curve with reproducible slope and shape (right). Repetition of this procedure during application of messenger may then reveal change in position and slope of curve: lowering of curve indicates hyperpolarization; reduction of slope (as schematized for GABA effects) indicates reduced input resistance, or increased conductance. Increased slope signifies reduced conductance. Note that control curve intercepts GABA curve at ca. −75 mV, which would be reversal potential for hyperpolarizing GABA response. Use of K‐acetate‐filled electrode is assumed.

Figure 8. Figure 8.

More sophisticated methods for assessing changes in ionic conductance produced by neuromessengers. A: using current clamp, membrane potential is experimentally altered during transmitter action to determine potential at which there is no net ion flow and therefore no change in membrane potential, even though channels are open. This potential (produced by injecting steady DC into cell and assuming pipette contains no Cl) is the reversal potential for response. Ideally, reversal potential occurs at the equilibrium potential for ions (here, probably Cl) mediating response (ca. −75 mV for GABA, curve on right), but under certain circumstances, such as a remote location for active site, technical problems limit the agreement between equilibrium potential for ions involved and reversal potential for measured response. B: voltage clamp. In this example, voltage is held steady at −40 mV and current is measured while GABA is applied. Downward current deflections indicate inward current produced by command voltage pulses. Note that GABA application causes outward current (due to Cl influx) and increase in command current pulses (due to increased conductance). Repetition of this paradigm while clamping membrane potential at different voltage levels would indicate the reversal potential (shown on I‐V curve on right) to be the potential (−75 mV) at which GABA no longer could elicit current flow.

Figure 9. Figure 9.

Theoretical explanation of the method of fluctuation or noise analysis. A: top traces show activity of 10 computer‐simulated channels. Each channel undergoes transitions between open and closed states as a result of a Poisson process. All 10 channels operate independently but have same amplitude and average lifetime in open state. Bottom trace shows summed activity of these 10 channels. Fluctuating signal is produced; it reflects moment‐to‐moment variation in number of open channels present. B: typical observations made during application of fluctuation analysis to study of synaptic channels in biological membrane. Under voltage‐clamp conditions, application of agonist during period indicated by vertical bar results in change in DC membrane current (I). In addition variance of membrane current is seen to increase from prior to agonist to during agonist. This increase is readily apparent in AC trace, which represents condenser‐coupled, amplified varion of DC trace. Additional variance is assumed to reflect moment‐to‐moment changes in number of synaptic channels opened by agonist. C: kinetic properties of these channels can be estimated from power spectral density (PSD) of agonist‐induced current fluctuations. It is assumed that simple kinetic scheme of type used to generate simulated channels in A also controls operation of synaptic channels. Under this assumption, PSD of biological current noise is expected to be of Lorentzian form (C, smooth curve). Mean open time of agonist‐induced channels can then be calculated from half‐power frequency (fc, arrow) of Lorentz curve that affords best fit to observed spectral point. S(f), power spectral density function.

From Mathers and Barker 488
Figure 10. Figure 10.

Various patch‐clamp recording configurations (top) and representative single‐channel recordings from a patch on cultured CNS neuron (bottom). Single‐channel recordings can be made in 3 configurations: cell‐attached, inside‐out, or outside‐out patch. In addition membrane patch can be broken and recording of whole‐cell voltage or current obtained (whole‐cell recording). Single‐channel activity (bottom) is recorded as current flow through open channel and appears as boxlike events. In this patch (cell‐attached) at least 2 channel types were present, as indicated by amplitudes of events. Brief upward events are channel openings that were not fully resolved because of limitations of recording equipment. Open (O) and closed (C) states for largest events are indicated. A 70‐mV depolarization was applied to membrane patch. Analysis of channel activity in this patch indicated that both channel types were K+ channels, 1 with a single‐channel conductance of 100 pS (•) and the other of 20 pS (•).

D. L. Gruol, unpublished observations
Figure 11. Figure 11.

Simplified coronal section of spinal cord. Spinal cord neurons were 1st mammalian central neurons studied intracellularly (see refs. 97,225,226), and considerable electrophysiological data are still derived from various types of spinal cord (or associated spinal ganglionic) preparations. Spinal cord circuits are often considered to be relatively simple, yet more recent data show them to be extremely complicated. Diagram is greatly simplified to highlight neurons and pathways most often studied by synaptic physiologists and pharmacologists. Cell body and fiber locations are shown on left; laminae and other regions of gray matter are numbered on right. Gray matter of the spinal cord consists of dorsal and ventral horns; these are surrounded by fiber tracts (white matter) of ascending and descending fibers to and from brain and other spinal regions (not shown). Sensory information enters spinal cord via dorsal roots, along axons of the monopolar neurons whose cell bodies lie in dorsal root ganglia (dorsal root ganglia neurons). These axons then enter spinal cord and terminate either on neurons in ventral horn or on 1 of several cell types in the various laminae of dorsal horn. Those axons (from muscles) terminating on ventral horn neurons follow 1 of 3 projections: 1) monosynaptically, from Ia afferents onto agonistic motoneurons via excitatory terminals, with collaterals onto inhibitory interneurons that project to antagonistic motoneurons (constituting stretch reflex); 2) disynaptically, from Ib afferents that terminate on excitatory interneurons projecting to agonistic motoneurons, with collaterals terminating on inhibitory interneurons that project to antagonist motoneurons [inverse myotatic reflex, not shown; (see refs. 121,694)]; or 3) mono‐ or disynaptically from group II (muscle spindle) fibers that project either directly onto agonistic motoneurons or through inhibitory and excitatory interneurons to agonistic and antagonistic motoneurons (see refs. 121,694). Those pathways (e.g., from skin) that project onto dorsal horn neurons in turn transmit information in 1 of 2 ways: 1) fibers from dorsal horn cells cross midline and then ascend contralaterally to thalamus through spinothalamic tract or ascend in ipsilateral dorsal column to brain stem (not shown); or 2) dorsal horn cells receiving terminations of Aδ−, Aβ‐, and C‐fibers (from mechano‐, thermo‐, and pain receptors) project, probably through other interneurons, both to ipsilateral and contralateral motoneurons in ventral horns, and to spinothalamic tracts. Pathways to the motoneurons constitute part of flexor reflex, in which ipsilateral motoneurons to flexor muscles are activated by stimuli (e.g., pain), while the motoneurons innervating extensor muscles are inhibited. Reverse situation prevails contralaterally: flexors are relaxed and extensors activated. Sensory fibers entering the spinal cord through the dorsal roots probably use glutamate (Glu) (Ia, Ib, and II afferents), substance P, somatostatin (SS), vasoactive intestinal peptide (VIP), and cholecystokinin (CCK) (C‐fibers); the various interneurons in dorsal horn probably contain GABA and enkephalin, among others. In addition some descending fibers terminating in dorsal horn contain NE, DA, and serotonin [see the chapter by Zieglgänsberger in this Handbook; 481,501,694]. Several phenomena of pain perception probably involve integration in the dorsal horn substantia gelatinosa (SG). Elucidation of the interconnections of various interneurons and the histochemical delineation of presence of several putative neurotransmitters in this sensory region of spinal cord should lead to a greater understanding of nociception in general. Several ventral horn transmitters are also known. Thus motoneurons are cholinergic, Renshaw cells (RC) and some other inhibitory interneurons probably use glycine (Gly) as their transmitters, and others probably use GABA (see refs. 21,159,160,162,163,164,694). Glu or aspartate (Asp) would seem likely candidates as transmitters for some of the excitatory interneurons (see refs. 161,694). IN, interneuron; γMN, γ‐moto‐neuron; and αMN, α‐motoneuron.

Figure 12. Figure 12.

Simplified coronal section of hippocampus. Hippocampal formation is a well‐defined cortical structure extensively studied with anatomical, biochemical, and electrophysiological techniques (see refs. 98,465,611,758). We include Ammon's horn and the dentate in the hippocampus proper. Neuronal elements comprising the hippocampus are organized into well‐defined lamellar structures that are maintained throughout its extent. Principal neurons of the hippocampal formation are pyramidal (Pyr) neurons that are arranged in a linear fashion in Ammon's horn. Pyr cell layer has been subdivided into areas CA1–CA4 based on anatomical and physiological differences between Pyr neurons and the synaptic organization within these regions 98,393,465,611. However, some researchers do not consider CA2 and CA4 as distinct areas. Pyr neurons receive input from local interneurons, from each other, and from other CNS regions via fornix (leftmost input pathway), from medial septum, the perforant pathway from entorhinal cortex and the alvear commissural pathways from the contralateral hippocampus. Axons of the Pyr neurons form the only efferent pathway leaving the hippocampus. Axon collaterals from CA3 Pyr neurons, called Schaffer collaterals, also provide excitatory input to other CA3 Pyr neurons and to interneurons and Pyr neurons in CA1 and CA2 region of the hippocampus. Excitatory amino acids (Glu or Asp) are putative transmitters for this pathway (see refs. 141,143,758). Other neuron types have been identified in layers adjacent to the Pyr cell layer. Those most well characterized are basket (B) cells 393,465,611 that provide inhibitory input to the Pyr neurons. B cells are activated by axon collaterals originating from nearby Pyr neurons or from distant or even contralateral Pyr neurons and are thought to use GABA as their transmitter (see refs. 10,13,19,393,758). Some neuropeptides (e.g., opioids, VIP, CCK, and SS) have been identified within other hippocampal interneurons, but their function is not clear (see Peptides, p. 61, and Other Peptides …, p. 74). Serotonergic, noradrenergic, and cholinergic fibers have also been described in hippocampus (see refs. 72,73,402,451,452,467,496,500,501,638,695,746,756,758). Targets for these fibers are not yet fully characterized, although it seems likely from light‐microscopic studies that NE projects to CA3 Pyr cells and dentate granule cells (Gr) and that cholinergic fibers contact CA1 and CA3 Pyr neurons 406,451,452,496,500,501,695,746. Principal neurons in the dentate are intrinsic excitatory neurons known as Gr cells. These cells receive afferent input from fibers in the perforant and commissural pathways. Gr cells send axons to Pyr neurons and interneurons in hilar, CA3, and CA4 regions. Several putative neurotransmitters have been suggested for Gr cells, including amino acids (see refs. 142,143), opioids 280,492,737,738, and other peptides (see refs. 737,738). Catecholaminergic and serotonergic inputs are also present in dentate 406,500,501. M, molecular layer.

Courtesy of G. R. Siggins and S. J. Henriksen
Figure 13. Figure 13.

Principal neurons and afferents in cerebellar cortex. The cerebellum is another laminated region that has been extensively studied anatomically, biochemically, and physiologically. Its neuronal organization and synaptic connections have been well characterized 98,227,574 and are similar throughout the structure. In the cortical region, 5 neuronal types are present: Purkinje neurons (P), basket cells (Ba), stellate cells (St), Golgi cells (Gg), and granule cells (Gr). Gr cells are the only excitatory neurons; all others are inhibitory. Parallel fibers, formed by axons of Gr cells, provide excitatory input to the other 4 types of neurons in cerebellar cortex. Glu and Asp are putative transmitters for the Gr cells [see EXCITATORY AMINO ACIDS …, p. 34; 141,142,741]. Ba and St cells provide inhibitory input to the P neurons, whereas Gg cells provide feedback inhibition to Gr cells (not shown). P cell axons are the only efferent pathway from cerebellar cortex and provide inhibitory input to neurons of deep cerebellar nuclei (DCN), whose axons in turn exit the cerebellum for other regions of CNS. GABA is strong transmitter candidate for all of the inhibitory neurons in cerebellum, including P cell (see refs. 70,826). At least 3 afferent pathways transmit information from other regions of CNS to cerebellar cortex. Climbing fibers (CFs), which are axons originating from cell bodies in inferior olive, provide powerful, bursting excitatory input to P cells. CF axon collaterals also innervate the other 4 types of neurons in the cerebellar cortex and DCN 574. Mossy fibers (MFs), whose axons originate from neurons in several CNS regions (including pons), provide excitatory afferents to Gr cells and Gg cells of cerebellar cortex and to DCN (not shown). Glu, Asp, and serotonin (latter from dorsal raphe nucleus) are putative excitatory transmitters for these MF pathways (see refs. 89,141,142). A 3rd afferent input arises from neurons of nucleus locus coeruleus (LC) and provides inhibitory but enabling 80 input to P neurons; NE is the transmitter for this pathway [see NOREPINEPHRINE, p. 40; 345,347,700].

Courtesy of F. E. Bloom
Figure 14. Figure 14.

Membrane current fluctuations to iontophoretically applied GABA in cultured mouse spinal neuron voltage clamped (Vc) at −70 mV. Membrane current is displayed unfiltered on DC trace (B) and at 10 × gain on AC trace (C) filtered at 0.2–200 Hz. Variance associated with filtered signal, updated at 1‐s intervals, is displayed in D. Increasing iontophoretic currents cause inward current responses of increasing amplitude, each of which is associated with a thickening of the DC and AC traces and increases in membrane current variance. Largest changes in variance at beginning and end of current responses reflect relatively rapid changes in membrane current occurring at these times due to AC coupling. Arrowheads mark spontaneous inward current events, which have a fast rise time and exponential decay, suggesting they are synaptic in origin.

From Barker et al. 39
Figure 15. Figure 15.

Gly‐ and GABA‐receptor channel currents recorded from outside‐out membrane patches isolated from soma of 3 different spinal cord neurons in culture. Patches were isolated from neurons bathed in normal bath solution [in mM: 140 NaCl, 1 MgCl2, 1 CaCl2, 1 KCl, and 10 Na‐HEPES (pH 7.2)], which was then exchanged for a solution containing Tris+ as major cation. For most patches this procedure removed background current activity. Tris+ substitution did not alter either chemo‐sensitivity of membrane to Gly or to GABA or conductance properties of the activated Cl channels. Pipette solution facing intracellular side was (in mM): 140 KCl, 3 NaCl, 1 MgCl2, 11 K‐EGTA, and 10 K‐HEPES (pH 7.2). All recordings were done at room temperature (22°C–25°C). A: top trace shows absence of current activity before and after rapid application (indicated by bar) of 50 μM GABA to bath solution. GABA was then washed out, and bottom trace shows increase in current caused by addition of 50 μM Gly (day 35 neuron). B: outside‐out patch in which application of 20 μM Gly did not increase current, but after washout of Gly and application of 20 μM GABA, there was a large increase in current, which decreased with time so that current steps of ∼2 pA were discernible. Note much noisier appearance of GABA‐activated current compared with the Gly‐activated current in A, which reflects higher frequency of brief interruptions evident in single GABA‐receptor currents compared with Gly‐receptor currents. C: outside‐out patch in which application of 5 μM GABA (not shown) activated a relatively low frequency of current steps (2‐pA amplitude), but when concentrations of GABA was increased (bar) to 50 μM, additional channels were activated. At this higher concentration current response quickly and completely desensitized so that after 90 s, no current steps were evident. At this point, addition of 20 μM Gly to bath solution activated current steps of 3‐pA amplitude.

From Hamill et al. 324. Reprinted by permission from Nature, copyright 1983, Macmillan Journals Limited
Figure 16. Figure 16.

Multiphasic synaptic responses of CA1 Pyr neuron in hippocampal slice preparation to stimulation of stratum radiatum (SR). Top traces were recorded at resting membrane potential (RMP) and lower traces while membrane was artificially hyperpolarized by negative current injections in amounts (mV) shown. Vertical arrow indicates time of SR stimulation. Stimulus artifact is followed by short EPSP (seen best in left recording hyperpolarized by 5 mV); in recording at RMP, the EPSP is followed by an early (angled arrow) and a late (•) hyperpolarization. Note that early hyperpolarization, probably GABAergic IPSP (see refs. 9,10,19,195,529), is reduced and then inverted by increasing hyperpolarizations, whereas late hyperpolarization is nullified only at greater hyperpolarizations (−11 to −16 mV below RMP). These and other data (see refs. 195,529,766) suggest that the early IPSP has a reversal potential near the expected equilibrium potential for Cl, whereas late hyperpolarization [possibly late or slow IPSP 529,766] has a reversal potential more like that for K+. Spikes (spontaneous or those driven by strong SR stimulation) are attenuated by slow rise time of polygraph.

D. L. Gruol, unpublished observations
Figure 17. Figure 17.

Inhibitory effectiveness and Cl dependence of depolarizing IPSP (in rat hippocampal slice preparation). A: train (2 Hz) of depolarizing current pulses initiated full‐sized somatic action potentials.

(Amplitudes not accurately reproduced by pen recorder), which were blocked during most of stimulation‐evoked response. B: inhibition of spikes elicited by just suprathreshold EPSP during depolarizing IPSP 2. Unconditioned control responses are shown before and after inhibition (1 and 3, respectively). Same cell as in A. RPM = −59 mV. C: effects of iontophoretic GABA (1 M; horizontal bar) in normal medium without pentobarbital. 1, Application through pipette in stratum pyramidale. Current pulses of 100 ms were given at 1 Hz. (see Fig 1B, pt. 1 of ref. 10). RPM = −55 mV. 2 and 3, GABA iontophoresed in stratum radiatum of another cell. During depolarization, both directly and synaptically activated action potentials are blocked (2 and 3, respectively). RPM = −61 mV. Iontophoretic pipettes were lowered independently 200 μm into slice before impaling a cell. Ejection currents were 500 nA for 5 s. D: 2 cells from same slice bathed in pentobarbital. 1, Recorded with a 2 M K‐methylsulfate‐filled pipette; 2, recorded with a 3 M KCl‐filled pipette. Note large amplitude and prolonged time course of depolarizing response in 2, after Cl leakage into cell. Small reversed IPSPs visible on base line of KCl‐filled cell are able to trigger action potentials. Antidromic responses from same cells are shown for comparison (2 and 4). RPM= −59 mV (1 and 3) and −56 mV (2 and 4). Calibration for pen traces in A, B, and D: 5 mV, 2 s; C: 5 mV, 10 s. From Alger and Nicoll 10. Reprinted by permission from Nature, copyright 1979, Macmillan Journals Limited
Figure 18. Figure 18.

Responses of a single cultured brain neuron to L‐Glu (A, 100 μM), D,L‐kainate (B, 100 μM), D,L‐homocysteic acid (D,L‐HCA; C, 100 μM), and L‐Asp (D, 100 μM) demonstrating decreased conductance. All amino acids were applied (C2: 150 ms; D2: 450 ms; all others: 350 ms) in the presence of 3 μM tetrodotoxin (TTX) (no spikes were evoked with brief cathodal current pulses). Trace below each voltage recording indicates period of drug application (upward deflections). Constant current anodal pulses (50 ms; not shown) were used to assay Gm in (A1–D1). Recordings were at neuron's resting level, −60 mV. Calibrations: A2, C2, and D2: 4 s; D1 and A1: 8 s; and B1, B2, and C1: 20 s. C1: 40 mV; all others: 20 mV. A: responses to L‐Glu. Initial phase is large depolarization (∼18 mV) accompanied by rapid rise in conductance (Gm; reduction in voltage deflections). Later, Gm drops dramatically. Note small, irregular potentials on falling phase of A2. Longer applications of L‐Glu lead to full spikelike potentials (not shown). B: responses to D,L‐kainate. Entire response to D,L‐kainate is increase in Gm. Gm remains elevated even when membrane falls close to preapplication values. Note lack of spikelike potentials in B2. Shape of offset of D,L‐kainate response differs from other amino acids because of absence of decreased Gm. In addition the duration is significantly greater than those to L‐Glu and L‐Asp. C: responses to DL‐HCA. Responses were particularly large and lasted longer than those to L‐Glu and L‐Asp. Although initial depolarization is associated with increase in Gm, response quickly reverts to surprisingly large decrease in Gm. Gain of C1 was reduced because of the size of voltage deflections. C2: generation of large spikelike potentials occurring during falling phase of response. D: responses to L‐Asp. This amino acid induces a depolarization comparable to that evoked by L‐Glu, but in this case only a decrease in Gm was observed. Spikelike potentials are also seen (D2). Higher concentrations of L‐Asp (500 μM) also increased Gm (not shown).

From MacDonald and Wojtowicz 471
Figure 19. Figure 19.

N‐methyl, D‐aspartate (NMA) produces dose‐related biphasic conductance change. A: responses of cell to increasing ionophoretic currents of NMA. Small NMA depolarizations are accompanied by apparent rise in input resistance, whereas large responses are dominated by fall in input resistance. B: NMA pulses (150‐ms wide), with and without constant current negative pulses to measure input resistance. C: in another cell, a 40‐ms, 180‐nA NMA pulse produces an apparent rise in input resistance, as measured with positive current pulses, with no underlying depolarization. A−C from different cells.

From Dingledine 192
Figure 20. Figure 20.

I−V curve for cultured spinal cord neuron under voltage clamp (2‐electrode clamp). RMP was −38 mV and both depolarizing and hyperpolarizing command steps were employed to construct curve. Control curve (•) was performed in bathing solution supplemented with TTX (2 μM) and repeated during constant microperfusion with L‐aspartic acid (500 μM; •). Net inward current (negative by convention) was evoked by L‐Asp. Slope of this steady‐state relationship was reduced in range from −70 to −30 mV.

From MacDonald et al. 470
Figure 21. Figure 21.

Intracellular recordings from rat cerebellar Purkinje cells in vivo. A: recording and electrophoretic setup: 3‐barreled micropipette with a Purkinje cell. Intracellular electrode protrudes beyond orifices of the 2 extracellular microelectrophoretic barrels. B: multispiked spontaneous climbing fiber discharge obtained during intracellular recording from a Purkinje cell. RMP (in mV) is given in parentheses. Calibration: 20 ms and 25 mV. C: changes in membrane potential and membrane resistance of 4 Purkinje cells in response to drugs. All specimens in each row of records are from same cell. Bar above each record indicates extracellular electrophoresis of indicated drug (100–150 nA). RMP (in mV) is given in parentheses below each recording. Calibration: 10 s and 20 mV for NE, dibutyryl cAMP (DB), and cAMP; 5 s and 10 mV for GABA. Right: effective input resistance was judged by size of pulses resulting from passage of a brief constant current (1‐nA) pulse across the membrane before, during, and after electrophoresis of respective drugs (1 mV = 1 MΩ). Discontinuities in fast transients of pulses (and loss of spikes) result from loss of high frequencies (>10 kHz) and from chopped nature of the frequency‐modulated magnetic tape recording used. All pulse records were graphically normalized to same base‐line level. Calibration: 80 ms and 15 mV for all pulse records.

From Siggins et al. 716. Copyright 1971 by the American Association for the Advancement of Science
Figure 22. Figure 22.

Intracellular recording from a CA1 pyramidal neuron in hippocampal slice in vitro preparation. Isoproterenol (IP) reduced amplitude and duration of afterhyperpolarization (AHP) evoked by depolarizing current pulses that generated action potentials (attenuated by polygraph). AHP is a hyperpolarization and inhibition of spontaneous activity after termination of depolarizing current pulse (arrow). Amplitude and duration of AHP increases as amplitude of depolarizing pulse increases. Note also activity increase produced by IP.

From Gruol and Siggins 302
Figure 23. Figure 23.

Intracellular recordings show effect of adrenergic agonists on membrane potential and spontaneous activity of CA1 pyramidal neurons in hippocampal slice preparation of rat. A: typical response of CA1 pyramidal neuron to superfusion of IP. In all neurons tested, IP evoked an increase in subthreshold activity (indicated by thickness of base line) and action‐potential generation (large upward deflections). Superfusion of IP began at downward arrow and stopped at end of bracket. Delay to onset of increase in activity is partially due to “dead time” (1–2 min) of perfusion system. Increase in activity evoked by IP was prolonged in duration and far outlasted estimated washout time (2–4 min). Note depolarization evoked by IP. B: typical response of CA1 pyramidal neurons to the α‐adrenergic agonist clonidine. NE decreased spontaneous activity and produced a small hyperpolarization. Concentrations of 2–10 μM were usually required to produce this effect. This is same neuron as in A, where superfusion with β‐agonist IP increased spontaneous activity. Clonidine mimicked action of NE in this neuron, but in other neurons NE was excitatory whereas clonidine was inhibitory. These data suggest that CA1 pyramidal neurons have both α− and β‐adrenergic receptors.

D.L. Gruol and G.R. Siggins, unpublished observations
Figure 24. Figure 24.

Suggested locus of NE action at membrane of hippocampal pyramidal cell. NE, in binding with β‐adrenergic receptor (triangular notches) on neuronal membrane, could block either of 2 ion channels: 1) the Ca2+ channels that are voltage‐sensitive channels opened by a depolarization such as occurs during an action potential; or 2) the Ca2+‐dependent K+ channels that are opened by influx of Ca2+. Data showing no NE effect on Ca2+ spike during TTX treatment (see refs. 302,477) suggest that 2nd possibility is more likely. Because NE‐induced inactivation of the Ca2+‐dependent K+ channels would retard repolarization of membrane during an action potential, more Ca2+ might then enter the cell during this state of depolarization by virtue of voltage‐dependence of Ca2+ channel.

Figure 25. Figure 25.

NE activates K+ conductance in the locus coeruleus in vitro. A: pressure application of NE [50 ms, 10 psi (68 kPa); ▴] produced hyperpolarization and increased membrane conductance. Top traces (in 1 and 2) are of membrane potential and bottom traces are of membrane current. Parts 1 and 2 are from same neurons. NE was applied twice in each trace. Full amplitude of NE‐induced hyperpolarization is seen after 1st application. After 2nd application, membrane potential was maintained at rest with manual‐clamp technique. Twice as much NE was applied in 2 as in 1. B: determination of NE reversal potential in solutions of varying K+ content. A single pressure‐ejection pulse [50 ms, 10 psi (68 kPa)] was applied at each arrowhead. Left: normal, 2.5 mM K+ solution; as membrane was hyperpolarized, NE potential decreased and reversed at −100 mV. Middle: 4.5 mM K+ solution; reversal potential decreased to −85 mV. Right: 10.5 mM K+ solution; reversal potential was −65 mV.

From Egan et al. 229
Figure 26. Figure 26.

Reversal potential of IPSP in locus coeruleus in vitro. A: synaptic potentials were evoked by single pulse (30 V, 0.5 ms; arrow) at different RMPs. At −67 mV the IPSP was hyperpolarizing, at −105 mV it was nullified, and at −127 mV it was depolarizing. Durations of EPSP and IPSP were reduced at more negative potentials, probably because of membrane rectification at these potentials. B: amplitude of IPSP in another cell is plotted as function of membrane potential in solutions with different K+ content. •, Normal K+ concentration (2.5 mM); •, 4.5 mM K+; and ▴, 10.5 mM K+. Estimated reversal potentials were −121, −105, and −92 mV. (These reversal potentials were the most negative of cells included in Table 1 of ref. 229.)

From Egan et al. 229
Figure 27. Figure 27.

Effects of iontophoretic DA and protons (H+) on cat caudate neurons in vivo. A: iontophoretically applied DA depolarizes membrane of cat caudate neuron and decreases its firing rate, whereas Na+ ejected at twice the current intensity and same pH used for DA had no comparable effect. B: H+ ejected from barrel containing HCl at pH of 0.5 produces long‐lasting increases of firing rate of another caudate neuron (ratemeter output). Excitatory effect of Glu is much faster in onset and offset, whereas DA had no effect on firing rate of this cell displaying no spontaneous action potentials. 2/1 and 69Q2, Experiment and cell numbers.

From Herrling and Hull 337
Figure 28. Figure 28.

DA‐induced effects with short (50 μm) distances between tips of recording and iontophoretic electrodes on 2 different cells of cat caudate in vivo. A: DA induces hyperpolarizations. B: if DA is applied for longer period on another cell, initial hyperpolarization is followed by slow depolarization. Single arrow, cortically evoked EPSP; double arrows, cortically evoked EPSP with superimposed action potential. Glu, glutamate.

From Herrling and Hull 337
Figure 29. Figure 29.

Effect of carbachol on current relaxations generated by hippocampal CA1 neuron (in slice preparation) at rest and at depolarized holding potential. Traces are currents initiated by hyperpolarizing voltage‐clamp step (ΔV) of 14 mV from holding level of −42 mV (top row) and from close to resting potential, −62 mV (bottom row), before, during, and 10 min after brief (2 min) bath application of carbachol (48 μM). During drug administration an inward current developed at the more positive holding potential, as indicated by positioning of traces, but not when cell was held at −62 mV. Note approximately ohmic behavior of cell at latter potential, whereas a similar current response was only observed in presence of carbachol at depolarized level, i.e., when time‐dependent inward relaxation had been abolished. TTX (0.5 μM) was included in bathing medium to abolish Na+‐dependent action potentials; bath temperature was 23°C.

From Halliwell and Adams 323
Figure 30. Figure 30.

Effects of ACh on rat CA1 pyramidal cells in hippocampal slice preparation. All responses are from same cell. A: chart record of control AHP after 60‐ms direct depolarizing current pulse (trace 1) and film record of response to 600‐ms pulse (trace 2). Current record is positioned below voltage record. B: ACh (200 μM) superfusion depolarized membrane and increased cell's input resistance. C: blockade of AHP (trace 1) and accommodation (trace 2) in presence of ACh. D: addition of atropine (0.5 μM) in presence of ACh reversed effects of ACh. E: atropine also reversed effect of ACh on AHP (trace 1) and on accommodation (traces 2 and 3). Current pulse in trace 3 was identical to those of trace 2 in A and C. Current pulse was increased in trace 2 to match depolarization evoked in presence of ACh (trace 2) in C. Gain in trace 1 in A applies to all chart records. Time calibration for trace 1 in A also applies to trace 1 in C and E, and time calibration in D and B are the same. Calibration (time, volts, and current) for trace 2 in A applies to all film records. RMP = −57 mV.

From Cole and Nicoll 134. Copyright 1983 by the American Association for the Advancement of Science
Figure 31. Figure 31.

Hippocampal slice preparation. Effect of exogenous ACh mimicked by stimulating stratum oriens. Records from 4 different cells are shown (A‐D). A: immediately after train of stimuli (S), cell hyperpolarizes (condensed on time scale) and then depolarizes (arrow in trace 1). In trace 2, slow depolarizaton was manually voltage clamped, which clearly shows increase in size of hyperpolarizing current pulses. RMP = −53 mV. B: slow depolarization (trace 1) was completely blocked by 1 μM atropine (trace 2). RMP= −59 mV. Calibration in A also applies to B. C: AHPs before pathway stimulation and 4 s after stimulus. In control solution (trace 1) AHP was slightly depressed after stimulus. After addition of 1 μM eserine (trace 2), AHP was greatly reduced by stimulus (arrow); AHP was restored to control amplitude by 1 μM atropine (trace 3). RMP = −56 mV. D: identical protocol as in C, except showing accommodation during 600‐ms pulse before and after pathway stimulation. RMP = −64 mV.

From Cole and Nicoll 134. Copyright 1983 by the American Association for the Advancement of Science
Figure 32. Figure 32.

Effect of adenosine on membrane potential and input resistance of pyramidal neuron in rat hippocampal slice. A: low concentration (5 μM) of adenosine (•) has no detectable effect on input resistance measured from electrotonic potentials (bottom inset oscillographs in 1 and 2) generated by current injection (top inset oscillographs in 1 and 2) through a bridge circuit and recording electrode. 1: Control; 2, adenosine superfusion; I‐V curve during adenosine superfusion is superimposable with that of control period (x). B: higher adenosine concentration (20 μM) reduces input resistance during hyperpolarization. Right, current and electrotonic potentials before (1), during (2), and after (3) adenosine perfusion. Calibration: 20 mV, 50 ms. Spiking after offset of current pulse was reduced by adenosine. All voltage measurements were taken from later, flat portions of electrotonic potentials. Curves were normalized to RMP. C: 20 μM adenosine hyperpolarizes and inhibits spontaneous discharge of pyramidal neuron via postsynaptic action; slice was continuously perfused with 10 mM MgCl2 to prevent transmitter release. Spikes were attenuated by slow frequency response of polygraph. Brackets show duration of adenosine (plus 10 mM MgCl2) perfusion. Note rebound depolarization often seen after withdrawal of adenosine.

From Siggins and Schubert 717
Figure 33. Figure 33.

Effects of adenosine on Ca2+ spikes from TTX‐treated hippocampal neurons in slice preparation. All traces have been digitized (0.2‐ to 0.4‐ms sample time). A: individual responses before (A) and after (B‐D) 20 μM adenosine. Traces B‐D are successive responses at 15−, 30−, and 60‐s intervals after addition of adenosine to recording chamber. Spike seen in traces A and B is largely abolished in trace C; trace D shows no active response to current injection. Depolarizing current pulse was held constant (1.0 nA) for all traces. B: superimposed voltage traces from cell before (A), during (B), and after (C) perfusion with 20 μM adenosine. Depolarizing current was 0.8 nA (top traces) and 1.1 nA (bottom). C: superimposed records showing response of cell (top traces) to depolarizing current injection before (C1), and during (C2) treatment with 20 μM adenosine. Amount of current injected is illustrated by current monitor (bottom trace). Note that calcium spike can be elicited in presence of adenosine, but that considerably higher currents are required (top 2 traces in C2). Calibration: 10 mV (2 nA for current traces) and 20 ms for each panel.

From Proctor and Dunwiddie 605
Figure 34. Figure 34.

Possible sites of action of nucleosides or nucleotides in nervous system. XYP is a molecule in which X could be any purine or pyramidine base and Y is number of phosphate groups, from 0 to 3. These substances could have several roles: 1) as local modulators released from glia or neurons and acting either pre‐ or postsynaptically; 2) classic transmitters; 3) presynaptic modulators of transmitter release; 4) cotransmitters with some other transmitter such as ACh; 5) 2nd messengers acting at several possible sites, including at some nucleotide cyclase, which would then generate a 3rd messenger such as cAMP or cGMP; and 6) unconventional transmitters acting on voltage‐dependent conductance such as M‐current (IM).

Figure 35. Figure 35.

Enkephalin opens K+ channels in locus coeruleus neurons in vitro. A: D‐Ala, D‐Leu‐enkephalin (DADL) was applied by pressure [arrow; 50‐ms pulse, 70 kN/m2 (10 psi)] at various membrane potentials. The K+ concentration was as indicated (in mM). At elevated K+ concentrations, enkephalin response reversed polarity with membrane hyperpolarization. Marked rectification occurred with hyperpolarization and elevated K+ concentration. B: relationship between hyperpolarization amplitude and membrane potential. C: enkephalin reversal potential (Erev) calculated from many experiments such as those in A and B. Dashed line, Erev = RT/F ln([K]o/[K]i), where [K]i, intracellular K+ concentration, was 145 mM. [In normal K+ solutions (2.5 mM), it was usually not possible to reverse enkephalin hyperpolarization; therefore extrapolated values were used.] Good agreement between observed values of Erev and those predicted from Nernst equation (Eq. 4) suggests a predominantly somatic site of action; large steady‐state conductance increase that accompanied membrane hyperpolarization would effectively remove dendritic contribution.

From Williams et al. 814. Reprinted by permission from Nature, copyright 1982, Macmillan Journals Limited
Figure 36. Figure 36.

Stratum radiatum (Rad) stimulation produces EPSP/IPSP sequence in CA1 pyramidal cells (A1) of hippocampal slice. Application of 100 μM D‐Ala, D‐Leu‐enkephalin (DADL) reduces IPSP (A2) and leads to depolarization and burst of spike activity after EPSP (A3). Alvear (Alv) stimulation produces an IPSP (B1) that is blocked by 100 μM DADL (B2). Recovery can occur within 10 min (B3). However, in presence of 2 μM naloxone, alvear‐induced IPSP (C1) is not blocked by repeated applications of 100 μM enkephalin (C2). •, Time of stimulation. Calibration: A, 10 mV and 40 ms; B and C, 5 mV and 40 ms.

From Masukawa and Prince 486
Figure 37. Figure 37.

COOH‐terminal portion of porcine prodynorphin sequence (residues 171–256) deduced by Kakidana et al. 390. Three [Leu5]enkephalin‐containing segments are bracketed by putative peptide‐processing signal “Lys‐Arg”. α‐Neo‐ and β‐neoendorphin overlap as prodynorphin residues 175–184 and 175–183, respectively. Dynorphin A(1–17) corresponds to full sequence of 2nd segment. Dynorphin A(1–8) is produced by unusual cleavage to left of Arg217. Similar cleavage to left of Arg241 produces dynorphin B from 3rd [Leu5]enkephalin‐containing segment. Presence of full sequence of dynorphin B(1–29) in brain tissue has not been reported.

From Chavkin et al. 127
Figure 38. Figure 38.

Substance P (SP) responses of cultured spinal cord neuron. Clear reversal of SP‐response polarity was not obtained. SP and eledoisin‐related peptide response amplitude varied as function of membrane potential, but responses did not reverse polarity. SP‐response amplitude increased when membrane was depolarized and decreased when it was hyperpolarized by 1 of 2 intracellular micropipettes. Response amplitude did not vary as linear function of membrane potential at potentials <‐50 mV or >‐95 mV in this neuron. Although activation of voltage‐dependent conductances might reduce voltage response recorded at potentials <‐50 mV, it was not clear why no reversal of response polarity was seen. Even when this neuron was hyperpolarized to very negative potentials (−106 and −114 mV), SP responses did not reverse polarity. RPe, extrapolated reversal potential.

From Nowak and Macdonald 551
Figure 39. Figure 39.

Structure of the presumed somatostatin (SS) prepro‐hormone SS‐28 and 2 fragments, SS‐28(1–12) and original SS‐14, or SS‐28(15–28), derived from SS‐28. Heavy brackets with arrows denote the 2 fragments; light bracket indicates ring structure of SS‐28 and SS‐14. Note that SS‐28 is NH2‐terminally extended version of SS‐14. Cleavage of SS‐28 to SS‐14 would also appear to liberate SS‐28(1–12) and an Arg‐Lys dipeptide (see refs. 63,64).

Figure 40. Figure 40.

Effect of SS on CA1 pyramidal neuron, recorded with intracellular microelectrode in rat hippocampal slice preparation. Slices were 250 μm thick. Recording microelectrodes were filled with 1 M K‐acetate (tip resistance = 90–150 MΩ). Center of a 4‐barreled microelectrode was filled with 3 mM solution of SS in 5 mM Na‐acetate (pH 7.0), and the peripheral barrels were filled with 1 M NaCl, 1 M Na acetate, 0.5 M ACh chloride (pH 4.5), or 1 M monosodium l‐Glu (pH 6.7). A backing current of 25 nA and appropriate polarity was applied to each barrel throughout experiment. In some experiments synthetic SS was used with similar results. SS was tested only on cells from which stable intracellular RMP recordings > −60 mV were obtained. A: SS [somatomedin‐release inhibitory factor (SRIF)], applied by passage of 196 nA negative current (−ve), caused depolarization of 21 mV (RMP of −75 mV). Termination of application resulted in release of some voltage coupling between recording and iontophoretic electrodes, which masked an even greater depolarization. Depolarizing pulses and both the evoked and spontaneously occurring action potentials disappeared when photographic conditions suitable for reproduction of changes in DC level at high gain were selected. B: same record at a faster sweep speed shows effect of SS on excitability of cell at times a‐d indicated in A. Bottom trace shows depolarizing current pulse injected through recording electrode, adjusted so as to evoke 2 action potentials in resting conditions. C: similar records show excitation of cell by l‐Glu (GLUT).

From Dodd and Kelly 203. Reprinted by permission from Nature, copyright 1978, Macmillan Journals Limited
Figure 41. Figure 41.

Effect of SS analogue in medium with high Mg2+. A: membrane potential response of pyramidal cell in hippocampal slice. Cell displays reversible hyperpolarization in response to addition of SS nonapeptide agonist Ac‐Cys3‐Des AA1,2,4,5,12 (D‐Trp8, D‐Cys14)‐SS to perfusate. Artificial cerebrospinal fluid contains 12 mM MgCl2, which eliminated stratum radiatum‐induced action potentials. B: voltage deflections in response to depolarizing and hyperpolarizing current before, during, and after SS‐analogue perfusion. C: I‐V curves constructed from hyperpolarizing potentials. Decrease in slope of the I‐ V curve indicates that SS analogue increased conductance in this cell.

From Pittman and Siggins 601
Figure 42. Figure 42.

SS‐28 hyperpolarizes and inhibits activity of hippocampal pyramidal cell in vitro with a slight decrease in input resistance. SS‐28 was superfused onto hippocampal slice preparation during intracellular recording. Deflections below the membrane potential line are electrotonic potentials resulting from intracellular current injection through a bridge circuit. Note small reduction in these potentials during SS‐28‐induced hyperpolarizations. This apparent reduction in input resistance is also indicated by slight diminution of slope of I‐V curve generated during SS‐28 superfusion [membrane potential normalized to RMP (Vm) during SS‐28]. Note slow onset and long duration of action of SS‐28 and reduction in subthreshold (thin vertical lines above RMP) and spike activity (thick vertical lines). Spikes were attenuated by slow rise time of polygraph.

D. L. Gruol and G. R. Siggins, unpublished observations
Figure 43. Figure 43.

Effect of corticotropin‐releasing factor (CRF) on current‐evoked bursts of action potentials and associated AHPs in CA1 pyramidal cell of hippocampal slice. A: 0.5 μM CRF reduces AHPs that follow trains of spikes generated by passing depolarizing current pulses of 3 different intensities through active bridge circuit and recording electrode. Late components of AHPs are completely lost during CRF superfusion, despite greater number of spikes generated by equivalent current strengths. CRF also increases spontaneous firing and depolarizes the cell (membrane potentials are indicated for each trace). B: effect of CRF in presence of TTX (100 nM). TTX alone (left) abolishes early, fast‐action potentials, leaving slower Ca2+ spike. CRF (400 nM) added to TTX solution (middle) nearly abolishes AHP (arrows), but did not significantly alter Ca2+ spike. Recovery is shown on right. V, voltage trace; I, current trace. Current pulse = 0.2 nA; calibration: 10 mV and 0.2 s.

From Aldenhoff et al. 8. Copyright 1983 by the American Association for the Advancement of Science
Figure 44. Figure 44.

Suggested locus of CRF action on hippocampal pyramidal cell membrane. CRF molecule, in binding with a CRF receptor (triangular notches) on neuronal membrane, could block either of 2 ion channels: 1) Ca2+‐channels that are voltage‐sensitive channels opened by depolarization, e.g., during action potential, or 2) Ca2+‐dependent K+ channels, which are opened by Ca2+ influx. Data showing lack of a CRF effect on Ca2+ spike during TTX treatment 8,707 suggests that the 2nd possibility is more likely. Because CRF inactivation of Ca2+‐dependent K+ channels would retard repolarization of membrane during an action potential, more Ca2+ might enter the cell during this state of depolarization by virtue of voltage dependence of Ca2+ channel.

From Siggins et al. 705
Figure 45. Figure 45.

Comparison of oxytocin/vasopressin related neuropeptide structures.

Figure 46. Figure 46.

Two major brain slice recording methodologies currently used. In both cases slice is usually placed on nylon net (open circles), on an electron microscope grid, or on filter paper. Recording chamber (shown in cross‐section) may be circular or rectangular. A: static or perifusion (microdrop) method involves filling recording chamber with warm artificial CSF (ACSF; gassed with carbogen) only up to edges of the slice. Recording may be done without perfusion because space above slice is in contact with warm carbogen gas saturated with H2O; H2O‐gas interface at top of slice allows passage of O2 into tissue. However, “perifusion” may also be performed, wherein the inflow and outflow of ACSF are perfectly balanced so that the level of ACSF never changes and top of slice remains exposed to carbogen. Microdrop involves pressure application of drugs from pipettes to top, exposed surface of slice. B: superfusion method involves covering the entire slice with warm ACSF saturated with carbogen; fast flow rate (2–5 ml/min) is used to assure adequate oxygenation of tissue. Outflow must balance inflow so fluid level does not change; otherwise, tissue movement and changes in recording characteristics could occur. Drug pipettes in both types of chambers could also be used to pass drugs by iontophoresis as well as by pressure. Several laboratories now use a slice recording method that combines some properties of both methods (see refs. 312,317,318).

Figure 47. Figure 47.

Hypothetical integrative interactions between conventional and newly described synaptic messages. Input T: a conventional test EPSP produced by conductance increase; inputs A‐D: conditioning postsynaptic potentials (PSPs) that can modify the test EPSP, as well as each other. Inputs A, C, and E are EPSPs and B and D are IPSPs; inputs A and B are conventional (i.e., associated with increased conductance or decreased resistance), whereas inputs C‐E are associated with decreased conductance. Results of synaptic interactions, recorded in the soma, are shown as possible summations of various PSP types, and partially result from effects of changes in input resistance and consequently the space constant. “A, then T” shows decremental effects of conduction of A EPSP down dendrites to cell body; the initial EPSP is greatly attenuated compared with potential generated at dendritic sites. Also note shunting of secondary T EPSP by reduced resistance and space constant brought about by A input, even though both PSPs are depolarizing and should summate to some extent. “B, then T” shows similar shunting effect, combined with summation of opposite values. However, in “C, then T”, the increased input resistance (Ri) results in a larger T EPSP, overcoming to some extent effects of decremental conduction down the dendrite. Similar augmentation of T might occur even with hyperpolarization, if IPSP is associated with an increase in Ri (“D, then T”). When EPSPs with increased Ri occur nearly simultaneously (“E, then C”), the 1st may nullify 2nd, especially if same conductances are involved (see the chapter by North in this Handbook). In “E, then D” it is not clear what would result, although logic suggests that summation of opposite signs would nullify the responses. See Fig. 48 for interactive effects of other newly described messages. Vm, membrane potential.

G. R. Siggins, unpublished observations
Figure 48. Figure 48.

Different functional types of interactive messages. Column A: responses displayed over a long time course (0.5–600 s); column B: typical electrotonic potential shape produced by intracellular current injection before (left) and during (right) application of messenger. Curves above long recording traces in A represent sizes and shapes of EPSPs (1–3) and changes in spikes (top trace) and AHPs (bottom trace) (4) before and after application of representative messenger. Sharp downward deflections in traces in A indicate repetitively generated electrotonic potentials, of which those in column B (expanded in time) are representative. Note that shunting messenger in A1 reduces size of another (conductance increase) EPSP, even when both are depolarizing. In fact, both depolarization (by driving membrane potential toward Eepsp) and the reduced input resistance would tend to reduce the conventional EPSP. Furthermore, electrotonic potentials in B1 indicate that decreased input resistance would also shorten membrane rise time (note increased “squareness” of potential with messenger), effectively shortening EPSP. Messengers that increase input resistance in A2 (larger downward deflections) would enlarge EPSPs with conductance increases (and may constitute an enabling mechanism) by increasing voltage drop across membrane and by increasing membrane time constant (note longer time required for electrotonic potential in B2 to reach steady state). However, a depolarizing enabling messenger might still reduce the EPSP by driving membrane potential substantially nearer Eepsp, thereby reducing driving force for ion flow. Note long time course (see calibrations of enabling responses compared with those of shunting messengers). Modulating messengers in A3 might disenable (reduce) or enhance (not shown) other synaptic inputs, without altering membrane properties (B3). This enhancement could also be a 2nd mechanism for enabling. Messengers that alter voltage‐dependent conductances in A4 might prolong (or shorten) spike (above trace) by enhancing the voltage‐dependent Ca2+ conductance, or by decreasing K+ conductances, but without altering resting membrane properties (B4). AHPs (below trace) might also be reduced or augmented by similar effects on voltage‐dependent Ca2+ or K+ conductances, again without any effect on resting properties.



Figure 1.

Comparison of conventional and newly recognized mechanisms for generation of action potentials and responses to transmitters. A: conventional mechanisms. When excitatory postsynaptic potentials (EPSPs; left) are large enough, threshold for spike initiation is exceeded, triggering the voltage‐dependent conductances that elicit rising and falling phases of action potential (middle). Classic EPSP is due to opening of nonspecific ion channels that allow flow of cations (Na+, K+). Membrane potential is driven (thin arrows) toward equilibrium potential (E) for EPSP (EEPSP), usually around −20−0 mV. Spike results from driving membrane potential toward sum of ENa, EK, and ECa. Afterhyperpolarization (AHP) is produced by opening of voltage‐dependent or Ca2+‐dependent K+ channels, driving membrane potential toward EK. Classic inhibitory postsynaptic potential (IPSP; right) is generated by opening of Cl or K+ channels, driving membrane potential toward ECl or EK (usually −70 to −100 mV). RMP, resting membrane potential; g, conductance. B: newly recognized synaptic mechanisms. Several alternative mechanisms have been proposed as the basis for synaptic events, including those schematized here. Some transmitter actions might be elicited by activation of electrogenic pumps (left); for example, pumping Na+ or Ca2+ out of cell would hyperpolarize membrane, whereas pumping K+ out would depolarize it. However, there is little strong evidence for this type of transmitter action. Generation of synaptic potentials by reducing ion conductance is on more solid experimental ground. Thus the slow EPSP in sympathetic ganglia may be partly due to closing of K+ channels, thus driving membrane potential away from EK (thick arrows) and toward equilibrium potential for next most diffusible ion(s). Slow IPSP might result from closing of Na+ channels (however, see chapter by North in this Handbook), driving membrane potential away from ENa and toward that of next most important ion (K+). In addition at least 1 class of ion channel associated with slow EPSP is a voltage‐dependent K+ channel: the M‐channel. M‐channel is only open over a finite range of potentials (probably −10 to −40 mV). Other voltage‐dependent loci of transmitter action are the conductances responsible for action potential or associated afterpotentials (right). Thus transmitters alter either voltage‐dependent Ca2+ conductance of spike, voltage‐dependent K+ conductance, or Ca2+‐dependent K+ conductance. Such actions would either shorten or lengthen Ca2+ component of spike, or decrease or augment AHP after spike.



Figure 2.

Sample records from cultured CNS neuron illustrating types of events that can be measured with extracellular (top trace), intracellular (middle trace), and single‐channel (bottom trace) recording techniques. For extracellular recordings, relatively large‐tipped microelectrode is placed close to neuronal surface, and neuronal activity is measured from an extracellular vantage point. This technique can detect action potentials but not subthreshold changes in membrane potential. Extra‐cellular recordings provide information about firing rate and pattern and type of event generating pattern. Recorded activity reflects both synaptic and intrinsic influences, single events being generated by integration of excitatory and inhibitory synaptic input, passive and active cable properties of membrane, endogenously generated activity (if present), and any metabolic or second‐messenger influences. For intracellular recording, fine‐tipped microelectrode is inserted through membrane and used to measure potential changes at membrane level. This technique can detect action potentials, synaptic potentials, and subthreshold changes in membrane potential. For single‐channel recording, relatively large‐tipped microelectrode is placed in close contact with neuronal membrane and measures activity of individual ion channels in membrane patch. Channel activity appears as current flow through open channel. Transitions between open and closed states are instantaneous, producing boxlike events. This technique provides information about individual ion channels that underlie chemical and electrical excitability and produce the events recorded with intracellular and extracellular recording techniques.

D. L. Gruol, unpublished observations


Figure 3.

Concentrations of ionic species responsible for electrical excitability, in extracellular and intracellular fluids of CNS. Values are expressed as mM in approximate range most often reported (see refs. 193,417). Width of channel or pore (gap in membrane) reflects relative resting permeability (p) for these ions. Note low permeability and intracellular concentration for Ca2+.



Figure 4.

Simple equivalent electrical circuit analogy of neuronal membrane. Battery (E) generates unequal distribution of electrical charge (positive outside, negative inside), in analogy to action of ion pumps and selective ionic permeability. Resistor (R) symbolizes resistance (limited permeability or conductance) in membrane to flow of ions across it. Capacitor (C) signifies capacity of membrane to store a charge for finite periods of time.



Figure 5.

Ion channels and related processes contributing to electrical activity of excitable cells. Most activity is generated by flow of ions through such conductance channels. Channels that are always open (top) are termed leak channels; they largely generate the resting membrane potential (RMP). Receptor‐mediated channels are depicted on right. These channels are activated or inactivated by chemicals (mostly hormones or neurotransmitters) and are thought to account for conventional non‐voltage‐dependent responses evoked by activation of synaptic pathways. Traditional voltage‐dependent channels (left) open only at certain membrane potentials; they contribute to the generation of an action potential, after the membrane potential is brought to threshold (trigger) level of depolarization by injected current or chemical activation of passive conductances (see Fig. 1A). One result of voltage‐sensitive conductance is shown at bottom left; entry of Ca2+ during action potential triggers efflux of K+, resulting in membrane repolarization (and sometimes hyperpolarizing afterpotential) at conclusion of action potential. Such ion‐sensitive channels are opened only when particular ions (e.g., Ca2+) are present. Existence of voltage‐sensitive, receptor‐mediated ion channels (bottom left) is a relatively new concept; as with conventional synapses, such channels may be opened or closed by neurotransmitters, but only at certain membrane potentials. Because many voltage‐dependent conductances are associated with action‐potential mechanisms, activation of such receptors might be expected to alter properties of the spike or its afterpotentials. Activation or inhibition of electrogenic ion pumps (bottom right) could also contribute to receptor‐ or nonreceptor‐coupled changes in membrane potential. Generation of cyclic nucleotides by nucleotide cyclases, possibly through activation of transmitter receptors, could open or close ion channels directly (or perhaps via protein phosphorylation) or alter voltage‐sensitive conductances or membrane pumps, thus significantly altering neuronal excitability. Equivalent channel effects or mechanisms could apply to presynaptic actions of transmitters (see nerve terminal, top right).

Adapted from Siggins 701


Figure 6.

Possible interactions of cAMP‐mediated synaptic receptor mechanisms and their consequent actions on neuronal physiology. Neuron receives synaptic contact from dopaminergic and noradrenergic terminals. Both types of terminals have been described on separate cell types but not on same cell. Both catecholamines activate adenylate cyclase, dopamine (DA) via DA receptor, and norepinephrine (NE) via β‐receptor. cAMP that is formed and survives phosphodiesterase catabolism can activate cAMP‐dependent protein kinases. Two such kinase actions are illustrated: 1) to phosphorylate or dephosphorylate proteins in synaptic membrane, which could then alter resistive or capacitative properties of membrane (same symbols at bottom left as in Fig. 4); and 2) to phosphorylate nuclear histones, thereby changing extent to which genetic properties are expressed in current metabolic properties of cell. Latter mechanism might also alter electrical or pharmacological properties of cell membrane through classic DNA‐RNA protein route over longer time course of action. Prostaglandins of E series inhibit responses to NE in hippocampal pyramidal neurons 685 and in cerebellar Purkinje neurons 347,711; prostaglandins potentiate DA in caudate nucleus 712,714. Antipsychotic drugs can block both NE and DA, depending on target cell.



Figure 7.

Traditional methods for assessing changes in ionic conductance produced by neuromessengers. A: (left) brief injection (20–200 ms) of rectangular negative current (I) pulse through recording barrel is recorded as hyperpolarizing voltage (V) deflection, which, after nulling of effects of current on electrode resistances with balanced‐bridge circuit, reflects impedance properties of membrane. Slowness of falling and rising phases of this voltage pulse (electro‐tonic potential) represents time constant of cell, which is proportional to product of specific membrane resistance and membrane capacitance. Amplitude of electrotonic potential is proportional to membrane resistance; this value and that of input current allow calculation of input resistance (sum of membrane resistance for whole cell membrane and other, smaller resistances, e.g., that of cytoplasm). If such a current pulse is injected repetitively at regular intervals (1–5 s) and if recording time base is retarded, pulses appear as short downward deflections (A, right). Application of conventional inhibitory transmitter, such as GABA, during sequence reveals expected hyperpolarization, and resultant reduction in size of downward deflections suggests a reduction of input resistance (increase of membrane conductance). However, this simple method may not be adequate for analysis of drug effects on those neurons that rectify in response to injection of some current values or with drugs having voltage‐dependent effects. In both cases, nonlinear I‐ V curves would result. In these cases, construction of I‐V curves before and during messenger administration is preferable and also allows estimation of the reversal potential for messenger effect (see Fig. 8). B: multiple pulses of both polarities and various intensities (usually from 0.05 to 2 nA) are injected into cell at regular intervals, and resultant electrotonic potentials are often displayed as shown on left, using repeated fast sweeps of oscilloscope. “Sag” in the electrotonic potential seen after earlier “hump” is representative of certain central neurons such as hippocampal pyramids (see refs. 601,717,718) and probably reflects time‐dependent activation of some current. Plotting the size of these potentials (either at peak of hump or at steady state) against current used to generate each potential yields IV curve with reproducible slope and shape (right). Repetition of this procedure during application of messenger may then reveal change in position and slope of curve: lowering of curve indicates hyperpolarization; reduction of slope (as schematized for GABA effects) indicates reduced input resistance, or increased conductance. Increased slope signifies reduced conductance. Note that control curve intercepts GABA curve at ca. −75 mV, which would be reversal potential for hyperpolarizing GABA response. Use of K‐acetate‐filled electrode is assumed.



Figure 8.

More sophisticated methods for assessing changes in ionic conductance produced by neuromessengers. A: using current clamp, membrane potential is experimentally altered during transmitter action to determine potential at which there is no net ion flow and therefore no change in membrane potential, even though channels are open. This potential (produced by injecting steady DC into cell and assuming pipette contains no Cl) is the reversal potential for response. Ideally, reversal potential occurs at the equilibrium potential for ions (here, probably Cl) mediating response (ca. −75 mV for GABA, curve on right), but under certain circumstances, such as a remote location for active site, technical problems limit the agreement between equilibrium potential for ions involved and reversal potential for measured response. B: voltage clamp. In this example, voltage is held steady at −40 mV and current is measured while GABA is applied. Downward current deflections indicate inward current produced by command voltage pulses. Note that GABA application causes outward current (due to Cl influx) and increase in command current pulses (due to increased conductance). Repetition of this paradigm while clamping membrane potential at different voltage levels would indicate the reversal potential (shown on I‐V curve on right) to be the potential (−75 mV) at which GABA no longer could elicit current flow.



Figure 9.

Theoretical explanation of the method of fluctuation or noise analysis. A: top traces show activity of 10 computer‐simulated channels. Each channel undergoes transitions between open and closed states as a result of a Poisson process. All 10 channels operate independently but have same amplitude and average lifetime in open state. Bottom trace shows summed activity of these 10 channels. Fluctuating signal is produced; it reflects moment‐to‐moment variation in number of open channels present. B: typical observations made during application of fluctuation analysis to study of synaptic channels in biological membrane. Under voltage‐clamp conditions, application of agonist during period indicated by vertical bar results in change in DC membrane current (I). In addition variance of membrane current is seen to increase from prior to agonist to during agonist. This increase is readily apparent in AC trace, which represents condenser‐coupled, amplified varion of DC trace. Additional variance is assumed to reflect moment‐to‐moment changes in number of synaptic channels opened by agonist. C: kinetic properties of these channels can be estimated from power spectral density (PSD) of agonist‐induced current fluctuations. It is assumed that simple kinetic scheme of type used to generate simulated channels in A also controls operation of synaptic channels. Under this assumption, PSD of biological current noise is expected to be of Lorentzian form (C, smooth curve). Mean open time of agonist‐induced channels can then be calculated from half‐power frequency (fc, arrow) of Lorentz curve that affords best fit to observed spectral point. S(f), power spectral density function.

From Mathers and Barker 488


Figure 10.

Various patch‐clamp recording configurations (top) and representative single‐channel recordings from a patch on cultured CNS neuron (bottom). Single‐channel recordings can be made in 3 configurations: cell‐attached, inside‐out, or outside‐out patch. In addition membrane patch can be broken and recording of whole‐cell voltage or current obtained (whole‐cell recording). Single‐channel activity (bottom) is recorded as current flow through open channel and appears as boxlike events. In this patch (cell‐attached) at least 2 channel types were present, as indicated by amplitudes of events. Brief upward events are channel openings that were not fully resolved because of limitations of recording equipment. Open (O) and closed (C) states for largest events are indicated. A 70‐mV depolarization was applied to membrane patch. Analysis of channel activity in this patch indicated that both channel types were K+ channels, 1 with a single‐channel conductance of 100 pS (•) and the other of 20 pS (•).

D. L. Gruol, unpublished observations


Figure 11.

Simplified coronal section of spinal cord. Spinal cord neurons were 1st mammalian central neurons studied intracellularly (see refs. 97,225,226), and considerable electrophysiological data are still derived from various types of spinal cord (or associated spinal ganglionic) preparations. Spinal cord circuits are often considered to be relatively simple, yet more recent data show them to be extremely complicated. Diagram is greatly simplified to highlight neurons and pathways most often studied by synaptic physiologists and pharmacologists. Cell body and fiber locations are shown on left; laminae and other regions of gray matter are numbered on right. Gray matter of the spinal cord consists of dorsal and ventral horns; these are surrounded by fiber tracts (white matter) of ascending and descending fibers to and from brain and other spinal regions (not shown). Sensory information enters spinal cord via dorsal roots, along axons of the monopolar neurons whose cell bodies lie in dorsal root ganglia (dorsal root ganglia neurons). These axons then enter spinal cord and terminate either on neurons in ventral horn or on 1 of several cell types in the various laminae of dorsal horn. Those axons (from muscles) terminating on ventral horn neurons follow 1 of 3 projections: 1) monosynaptically, from Ia afferents onto agonistic motoneurons via excitatory terminals, with collaterals onto inhibitory interneurons that project to antagonistic motoneurons (constituting stretch reflex); 2) disynaptically, from Ib afferents that terminate on excitatory interneurons projecting to agonistic motoneurons, with collaterals terminating on inhibitory interneurons that project to antagonist motoneurons [inverse myotatic reflex, not shown; (see refs. 121,694)]; or 3) mono‐ or disynaptically from group II (muscle spindle) fibers that project either directly onto agonistic motoneurons or through inhibitory and excitatory interneurons to agonistic and antagonistic motoneurons (see refs. 121,694). Those pathways (e.g., from skin) that project onto dorsal horn neurons in turn transmit information in 1 of 2 ways: 1) fibers from dorsal horn cells cross midline and then ascend contralaterally to thalamus through spinothalamic tract or ascend in ipsilateral dorsal column to brain stem (not shown); or 2) dorsal horn cells receiving terminations of Aδ−, Aβ‐, and C‐fibers (from mechano‐, thermo‐, and pain receptors) project, probably through other interneurons, both to ipsilateral and contralateral motoneurons in ventral horns, and to spinothalamic tracts. Pathways to the motoneurons constitute part of flexor reflex, in which ipsilateral motoneurons to flexor muscles are activated by stimuli (e.g., pain), while the motoneurons innervating extensor muscles are inhibited. Reverse situation prevails contralaterally: flexors are relaxed and extensors activated. Sensory fibers entering the spinal cord through the dorsal roots probably use glutamate (Glu) (Ia, Ib, and II afferents), substance P, somatostatin (SS), vasoactive intestinal peptide (VIP), and cholecystokinin (CCK) (C‐fibers); the various interneurons in dorsal horn probably contain GABA and enkephalin, among others. In addition some descending fibers terminating in dorsal horn contain NE, DA, and serotonin [see the chapter by Zieglgänsberger in this Handbook; 481,501,694]. Several phenomena of pain perception probably involve integration in the dorsal horn substantia gelatinosa (SG). Elucidation of the interconnections of various interneurons and the histochemical delineation of presence of several putative neurotransmitters in this sensory region of spinal cord should lead to a greater understanding of nociception in general. Several ventral horn transmitters are also known. Thus motoneurons are cholinergic, Renshaw cells (RC) and some other inhibitory interneurons probably use glycine (Gly) as their transmitters, and others probably use GABA (see refs. 21,159,160,162,163,164,694). Glu or aspartate (Asp) would seem likely candidates as transmitters for some of the excitatory interneurons (see refs. 161,694). IN, interneuron; γMN, γ‐moto‐neuron; and αMN, α‐motoneuron.



Figure 12.

Simplified coronal section of hippocampus. Hippocampal formation is a well‐defined cortical structure extensively studied with anatomical, biochemical, and electrophysiological techniques (see refs. 98,465,611,758). We include Ammon's horn and the dentate in the hippocampus proper. Neuronal elements comprising the hippocampus are organized into well‐defined lamellar structures that are maintained throughout its extent. Principal neurons of the hippocampal formation are pyramidal (Pyr) neurons that are arranged in a linear fashion in Ammon's horn. Pyr cell layer has been subdivided into areas CA1–CA4 based on anatomical and physiological differences between Pyr neurons and the synaptic organization within these regions 98,393,465,611. However, some researchers do not consider CA2 and CA4 as distinct areas. Pyr neurons receive input from local interneurons, from each other, and from other CNS regions via fornix (leftmost input pathway), from medial septum, the perforant pathway from entorhinal cortex and the alvear commissural pathways from the contralateral hippocampus. Axons of the Pyr neurons form the only efferent pathway leaving the hippocampus. Axon collaterals from CA3 Pyr neurons, called Schaffer collaterals, also provide excitatory input to other CA3 Pyr neurons and to interneurons and Pyr neurons in CA1 and CA2 region of the hippocampus. Excitatory amino acids (Glu or Asp) are putative transmitters for this pathway (see refs. 141,143,758). Other neuron types have been identified in layers adjacent to the Pyr cell layer. Those most well characterized are basket (B) cells 393,465,611 that provide inhibitory input to the Pyr neurons. B cells are activated by axon collaterals originating from nearby Pyr neurons or from distant or even contralateral Pyr neurons and are thought to use GABA as their transmitter (see refs. 10,13,19,393,758). Some neuropeptides (e.g., opioids, VIP, CCK, and SS) have been identified within other hippocampal interneurons, but their function is not clear (see Peptides, p. 61, and Other Peptides …, p. 74). Serotonergic, noradrenergic, and cholinergic fibers have also been described in hippocampus (see refs. 72,73,402,451,452,467,496,500,501,638,695,746,756,758). Targets for these fibers are not yet fully characterized, although it seems likely from light‐microscopic studies that NE projects to CA3 Pyr cells and dentate granule cells (Gr) and that cholinergic fibers contact CA1 and CA3 Pyr neurons 406,451,452,496,500,501,695,746. Principal neurons in the dentate are intrinsic excitatory neurons known as Gr cells. These cells receive afferent input from fibers in the perforant and commissural pathways. Gr cells send axons to Pyr neurons and interneurons in hilar, CA3, and CA4 regions. Several putative neurotransmitters have been suggested for Gr cells, including amino acids (see refs. 142,143), opioids 280,492,737,738, and other peptides (see refs. 737,738). Catecholaminergic and serotonergic inputs are also present in dentate 406,500,501. M, molecular layer.

Courtesy of G. R. Siggins and S. J. Henriksen


Figure 13.

Principal neurons and afferents in cerebellar cortex. The cerebellum is another laminated region that has been extensively studied anatomically, biochemically, and physiologically. Its neuronal organization and synaptic connections have been well characterized 98,227,574 and are similar throughout the structure. In the cortical region, 5 neuronal types are present: Purkinje neurons (P), basket cells (Ba), stellate cells (St), Golgi cells (Gg), and granule cells (Gr). Gr cells are the only excitatory neurons; all others are inhibitory. Parallel fibers, formed by axons of Gr cells, provide excitatory input to the other 4 types of neurons in cerebellar cortex. Glu and Asp are putative transmitters for the Gr cells [see EXCITATORY AMINO ACIDS …, p. 34; 141,142,741]. Ba and St cells provide inhibitory input to the P neurons, whereas Gg cells provide feedback inhibition to Gr cells (not shown). P cell axons are the only efferent pathway from cerebellar cortex and provide inhibitory input to neurons of deep cerebellar nuclei (DCN), whose axons in turn exit the cerebellum for other regions of CNS. GABA is strong transmitter candidate for all of the inhibitory neurons in cerebellum, including P cell (see refs. 70,826). At least 3 afferent pathways transmit information from other regions of CNS to cerebellar cortex. Climbing fibers (CFs), which are axons originating from cell bodies in inferior olive, provide powerful, bursting excitatory input to P cells. CF axon collaterals also innervate the other 4 types of neurons in the cerebellar cortex and DCN 574. Mossy fibers (MFs), whose axons originate from neurons in several CNS regions (including pons), provide excitatory afferents to Gr cells and Gg cells of cerebellar cortex and to DCN (not shown). Glu, Asp, and serotonin (latter from dorsal raphe nucleus) are putative excitatory transmitters for these MF pathways (see refs. 89,141,142). A 3rd afferent input arises from neurons of nucleus locus coeruleus (LC) and provides inhibitory but enabling 80 input to P neurons; NE is the transmitter for this pathway [see NOREPINEPHRINE, p. 40; 345,347,700].

Courtesy of F. E. Bloom


Figure 14.

Membrane current fluctuations to iontophoretically applied GABA in cultured mouse spinal neuron voltage clamped (Vc) at −70 mV. Membrane current is displayed unfiltered on DC trace (B) and at 10 × gain on AC trace (C) filtered at 0.2–200 Hz. Variance associated with filtered signal, updated at 1‐s intervals, is displayed in D. Increasing iontophoretic currents cause inward current responses of increasing amplitude, each of which is associated with a thickening of the DC and AC traces and increases in membrane current variance. Largest changes in variance at beginning and end of current responses reflect relatively rapid changes in membrane current occurring at these times due to AC coupling. Arrowheads mark spontaneous inward current events, which have a fast rise time and exponential decay, suggesting they are synaptic in origin.

From Barker et al. 39


Figure 15.

Gly‐ and GABA‐receptor channel currents recorded from outside‐out membrane patches isolated from soma of 3 different spinal cord neurons in culture. Patches were isolated from neurons bathed in normal bath solution [in mM: 140 NaCl, 1 MgCl2, 1 CaCl2, 1 KCl, and 10 Na‐HEPES (pH 7.2)], which was then exchanged for a solution containing Tris+ as major cation. For most patches this procedure removed background current activity. Tris+ substitution did not alter either chemo‐sensitivity of membrane to Gly or to GABA or conductance properties of the activated Cl channels. Pipette solution facing intracellular side was (in mM): 140 KCl, 3 NaCl, 1 MgCl2, 11 K‐EGTA, and 10 K‐HEPES (pH 7.2). All recordings were done at room temperature (22°C–25°C). A: top trace shows absence of current activity before and after rapid application (indicated by bar) of 50 μM GABA to bath solution. GABA was then washed out, and bottom trace shows increase in current caused by addition of 50 μM Gly (day 35 neuron). B: outside‐out patch in which application of 20 μM Gly did not increase current, but after washout of Gly and application of 20 μM GABA, there was a large increase in current, which decreased with time so that current steps of ∼2 pA were discernible. Note much noisier appearance of GABA‐activated current compared with the Gly‐activated current in A, which reflects higher frequency of brief interruptions evident in single GABA‐receptor currents compared with Gly‐receptor currents. C: outside‐out patch in which application of 5 μM GABA (not shown) activated a relatively low frequency of current steps (2‐pA amplitude), but when concentrations of GABA was increased (bar) to 50 μM, additional channels were activated. At this higher concentration current response quickly and completely desensitized so that after 90 s, no current steps were evident. At this point, addition of 20 μM Gly to bath solution activated current steps of 3‐pA amplitude.

From Hamill et al. 324. Reprinted by permission from Nature, copyright 1983, Macmillan Journals Limited


Figure 16.

Multiphasic synaptic responses of CA1 Pyr neuron in hippocampal slice preparation to stimulation of stratum radiatum (SR). Top traces were recorded at resting membrane potential (RMP) and lower traces while membrane was artificially hyperpolarized by negative current injections in amounts (mV) shown. Vertical arrow indicates time of SR stimulation. Stimulus artifact is followed by short EPSP (seen best in left recording hyperpolarized by 5 mV); in recording at RMP, the EPSP is followed by an early (angled arrow) and a late (•) hyperpolarization. Note that early hyperpolarization, probably GABAergic IPSP (see refs. 9,10,19,195,529), is reduced and then inverted by increasing hyperpolarizations, whereas late hyperpolarization is nullified only at greater hyperpolarizations (−11 to −16 mV below RMP). These and other data (see refs. 195,529,766) suggest that the early IPSP has a reversal potential near the expected equilibrium potential for Cl, whereas late hyperpolarization [possibly late or slow IPSP 529,766] has a reversal potential more like that for K+. Spikes (spontaneous or those driven by strong SR stimulation) are attenuated by slow rise time of polygraph.

D. L. Gruol, unpublished observations


Figure 17.

Inhibitory effectiveness and Cl dependence of depolarizing IPSP (in rat hippocampal slice preparation). A: train (2 Hz) of depolarizing current pulses initiated full‐sized somatic action potentials.

(Amplitudes not accurately reproduced by pen recorder), which were blocked during most of stimulation‐evoked response. B: inhibition of spikes elicited by just suprathreshold EPSP during depolarizing IPSP 2. Unconditioned control responses are shown before and after inhibition (1 and 3, respectively). Same cell as in A. RPM = −59 mV. C: effects of iontophoretic GABA (1 M; horizontal bar) in normal medium without pentobarbital. 1, Application through pipette in stratum pyramidale. Current pulses of 100 ms were given at 1 Hz. (see Fig 1B, pt. 1 of ref. 10). RPM = −55 mV. 2 and 3, GABA iontophoresed in stratum radiatum of another cell. During depolarization, both directly and synaptically activated action potentials are blocked (2 and 3, respectively). RPM = −61 mV. Iontophoretic pipettes were lowered independently 200 μm into slice before impaling a cell. Ejection currents were 500 nA for 5 s. D: 2 cells from same slice bathed in pentobarbital. 1, Recorded with a 2 M K‐methylsulfate‐filled pipette; 2, recorded with a 3 M KCl‐filled pipette. Note large amplitude and prolonged time course of depolarizing response in 2, after Cl leakage into cell. Small reversed IPSPs visible on base line of KCl‐filled cell are able to trigger action potentials. Antidromic responses from same cells are shown for comparison (2 and 4). RPM= −59 mV (1 and 3) and −56 mV (2 and 4). Calibration for pen traces in A, B, and D: 5 mV, 2 s; C: 5 mV, 10 s. From Alger and Nicoll 10. Reprinted by permission from Nature, copyright 1979, Macmillan Journals Limited


Figure 18.

Responses of a single cultured brain neuron to L‐Glu (A, 100 μM), D,L‐kainate (B, 100 μM), D,L‐homocysteic acid (D,L‐HCA; C, 100 μM), and L‐Asp (D, 100 μM) demonstrating decreased conductance. All amino acids were applied (C2: 150 ms; D2: 450 ms; all others: 350 ms) in the presence of 3 μM tetrodotoxin (TTX) (no spikes were evoked with brief cathodal current pulses). Trace below each voltage recording indicates period of drug application (upward deflections). Constant current anodal pulses (50 ms; not shown) were used to assay Gm in (A1–D1). Recordings were at neuron's resting level, −60 mV. Calibrations: A2, C2, and D2: 4 s; D1 and A1: 8 s; and B1, B2, and C1: 20 s. C1: 40 mV; all others: 20 mV. A: responses to L‐Glu. Initial phase is large depolarization (∼18 mV) accompanied by rapid rise in conductance (Gm; reduction in voltage deflections). Later, Gm drops dramatically. Note small, irregular potentials on falling phase of A2. Longer applications of L‐Glu lead to full spikelike potentials (not shown). B: responses to D,L‐kainate. Entire response to D,L‐kainate is increase in Gm. Gm remains elevated even when membrane falls close to preapplication values. Note lack of spikelike potentials in B2. Shape of offset of D,L‐kainate response differs from other amino acids because of absence of decreased Gm. In addition the duration is significantly greater than those to L‐Glu and L‐Asp. C: responses to DL‐HCA. Responses were particularly large and lasted longer than those to L‐Glu and L‐Asp. Although initial depolarization is associated with increase in Gm, response quickly reverts to surprisingly large decrease in Gm. Gain of C1 was reduced because of the size of voltage deflections. C2: generation of large spikelike potentials occurring during falling phase of response. D: responses to L‐Asp. This amino acid induces a depolarization comparable to that evoked by L‐Glu, but in this case only a decrease in Gm was observed. Spikelike potentials are also seen (D2). Higher concentrations of L‐Asp (500 μM) also increased Gm (not shown).

From MacDonald and Wojtowicz 471


Figure 19.

N‐methyl, D‐aspartate (NMA) produces dose‐related biphasic conductance change. A: responses of cell to increasing ionophoretic currents of NMA. Small NMA depolarizations are accompanied by apparent rise in input resistance, whereas large responses are dominated by fall in input resistance. B: NMA pulses (150‐ms wide), with and without constant current negative pulses to measure input resistance. C: in another cell, a 40‐ms, 180‐nA NMA pulse produces an apparent rise in input resistance, as measured with positive current pulses, with no underlying depolarization. A−C from different cells.

From Dingledine 192


Figure 20.

I−V curve for cultured spinal cord neuron under voltage clamp (2‐electrode clamp). RMP was −38 mV and both depolarizing and hyperpolarizing command steps were employed to construct curve. Control curve (•) was performed in bathing solution supplemented with TTX (2 μM) and repeated during constant microperfusion with L‐aspartic acid (500 μM; •). Net inward current (negative by convention) was evoked by L‐Asp. Slope of this steady‐state relationship was reduced in range from −70 to −30 mV.

From MacDonald et al. 470


Figure 21.

Intracellular recordings from rat cerebellar Purkinje cells in vivo. A: recording and electrophoretic setup: 3‐barreled micropipette with a Purkinje cell. Intracellular electrode protrudes beyond orifices of the 2 extracellular microelectrophoretic barrels. B: multispiked spontaneous climbing fiber discharge obtained during intracellular recording from a Purkinje cell. RMP (in mV) is given in parentheses. Calibration: 20 ms and 25 mV. C: changes in membrane potential and membrane resistance of 4 Purkinje cells in response to drugs. All specimens in each row of records are from same cell. Bar above each record indicates extracellular electrophoresis of indicated drug (100–150 nA). RMP (in mV) is given in parentheses below each recording. Calibration: 10 s and 20 mV for NE, dibutyryl cAMP (DB), and cAMP; 5 s and 10 mV for GABA. Right: effective input resistance was judged by size of pulses resulting from passage of a brief constant current (1‐nA) pulse across the membrane before, during, and after electrophoresis of respective drugs (1 mV = 1 MΩ). Discontinuities in fast transients of pulses (and loss of spikes) result from loss of high frequencies (>10 kHz) and from chopped nature of the frequency‐modulated magnetic tape recording used. All pulse records were graphically normalized to same base‐line level. Calibration: 80 ms and 15 mV for all pulse records.

From Siggins et al. 716. Copyright 1971 by the American Association for the Advancement of Science


Figure 22.

Intracellular recording from a CA1 pyramidal neuron in hippocampal slice in vitro preparation. Isoproterenol (IP) reduced amplitude and duration of afterhyperpolarization (AHP) evoked by depolarizing current pulses that generated action potentials (attenuated by polygraph). AHP is a hyperpolarization and inhibition of spontaneous activity after termination of depolarizing current pulse (arrow). Amplitude and duration of AHP increases as amplitude of depolarizing pulse increases. Note also activity increase produced by IP.

From Gruol and Siggins 302


Figure 23.

Intracellular recordings show effect of adrenergic agonists on membrane potential and spontaneous activity of CA1 pyramidal neurons in hippocampal slice preparation of rat. A: typical response of CA1 pyramidal neuron to superfusion of IP. In all neurons tested, IP evoked an increase in subthreshold activity (indicated by thickness of base line) and action‐potential generation (large upward deflections). Superfusion of IP began at downward arrow and stopped at end of bracket. Delay to onset of increase in activity is partially due to “dead time” (1–2 min) of perfusion system. Increase in activity evoked by IP was prolonged in duration and far outlasted estimated washout time (2–4 min). Note depolarization evoked by IP. B: typical response of CA1 pyramidal neurons to the α‐adrenergic agonist clonidine. NE decreased spontaneous activity and produced a small hyperpolarization. Concentrations of 2–10 μM were usually required to produce this effect. This is same neuron as in A, where superfusion with β‐agonist IP increased spontaneous activity. Clonidine mimicked action of NE in this neuron, but in other neurons NE was excitatory whereas clonidine was inhibitory. These data suggest that CA1 pyramidal neurons have both α− and β‐adrenergic receptors.

D.L. Gruol and G.R. Siggins, unpublished observations


Figure 24.

Suggested locus of NE action at membrane of hippocampal pyramidal cell. NE, in binding with β‐adrenergic receptor (triangular notches) on neuronal membrane, could block either of 2 ion channels: 1) the Ca2+ channels that are voltage‐sensitive channels opened by a depolarization such as occurs during an action potential; or 2) the Ca2+‐dependent K+ channels that are opened by influx of Ca2+. Data showing no NE effect on Ca2+ spike during TTX treatment (see refs. 302,477) suggest that 2nd possibility is more likely. Because NE‐induced inactivation of the Ca2+‐dependent K+ channels would retard repolarization of membrane during an action potential, more Ca2+ might then enter the cell during this state of depolarization by virtue of voltage‐dependence of Ca2+ channel.



Figure 25.

NE activates K+ conductance in the locus coeruleus in vitro. A: pressure application of NE [50 ms, 10 psi (68 kPa); ▴] produced hyperpolarization and increased membrane conductance. Top traces (in 1 and 2) are of membrane potential and bottom traces are of membrane current. Parts 1 and 2 are from same neurons. NE was applied twice in each trace. Full amplitude of NE‐induced hyperpolarization is seen after 1st application. After 2nd application, membrane potential was maintained at rest with manual‐clamp technique. Twice as much NE was applied in 2 as in 1. B: determination of NE reversal potential in solutions of varying K+ content. A single pressure‐ejection pulse [50 ms, 10 psi (68 kPa)] was applied at each arrowhead. Left: normal, 2.5 mM K+ solution; as membrane was hyperpolarized, NE potential decreased and reversed at −100 mV. Middle: 4.5 mM K+ solution; reversal potential decreased to −85 mV. Right: 10.5 mM K+ solution; reversal potential was −65 mV.

From Egan et al. 229


Figure 26.

Reversal potential of IPSP in locus coeruleus in vitro. A: synaptic potentials were evoked by single pulse (30 V, 0.5 ms; arrow) at different RMPs. At −67 mV the IPSP was hyperpolarizing, at −105 mV it was nullified, and at −127 mV it was depolarizing. Durations of EPSP and IPSP were reduced at more negative potentials, probably because of membrane rectification at these potentials. B: amplitude of IPSP in another cell is plotted as function of membrane potential in solutions with different K+ content. •, Normal K+ concentration (2.5 mM); •, 4.5 mM K+; and ▴, 10.5 mM K+. Estimated reversal potentials were −121, −105, and −92 mV. (These reversal potentials were the most negative of cells included in Table 1 of ref. 229.)

From Egan et al. 229


Figure 27.

Effects of iontophoretic DA and protons (H+) on cat caudate neurons in vivo. A: iontophoretically applied DA depolarizes membrane of cat caudate neuron and decreases its firing rate, whereas Na+ ejected at twice the current intensity and same pH used for DA had no comparable effect. B: H+ ejected from barrel containing HCl at pH of 0.5 produces long‐lasting increases of firing rate of another caudate neuron (ratemeter output). Excitatory effect of Glu is much faster in onset and offset, whereas DA had no effect on firing rate of this cell displaying no spontaneous action potentials. 2/1 and 69Q2, Experiment and cell numbers.

From Herrling and Hull 337


Figure 28.

DA‐induced effects with short (50 μm) distances between tips of recording and iontophoretic electrodes on 2 different cells of cat caudate in vivo. A: DA induces hyperpolarizations. B: if DA is applied for longer period on another cell, initial hyperpolarization is followed by slow depolarization. Single arrow, cortically evoked EPSP; double arrows, cortically evoked EPSP with superimposed action potential. Glu, glutamate.

From Herrling and Hull 337


Figure 29.

Effect of carbachol on current relaxations generated by hippocampal CA1 neuron (in slice preparation) at rest and at depolarized holding potential. Traces are currents initiated by hyperpolarizing voltage‐clamp step (ΔV) of 14 mV from holding level of −42 mV (top row) and from close to resting potential, −62 mV (bottom row), before, during, and 10 min after brief (2 min) bath application of carbachol (48 μM). During drug administration an inward current developed at the more positive holding potential, as indicated by positioning of traces, but not when cell was held at −62 mV. Note approximately ohmic behavior of cell at latter potential, whereas a similar current response was only observed in presence of carbachol at depolarized level, i.e., when time‐dependent inward relaxation had been abolished. TTX (0.5 μM) was included in bathing medium to abolish Na+‐dependent action potentials; bath temperature was 23°C.

From Halliwell and Adams 323


Figure 30.

Effects of ACh on rat CA1 pyramidal cells in hippocampal slice preparation. All responses are from same cell. A: chart record of control AHP after 60‐ms direct depolarizing current pulse (trace 1) and film record of response to 600‐ms pulse (trace 2). Current record is positioned below voltage record. B: ACh (200 μM) superfusion depolarized membrane and increased cell's input resistance. C: blockade of AHP (trace 1) and accommodation (trace 2) in presence of ACh. D: addition of atropine (0.5 μM) in presence of ACh reversed effects of ACh. E: atropine also reversed effect of ACh on AHP (trace 1) and on accommodation (traces 2 and 3). Current pulse in trace 3 was identical to those of trace 2 in A and C. Current pulse was increased in trace 2 to match depolarization evoked in presence of ACh (trace 2) in C. Gain in trace 1 in A applies to all chart records. Time calibration for trace 1 in A also applies to trace 1 in C and E, and time calibration in D and B are the same. Calibration (time, volts, and current) for trace 2 in A applies to all film records. RMP = −57 mV.

From Cole and Nicoll 134. Copyright 1983 by the American Association for the Advancement of Science


Figure 31.

Hippocampal slice preparation. Effect of exogenous ACh mimicked by stimulating stratum oriens. Records from 4 different cells are shown (A‐D). A: immediately after train of stimuli (S), cell hyperpolarizes (condensed on time scale) and then depolarizes (arrow in trace 1). In trace 2, slow depolarizaton was manually voltage clamped, which clearly shows increase in size of hyperpolarizing current pulses. RMP = −53 mV. B: slow depolarization (trace 1) was completely blocked by 1 μM atropine (trace 2). RMP= −59 mV. Calibration in A also applies to B. C: AHPs before pathway stimulation and 4 s after stimulus. In control solution (trace 1) AHP was slightly depressed after stimulus. After addition of 1 μM eserine (trace 2), AHP was greatly reduced by stimulus (arrow); AHP was restored to control amplitude by 1 μM atropine (trace 3). RMP = −56 mV. D: identical protocol as in C, except showing accommodation during 600‐ms pulse before and after pathway stimulation. RMP = −64 mV.

From Cole and Nicoll 134. Copyright 1983 by the American Association for the Advancement of Science


Figure 32.

Effect of adenosine on membrane potential and input resistance of pyramidal neuron in rat hippocampal slice. A: low concentration (5 μM) of adenosine (•) has no detectable effect on input resistance measured from electrotonic potentials (bottom inset oscillographs in 1 and 2) generated by current injection (top inset oscillographs in 1 and 2) through a bridge circuit and recording electrode. 1: Control; 2, adenosine superfusion; I‐V curve during adenosine superfusion is superimposable with that of control period (x). B: higher adenosine concentration (20 μM) reduces input resistance during hyperpolarization. Right, current and electrotonic potentials before (1), during (2), and after (3) adenosine perfusion. Calibration: 20 mV, 50 ms. Spiking after offset of current pulse was reduced by adenosine. All voltage measurements were taken from later, flat portions of electrotonic potentials. Curves were normalized to RMP. C: 20 μM adenosine hyperpolarizes and inhibits spontaneous discharge of pyramidal neuron via postsynaptic action; slice was continuously perfused with 10 mM MgCl2 to prevent transmitter release. Spikes were attenuated by slow frequency response of polygraph. Brackets show duration of adenosine (plus 10 mM MgCl2) perfusion. Note rebound depolarization often seen after withdrawal of adenosine.

From Siggins and Schubert 717


Figure 33.

Effects of adenosine on Ca2+ spikes from TTX‐treated hippocampal neurons in slice preparation. All traces have been digitized (0.2‐ to 0.4‐ms sample time). A: individual responses before (A) and after (B‐D) 20 μM adenosine. Traces B‐D are successive responses at 15−, 30−, and 60‐s intervals after addition of adenosine to recording chamber. Spike seen in traces A and B is largely abolished in trace C; trace D shows no active response to current injection. Depolarizing current pulse was held constant (1.0 nA) for all traces. B: superimposed voltage traces from cell before (A), during (B), and after (C) perfusion with 20 μM adenosine. Depolarizing current was 0.8 nA (top traces) and 1.1 nA (bottom). C: superimposed records showing response of cell (top traces) to depolarizing current injection before (C1), and during (C2) treatment with 20 μM adenosine. Amount of current injected is illustrated by current monitor (bottom trace). Note that calcium spike can be elicited in presence of adenosine, but that considerably higher currents are required (top 2 traces in C2). Calibration: 10 mV (2 nA for current traces) and 20 ms for each panel.

From Proctor and Dunwiddie 605


Figure 34.

Possible sites of action of nucleosides or nucleotides in nervous system. XYP is a molecule in which X could be any purine or pyramidine base and Y is number of phosphate groups, from 0 to 3. These substances could have several roles: 1) as local modulators released from glia or neurons and acting either pre‐ or postsynaptically; 2) classic transmitters; 3) presynaptic modulators of transmitter release; 4) cotransmitters with some other transmitter such as ACh; 5) 2nd messengers acting at several possible sites, including at some nucleotide cyclase, which would then generate a 3rd messenger such as cAMP or cGMP; and 6) unconventional transmitters acting on voltage‐dependent conductance such as M‐current (IM).



Figure 35.

Enkephalin opens K+ channels in locus coeruleus neurons in vitro. A: D‐Ala, D‐Leu‐enkephalin (DADL) was applied by pressure [arrow; 50‐ms pulse, 70 kN/m2 (10 psi)] at various membrane potentials. The K+ concentration was as indicated (in mM). At elevated K+ concentrations, enkephalin response reversed polarity with membrane hyperpolarization. Marked rectification occurred with hyperpolarization and elevated K+ concentration. B: relationship between hyperpolarization amplitude and membrane potential. C: enkephalin reversal potential (Erev) calculated from many experiments such as those in A and B. Dashed line, Erev = RT/F ln([K]o/[K]i), where [K]i, intracellular K+ concentration, was 145 mM. [In normal K+ solutions (2.5 mM), it was usually not possible to reverse enkephalin hyperpolarization; therefore extrapolated values were used.] Good agreement between observed values of Erev and those predicted from Nernst equation (Eq. 4) suggests a predominantly somatic site of action; large steady‐state conductance increase that accompanied membrane hyperpolarization would effectively remove dendritic contribution.

From Williams et al. 814. Reprinted by permission from Nature, copyright 1982, Macmillan Journals Limited


Figure 36.

Stratum radiatum (Rad) stimulation produces EPSP/IPSP sequence in CA1 pyramidal cells (A1) of hippocampal slice. Application of 100 μM D‐Ala, D‐Leu‐enkephalin (DADL) reduces IPSP (A2) and leads to depolarization and burst of spike activity after EPSP (A3). Alvear (Alv) stimulation produces an IPSP (B1) that is blocked by 100 μM DADL (B2). Recovery can occur within 10 min (B3). However, in presence of 2 μM naloxone, alvear‐induced IPSP (C1) is not blocked by repeated applications of 100 μM enkephalin (C2). •, Time of stimulation. Calibration: A, 10 mV and 40 ms; B and C, 5 mV and 40 ms.

From Masukawa and Prince 486


Figure 37.

COOH‐terminal portion of porcine prodynorphin sequence (residues 171–256) deduced by Kakidana et al. 390. Three [Leu5]enkephalin‐containing segments are bracketed by putative peptide‐processing signal “Lys‐Arg”. α‐Neo‐ and β‐neoendorphin overlap as prodynorphin residues 175–184 and 175–183, respectively. Dynorphin A(1–17) corresponds to full sequence of 2nd segment. Dynorphin A(1–8) is produced by unusual cleavage to left of Arg217. Similar cleavage to left of Arg241 produces dynorphin B from 3rd [Leu5]enkephalin‐containing segment. Presence of full sequence of dynorphin B(1–29) in brain tissue has not been reported.

From Chavkin et al. 127


Figure 38.

Substance P (SP) responses of cultured spinal cord neuron. Clear reversal of SP‐response polarity was not obtained. SP and eledoisin‐related peptide response amplitude varied as function of membrane potential, but responses did not reverse polarity. SP‐response amplitude increased when membrane was depolarized and decreased when it was hyperpolarized by 1 of 2 intracellular micropipettes. Response amplitude did not vary as linear function of membrane potential at potentials <‐50 mV or >‐95 mV in this neuron. Although activation of voltage‐dependent conductances might reduce voltage response recorded at potentials <‐50 mV, it was not clear why no reversal of response polarity was seen. Even when this neuron was hyperpolarized to very negative potentials (−106 and −114 mV), SP responses did not reverse polarity. RPe, extrapolated reversal potential.

From Nowak and Macdonald 551


Figure 39.

Structure of the presumed somatostatin (SS) prepro‐hormone SS‐28 and 2 fragments, SS‐28(1–12) and original SS‐14, or SS‐28(15–28), derived from SS‐28. Heavy brackets with arrows denote the 2 fragments; light bracket indicates ring structure of SS‐28 and SS‐14. Note that SS‐28 is NH2‐terminally extended version of SS‐14. Cleavage of SS‐28 to SS‐14 would also appear to liberate SS‐28(1–12) and an Arg‐Lys dipeptide (see refs. 63,64).



Figure 40.

Effect of SS on CA1 pyramidal neuron, recorded with intracellular microelectrode in rat hippocampal slice preparation. Slices were 250 μm thick. Recording microelectrodes were filled with 1 M K‐acetate (tip resistance = 90–150 MΩ). Center of a 4‐barreled microelectrode was filled with 3 mM solution of SS in 5 mM Na‐acetate (pH 7.0), and the peripheral barrels were filled with 1 M NaCl, 1 M Na acetate, 0.5 M ACh chloride (pH 4.5), or 1 M monosodium l‐Glu (pH 6.7). A backing current of 25 nA and appropriate polarity was applied to each barrel throughout experiment. In some experiments synthetic SS was used with similar results. SS was tested only on cells from which stable intracellular RMP recordings > −60 mV were obtained. A: SS [somatomedin‐release inhibitory factor (SRIF)], applied by passage of 196 nA negative current (−ve), caused depolarization of 21 mV (RMP of −75 mV). Termination of application resulted in release of some voltage coupling between recording and iontophoretic electrodes, which masked an even greater depolarization. Depolarizing pulses and both the evoked and spontaneously occurring action potentials disappeared when photographic conditions suitable for reproduction of changes in DC level at high gain were selected. B: same record at a faster sweep speed shows effect of SS on excitability of cell at times a‐d indicated in A. Bottom trace shows depolarizing current pulse injected through recording electrode, adjusted so as to evoke 2 action potentials in resting conditions. C: similar records show excitation of cell by l‐Glu (GLUT).

From Dodd and Kelly 203. Reprinted by permission from Nature, copyright 1978, Macmillan Journals Limited


Figure 41.

Effect of SS analogue in medium with high Mg2+. A: membrane potential response of pyramidal cell in hippocampal slice. Cell displays reversible hyperpolarization in response to addition of SS nonapeptide agonist Ac‐Cys3‐Des AA1,2,4,5,12 (D‐Trp8, D‐Cys14)‐SS to perfusate. Artificial cerebrospinal fluid contains 12 mM MgCl2, which eliminated stratum radiatum‐induced action potentials. B: voltage deflections in response to depolarizing and hyperpolarizing current before, during, and after SS‐analogue perfusion. C: I‐V curves constructed from hyperpolarizing potentials. Decrease in slope of the I‐ V curve indicates that SS analogue increased conductance in this cell.

From Pittman and Siggins 601


Figure 42.

SS‐28 hyperpolarizes and inhibits activity of hippocampal pyramidal cell in vitro with a slight decrease in input resistance. SS‐28 was superfused onto hippocampal slice preparation during intracellular recording. Deflections below the membrane potential line are electrotonic potentials resulting from intracellular current injection through a bridge circuit. Note small reduction in these potentials during SS‐28‐induced hyperpolarizations. This apparent reduction in input resistance is also indicated by slight diminution of slope of I‐V curve generated during SS‐28 superfusion [membrane potential normalized to RMP (Vm) during SS‐28]. Note slow onset and long duration of action of SS‐28 and reduction in subthreshold (thin vertical lines above RMP) and spike activity (thick vertical lines). Spikes were attenuated by slow rise time of polygraph.

D. L. Gruol and G. R. Siggins, unpublished observations


Figure 43.

Effect of corticotropin‐releasing factor (CRF) on current‐evoked bursts of action potentials and associated AHPs in CA1 pyramidal cell of hippocampal slice. A: 0.5 μM CRF reduces AHPs that follow trains of spikes generated by passing depolarizing current pulses of 3 different intensities through active bridge circuit and recording electrode. Late components of AHPs are completely lost during CRF superfusion, despite greater number of spikes generated by equivalent current strengths. CRF also increases spontaneous firing and depolarizes the cell (membrane potentials are indicated for each trace). B: effect of CRF in presence of TTX (100 nM). TTX alone (left) abolishes early, fast‐action potentials, leaving slower Ca2+ spike. CRF (400 nM) added to TTX solution (middle) nearly abolishes AHP (arrows), but did not significantly alter Ca2+ spike. Recovery is shown on right. V, voltage trace; I, current trace. Current pulse = 0.2 nA; calibration: 10 mV and 0.2 s.

From Aldenhoff et al. 8. Copyright 1983 by the American Association for the Advancement of Science


Figure 44.

Suggested locus of CRF action on hippocampal pyramidal cell membrane. CRF molecule, in binding with a CRF receptor (triangular notches) on neuronal membrane, could block either of 2 ion channels: 1) Ca2+‐channels that are voltage‐sensitive channels opened by depolarization, e.g., during action potential, or 2) Ca2+‐dependent K+ channels, which are opened by Ca2+ influx. Data showing lack of a CRF effect on Ca2+ spike during TTX treatment 8,707 suggests that the 2nd possibility is more likely. Because CRF inactivation of Ca2+‐dependent K+ channels would retard repolarization of membrane during an action potential, more Ca2+ might enter the cell during this state of depolarization by virtue of voltage dependence of Ca2+ channel.

From Siggins et al. 705


Figure 45.

Comparison of oxytocin/vasopressin related neuropeptide structures.



Figure 46.

Two major brain slice recording methodologies currently used. In both cases slice is usually placed on nylon net (open circles), on an electron microscope grid, or on filter paper. Recording chamber (shown in cross‐section) may be circular or rectangular. A: static or perifusion (microdrop) method involves filling recording chamber with warm artificial CSF (ACSF; gassed with carbogen) only up to edges of the slice. Recording may be done without perfusion because space above slice is in contact with warm carbogen gas saturated with H2O; H2O‐gas interface at top of slice allows passage of O2 into tissue. However, “perifusion” may also be performed, wherein the inflow and outflow of ACSF are perfectly balanced so that the level of ACSF never changes and top of slice remains exposed to carbogen. Microdrop involves pressure application of drugs from pipettes to top, exposed surface of slice. B: superfusion method involves covering the entire slice with warm ACSF saturated with carbogen; fast flow rate (2–5 ml/min) is used to assure adequate oxygenation of tissue. Outflow must balance inflow so fluid level does not change; otherwise, tissue movement and changes in recording characteristics could occur. Drug pipettes in both types of chambers could also be used to pass drugs by iontophoresis as well as by pressure. Several laboratories now use a slice recording method that combines some properties of both methods (see refs. 312,317,318).



Figure 47.

Hypothetical integrative interactions between conventional and newly described synaptic messages. Input T: a conventional test EPSP produced by conductance increase; inputs A‐D: conditioning postsynaptic potentials (PSPs) that can modify the test EPSP, as well as each other. Inputs A, C, and E are EPSPs and B and D are IPSPs; inputs A and B are conventional (i.e., associated with increased conductance or decreased resistance), whereas inputs C‐E are associated with decreased conductance. Results of synaptic interactions, recorded in the soma, are shown as possible summations of various PSP types, and partially result from effects of changes in input resistance and consequently the space constant. “A, then T” shows decremental effects of conduction of A EPSP down dendrites to cell body; the initial EPSP is greatly attenuated compared with potential generated at dendritic sites. Also note shunting of secondary T EPSP by reduced resistance and space constant brought about by A input, even though both PSPs are depolarizing and should summate to some extent. “B, then T” shows similar shunting effect, combined with summation of opposite values. However, in “C, then T”, the increased input resistance (Ri) results in a larger T EPSP, overcoming to some extent effects of decremental conduction down the dendrite. Similar augmentation of T might occur even with hyperpolarization, if IPSP is associated with an increase in Ri (“D, then T”). When EPSPs with increased Ri occur nearly simultaneously (“E, then C”), the 1st may nullify 2nd, especially if same conductances are involved (see the chapter by North in this Handbook). In “E, then D” it is not clear what would result, although logic suggests that summation of opposite signs would nullify the responses. See Fig. 48 for interactive effects of other newly described messages. Vm, membrane potential.

G. R. Siggins, unpublished observations


Figure 48.

Different functional types of interactive messages. Column A: responses displayed over a long time course (0.5–600 s); column B: typical electrotonic potential shape produced by intracellular current injection before (left) and during (right) application of messenger. Curves above long recording traces in A represent sizes and shapes of EPSPs (1–3) and changes in spikes (top trace) and AHPs (bottom trace) (4) before and after application of representative messenger. Sharp downward deflections in traces in A indicate repetitively generated electrotonic potentials, of which those in column B (expanded in time) are representative. Note that shunting messenger in A1 reduces size of another (conductance increase) EPSP, even when both are depolarizing. In fact, both depolarization (by driving membrane potential toward Eepsp) and the reduced input resistance would tend to reduce the conventional EPSP. Furthermore, electrotonic potentials in B1 indicate that decreased input resistance would also shorten membrane rise time (note increased “squareness” of potential with messenger), effectively shortening EPSP. Messengers that increase input resistance in A2 (larger downward deflections) would enlarge EPSPs with conductance increases (and may constitute an enabling mechanism) by increasing voltage drop across membrane and by increasing membrane time constant (note longer time required for electrotonic potential in B2 to reach steady state). However, a depolarizing enabling messenger might still reduce the EPSP by driving membrane potential substantially nearer Eepsp, thereby reducing driving force for ion flow. Note long time course (see calibrations of enabling responses compared with those of shunting messengers). Modulating messengers in A3 might disenable (reduce) or enhance (not shown) other synaptic inputs, without altering membrane properties (B3). This enhancement could also be a 2nd mechanism for enabling. Messengers that alter voltage‐dependent conductances in A4 might prolong (or shorten) spike (above trace) by enhancing the voltage‐dependent Ca2+ conductance, or by decreasing K+ conductances, but without altering resting membrane properties (B4). AHPs (below trace) might also be reduced or augmented by similar effects on voltage‐dependent Ca2+ or K+ conductances, again without any effect on resting properties.

References
 1. Abe, H., M. Inoue, T. Matsuo, and N. Ogata. The effects of vasopressin on electrical activity in the guinea‐pig supraoptic nucleus in vitro. J. Physiol. London 337: 665–685, 1983.
 2. Adams, M. E., and, M. O'Shea. Peptide cotransmitter at a neuromuscular junction. Science 221: 286–289, 1983.
 3. Adams, P. R., and D. A. Brown. Synaptic inhibition of the M‐current: slow excitatory post‐synaptic potential mechanism in bullfrog sympathetic neurones. J. Physiol. London 332: 263–272, 1982.
 4. Adams, P. R., D. A. Brown, and A. Constanti. M‐currents and other potassium currents in bullfrog sympathetic neurones. J. Physiol. London 330: 537–572, 1982.
 5. Adams, P. R., D. A. Brown, and A. Constanti. Pharmacological inhibition of the M‐current. J. Physiol. London 332: 223–262, 1982.
 6. Aghajanian, G. K., and B. S. Bunney. Dopamine “autoreceptors:” pharmacological characterization by microiontophoretic single cell recording studies. Naunyn‐Schmiedebergs Arch. Pharmakol. 297: 1–7, 1977.
 7. Aghajanian, G. K., and C. P. VanderMaelen. α2‐Adrenoreceptor‐mediated hyperpolarization of locus coeruleus neurons: intracellular studies in vivo. Science 215: 1394–1396, 1982.
 8. Aldenhoff, J. B., D. L. Gruol, J. Rivier, W. Vale, and G. R. Siggins. Corticotropin releasing factor decreases postburst hyperpolarizations and excites hippocampal neurons. Science 221: 875–877, 1983.
 9. Alger, B. E., C. E. Jahr, and R. A. Nicoll. Electrophysiological analysis of gabaergic local circuit neurons in the central nervous system. In: GABA and Benzodiazepine Receptors, edited by E. Costa, G. DiChiara, and G. L. Gessa. New York: Raven, 1981, p. 77–91.
 10. Alger, B. E., and R. A. Nicoll. GABA‐mediated biphasic inhibitory responses in hippocampus. Nature London 281: 315–317, 1979.
 11. Alger, B. E., and R. A. Nicoll. Epileptiform burst after hyperpolarization: calcium‐dependent potassium potential in hippocampal CA1 pyramidal cells. Science 210: 1122–1124, 1980.
 12. Alger, B. E., and R. A. Nicoll. Spontaneous inhibitory post‐synaptic potentials in hippocampus: mechanism for tonic inhibition. Brain Res. 200: 195–200, 1980.
 13. Alger, B. E., and R. A. Nicoll. Feed‐forward dendritic inhibition in rat hippocampal pyramidal cells studied in vitro. J. Physiol. London 328: 105–123, 1982.
 14. Alger, B. E., and R. A. Nicoll. Pharmacological evidence for two kinds of GABA receptor on rat hippocampal pyramidal cells studied in vitro. J. Physiol. London 328: 125–141, 1982.
 15. Allen, G. I., J. Eccles, R. A. Nicoll, T. Oshima, and F. J. Rubia. The ionic mechanisms concerned in generating the i.p.s.p.s of hippocampal pyramidal cells. Proc. R. Soc. London Ser. B 198: 363–384, 1977.
 16. Allen, Y. S., T. E. Adrian, J. M. Allen, K. Tatemoto, T. J. Crow, S. R. Bloom, and J. M. Polak. Neuropeptide Y distribution in the rat brain. Science 221: 877–879, 1983.
 17. Alvarez‐Leefmans, F. J., and A. R. Gardner‐Medwin. Influences of the septum on the hippocampal dentate area which are unaccompanied by field potentials (Abstract). J. Physiol. London 249: 14P–16P, 1975.
 18. Andersen, P., H. Bruland, and B. R. Kaada. Activation of field CA1 of the hippocampus by septal stimulation. Acta Physiol. Scand. 51: 29–40, 1961.
 19. Andersen, P., R. Dingledine, L. Gjerstad, I. A. Langmoen, and A. M. Laursen. Two different responses of hippocampal pyramidal cells to application of γ‐amino butyric acid. J. Physiol. London 305: 279–296, 1980.
 20. Andrade, R., and G. K. Aghajanian. Locus coeruleus activity in vitro: intrinsic regulation by a calcium‐dependent potassium conductance but not α2‐adrenoceptors. J. Neurosci. 4: 161–170, 1984.
 21. Aprison, M. H. Glycine as a neurotransmitter. In: Psychopharmacology: A Generation of Progress, edited by M. A. Lipton, A. DiMascio, and K. F. Killam. New York: Raven, 1978, p. 333–346.
 22. Aronin, N., M. DiFiglia, A. S. Liotta, and J. B. Martin. Ultrastructural localization and biochemical features of immunoreactive LEU‐enkephalin in monkey dorsal horn. J. Neurosci. 1: 561–577, 1981.
 23. Ashe, J. H., and B. Libet. Modulation of slow postsynaptic potentials by dopamine, in rabbit sympathetic ganglion. Brain Res. 217: 93–106, 1981.
 24. Assaf, S. Y., S. T. Mason, and J. J. Miller. Noradrenergic modulation of neuronal transmissions between the entorhinal cortex and the dentate gyrus of the rat (Abstract). J. Physiol. London 292: 52P, 1979.
 25. Ault, B., R. H. Evans, A. A. Francis, D. J. Oakes, and J. C. Watkins. Selective depression of excitatory amino acid induced depolarisations by magnesium ions in isolated spinal cord preparation. J. Physiol. London 307: 413–428, 1980.
 26. Ault, B., and J. V. Nadler. Baclofen selectively inhibits transmission at synapses made by axons of CA3 pyramidal cells in the hippocampal slice. J. Pharmacol. Exp. Ther. 223: 291–297, 1982.
 27. Ault, B., and J. V. Nadler. Anticonvulsant‐like actions of baclofen in the rat hippocampal slice. Br. J. Pharmacol. 78: 701–708, 1983.
 28. Baker, P. F. Calcium pumps and electrogenesis. In: Advances in Physiological Sciences. Physiology of Excitable Membranes, edited by J. Salanki. New York: Pergamon, 1981, vol. 4, p. 127–134.
 29. Bakhit, C., R. Benoit, and F. E. Bloom. Release of somatostatin‐28(1–12) from rat hypothalamus in vitro. Nature London 301: 524–526, 1983.
 30. Bannon, M. J., P. J. Elliott, J. E. Alpert, M. Goedert, S. D. Iversen, and L. L. Iversen. Role of endogenous substance P in stress‐induced activation of mesocortical dopamine neurones. Nature London 306: 791–792, 1983.
 31. Baraban, J. M., and G. K. Aghajanian. Suppression of firing activity of 5‐HT neurons in the dorsal raphe by alpha‐adrenoceptor antagonists. Neuropharmacology 19: 355–363, 1980.
 32. Baraban, J. M., S. H. Snyder, and B. E. Alger. Protein kinase C regulates ionic conductance in hippocampal pyramidal neurons: electrophysiological effects of phorbol esters. Proc. Natl. Acad. Sci. USA 82: 2538–2542, 1985.
 33. Barber, R., and K. Saito. Light microscopic visualization of GAD and GABA‐T in immunocytochemical preparation of rodent CNS. In: GABA in Nervous System Function, edited by E. Roberts, T. N. Chase, and D. B. Tower. New York: Raven, 1976, p. 113–132
 34. Barber, R. P., J. E. Vaughn, K. Saito, B. J. McLaughlin, and E. Roberts. GABAergic terminals are presynaptic to primary afferent terminals in the substantia gelatinosa of the rat spinal cord. Brain Res. 141: 35–55, 1978.
 35. Barker, J. L. Peptides: roles in neuronal excitability. Physiol. Rev. 56: 435–452, 1976.
 36. Barker, J. L., B. Dufy, D. G. Owen, and M. Segal. Excitable membrane properties of cultured central nervous system neurons and clonal pituitary cells. Cold Spring Harbor Symp. Quant. Biol. 48: 259–268, 1983.
 37. Barker, J. L., E. Gratz, D. G. Owen, and R. E. Study. Pharmacological effects of clinically important drugs on the excitability of cultured mouse spinal neurons. In: Actions and Interactions of GABA and Benzodiazepines, edited by N. G. Bowery. New York: Raven, 1984, p. 203–216.
 38. Barker, J. L. D. L., Gruol, L.‐Y., M. Huang, J. F. MacDonald, and T. G. Smith, Jr. Peptide receptor functions on cultured spinal neurons. In: Regulation and Function of Neural Peptides, edited by M. Trabucchi and E. Costa. New York: Raven, 1980, p. 409–423.
 39. Barker, J. L., and D. A. Mathers. GABA analogues activate channels of different duration on cultured mouse spinal neurons. Science 212: 358–361, 1981.
 40. Barker, J. L., R. N. McBurney, and J. F. MacDonald. Fluctuation analysis of neutral amino acid responses in cultured mouse spinal neurones. J. Physiol. London 322: 365–387, 1982.
 41. Barker, J. L., J. H. Neale, T. G. Smith, Jr., and R. L. Macdonald. Opiate peptide modulation of amino acid responses suggests novel form of neuronal communication. Science 199: 1451–1453, 1978.
 42. Barker, J. L., and B. R. Ransom. Amino acid pharmacology of mammalian central neurones grown in tissue culture. J. Physiol. London 280: 331–354, 1978.
 43. Barker, J. L., T. G. Smith, Jr., and J. H. Neale. Multiple membrane actions of enkephalin revealed using cultured spinal neurons. Brain Res. 154: 153–158, 1978.
 44. Barker, J. L., J.‐D. Vincent, and J. F. MacDonald. Substance P pharmacology of cultured mouse spinal neurons. In: Neuropeptides and Neural Transmission, edited by C. Ajmone‐Marsan and W. Z. Traczyk. New York: Raven, 1980, p. 93–103.
 45. Basile, A. S., and T. V. Dunwiddie. Alpha and beta receptor mediated effects on Purkinje cell spontaneous activity in the in vitro rat cerebellar slice. Soc. Neurosci. Abstr. 9: 877 1983.
 46. Basile, A. S., and T. V. Dunwiddie. Norepinephrine elicits both excitatory and inhibitory responses from Purkinje cells in the in vitro rat cerebellar slice. Brain Res. 296: 15–25, 1984.
 47. Baylor, D. A., and M. G. F. Fuortes. Electrical responses of single cones in the retina of the turtle. J. Physiol. London 207: 77–92, 1970.
 48. Bayon, A., L. Koda, E. Battenberg, and F. E. Bloom. Redistribution of endorphin and enkephalin immunoreactivity in the rat brain and pituitary after in vivo treatment with colchicine or cytochalasin B. Brain Res. 183: 103–111, 1980.
 49. Bayon, A., W. J. Shoemaker, J. F. McGinty, and F. Bloom. Immunodetection of endorphins and enkephalins: a search for reliability. Int. Rev. Neurobiol. 24: 51–92, 1983.
 50. Belcher, G., and R. W. Ryall. Substance P and Renshaw cells: a new concept of inhibitory synaptic interactions. J. Physiol. London 272: 105–119, 1977.
 51. Benardo, L. S., L. M. Masukawa, and D. A. Prince. Electrophysiology of isolated hippocampal pyramidal dendrites. J. Neurosci. 2: 1614–1622, 1982.
 52. Benardo, L. S., and D. A. Prince. Acetylcholine induced modulation of hippocampal pyramidal neurons. Brain Res. 211: 227–234, 1981.
 53. Benardo, L. S., and D. A. Prince. Cholinergic excitation of mammalian hippocampal pyramidal cells. Brain Res. 249: 315–331, 1982.
 54. Benardo, L. S., and D. A. Prince. Cholinergic pharmacology of mammalian hippocampal pyramidal cells. Neuroscience 7: 1703–1712, 1982.
 55. Benardo, L. S., and D. A. Prince. Dopamine action on hippocampal pyramidal cells. J. Neurosci. 2: 415–423, 1982.
 56. Benardo, L. S., and D. A. Prince. Dopamine modulates a Ca2+‐activated potassium conductance in mammalian hippocampal pyramidal cells. Nature London 297: 76–79, 1982.
 57. Benardo, L. S., and D. A. Prince. Ionic mechanisms of cholinergic excitation in mammalian hippocampal pyramidal cells. Brain Res. 249: 333–344, 1982.
 58. Ben‐Ari, Y., E. Cherubini, and C. Rovira. Effects of muscarinic and nicotinic agents upon dendritic and somatic field potentials recorded simultaneously in the rat ammon's horn in vivo (Abstract). J. Physiol. London 325: 27P–28P, 1981.
 59. Ben‐Ari, Y., R. Dingledine, I. Kanazawa, and J. S. Kelly. Inhibitory effects of acetylcholine on neurones in the feline nucleus reticularis thalami. J. Physiol. London 261: 647–671, 1976.
 60. Ben‐Ari, Y., and, K. Krnjević. Actions of GABA on hippocampal neurons with special reference to the etiology of epilepsy. In: Neurotransmitters, Seizures, and Epilepsy, edited by P. L. Morselli, K. G. Lloyd, W. Loscher, B. Meldram, and E. H. Reynolds. New York: Raven, 1981, p. 6–73.
 61. Ben‐Ari, Y., K. Krnjević, R. J. Reiffenstein, and W. Reinhardt. Inhibitory conductance changes and action of γ‐aminobutyrate in rat hippocampus. Neuroscience 6: 2445–2463, 1981.
 62. Ben‐Ari, Y., K. Krnjević, W. Reinhardt, and N. Ropert. Intracellular observations on the disinhibitory action of acetylcholine in the hippocampus. Neuroscience 6: 2475–2484, 1981.
 63. Bennet, J. P., Jr., and S. H. Snyder. Angiotensin II binding to mammalian brain membrane. J. Biol. Chem. 251: 7423–7430, 1976.
 64. Bennett, G. W., J. A. Edwardson, D. Marcano de Cotte, M. Berelowitz, B. L. Pimstone, and S. Kronheim. Release of somatostatin from rat brain synaptosomes. J. Neurochem. 32: 1127–1130, 1979.
 65. Bennett, M. V. L. Electrical transmission: a functional analysis and comparison to chemical transmission. In: Handbook of Physiology. The Nervous System, edited by J. M. Brookhart and V. B. Mountcastle. Bethesda, MD: Am. Physiol. Soc., 1977, sect. 1, vol. I, pt. 1, chapt. 11, p. 357–416.
 66. Benoit, R., P. Bohlen, N. Ling, F. Esch, A. Baird, S. Y. Ying, W. B. Wehrenberg, R. Guillemin, J. H. Morrison, C. Bakhit, L. Koda, and F. E. Bloom. Somatostatin‐28(1–12)‐like peptides. In: Proc. Int. Symp. Somatostatin, 3rd, edited by Y. C. Patel. New York: Academic, in press.
 67. Benoit, R., N. Ling, B. Alford, and R. Guillemin. Seven peptides derived from pro‐somatostatin in rat brain. Biochem. Biophys. Res. Commun. 107: 944–950, 1982.
 68. Bergman, H., S. Glusman, R. M. Harris‐Warrick, E. A. Kravitz, I. Nussinovitch, and R. Rahamimoff. Noradrenaline augments tetanic potentiation of transmitter release by a calcium dependent process. Brain Res. 214: 200–204, 1981.
 69. Bernardi, G., M. G. Marciani, C. Morocutti, F. Pavone, and P. Stantione. The action of dopamine on rat caudate nucleus neurons intracellularly recorded. Neurosci. Lett. 8: 235–240, 1978.
 70. Bevan, P., C. M. Bradshaw, R. Y. K. Pun, N. T. Slater, and E. Szabadi. Responses of single cortical neurones to noradrenaline and dopamine. Neuropharmacology 17: 611–617, 1978.
 71. Bird, S. J., and J. K. Aghajanian. The cholinergic pharmacology of hippocampal pyramidal cells: a microiontophoretic study. Neuropharmacology 15: 273–282, 1976.
 72. Biscoe, T. J., and D. W. Straughan. Micro‐electrophoretic studies of neurones in the cat hippocampus. J. Physiol. London 183: 341–359, 1966.
 73. Bisti, S., G. Isoif, and P. Strata. Suppression of inhibition in the cerebellar cortex by picrotoxin and bicuculline. Brain Res. 28: 591–593, 1971.
 74. Black, A. C., Jr., J. Y. Jew, J. R. West, J. K. Wamsley, D. Sandquist, and T. H. Williams. Catecholamines and cyclic nucleotides in the modulation of superior cervical ganglia. Neurosci. Biobehav. Rev. 3: 125–135, 1978.
 75. Blackstad, T. W. Ultrastructural studies on the hippocampal region. In: Progress in Brain Research. The Rhinencephalon and Related Structures, edited by W. Bargmann and J. P. Schadé. Amsterdam: Elsevier, 1963, vol. 3, p. 122–148.
 76. Blackstad, T. W., K. Fuxe, and, T. Hökfelt. Noradrenaline nerve terminals in the hippocampal region of the rat and the guinea pig. Z. Zellforsch. Mikrosk. Anat. 78: 463–473, 1967.
 77. Bland, B. H., G. K. Kostopoulos, and J. W. Phillis. Acetylcholine sensitivity of hippocampal formation neurons. Can. J. Physiol. Pharmacol. 52: 966–971, 1974.
 78. Bloom, F. E. Amino acids and polypeptides in neuronal function. Neurosci. Res. Program Bull. 10: 122–251, 1972.
 79. Bloom, F. E. To spritz or not to spritz: the doubtful value of aimless iontophoresis. Life Sci. 14: 1819–1834, 1974.
 80. Bloom, F. E. Amino receptors in CNS. I. Norepinephrine. In: Handbook of Psychopharmacology. Biogenic Amine Receptors, edited by L. L. Iversen, S. D. Iversen, and S. H. Snyder. New York: Plenum, 1975, vol. 6, p. 1–22.
 81. Bloom, F. E. The role of cyclic nucleotides in central synaptic function. Rev. Physiol. Biochem. Pharmacol. 74: 1–103, 1975.
 82. Bloom, F. E. Central noradrenergic systems: physiology and pharmacology. In: Psychopharmacology: A Generation of Progress, edited by M. A. Lipton, A. DiMascio, and K. F. Killam. New York: Raven, 1978, p. 131–141.
 83. Bloom, F. E. Central noradrenergic neurons: structure‐function considerations. In: Catecholamines: Basic and Clinical Frontiers, edited by E. Usdin. New York: Pergamon, 1979, p. 609–618.
 84. Bloom, F. E. Chemical integrative processes in the central nervous system. In: Neurosciences: Fourth Study Program, edited by F. O. Schmitt and F. G. Worden. Cambridge, MA: MIT Press, 1979, p. 51–58.
 85. Bloom, F. E. Contrasting principles of synaptic physiology: peptidergic and non‐peptidergic neurons. In: The Peptidergic Neuron, edited by K. Fuxe, T. Hökfelt, and R. Luft. New York: Plenum, 1979, vol. 2, p. 173–187.
 86. Bloom, F. E. Is there a neurotransmitter code in the brain? In: Advances in Pharmacology and Therapeutics. Neurotransmitters, edited by P. Simon. New York: Pergamon, 1979, vol. 2, p. 205–216.
 87. Bloom, F. E. The functional significance of neurotransmitter diversity. Am. J. Physiol. 246 (Cell Physiol. 15): C184–C194, 1984.
 88. Bloom, F. E., J. Aldenhoff, C. Bakhit, R. Benoit, E. French, D. Gruol, L. Koda, J. Morrison, and G. Siggins. Gut peptides in the brain: implications for therapeutics. In: Proc. Second World Conf. Clin. Pharmacol. Ther., edited by L. Lemberger and M. M. Reidenburg. Bethesda, MD: Am. Soc. Pharmacol. Exp. Ther., 1984, p. 754–768.
 89. Bloom, F. E., E. L. F. Battenberg, J. Rivier, and W. Vale. Corticotropin releasing factor (CRF): immunoreactive neurones and fibers in rat hypothalamus. Regul. Pept. 4: 43–48, 1982.
 90. Bloom, F. E., and B. J. Hoffer. Norepinephrine as a central synaptic transmitter. In: Frontiers in Catecholamine Research, edited by E. Usdin and S. H. Snyder. New York: Pergamon, 1974, p. 637–642. (Proc. Int. Catechol. Symp., 3rd, Strasbourg, 1973.)
 91. Bloom, F. E., B. J. Hoffer, and G. R. Siggins. Studies on norepinephrine‐containing afferents to Purkinje cells of rat cerebellum. I. Localization of the fibers and their synapses. Brain Res. 25: 501–521, 1971.
 92. Bloom, F. E., B. J. Hoffer, G. R. Siggins, J. L. Barker, and R. A. Nicoll. Effects of serotonin on central neurons: microiontophoretic administration. Federation Proc. 31: 97–106, 1972.
 93. Bloom, F. E., H. Krebs, J. Nicholson, and V. Pickel. The noradrenergic innervation of cerebellar Purkinje cells: localization, function, synaptogenesis, and axonal sprouting of locus coeruleus. In: Dynamics of Degeneration and Growth in Neurons, edited by K. Fuxe, L. Olson, and Y. Zotterman. New York: Pergamon, 1974, p. 413–423.
 94. Bloom, F. E., G. R. Siggins, and B. J. Hoffer. Interpreting the failures to confirm the depression of cerebellar Purkinje cells by cyclic AMP. Science 185: 627–629, 1974.
 95. Bloom, F. E., G. R. Siggins, B. J. Hoffer, M. Segal, and A. P. Oliver. Cyclic nucleotides in the central synaptic actions of catecholamines. In: Advances in Cyclic Nucleotide Research. Proceedings, edited by G. Drummond, P. Greengard, and G. Robison. New York: Raven, 1975, vol. 5, p. 603–618.
 96. Bowery, N. G. Baclofen: 10 years on. Trends Pharmacol. Sci. 3: 400–403, 1982.
 97. Bowery, N. G., D. R. Hill, A. L. Hudson, A. Doble, D. N. Middlemiss, J. Shaw, and M. Turnbull. (—)Baclofen decreases neurotransmitter release in the mammalian CNS by an action at a novel GABA receptor. Nature London 283: 92–94, 1980.
 98. Bradley, P. B., and A. Dray. Morphine and neurotransmitter substances: microiontophoretic study in the rat brain stem. Br. J. Pharmacol. 50: 47–55, 1974.
 99. Bradshaw, C. M., R. Y. K. Pun, N. T. Slater, and E. Szabadi. Comparison of the effects of methoxamine with those of noradrenaline and phenylephrine on single cerebral cortical neurones. Br. J. Pharmacol. 73: 47–54, 1981.
 100. Brazeau, P., W. Vale, R. Burgus, N. Ling, M. Butcher, J. Rivier, and R. Guillemin. Hypothalamic polypeptide that inhibits the secretion of immunoreactive pituitary growth hormone. Science 179: 77–79, 1973.
 101. Brock, L. G., J. S. Coomb, and J. C. Eccles. The recording of potentials from neurones with an intracellular electrode. J. Physiol. London 117: 431–460, 1952.
 102. Brodal, A. Neurological Anatomy in Relation to Clinical Medicine. New York: Oxford Univ. Press, 1981.
 103. Brookes, N. Actions of glutamate on dissociated mammalian spinal neurons in vitro. Dev. Neurosci. J. 1: 203–215, 1978.
 104. Brown, A. G., R. E. W. Fyffe, and D. J. Maxwell. Ultrastructure of physiologically identified hair follicle afferent fibres in cat spinal cord (Abstract). J. Physiol. London 318: 18P–19P, 1981.
 105. Brown, D. A., and P. R. Adams. Muscarinic suppression of a novel voltage‐sensitive K+ current in a vertebrate neurone. Nature London 283: 673–676, 1980.
 106. Brown, D. A., and M. P. Caulfield. Hyperpolarizing α2‐adrenoceptors in rat sympathetic ganglia. Br. J. Pharmacol. 65: 435–445, 1979.
 107. Brown, D. A., A. Constanti, and S. Marsh. Angiotensin mimics the action of muscarinic agonists on rat sympathetic neurones. Brain Res. 193: 614–619, 1980.
 108. Brown, D. A., and C. N. Scholfield. Depolarisation of neurones in the isolated olfactory cortex of the guinea pig by γ‐aminobutyric acid. Br. J. Pharmacol. 65: 339–345, 1979.
 109. Brown, M. R., L. A. Fisher, J. Spiess, C. Rivier, J. Rivier, and W. Vale. Corticotropin‐releasing factor: actions on the sympathetic nervous system and metabolism. Endocrinology 111: 928–931, 1982.
 110. Brown, T. H., and D. Johnston. Voltage‐clamp analysis of mossy fiber synaptic input to hippocampal neurons. J. Neurophysiol. 50: 487–507, 1983.
 111. Brownsein, M., A. Arimura, H. Sato, A. V. Schally, and J. S. Kizer. The regional distribution of somatostatin in the rat brain. Endocrinology 96: 1456–1461, 1975.
 112. Bruggencate, G. T., and I. Engberg. Iontophoretic studies in deiters nucleus of the inhibitory actions of GABA and related amino acids and the interactions of strychnine and picrotoxin. Brain Res. 25: 431–448, 1971.
 113. Buijs, R. M., and D. F. Swaab. Immuno‐electron microscopical demonstration of vasopressin and oxytocin synapses in the limbic system of the rat. Cell Tissue Res. 204: 353–365, 1979.
 114. Buijs, R. M., and J. J. Van Heerikhuize. Vasopressin and oxytocin release in the brain—a synaptic event. Brain Res. 252: 71–76, 1982.
 115. Bunney, B. S., A. A. Grace, D. W. Hommer, and L. R. Skirboll. Effect of cholecystokinin on the activity of midbrain dopaminergic neurons. In: Regulatory Peptides: From Molecular Biology to Function, edited by E. Costa and M. Trabucchi. New York: Raven, 1982, p. 429–436.
 116. Burke, R. E., and P. Rudomin. Spinal neurons and synapses. In: Handbook of Physiology. The Nervous System, edited by J. M. Brookhart and V. B. Mountcastle. Bethesda, MD: Am. Physiol. Soc., 1977, sect. 1, vol. I, pt. 2, chapt. 24, p. 877–944.
 117. Burnstock, G. Purinergic nerves. Pharmacol. Rev. 24: 509–581, 1972.
 118. Burnstock, G. Do some nerve cells release more than one transmitter? Neuroscience 1: 239–248, 1976.
 119. Burnstock, G. Purine nucleotides and nucleosides as neurotransmitters or neuromodulators in the central nervous system. In: Neuroregulatory and Psychiatric Disorders, edited by E. Usdin, D. A. Hamburg, and J. D. Barchas. New York: Oxford Univ. Press, 1977, p. 470–477.
 120. Burnstock, G. A basis for distinguishing two types of purinergic receptors. In: Cell Membrane Receptors for Drugs and Hormones: A Multidisciplinary Approach, edited by R. W. Straub and L. Bolis. New York: Raven, 1978, p. 107–118.
 121. Burnstock, G., M. E. Holman, and C. L. Prosser. Electrophysiology of smooth muscle. Physiol. Rev. 43: 482–527, 1963.
 122. Bylund, D. B., and D. C. U'Prichard. Characterization of α1‐ and α2‐adrenergic receptors. Int. Rev. Neurobiol. 24: 343 1983.
 123. Caesar, R., G. Edwards, and H. Ruska. Architecture and nerve supply of mammalian smooth muscle tissue. J. Biophys. Biochem. Cytol. 3: 867–877, 1957.
 124. Carpenter, F. G., and J. C. Tankersley. Response of a parasympathetic neuroeffector system to motor nerve stimulation. Am. J. Physiol. 196: 1185–1188, 1959.
 125. Carpenter, M. B. Human Neuroanatomy. Baltimore, MD: Williams & Wilkins, 1976.
 126. Carstens, E., I. Tulloch, W. Zieglgänsberger, and M. Zimmermann. Presynaptic excitability changes induced by morphine in single cutaneous afferent C‐ and A‐fibers. Pfluegers Arch. Gesamte Physiol. Menschen Tiere 379: 143–147, 1979.
 127. Cedarbaum, J. M., and G. K. Aghajanian. Activation of locus coeruleus neurons by peripheral stimuli: modulation by a collateral inhibitory mechanism. Life Sci. 23: 1383–1392, 1978.
 128. Changaris, D. G., W. B. Severs, and L. C. Keil. Localization of angiotensin in rat brain. J. Histochem. Cytochem. 26: 593–607, 1978.
 129. Charpak, S., W. E. Armstrong, M. Mühlethaler, and J. J. Dreifuss. Stimulatory action of oxytocin on neurones of the dorsal motor nucleus of the vagus nerve. Brain Res. 300: 83–89, 1984.
 130. Chasin, M., F. Mamrak, S. G. Samaniego. Preparation and properties of a cell‐free, hormonally responsive adenylate cyclase from guinea pig brain. J. Neurochem. 22: 1031–1038, 1974.
 131. Chavkin, C., C. Bakhit, E. Weber, and F. E. Bloom. Relative contents and concomitant release of prodynorphin/neoendorphin‐derived peptide in rat hippocampus. Proc. Natl. Acad. Sci. USA 80: 7669–7673, 1983.
 132. Chavkin, C., S. J. Henriksen, G. R. Siggins, and F. E. Bloom. Selective inactivation of opioid receptors in rat hippocampus demonstrates that dynorphin A and B act on μ‐receptors in CA1. Brain Res. 331: 366–370, 1985.
 133. Chiodo, L. A., and B. S. Bunney. Proglumide: selective antagonism of excitatory effects of cholecystokinin in central nervous system. Science 219: 1449–1451, 1983.
 134. Choi, D. W., D. H. Farb, and G. D. Fischbach. Chlordiazepoxide selectively potentiates GABA conductance of spinal cord and sensory neurons in cell culture. J. Neurophysiol. 45: 621–631, 1981.
 135. Choi, D. W., D. H. Farb, and G. D. Fischbach. GABA‐mediated synaptic potentials in chick spinal cord and sensory neurons. J. Neurophysiol. 45: 632–643, 1981.
 136. Choi, D. W., and G. D. Fischbach. GABA conductance of chick spinal cord and dorsal root ganglion neurons in cell culture. J. Neurophysiol. 45: 605–620, 1981.
 137. Clapham, D. E., and E. Neher. Substance P reduces acetylcholine‐induced currents in isolated bovine chromaffin cells. J. Physiol. London 347: 255–277, 1984.
 138. Cole, A. E., and R. A. Nicoll. Acetylcholine mediates a slow synaptic potential in hippocampal pyramidal cells. Science 221: 1299–1301, 1983.
 139. Collingridge, G. L., S. J. Kehl, and H. McLennan. The antagonism of amino acid‐induced excitations of rat hippocampal CA1 neurones in vitro. J. Physiol. London 334: 19–31, 1983.
 140. Collingridge, G. L., S. J. Kehl, and H. McLennan. Excitatory amino acids in synaptic transmission in the Schaffer collateral‐commissural pathway of the rat hippocampus. J. Physiol. London 334: 33–46, 1983.
 141. Constanti, A., and M. Galvan. M‐current in voltage‐clamped olfactory cortex neurones. Neurosci. Lett. 39: 65–70, 1983.
 142. Cooper, P. E., M. H. Fernstrom, O. P. Rorstad, S. E. Leeman, and J. B. Martin. The regional distribution of somatostatin, substance P and neurotensin in human brain. Brain Res. 218: 219–232, 1981.
 143. Corey, D. P. Patch clamp: current excitement in membrane physiology. Neurosci. Commentaries 1: 99–110, 1983.
 144. Corrigall, W. A. Opiates and the hippocampus: a review of the functional and morphological evidence. Pharmacol. Biochem. Behav. 18: 255–262, 1983.
 145. Cotman, C. W. Acidic amino acids as excitatory transmitters. In: Regulatory Mechanisms of Synaptic Transmission, edited by R. Tapia and C. W. Cotman. New York: Plenum, 1981, p. 43–57.
 146. Cotman, C. W., A. Foster, and T. Lanthorn. An overview of glutamate as a neurotransmitter. In: Glutamate as a Neurotransmitter, edited by G. Di Chiara and G. L. Gessa. New York: Raven, 1981, p. 1–27. (Adv. Biochem. Psychopharmacol. Ser., vol. 27)
 147. Cotman, C. W., and J. V. Nadler. Glutamate and aspartate as hippocampal transmitters: biochemical and pharmacological evidence. In: Glutamate: Transmitter in the Central Nervous System, edited by P. J. Roberts, J. Storm‐Mathisen, and G. A. R. Johnston. New York: Wiley, 1981, p. 117–154.
 148. Cox, B. M., A. Goldstein, and C. H. Li. Opioid activity of a peptide, β‐lipotropin(61–91), derived from β‐lipotropin. Proc. Natl. Acad. Sci. USA 73: 1821–1823, 1976.
 149. Crain, S. M., B. Crain, E. R. Peterson, and E. J. Simon. Selective depression by opioid peptides of sensory‐evoked dorsal horn network responses in organized spinal cord cultures. Brain Res. 157: 196–201, 1978.
 150. Crain, S. M., E. R. Peterson, B. Crain, and E. J. Simon. Selective opiate depression of sensory‐evoked synaptic networks in dorsal horn regions of spinal cord cultures. Brain Res. 133: 162–166, 1977.
 151. Crawford, J. M., D. R. Curtis, P. E. Voorhoeve, and V. J. Wilson. Excitation of cerebellar neurones by acetylcholine. Nature London 200: 579–580, 1963.
 152. Crepel, F., and, N. Delhaye‐Bouchaud. Intracellular analyses of synaptic potentials in cerebellar Purkinje cells of the rat. Brain Res. 155: 176–181, 1978.
 153. Crepel, F., and S. S. Dhanjal. Cholinergic mechanisms and neurotransmission in the cerebellum of the rat. An in vitro study. Brain Res. 244: 59–68, 1982.
 154. Crepel, F., S. S. Dhanjal, and T. A. Sears. Effect of glutamate, aspartate and related derivatives on cerebellar Purkinje cell dendrites in the rat: an in vitro study. J. Physiol. London 329: 297–317, 1982.
 155. Crepel, F., J. L. Dupont, and R. Gardette. Voltage clamp analysis of the effect of excitatory amino acids and derivates on Purkinje cell dendrites in rat cerebellar slices maintained in vitro. Brain Res. 279: 311–315, 1983.
 156. Crill, W. E., and P. C. Schwindt. Active currents in mammalian central neurons. Trends Neurosci. 6: 236–240, 1983.
 157. Crunelli, V., S. Forda, G. L. Collingridge, and J. S. Kelly. Intracellular recorded synaptic antagonism in the rat dentate gyrus. Nature London 300: 450–452, 1982.
 158. Crutcher, K. A., and J. N. Davis. Hippocampal α‐ and β‐adrenergic receptors: comparison of [3H]dihydroalprenolol and [3H]WB 4101 binding with noradrenergic innervation in the rat. Brain Res. 182: 107–117, 1980.
 159. Cuello, A. C., and I. Kanazawa. The distribution of substance P immunoreactive fibers in the rat central nervous system. J. Comp. Neurol. 178: 129–156, 1978.
 160. Cuello, A. C., and M. V. Sofroniew. The anatomy of the CNS cholinergic neurons. Trends Neurosci. 7: 74–78, 1984.
 161. Cullheim, S., and J.‐O. Kellerth. Two kinds of recurrent inhibition of cat spinal α‐motoneurones as differentiated pharmacologically. J. Physiol. London 312: 209–224, 1981.
 162. Curtis, D. R. Microelectrophoresis. In: Physical Techniques in Biological Research: Electrophysiological Methods, edited by W. L. Nastuk. New York: Academic, 1964, vol. 5, pt. A, p. 144–190.
 163. Curtis, D. R. Bicuculline, GABA and central inhibition. Proc. Aust. Assoc. Neurol. 9: 145–153, 1973.
 164. Curtis, D. R. Gabaergic transmission in the mammalian central nervous system. In: GABA‐Neurotransmitters, edited by P. Krogsgaard‐Larsen, J. Scheel‐Krüger, and H. Kofod. New York: Academic, 1978, p. 17–27.
 165. Curtis, D. R. Problems in the evaluation of glutamate as a central nervous system transmitter. In: Glutamic Acid: Advances in Biochemistry and Physiology, edited by L. J. Filer, Jr., S. Garattini, M. R. Kare, W. A. Reynolds, and R. J. Wurtman. New York: Raven, 1979, p. 163–175.
 166. Curtis, D. R., and R. M. Eccles. The excitation of Renshaw cells by pharmacological agents applied electrophoretically. J. Physiol. London 141: 435–445, 1958.
 167. Curtis, D. R., L. Hösli, G. A. R. Johnston, and I. H. Johnson. The hyperpolarization of spinal motoneurones by glycine and related amino acids. Exp. Brain Res. 5: 235–258, 1968.
 168. Curtis, D. R., and G. A. R. Johnston. Amino acid transmitters in the mammalian central nervous system. Ergeb. Physiol. Biol. Chem. Exp. Pharmakol. 69: 97–188, 1974.
 169. Curtis, D. R., D. Lodge, J. C. Bornstein, and M. J. Peet. Selective effects of (—)‐baclofen on spinal synaptic transmission in the cat. Exp. Brain Res. 42: 158–170, 1981.
 170. Curtis, D. R., J. W. Phillis, and J. C. Watkins. Cholinergic and non‐cholinergic transmission in the mammalian spinal cord. J. Physiol. London 158: 296–323, 1961.
 171. Dahl, D., W. H. Bailey, and J. Winson. Effect of norepinephrine depletion of hippocampus on neuronal transmission from perforant pathway through dentate gyrus. J. Neurophysiol. 49: 123–133, 1983.
 172. Dale, H. H. Chemical transmission of the effects of nerve impulses. Br. Med. J. 1: 835–841, 1934.
 173. Dale, H. H. Acetylcholine as a chemical transmitter of the effects of nerve impulses. J. Mt. Sinai Hosp. NY 4: 401–429, 1938.
 174. Dale, H. H. The action of certain esters and ethers of choline and their relation to muscarine. J. Pharmacol. Exp. Ther. 6: 147–190, 1941.
 175. Daly, J. W. The nature of receptors regulating the formation of cyclic AMP in brain tissue. Life Sci. 18: 1349–1358, 1976.
 176. Daly, J. W. Cyclic Nucleotides in the Nervous System. New York: Plenum, 1977.
 177. Daly, J. W., R. F. Bruns, and S. H. Snyder. Adenosine receptors in the central nervous system: relationship to the central actions of methylxanthines. Life Sci. 28: 2083–2097, 1981.
 178. Daly, J. W., W. Padgett, C. R. Creveling, D. Cantacuzene, and K. L. Kirk. Cyclic AMP‐generating systems: regional differences in activation by adrenergic receptors in rat brain. J. Neurosci. 1: 49–59, 1981.
 179. Daly, J. W., W. Padgett, Y. Nimitkitpaisan, C. R. Creveling, D. Cantacuzene, and K. L. Kirk. Fluoronorepinephrines: specific agonists for the activation of alpha and beta adrenergic‐sensitive cyclic AMP‐generating systems in brain slices. J. Pharmacol. Exp. Ther. 212: 382–389, 1980.
 180. Davidoff, R. A. Amino acids and presynaptic inhibition. In: Amino Acid Neurotransmitters, edited by F. V. DeFeudis and P. Mandel. New York: Raven, 1981, p. 249–255.
 181. Davies, J. Effects of morphine and naloxone on Renshaw cells and spinal interneurones in morphine dependent and non‐dependent rats. Brain Res. 113: 311–326, 1976.
 182. Davies, J., and A. Dray. Substance P in the substantia nigra. Brain Res. 107: 623–627, 1976.
 183. Davies, J., and J. C. Watkins. Pharmacology of glutamate and aspartate antagonists on cat spinal neurons. In: Glutamate as a Neurotransmitter, edited by G. Di Chiara and G. L. Gessa. New York: Raven, 1981, p. 275–284. (Adv. Biochem. Psychopharmacol. Ser., vol. 27.)
 184. Davis, J. N., C. D. Arnett, E. Hoyler, L. P. Stalvey, J. W. Daly, and P. Skolnick. Brain α‐adrenergic receptors: comparison of [3H]WB 4101 binding with norepinephrine‐stimulated cyclic AMP accumulation in rat cerebral cortex. Brain Res. 159: 125–135, 1978.
 185. DeFeudis, F. V. Amino acids as central neurotransmitters. Annu. Rev. Pharmacol. 15: 105–130, 1975.
 186. DeFeudis, F. V. Binding and iontophoretic studies on centrally active amino acids—a search for physiological receptors. Int. Rev. Neurobiol. 21: 129–216, 1979.
 187. Delfs, J. R., and M. A. Dichter. Effects of somatostatin on mammalian cortical neurons in culture: physiological actions and unusual dose‐response characteristics. J. Neurosci. 3: 1176–1188, 1983.
 188. De Mey, J., G. Burnstock, and P. M. Vanhoutte. Modulation of the evoked release of noradrenaline in canine saphenous vein via presynaptic receptors for adenosine but not ATP Eur. J. Pharmacol. 55: 401–405, 1979.
 189. DeMontigny, C., R. Y. Wang, T. A. Reader, and G. K. Aghajanian. Monoaminergic denervation of the rat hippocampus: microiontophoretic studies on pre‐ and postsynaptic supersensitivity to norepinephrine and serotonin. Brain Res. 200: 363–376, 1980.
 190. Deterre, P., D. Paupardin‐Tritsch, J. Bockaert, and H. M. Gerschenfeld. cAMP‐mediated decrease in K+ conductance evoked by serotonin and dopamine in the same neuron: a biochemical and physiological single‐cell study. Proc. Natl. Acad. Sci. USA 79: 7934–7938, 1982.
 191. De Wied, D. Behavioural actions of neurohypophysial peptides. Proc. R. Soc. London Ser. B 210: 183–195, 1980.
 192. Dichter, M. A. Physiological identification of GABA as the inhibitory transmitter for mammalian cortical neurones in cell culture. Brain Res. 190: 111–121, 1980.
 193. Dichter, M. A., and J. R. Delfs. Somatostatin and cortical neurons in cell culture. In: Neurosecretion and Brain Peptides: Implications for Brain Function and Neurological Disease, edited by J. B. Martin, S. Reichlin, and K. L. Bick. New York: Raven, 1981, p. 145–157.
 194. Dillier, N., J. Laszlo, B. Muller, W. P. Koella, and H.‐R. Olpe. Activation of an inhibitory noradrenergic pathway projecting from the locus coeruleus to the cingulate cortex of the rat. Brain Res. 154: 61–68, 1978.
 195. Dingledine, R. Enkephalin effect of calcium spikes, N‐methyl‐aspartate responses and synaptic inhibition in the rat hippocampal slice. Soc. Neurosci. Abstr. 7: 577 1981.
 196. Dingledine, R. Possible mechanisms of enkephalin action on hippocampal CA1 pyramidal neurons. J. Neurosci. 1: 1022–1035, 1981.
 197. Dingledine, R. N‐methyl aspartate activates voltage‐dependent calcium conductance in rat hippocampal pyramidal cells. J. Physiol. London 343: 385–405, 1983.
 198. Dingledine, R., J. Dodd, and J. S. Kelly. The in vitro brain slice as a useful neurophysiological preparation for intracellular recording. J. Neurosci. Methods 2: 323–362, 1980.
 199. Dingledine, R., and J. S. Kelly. Brain stem stimulation and the acetylcholine‐evoked inhibition of neurones in the feline nucleus reticularis thalami J. Physiol. London 271: 135–154, 1977.
 200. Dingledine, R., and I. A. Langmoen. Conductance changes and inhibitory actions of hippocampal recurrent IPSPs. Brain Res. 185: 277–287, 1980.
 201. Dionne, V. E. Noise analysis. Tech. Cell. Physiol. Part II: 1–19, 1982.
 202. Dismukes, K. New look at the aminergic nervous system. Nature London 269: 557–558, 1977.
 203. Djørup, A., H. Jahnsen, and A. M. Laursen. The dendritic response to GABA in CA1 of the hippocampal slice. Brain Res. 219: 196–201, 1981.
 204. Dockray, G. J. Immunochemical evidence of cholecystokinin‐like peptides in brain. Nature London 264: 568–570, 1976.
 205. Dockray, G. J. Molecular evolution of gut hormones: application of comparative studies on the regulation of digestion. Gastroenterology 72: 344–358, 1977.
 206. Dockray, G. J. Cholecystokinins in rat cerebral cortex: identification, purification and characterization by immunochemical methods. Brain Res. 188: 155–165, 1980.
 207. Dodd, J., R. Dingledine, and J. S. Kelly. The excitatory action of acetylcholine on hippocampal neurones of the guinea pig and rat maintained in vitro. Brain Res. 207: 109–127, 1981.
 208. Dodd, J., and J. S. Kelly. Is somatostatin an excitatory transmitter in the hippocampus? Nature London 273: 674–675, 1978.
 209. Dodd, J., and J. S. Kelly. The actions of cholecystokinin and related peptides on pyramidal neurones of the mammalian hippocampus. Brain Res. 205: 337–350, 1981.
 210. Dodd, J., J. S. Kelly, and S. I. Said. Excitation of CA1 neurones of the rat hippocampus by the octacosapeptide, vasoactive intestinal polypeptide (VIP). Br. J. Pharmacol. 66: 125P, 1979.
 211. Dolphin, A., M. Hamont, and J. Bockaert. The resolution of dopamine and β1‐ and β2‐adrenergic‐sensitive adenylate cyclase activites in homogenates of cat cerebellum, hippocampus and cerebral cortex. Brain Res. 179: 305–317, 1979.
 212. Dowdall, M. J., A. F. Boyne, and V. P. Whittaker. ATP: a constituent of cholinergic synaptic vesicles. Biochem. J. 140: 1–12, 1974.
 213. Dreifuss, J. J., M. C. Harris, and E. Tribollet. Excitation and phasically firing hypothalamic supraoptic neurones by carotid occlusion in rats. J. Physiol. London 257: 337–354, 1976.
 214. Dreifuss, J. J., J. S. Kelly, and, K. Krnjević. Cortical inhibition and γ‐aminobutyric acid. Exp. Brain Res. 9: 137–154, 1969.
 215. Dreifuss, J. J., and, M. Mühlethaler. Vasopressin‐induced modulation of firing rate in hippocampal slices (Abstract). J. Physiol. London 318: 18P, 1981.
 216. Dreifuss, J. J., and, M. Mühlethaler. Neurohypophysial hormones: primary effects on non‐pyramidal neurones in rat hippocampal slices (Abstract). J. Physiol. London 334: 87P, 1983.
 217. Drummond, A. H., J. A. Benson, and I. B. Levitan. Serotonin‐induced hyperpolarization of an identified Aplysia neuron is mediated by cyclic AMP. Proc. Natl. Acad. Sci. USA 77: 5013–5017, 1980.
 218. Dudar, J. D. The effect of septal nuclei stimulation on the release of acetylcholine from the rabbit hippocampus. Brain Res. 83: 123–133, 1975.
 219. Dun, N. J., and A. G. Karczmar. Actions of substance P on sympathetic neurons. Neuropharmacology 42: 215–218, 1979.
 220. Dunlap, K. Two types of γ‐aminobutyric acid receptor on embryonic sensory neurones. Br. J. Pharmacol. 74: 579–585, 1981.
 221. Dunlap, K. Opioid peptides decrease calcium‐dependent action potential duration of mouse dorsal root ganglion neurons in cell‐culture. Brain Res. 239: 315–321, 1982.
 222. Dunlap, K., and G. D. Fischbach. Neurotransmitters decrease the calcium component of sensory neurone action potentials. Nature London 276: 837–839, 1978.
 223. Dunlap, K., and G. D. Fischbach. Neurotransmitters decrease the calcium conductance activated by depolarization of embryonic chick sensory neurones. J. Physiol. London 317: 519–535, 1981.
 224. Dunwiddie, T. V., and B. J. Hoffer. Adenine nucleotides and synaptic transmission in the in vitro rat hippocampus. Br. J. Pharmacol. 69: 59–68, 1980.
 225. Dunwiddie, T. V., and B. J. Hoffer. The role of cyclic nucleotides in the nervous system. In: Handbook of Experimental Pharmacology. Cyclic Nucleotides, edited by J. W. Kebabian and J. A. Nathanson. Berlin: Springer‐Verlag, 1982, vol. 58, pt. 2, p. 389–463.
 226. Dunwiddie, T., A. Mueller, M. Palmer, J. Stewart, and B. Hoffer. Electrophysiological interactions of enkephalins with neuronal circuitry in the rat hippocampus. I. Effects on pyramidal cell activity. Brain Res. 184: 311–330, 1980.
 227. DuPont, J. L., F. Crepel, and, N. Delhaye‐Bouchaud. Influence of bicuculline and picrotoxin on reversal properties of excitatory synaptic potentials in cerebellar Purkinje cells of the rat. Brain Res. 173: 577–580, 1979.
 228. Dutar, P., Y. Lamour, and A. Jobert. Excitatory effects of two peptides on rat cortical neurons: laminar distribution and correlation with the effects of acetylcholine. C. R. Seances Acad. Sci. Ser. III 295: 247–250, 1982.
 229. DuVigneaud, V. Trials of sulfur research: from insulin to oxytocin. Science 123: 967–974, 1956.
 230. Eccles, J. C. The Physiology of Nerve Cells. Baltimore: Johns Hopkins Univ. Press, 1957.
 231. Eccles, J. C. The Physiology of Synapses. New York: Academic, 1964.
 232. Eccles, J. C., M. Ito, and J. Szentagothai. The Cerebellum as a Neuronal Machine, New York: Springer‐Verlag, 1967.
 233. Eccles, R. M., and B. Libet. Origin and blockade of the synaptic responses of curarized sympathetic ganglia. J. Physiol. London 157: 484–503, 1961.
 234. Egan, T. M., G. Henderson, R. A. North, and J. T. Williams. Noradrenaline‐mediated synaptic inhibition in rat locus coeruleus neurones. J. Physiol. London 345: 477–488, 1983.
 235. Ehlers, C. L., S. J. Henriksen, M. Wang, J. Rivier, W. Vale, and F. E. Bloom. Corticotropin releasing factor produces increases in brain excitability and convulsive seizures in rats. Brain Res. 278: 332–336, 1983.
 236. Elde, R., T. Hökfelt, O. Johannson, and L. Terenius. Immunohistochemical studies using antibodies to leu‐enkephalin: initial observations on the nervous system of the rat. Neuroscience 1: 349–351, 1976.
 237. Elliott, T. R. The action of adrenaline. J. Physiol. London 32: 401–467, 1905.
 238. Emson, P. C., J. Fahrenkrug, O. B. Schaffalitzky de Muckadell, T. M. Jessell, and L. L. Iversen. Vasoactive intestinal polypeptide (VIP): vesicular localization and potassium evoked release from rat hypothalamus. Brain Res. 143: 174–178, 1978.
 239. Engberg, G., T. H. Svensson, S. Rosell, and K. Folkers. A synthetic peptide as an antagonist of substance P. Nature London 293: 222–223, 1981.
 240. Engberg, I., J. A. Flatman, and K. Kadzielawa. Lack of specificity of motoneurones responses to microiontophoretically applied phenolic amines. Acta Physiol. Scand. 96: 137–139, 1976.
 241. Engberg, I., and K. C. Marshall. Mechanisms of noradrenaline hyperpolarization in spinal cord motoneurons of the cat. Acta Physiol. Scand. 83: 142–144, 1971.
 242. Engel, E., V. Barcilon, and R. S. Eisenberg. The interpretation of current‐voltage relations recorded from a spherical cell with a single microelectrode. Biophys. J. 12: 384–403, 1972.
 243. Enna, S. J., and J. P. Gallagher. Biochemical and electrophysiological characteristics of mammalian GABA receptors. Int. Rev. Neurobiol. 24: 181–214, 1983.
 244. Epstein, A. N., J. T. Fitzsimons, and B. J. Rolls. Drinking induced by injection of angiotensin II into the brain of the rat. J. Physiol. London 210: 457–474, 1970.
 245. Erspamer, V. The tachykinin peptide family. Trends Neurosci. 4: 267–269, 1981.
 246. Evans, R. H., and R. G. Hill. Effects of excitatory and inhibitory peptides on isolated spinal cord preparations. In: Iontophoresis and Transmitter Mechanisms in the Mammalian Central Nervous System, edited by R. W. Ryall and J. S. Kelly. Amsterdam: Elsevier/North‐Holland Biomed., 1978, p. 101–103.
 247. Evarts, E. V. Role of motor cortex in voluntary movements in primates. In: Handbook of Physiology. The Nervous System, edited by J. B. Brookhart and V. B. Mountcastle. Bethesda, MD: Am. Physiol. Soc., 1981, sect. 1, vol. II, pt. 2, chapt. 23, p. 1083–1120.
 248. Fagni, L., M. Baudry, and G. Lynch. Classification and properties of acidic amino acid receptors in hippocampus. I. Electrophysiological studies of an apparent desensitization and interactions with drugs which block transmission. J. Neurosci. 3: 1538–1546, 1983.
 249. Felix, D., and Akert, J. The effect of angiotensin II on neurones of the cat subfornical organ. Brain Res. 350–353, 1974.
 250. Ferron, A., G. R. Siggins, and F. E. Bloom. Vasoactive intestinal polypeptide (VIP) alters firing of cortical neurons and interacts with norepinephrine. Soc. Neurosci. Abstr. 10: 535 1984.
 251. Ffrench‐Mullen, J. M. H., R. Zacsek, K. J. Koller, J. T. Coyle, and D. O. Carpenter. Actions of N‐acetylspartylglutamate on mammalian neurons. Soc. Neurosci. Abstr., 9: 444 1983.
 252. Fischbach, G. D., and P. G. Nelson. Cell culture in neurobiology. In: Handbook of Physiology. The Nervous System, edited by J. M. Brookhart and V. B. Mountcastle. Bethesda, MD: Am. Physiol. Soc., 1977, sect. 1, vol. I, pt. 2, chapt. 20, p. 719–774.
 253. Fischman, A. J., and Moldow, R. L. Extrahypothalamic distribution of CRF‐like immunoreactivity in the rat brain. Peptides 1: 149–153, 1982.
 254. Flatman, J. A., I. Engberg, and J. D. C. Lambert. Reversibility of Ia EPSP investigated with intracellularly iontophoresed QX‐222. J. Neurophysiol. 48: 419–430, 1982.
 255. Flatman, J. A., J. D. C. Lambert, and I. Engberg. Some solutions to problems encountered on attempting to depolarize cat motoneurones sufficiently to reverse Ia e.p.s.p.s. Adv. Physiol. Sci. USSR 1: 83–91, 1981.
 256. Flatman, J. A., P. C. Schwindt, W. E. Crill, and C. E. Stafstrom. Multiple actions of N‐methyl‐D‐aspartate on cat neocortical neurons in vitro. Brain Res. 266: 169–173, 1983.
 257. Florey, E. Neurotransmitters and modulators in the animal kingdom. Federation Proc. 26: 1164–1178, 1967.
 258. Folkow, B. Impulse frequency in sympathetic vasomotor fibres correlated to the release and elimination of the transmitter. Acta Physiol. Scand. 25: 49–76, 1952.
 259. Fonnum, F. Localization of cholinergic and γ‐aminobutyric acid containing pathways in brain. In: Metabolic Compartmentation in the Brain, edited by R. Balázs and J. E. Cremer. New York: Wiley, 1972.
 260. Fonnum, F., and, D. Malthe‐Sorenssen. Localization of glutamate neurons. In: Glutamate: Transmitter in the Central Nervous System, edited by P. J. Roberts, J. Storm‐Mathisen, and G. A. R. Johnston. New York: Wiley, 1981, p. 205–222.
 261. Foote, S. L., F. E. Bloom, and, G. Aston‐Jones. Nucleus locus ceruleus: new evidence of anatomical and physiological specificity. Physiol. Rev. 63: 844–914, 1983.
 262. Foote, S. L., R. Freedman, and A. P. Oliver. Effects of putative neurotransmitters on neuronal activity in monkey auditory cortex. Brain Res. 86: 229–242, 1975.
 263. Fox, S., K. Krnjević, M. E. Morris, E. Puil, and R. Werman. Action of baclofen on mammalian synaptic transmission. Neuroscience 3: 495–515, 1978.
 264. Fox, S. E., and J. B. Ranck Jr.. Electrophysiological characteristics of hippocampal complex‐spike and theta cells. Exp. Brain Res. 41: 399–410, 1981.
 265. Frankhuyzen, A. L., and A. H. Mulder. Noradrenaline inhibits depolarization‐induced 3H‐serotonin release from slices of rat hippocampus. Eur. J. Pharmacol. 63: 179–182, 1980.
 266. Frankhuyzen, A. L., and A. H. Mulder. Pharmacological characterization of presynaptic α‐adrenoceptors modulating [3H]noradrenaline and [3H]5‐hydroxytryptamine release from slices of the hippocampus of the rat. Eur. J. Pharmacol. 81: 97–106, 1982.
 267. Frederickson, R. C. A. Enkephalin pentapeptides—a review of current evidence for a physiological role in vertebrate neurotransmission. Life Sci. 21: 23–40, 1977.
 268. Fredholm, B. B., and P. Hedgvist. Modulation of neurotransmission by purine nucleotides and nucleosides. Biochem. Pharmacol. 29: 1635–1643, 1980.
 269. Freedman, R., and B. J. Hoffer. Phenothiazine antagonism of the noradrenergic inhibition of cerebellar Purkinje neurons. J. Neurobiol. 6: 277–288, 1975.
 270. Freedman, R., B. J. Hoffer, D. Puro, and D. J. Woodward. Noradrenaline modulation of the responses of the cerebellar Purkinje cell to afferent synaptic input. Br. J. Pharmacol. 57: 603–605, 1976.
 271. Freedman, R., B. J. Hoffer, and D. J. Woodward. A quantitative microiontophoretic analysis of the responses of central neurons to noradrenaline: interactions with cobalt, manganese, verapamil and dichloroisoprenaline. Br. J. Pharmacol. 54: 529–539, 1975.
 272. Freedman, R., B. J. Hoffer, D. J. Woodward, and D. Puro. Interaction of norepinephrine with cerebellar activity evoked by mossy and climbing fibers. Exp. Neurol. 55: 269–288, 1977.
 273. French, E. D., and G. R. Siggins. An iontophoretic survey of opioid peptide actions in the rat limbic system: in search of opiate epileptogenic mechanisms. Regul. Pept. 1: 127–146, 1980.
 274. Freund‐Mercier, M. J., and P. Richard. Excitatory effects of intraventricular injections of oxytocin on the milk‐ejection reflex in the rat. Neurosci. Lett. 23: 193–198, 1981.
 275. Fujita, Y. Evidence for the existence of inhibitory postsynaptic potentials in dendrites and their functional significance in hippocampal pyramidal cells of adult rabbits. Brain Res. 175: 59–69, 1979.
 276. Furness, J. B., and M. Costa. Actions of somatostatin on excitatory and inhibitory nerves in the intestine. Eur. J. Pharmacol. 56: 69–74, 1979.
 277. Fuxe, K., D. Ganten, T. M. Hökfelt, and P. Bolme. Immunohistochemical evidence for the existence of angiotensin II‐containing nerve terminals in the brain and spinal cord in the rat. Neurosci. Lett. 2: 229–234, 1976.
 278. Fuxe, K., T. Hökfelt, S. I. Said, and V. Mutt. Vasoactive intestinal polypeptide and the nervous system: immunohistochemical evidence for localization in central and peripheral neurons, particularly intracortical neurons of the cerebral cortex. Neurosci. Lett. 5: 241–246, 1977.
 279. Gähwiler, B. H. Inhibitory action of noradrenaline and cyclic adenosine monophosphate in explants of rat cerebellum. Nature London 259: 483–484, 1976.
 280. Gähwiler, B. H. Excitatory action of opioid peptides and opiates on cultured hippocampal pyramidal cells. Brain Res. 194: 193–203, 1980.
 281. Gähwiler, B. H. Facilitation by acetylcholine of tetrodotoxin‐resistant spikes in rat hippocampal pyramidal cells. Neuroscience 11: 381–388, 1984.
 282. Gähwiler, B. H., and D. A. Brown. GABAB‐receptor‐activated K+ current in voltage‐clamped CA3 pyrimidal cells in hippocampal cultures. Proc. Natl. Acad. Sci. USA 82: 1558–1562, 1985.
 283. Gähwiler, B. H., and J. J. Dreifuss. Transition from random to phasic firing induced in neurons cultured from the hypothalamic supraoptic area. Brain Res. 193: 415–425, 1980.
 284. Gähwiler, B. H., and J. J. Dreifuss. Multiple actions of acetylcholine on hippocampal pyramidal cells in organotypic explant cultures. Neuroscience 7: 1243–1256, 1982.
 285. Gähwiler, B. H., and P. L. Herrling. Effects of opioid peptides on synaptic potentials in explants of rat hippocampus. Regul. Pept. 1: 317–326, 1981.
 286. Galindo, A., K. Krnjević, and S. Schwartz. Micro‐iontophoretic studies on neurones in the cuneate nucleus. J. Physiol. London 192: 359–377, 1967.
 287. Gall, C., N. Brecha, H. J. Karten, and K. Chang. Localization of enkephalin‐like immunoreactivity to identified axonal and neuronal populations of the rat hippocampus. J. Comp. Neurol. 198: 335–350, 1981.
 288. Gallagher, J. P., H. Higashi, and S. Nishi. Characterisation and ionic basis of GABA‐induced depolarisations recorded in vitro from cat primary afferent neurones. J. Physiol. London 275: 263–282, 1978.
 289. Gallagher, J. P., H. Inokuchi, and P. Shinnick‐Gallagher. Dopamine depolarisation of mammalian primary afferent neurones. Nature London 283: 770–772, 1980.
 290. Galvan, M., and P. R. Adams. Control of calcium current in rat sympathetic neurons by norepinephrine. Brain Res. 244: 135–144, 1982.
 291. Gerfen, C. R., and P. E. Sawchenko. An anterograde neuroanatomical tracing method that shows the detailed morphology of neurons, their axons and terminals: immunohistochemical localization of an axonally transported plant lectin, phaseolus vulgaris leucoagglutinin. Brain Res. 290: 219–238, 1984.
 292. Giachetti, A., S. I. Said, R. C. Reynolds, and F. C. Koniges. Vasoactive intestinal polypeptides in brain: localization in and release from isolated nerve terminals. Proc. Natl. Acad. Sci. USA 74: 3424–3428, 1977.
 293. Giguere, V., F. Labrie, J. Cote, D. H. Coy, J. Sueiras‐Diaz, and A. V. Schally. Stimulation of cyclic AMP accumulation and corticotropin release by synthetic ovine corticotropin‐releasing factor in rat anterior pituitary cells: site of glucocorticoid action. Proc. Natl. Acad. Sci. USA 79: 3466–3469, 1982.
 294. Gilbey, M. P., J. H. Coote, S. Fleetwood‐Walker, and D. F. Peterson. The influence of the paraventriculo‐spinal pathway, and oxytocin and vasopressin on sympathetic preganglionic neurones. Brain Res. 251: 283–290, 1982.
 295. Ginsborg, B. L., and G. D. S. Hirst. The effects of adenosine on the release of the transmitter from the phrenic nerve of the rat. J. Physiol. London 224: 629–645, 1972.
 296. Godfraind, J. M., and R. Pumain. Cyclic adenosine monophosphate and norepinephrine: effect on Purkinje cells in rat cerebellar cortex. Science 174: 1257–1259, 1971.
 297. Gold, M. R., and A. R. Martin. γ‐Aminobutyric acid and glycine activate Cl‐channels having different characteristics in CNS neurones. Nature London 308: 639–641. 1984.
 298. Goldberg, N. D., S. B. Dietz, and A. G. O'Toole. Cyclic guanosine 3′,5′‐monophosphate in mammalian tissues and urine. J. Biol. Chem. 244: 4458–4466, 1969.
 299. Goldstein, A., S. Tachibana, L. I. Lowney, M. Hunkapiller, and L. Hood. Dynorphin‐(1–13), an extraordinarily potent opioid peptide. Proc. Natl. Acad. Sci. USA 76: 6666–6670, 1979.
 300. Gothert, N. Somatostatin selectively inhibits noradrenaline release from hypothalamic neurones. Nature London 288: 86–88, 1980.
 301. Gottlieb, D. L., and W. M. Cowan. On the distribution of axonal terminals containing spheroidal and flattened synaptic vesicles in the hippocampus and dentate gyrus of the rat and cat. Z. Zellforsch. Mikrosk. Anat. 129: 413–419, 1972.
 302. Goyal, R. K., S. I. Said, and S. Rattan. Influence of VIP antiserum on lower esophageal sphincter relaxations possible evidence for VIP as the inhibitory neurotransmitter (Abstract). Gastroenterology 76: 1142 1979.
 303. Greengard, P. Intracellular signals in the brain. Harvey Lect. Ser. 75: 277–331, 1980.
 304. Greenwood, R. S., S. E. Godar, T. A. Reaves, Jr., and J. N. Hayward. Cholecystokinin in hippocampal pathways. J. Comp. Neurobiol. 203: 335–350, 1981.
 305. Gruol, D. L. Cultured cerebellar neurons: endogenous and exogenous components of Purkinje cell activity and membrane response to putative transmitters. Brain Res. 263: 223–241, 1983.
 306. Gruol, D. L. Excitatory and inhibitory actions of putative neurotransmitters on cultured cerebellar neurons. Soc. Neurosci. Abstr. 9: 680 1983.
 307. Gruol, D. L., C. Chavkin, R. J. Valentino, and G. R. Siggins. Dynorphin‐A alters the excitability of pyramidal neurons of the rat hippocampus in vitro. Life Sci. 33: 533–536, 1983.
 308. Gruol, D. L., and G. R. Siggins. Norepinephrine (NE) and adrenergic agonists in low concentrations alter spontaneous activity and membrane potential of CA1 pyramidal neurons of the rat hippocampus in vitro. Soc. Neurosci. Abstr. 8: 797 1982.
 309. Gruol, D. L., and G. R. Siggins. Adrenergic agonists alter the intracellularly recorded spontaneous activity and postdischarge after‐hyperpolarization of pyramidal neurons of the rat hippocampus in vitro. In: Catecholamines: Neuropharmacology and Central Nervous System—Theoretical Aspects, edited by E. Usdin, A. Carlsson, A. Dahlström, and J. Engel. New York: Liss, 1984, p. 131–133.
 310. Gruol, D. L., G. R. Siggins, A. L. Padjen, and D. S. Forman. Explant cultures of adult amphibian sympathetic ganglia: electrophysiological and pharmacological investigation of neurotransmitter and nucleotide action. Brain Res. 223: 81–106, 1981.
 311. Gruol, D. L., and T. G. Smith. Opiate antagonism of glycine‐evoked membrane polarizations in cultured mouse spinal cord neurons. Brain Res. 223: 355–365, 1981.
 312. Guillemin, R. Somatostatin inhibits the release of acetylcholine induced electrically in the myenteric plexus. Endocrinology 99: 1653–1654, 1976.
 313. Guillemin, R., P. Brazeau, P. Böhlen, F. Esch, N. Ling, and W. B. Wehrenberg. Growth hormone‐releasing factor from a human pancreatic tumor that caused acromegaly. Science 218: 585–587, 1982.
 314. Guillemin, R., and B. Rosenberg. Humoral hypothalamic control of anterior pituitary: a study with combined tissue cultures. Endocrinology 57: 599–607, 1955.
 315. Guyenet, P. G., and G. K. Aghajanian. ACh, substance P and met‐enkephalin in the locus coeruleus: pharmacological evidence for independent sites of action. Eur. J. Pharmacol. 53: 319–328, 1979.
 316. Guyenet, P. B., E. A. Mroz, G. K. Aghajanian, and S. E. Leeman. Delayed iontophoretic ejection of substance P from glass micropipettes: correlation with time‐course of neuronal excitation in vivo. Neuropharmacology 18: 553–558, 1979.
 317. Gysling, K., and R. Y. Wang. Morphine‐induced activation of A10 dopamine neurons in the rat. Brain Res. 277: 119–127, 1983.
 318. Haas, H. L. Histamine hyperpolarizes hippocampal neurones in vitro. Neurosci. Lett. 22: 75–78, 1981.
 319. Haas, H. L. Cholinergic disinhibition in hippocampal slices of the rat. Brain Res. 233: 200–204, 1982.
 320. Haas, H. L. Histamine potentiates neuronal excitation by blocking a calcium‐dependent potassium conductance. Agents Actions 14: 534–537, 1984.
 321. Haas, H. L., D. Felix, M. R. Celio, and T. Inagami. Angiotensin II in the hippocampus. A histochemical and electrophysiological study. Experientia 36: 1394–1395, 1980.
 322. Haas, H. L., D. Felix, and M. D. Davis. Angiotensin excites hippocampal pyramidal cells by two mechanisms. Cell. Mol. Neurobiol. 2: 21–32, 1982.
 323. Haas, H. L., and B. H. Gähwiler. Do enkephalins directly affect calcium‐spikes in hippocampal pyramidal cells? Neurosci. Lett. 19: 89–92, 1980.
 324. Haas, H. L., and A. Konnerth. Histamine and noradrenaline decrease calcium‐activated potassium conductance in hippocampal pyramidal cells. Nature London 302: 432–434, 1983.
 325. Haas, H. L., and R. W. Ryall. Is excitation by enkephalins of hippocampal neurones in the rat due to presynaptic facilitation or to disinhibition? J. Physiol. London 308: 315–330, 1980.
 326. Hablitz, J. J. Conductance changes induced by dl‐homocysteic acid and N‐methyl‐dl‐aspartic acid in hippocampal neurons. Brain Res. 247: 149–153, 1982.
 327. Hablitz, J. J., and D. Johnston. Endogenous nature of spontaneous bursting in hippocampal pyramidal neurons. Cell. Mol. Neurobiol. 1: 325–334, 1981.
 328. Hablitz, J. J., and I. A. Langmoen. Excitation of hippocampal pyramidal cells by glutamate in the guinea‐pig and rat. J. Physiol. London 325: 317–331, 1982.
 329. Haefely, W. E. Synaptic pharmacology of barbiturates and benzodiazepines. Agents Actions 7: 353–359, 1977.
 330. Halliwell, J. V., and P. R. Adams. Voltage‐clamp analysis of muscarinic excitation in hippocampal neurons. Brain Res. 250: 71–92, 1982.
 331. Hamill, O. P., J. Bormann, and B. Sakmann. Activation of multiple‐conductance state chloride channels in spinal neurones by glycine and GABA. Nature London 305: 805–807, 1983.
 332. Hamill, O. P., A. Marty, E. Neher, B. Sakmann, and F. J. Sigworth. Improved patch clamp techniques for high‐resolution current recording from cells and cell‐free membrane patches. Pfluegers Arch. 391: 85–100, 1981.
 333. Handelmann, G. E., D. K. Meyer, M. C. Beinfeld, and W. H. Oertel. CCK‐containing terminals in the hippocampus are derived from intrinsic neurons: an immunohistochemical and radioimmunological study. Brain Res. 224: 180–184, 1981.
 334. Hartzell, H. C. Mechanisms of slow postsynaptic potentials. Nature London 291: 539–544, 1981.
 335. Haskins, J. T., W. K. Samson, and R. L. Moss. Evidence for vasoactive intestinal polypeptide (VIP) altering the firing rate of preoptic, septal and midbrain central gray neurons. Regul. Pept. 3: 113–123, 1982.
 336. Henon, B. K., and D. A. McAfee. The ionic basis of adenosine receptor action on post‐ganglionic neurones in the rat. J. Physiol. London 336: 607–620, 1983.
 337. Henriksen, S. J., F. E. Bloom, F. McCoy, N. Ling, and R. Guillemin. β‐Endorphin induces nonconvulsive limbic seizures. Proc. Natl. Acad. Sci. USA 75: 5221–5225, 1978.
 338. Henriksen, S. J., G. Chouvet, and F. E. Bloom. In vivo cellular responses to electrophoretically applied dynorphin in the rat hippocampus. Life Sci. 31: 1785–1788, 1982.
 339. Henriksen, S. J., G. Chouvet, J. McGinty, and F. E. Bloom. Opioid peptides in the hippocampus: anatomical and physiological considerations. Ann. NY Acad. Sci. 398: 207–220, 1982.
 340. Henry, J. L. Effects of substance P on functionally identified units in cat spinal cord. Brain Res. 114: 431–451, 1976.
 341. Henry, J. L. Relation of substance P to pain transmission: neurophysiological evidence. In: Substance P in the Nervous System. London: Pitman, 1982, p. 206–224. (Ciba Found. Symp. 91.)
 342. Henry, J. L., K. Krnjević, and M. E. Morris. Substance P and spinal neurones. Can. J. Physiol. Pharmacol. 53: 423–432, 1975.
 343. Herrling, P. L. The effect of carbachol and acetylcholine on fornix evoked i.p.s.p.s recorded from cat hippocampal pyramidal cells in situ (Abstract). J. Physiol. London 318: 26P, 1981.
 344. Herrling, P. L. The membrane potential of cat hippocampal neurons recorded in vivo displays four different reaction‐mechanisms to iontophoretically applied transmitter agonists. Brain Res. 212: 331–343, 1981.
 345. Herrling, P. L., and C. D. Hull. Iontophoretically applied dopamine depolarizes and hyperpolarizes the membrane of cat caudate neurons. Brain Res. 192: 441–462, 1980.
 346. Herrling, P. L., R. Morris, and T. E. Salt. Effects of excitatory amino acids and their antagonists on membrane and action potentials of cat caudate neurones. J. Physiol. London 339: 207–222, 1983.
 347. Hicks, T. P., and H. McLennan. Amino acids and the synaptic pharmacology of granule cells in the dentate gyrus of the rat. Can. J. Physiol. Pharmacol. 57: 973–978, 1979.
 348. Hill, R. G., J. F. Mitchell, and C. M. Pepper. The excitation and depression of hippocampal neurones by iontophoretically applied enkephalins (Abstract). J. Physiol. London 272: 50P–51P, 1977.
 349. Hille, B. Ionic basis of resting and action potentials. In: Handbook of Physiology. The Nervous System, edited by J. M. Brookhart and V. B. Mountcastle. Bethesda, MD: Am. Physiol. Soc, 1977, sect. 1, vol. I, pt. 1, chapt. 4, p. 99–136.
 350. Hille, B. Ionic channels in excitable membranes. Current problems and biophysical approaches. Biophys. J. 22: 283–294, 1978.
 351. Hoffer, B., Å. Seiger, R. Freedman, L. Olson, and D. Taylor. Electrophysiology and cytology of hippocampal formation transplants in the anterior chamber of the eye. II. Cholinergic mechanism. Brain Res. 119: 107–132, 1977.
 352. Hoffer, B. J., and G. R. Siggins. Electrophysiological techniques for the study of hormone action in the central nervous system. In: Methods in Enzymology: Hormone Action, edited by B. W. O'Malley and J. G. Hardman. New York: Academic, 1975, vol. 39, pt. D, p. 429–440.
 353. Hoffer, B. J., G. R. Siggins, and F. E. Bloom. Studies on norepinephrine‐containing afferents to Purkinje cells of rat cerebellum. II. Sensitivity of Purkinje cells to norepinephrine and related substances administered by microiontophoresis. Brain Res. 25: 523–534, 1971.
 354. Hoffer, B. J., G. R. Siggins, A. P. Oliver, and F. E. Bloom. Cyclic AMP mediation of norepinephrine inhibition in rat cerebellar cortex: a unique class of synaptic responses. Ann. NY Acad. Sci. 185: 531–549, 1971.
 355. Hoffer, B. J., G. R. Siggins, A. P. Oliver, and F. E. Bloom. Activation of the pathway from locus coeruleus to rat cerebellar Purkinje neurons: pharmacological evidence of noradrenergic central inhibition. J. Pharmacol. Exp. Ther. 184: 553–569, 1973.
 356. Hoffer, B. J., G. R. Siggins, D. J. Woodward, and F. E. Bloom. Spontaneous discharge of Purkinje neurons after destruction of catecholamine‐containing afferents by 6‐hydroxydopamine. Brain Res. 30: 425–430, 1971.
 357. Hökfelt, T., S. Efendic, O. Johansson, R. Luft, and A. Arimura. Immunohistochemical localization of somatostatin (growth hormone release‐inhibiting factor) in the guinea pig brain. Brain Res. 80: 165–169, 1974.
 358. Hökfelt, T., R. Elde, O. Johansson, R. Luft, and A. Arimura. Immunohistochemical evidence for separate populations of somatostatin‐containing and substance P‐containing primary afferent neurons in the rat. Neuroscience 1: 131–136, 1976.
 359. Hökfelt, T., J. O. Kellerth, G. Nilsson, and B. Pernow. Experimental immunohistochemical studies on the localization and distribution of substance P in cat primary sensory neurones. Brain Res. 100: 235–252, 1975.
 360. Hökfelt, T., J. O. Kellerth, G. Nilsson, and B. Pernow. Substance P: localization in the central nervous system and in some primary sensory neurons. Science 190: 889–890, 1975.
 361. Hökfelt, T., A. Ljungdahl, H. Steinbusch, A. Verhofstad, G. Nilsson, E. Brodin, B. Pernow, and M. Goldstein. Immunohistochemical evidence of substance P‐like immunoreactivity in some 5‐hydroxytryptamine‐containing neurons in the rat central nervous system. Neuroscience 3: 517–538, 1978.
 362. Hökfelt, T., and U. Ungerstedt. Electron and fluorescence microscopic studies on the nucleus caudatus putamen of the rat after unilateral lesions of ascending nigro‐neostriatal dopamine neurons. Acta Physiol. Scand. 76: 415–426, 1969.
 363. Hori, N., C. R. Auker, D. J. Braitman, and D. O. Carpenter. Pharmacologic sensitivity of amino acid responses and synaptic activation of in vitro prepyriform neurons. J. Neurophysiol. 48: 1289–1301, 1982.
 364. Horn, J. P., and D. A. McAfee. Norepinephrine inhibits calcium‐dependent potentials in rat sympathetic neurons. Science 204: 1233–1235, 1979.
 365. Horn, J. P., and D. A. McAfee. Alpha‐adrenergic inhibition of calcium‐dependent potentials in rat sympathetic neurones. J. Physiol. London 301: 191–204, 1980.
 366. Hösli, L., E. Hösli, C. Zehntner, and H. Landolt. Effects of substance P on neurones and glial cells in cultured rat spinal cord. Neurosci. Lett. 24: 165–168, 1981.
 367. Hotson, J. R., and D. A. Prince. A calcium‐activated hyperpolarization follows repetitive firing in hippocampal neurons. J. Neurophysiol. 43: 409–419, 1980.
 368. Hounsgaard, J. Presynaptic inhibitory action of acetylcholine in area CA1 of the hippocampus. Exp. Neurol. 62: 787–797, 1978.
 369. Hudson, D. B., T. Valcana, G. Bean, and P. S. Timiras. Glutamic acid: a strong candidate as the neurotransmitter of the cerebellar granule cell. Neurochem. Res. 1: 73–81, 1976.
 370. Hughes, J., and H. W. Kosterlitz. Opioid peptides: introduction. Br. Med. Bull. 39: 1–3, 1983.
 371. Hughes, J., T. W. Smith, H. W. Kosterlitz, L. A. Fothergill, B. A. Morgan, and H. R. Morris. Identification of two related pentapeptides from the brain with potent opiate agonist activity. Nature London 258: 577–579, 1975.
 372. Huwyler, T., and D. Felix. Angiotensin II‐sensitive neurons in septal areas of the rat. Brain Res. 195: 187–195, 1980.
 373. Innis, R. B., F. M. A. Correa, G. R. Uhl, B. Schneider, and S. H. Snyder. Cholecystokinin octapeptide‐like immunoreactivity: histochemical localization in rat brain. Proc. Natl. Acad. Sci. USA 76: 521–525, 1979.
 374. Ioffe, S., V. Havlicek, H. Friesen, and V. Chernick. Effect of somatostatin (SRIF) and l‐glutamate on neurons of the sensorimotor cortex in awake habituated rabbits. Brain Res. 153: 414–418, 1978.
 375. Ishikawa, K., T. Ott, and J. L. McGaugh. Evidence for dopamine as a transmitter in dorsal hippocampus. Brain Res. 232: 222–226, 1982.
 376. Israel, M., B. Lesbats, F. M. Meunier, and J. Stinnakre. Postsynaptic release of adenosine triphosphate induced by single impulse transmitter action. Proc. R. Soc. London Ser. B 193: 461–468, 1976.
 377. Israel, M., and V. P. Whittaker. The isolation of mossy fibre endings from the granular layer of the cerebellar cortex. Experientia 27: 325–326, 1965.
 378. Itoh, N., K.‐I. Obata, N. Yanaihara, and H. Okamoto. Human preprovasoactive intestinal polypeptide contains a novel PHI‐27‐like peptide, PHM‐27. Nature London 304: 547–549, 1983.
 379. Ivanov, A. Y., and V. I. Skok. Slow inhibitory postsynaptic potentials and hyperpolarization evoked by noradrenaline in the neurones of mammalian sympathetic ganglion. J. Auton. Nerv. Syst. 1: 255–263, 1980.
 380. Iversen, L. L. Dopamine receptors in the brain. Science 188: 1084–1089, 1975.
 381. Iversen, L. L. Substance P. Br. Med. Bull. 38: 277–282, 1982.
 382. Iversen, L. L., S. D. Iversen, F. E. Bloom, C. Douglas, M. Brown, and W. Vale. Calcium‐dependent release of somatostatin and neurotensin from rat brain in vitro. Nature London 273: 161–163, 1978.
 383. Jackson, M. B., H. Lecar, D. A. Mathers, and J. L. Barker. Single channel currents activated by γ‐aminobutyric acid, muscimol, and (—)‐pentobarbital in cultured mouse spinal neurons. J. Neurosci. 2: 889–894, 1982.
 384. Jahnsen, H., and A. M. Laursen. The effects of a benzodiazepine on the hyperpolarizing and the depolarizing responses of hippocampal cells to GABA. Brain Res. 207: 214–217, 1981.
 385. Jahr, C. E., and R. A. Nicoll. Noradrenergic modulation of dendrodendritic inhibition in the olfactory bulb. Nature London 297: 227–279, 1982.
 386. Jan, Y. N., L. Y. Jan, and S. W. Kuffler. A peptide as a possible transmitter in sympathetic ganglia of the frog. Proc. Natl. Acad. Sci. USA 76: 1501–1505, 1979.
 387. Jan, Y. N., L. Y. Jan, and S. W. Kuffler. Further evidence for peptidergic transmission in sympathetic ganglia. Proc. Natl. Acad. Sci. USA 77: 5008–5012, 1980.
 388. Jeftinija, S., V. Miletic, and M. Randic. Cholecystokinin octapeptide excites dorsal horn neurones both in vivo and in vitro. Brain Res. 213: 231–236, 1981.
 389. Jeftinija, S., K. Murase, V. Nedeljkov, and M. Randic. Vasoactive intestinal polypeptide excites mammalian dorsal horn neurons both in vivo and in vitro. Brain Res. 243: 158–164, 1982.
 390. Jessel, T. M., and L. L. Iversen. Opiate analgesics inhibit substance P release from rat trigeminal nucleus. Nature London 268: 549–551, 1977.
 391. Jessen, K. R., and G. Burnstock. The enteric nervous system in tissue culture: a new mammalian model for the study of complex nervous networks. In: Trends in Autonomic Pharmacology, edited by S. Kalsner. Baltimore, MD: Urban & Schwarzenberg, 1982, vol. 2, p. 95–115.
 392. Johnson, J. L. The excitant amino acids glutamic and aspartic acid as transmitter candidates in the vertebrate central nervous system. Prog. Neurobiol. 10: 155–202, 1978.
 393. Johnston, D., and T. H. Brown. Interpretation of voltage‐clamp measurements in hippocampal neurons. J. Neurophysiol. 50: 464–486, 1983.
 394. Johnston, G. A. R. Physiologic pharmacology of GABA and its antagonists in the vertebrate nervous system. In: GABA in Nervous System Function, edited by E. Roberts, T. N. Chase, and D. B. Tower. New York: Raven, 1976, p. 395–411.
 395. Johnston, G. A. R. Neuropharmacology of amino acid inhibitory transmitters. Annu. Rev. Pharmacol. Toxicol. 18: 269–289, 1978.
 396. Johnston, G. A. R. Central nervous system receptors for glutamic acid. In: Glutamic Acid: Advances in Biochemistry and Physiology, edited by L. J. Filer Jr., S. Garattini, M. R. Kare, W. A. Reynolds, and R. J. Wurtman. New York: Raven, 1979, p. 177–185.
 397. Johnston, G. A. R., M. H. Hailstone, and C. G. Freeman. Baclofen: stereoselective inhibition of excitant amino acid release. J. Pharm. Pharmacol. 32: 230–231, 1980.
 398. Kakidani, H., Y. Furutani, H. Takahashi, M. Noda, Y. Morimoto, T. Hirose, M. Asai, S. Inayama, S. Nakanishi, and S. Numa. Cloning and sequence analysis of cDNA for porcine β‐neo‐endorphin/dynorphin precursor. Nature London 298: 245–249, 1982.
 399. Kalia, M., K. Fuxe, and, T. Hökfelt. Distribution of neuropeptide nerve terminals within the subnuclei of the nucleus of the tractus solitarius of the rat. Soc. Neurosci. Abstr. 9: 772 1983.
 400. Kandel, E. R., and J. H. Schwartz (editors). Principles of Neural Science. New York: Elsevier/North‐Holland, 1981.
 401. Kandel, E. R., W. A. Spencer, and F. J. Brinley, Jr. Electrophysiology of hippocampal neurons. I. Sequential invasion and synaptic organization. J. Neurophysiol. 24: 225–242, 1961.
 402. Katayama, Y., and S. Nishi. Voltage‐clamp analysis of peptidergic slow depolarisations in bullfrog sympathetic ganglion cells. J. Physiol. London 333: 305–313, 1982.
 403. Katayama, Y., and R. A. North. The action of somatostatin on neurones of the myenteric plexus of the guinea‐pig ileum. J. Physiol. London 303: 315–323, 1980.
 404. Katayama, Y., R. A. North, and J. T. Williams. The action of substance P on neurones of the myenteric plexus of the guinea‐pig small intestine. Proc. R. Soc. London Ser. B 206: 191–208, 1979.
 405. Katz, B. Nerve, Muscle, and Synapse. New York: McGraw‐Hill, 1966.
 406. Kebabian, J. W., and D. B. Calne. Multiple receptors for dopamine. Nature London 277: 93–96, 1979.
 407. Kelly, J. S. Electrophysiology of peptides in the central nervous system. Br. Med. Bull. 38: 283–290, 1982.
 408. Kerkut, G. A., and H. V. Wheal (editors). Electrophysiology of Isolated Mammalian CNS Preparations. London: Academic, 1981.
 409. Kitai, S. T., M. Sugimori, and J. D. Kocsis. Excitatory nature of dopamine in the nigro‐caudate pathway. Exp. Brain Res. 24: 351–363, 1976.
 410. Klainer, L. M., Y.‐M. Chi, S. L. Friedberg, T. W. Rall, and E. W. Sutherland. Adenyl cyclase. IV. The effects of neurohormones on the formation of adenosine 3′‐5′ phosphate by preparations from brain and other tissues. J. Biol. Chem. 237: 1239–1243, 1962.
 411. Klein, M., and E. R. Kandel. Presynaptic modulation of voltage‐dependent Ca2+ current: mechanism for behavioral sensitization in Aplysia californica. Proc. Natl. Acad. Sci. USA 75: 3512–3516, 1978.
 412. Kobayashi, R. M., M. Brown, and W. Vale. Regional distribution of neurotensin and somatostatin. Brain Res. 126: 584–588, 1977.
 413. Kobayashi, H., and B. Libet. Actions of noradrenaline and acetylcholine on sympathetic ganglion cells. J. Physiol. London 208: 353–372, 1970.
 414. Koda, L. Y., and F. E. Bloom. A light and electron microscopic study of noradrenergic terminals in the rat dentate gyrus. Brain Res. 120: 327–335, 1977.
 415. Koerner, J. F., and C. W. Cotman. Micromolar L‐2‐amino‐4‐phosphonobutyric acid selectively inhibits perforant path synapses from lateral entorhinal cortex. Brain Res. 216: 192–198, 1981.
 416. Koerner, J. F., and C. W. Cotman. Response of Schaffer collateral‐CA1 pyramidal cells synapses of the hippocampus to analogues of acidic amino acids. Brain Res. 251: 105–115, 1982.
 417. Köhler, C. Distribution and morphology of vasoactive intestinal polypeptide‐like immunoreactive neurons in regio superior of the rat hippocampal formation. Neurosci. Lett. 33: 265–270, 1982.
 418. Köhler, C., and, V. Chan‐Palay. The distribution of chole‐cystokinin‐like immunoreactive neurons and nerve terminals in the retrohippocampal region in the rat and guinea pig. J. Comp. Neurol. 210: 136–146, 1982.
 419. Konishi, S., and M. Otsuka. The effects of substance P and other peptides on spinal neurons of the frog. Brain Res. 65: 397–410, 1974.
 420. Konishi, S., and M. Otsuka. Excitatory action of hypothalamic substance P on spinal motoneurones of new‐born rats. Nature London 252: 734–735, 1974.
 421. Konishi, S., A. Tsunoo, and M. Otsuka. Substance P and non‐cholinergic excitatory synaptic transmission in guinea‐pig sympathetic ganglia. Proc. Jpn. Acad. 55: 525–530, 1979.
 422. Konishi, S., A. Tsunoo, N. Yanaihara, and M. Otsuka. Peptidergic excitatory and inhibitory synapses in mammalian sympathetic ganglia: roles of substance P and enkephalin. Biomed. Res. 1: 528–536, 1980.
 423. Konnerth, A., and U. Heinemann. Effects of GABA on presumed presynaptic Ca2+ entry in hippocampal slices. Brain Res. 270: 185–189, 1983.
 424. Korf, J., R. H. Roth, and G. K. Aghajanian. Alterations in turnover and endogenous levels of norepinephrine in cerebral cortex following electrical stimulation and acute axotomy of cerebral noradrenergic pathways. Eur. J. Pharmacol. 23: 276–282, 1973.
 425. Kretsinger, R. H. Calcium in neurobiology: a general theory of its function and evolution. In: Neurosciences: Fourth Study Program, edited by F. O. Schmitt and F. G. Worden. Cambridge, MA: MIT Press, 1979, p. 617–622.
 426. Krishtal, O. A., S. M. Marchenko, and V. I. Pidoplichko. Receptor for ATP in the membrane of mammalian sensory neurones. Neurosci. Lett. 35: 41–45, 1983.
 427. Krivoy, W., D. Kroeger, and E. Zimmermann. Additional evidence for a role of substance P in modulation of synaptic transmission. In: Substance P, edited by U. S. von Euler and B. Pernow. New York: Raven, 1977, p. 187–193.
 428. Krnjević, K. Inhibitory action of GABA and GABA‐mimetics on vertebrate neurons. In: GABA in Nervous System Function, edited by E. Roberts, T. N. Chase, and D. B. Tower. New York: Raven, 1976, p. 269–281.
 429. Krnjević, K. Effects of substance P on central neurons in cats. In: Substance P, edited by U. S. von Euler and B. Pernow. New York: Raven, 1977, p. 217–230.
 430. Krnjević, K. GABA and other transmitters in the cerebellum. Exp. Brain Res. Suppl. 6: 533–551, 1982.
 431. Krnjević, K., and J. W. Phillis. Iontophoretic studies of neurones in the mammalian cerebral cortex. J. Physiol. London 165: 274–304, 1963.
 432. Krnjević, K., R. Pumain, and L. Renaud. The mechanism of excitation of acetylcholine in the cerebral cortex. J. Physiol. London 215: 247–268, 1971.
 433. Krnjević, K., R. J. Reiffenstein, and N. Ropert. Disinhibitory action of acetylcholine in the rat's hippocampus: extracellular observations. Neuroscience 6: 2465–2474, 1981.
 434. Krnjević, K., and N. Ropert. Septo‐hippocampal pathway modulates hippocampal activity by a cholinergic mechanism. Can. J. Physiol. Pharmacol. 59: 911–914, 1981.
 435. Krnjević, K., and N. Ropert. Electrophysiological and pharmacological characteristics of facilitation of hippocampal population spikes by stimulation of the medial septum. Neuroscience 7: 2165–2183, 1982.
 436. Krnjević, K., and S. Schwartz. The action of γ‐aminobutyric acid on cortical neurons. Exp. Brain Res. 3: 320–336, 1967.
 437. Krnjević, K., and W. G. VanMeter. Cyclic nucleotides in spinal cells. Can. J. Physiol. Pharmacol. 54: 416–420, 1976.
 438. Krulich, L., A. P. S. Dhariwal, and S. M. McCann. Stimulatory and inhibitory effects of purified hypothalamic extracts on growth hormone release from rat pituitary in vitro. Endocrinology 83: 783–790, 1968.
 439. Kuba, K., and K. Koketsu. Ionic mechanism of slow excitatory postsynaptic potential in bullfrog sympathetic ganglion cells. Brain Res. 81: 338–342, 1974.
 440. Kuba, K., and K. Koketsu. Analysis of slow excitatory postsynaptic potential in bullfrog sympathetic ganglion cells. Jpn. J. Physiol. 26: 647–664, 1976.
 441. Kuffler, S. W. Slow synaptic responses in autonomic ganglia and the pursuit of a peptidergic transmitters. J. Exp. Biol. 89: 257–286, 1980.
 442. Kuroda, Y. Physiological roles of adenosine derivatives which are released during neurotransmission in mammalian brain. J. Physiol. Paris 74: 463–470, 1978.
 443. Kuroda, Y., and H. McIlwain. Uptake and release of [14C] adenine derivates at beds of mammalian cortical synaptosomes in a superfusion system. J. Neurochem. 22: 691–699, 1974.
 444. Lagendre, P., G. Simonnet, B. Dupouy, and J. D. Vincent. Transmitter‐like effects of angiotensin II on cultured spinal neurons. Soc. Neurosci. Abstr. 6: 842 1980.
 445. Lake, N., and L. M. Jordan. Failure to confirm cyclic AMP as second messenger for norepinephrine in rat cerebellum. Science 183: 663–664, 1974.
 446. Lambert, J. D. C., J. A. Flatman, and I. Engberg. Actions of excitatory amino acids on membrane conductance and potential in motoneurons. In: Glutamate as a Neurotransmitter, edited by G. Di Chiara and G. L. Gessa. New York: Raven, 1981, p. 205–216. (Adv. Biochem. Psychopharmacol. Ser., vol. 27.)
 447. Langer, S. Z. Presynaptic regulation of the release of catecholamines. Pharmacol. Rev. 32: 337–362, 1981.
 448. Langmoen, I. A., M. Segal, and P. Andersen. Mechanisms of norepinephrine actions on hippocampal pyramidal cells in vitro. Brain Res. 208: 349–362, 1981.
 449. Lanthorn, T. H., and C. W. Cotman. Baclofen selectively inhibits excitatory synaptic transmission in the hippocampus. Brain Res. 225: 171–178, 1981.
 450. Larsson, L.‐I., and J. F. Rehfeld. Localization and molecular heterogeneity of cholecystokinin in the central and peripheral nervous system. Brain Res. 165: 201–218, 1979.
 451. Latorre, R., R. Coronado, and C. Vergara. K+ channels gated by voltage and ions. Annu. Rev. Physiol. 46: 485–495, 1984.
 452. Lecar, H., and F. Sachs. Membrane noise analysis. In: Excitable Cells in Tissue Culture, edited by P. G. Nelson and M. Lieberman. New York: Plenum, 1981, p. 137–172.
 453. Lee, H. K., T. Dunwiddie, and B. Hoffer. Electrophysiological interactions of enkephalins with neuronal circuitry in the rat hippocampus. II. Effects on interneuron excitability. Brain Res. 184: 331–342, 1980.
 454. Lee, K. S., P. Schubert, H. Emmert, and G. W. Kreutzberg. Effect of adenosine versus adenine nucleotides on evoked potentials in a rat hippocampal slice preparation. Neurosci. Lett. 23: 309–314, 1981.
 455. Lee, K. S., P. Schubert, V. Gribkoff, B. Sherman, and G. Lynch. A combined in vivo/in vitro study of the presynaptic release of adenosine derivates in the hippocampus. J. Neurochem. 38: 80–83, 1982.
 456. Lembeck, F. Zur Frage der Zentralen Übertagung afferenter Impulse. III. Das Vorkommen und die Bedeutung der Substanz P in den dorsalen Würzelen des Rückenmarks. Naunyn‐Schmiedebergs Arch. Exp. Pathol. Pharmakol. 219: 107–213, 1953.
 457. Lembeck, F., K. Folkers, and J. Donnerer. Analgesic effects of antagonists of substance P. Biochem. Biophys. Res. Commun. 103: 1318–1321, 1981.
 458. Levy, R. A. The role of GABA in primary afferent depolarization. Prog. Neurobiol. Oxford 9: 211–267, 1977.
 459. Lewis, P. R., and C. C. D. Shute. The cholinergic limbic system: projections to hippocampal formation, medial cortex, nuclei of the ascending cholinergic reticular system and the subfornical organ and supra‐optic crest. Brain 90: 521–540, 1967.
 460. Lewis, P. R., C. C. D. Shute, and A. Silver. Confirmation from choline acetylase of a massive cholinergic innervation to the rat hippocampus. J. Physiol. London 191: 215–224, 1967.
 461. Libet, B. Long latent periods and further analysis of slow synaptic responses in sympathetic ganglia. J. Neurophysiol. 30: 494–514, 1967.
 462. Libet, B., and T. Tosaka. Dopamine as a synaptic transmitter and modulator in sympathetic ganglia: a different mode of synaptic action. Proc. Natl. Acad. Sci. USA 67: 667–673, 1970.
 463. Ling, N., R. Burgus, and R. Guillemin. Isolation, primary structure, and synthesis of α‐endorphin and γ‐endorphin, two peptides of hypothalamic‐hypophyseal origin with morphinomimetic activity. Proc. Natl. Acad. Sci. USA 73: 3942–3946, 1976.
 464. Llinás, R., and R. Hess. Tetrodotoxin‐resistant dendritic spikes in avian Purkinje cells. Proc. Natl. Acad. Sci. USA 73: 2520–2523, 1976.
 465. Llinás, R., and C. Nicholson. Electrophysiological properties of dendrites and somata in alligator Purkinje cells. J. Neurophysiol. 34: 532–551, 1971.
 466. Llinás, R., and C. Nicholson. Analysis of field potentials in the central nervous system. In: Handbook of Electroencephalography and Clinical Neurophysiology, edited by C. F. Stevens. Amsterdam: Elsevier, 1974, vol. 2, pt. B, p. 61–83.
 467. Llinás, R., and M. Sugimori. Calcium conductances in Purkinje cell dendrites: their role in development and integration. Prog. Brain Res. 51: 323–334, 1979.
 468. Llinás, R., and M. Sugimori. Electrophysiological properties of in vitro Purkinje cell dendrites in mammalian cerebellar slices. J. Physiol. London 305: 197–213, 1980.
 469. Loewi, O. Über humorale Übertragbarkeit der Herzennervenwirkung. I. Mitteilung. Pfluegers Arch. 189: 238–242, 1921.
 470. Londos, C., and J. Wolff. Two distinct adenosine sensitive sites on adenylate cyclase. Proc. Natl. Acad. Sci. USA 74: 5482–5486, 1977.
 471. Loren, I., J. Alumets, R. Hakanson, and F. Sundler. Distribution of gastrin and CCK‐like peptides in rat brain. Histochemistry 59: 249–257, 1979.
 472. Loren, I., P. C. Emson, J. Fahrenkrug, A. Björklund, J. Alumets, R. Håkanson, and F. Sundler. Distribution of vasoactive intestinal polypeptide in the rat and mouse brain. Neuroscience 4: 1953–1976, 1979.
 473. Lorente de Nó, R. Studies on the structure of the cerebral cortex. II. Continuation of the study of the ammonic system. J. Psychol. Neurol. 46: 113–117, 1934.
 474. Lovik, T. A. Enkephalin blocks primary afferent depolarization of Aδ tooth pulp afferents evoked by stimulation in nucleus raphe magnus in the decerebrate cat. Neurosci. Lett. 37: 273–278, 1983.
 475. Loy, R., D. A. Koziell, J. D. Lindsey, and R. Y. Moore. Noradrenergic innervation of the adult rat hippocampal formation. J. Comp. Neurol. 189: 699–710, 1980.
 476. Lundberg, J. M., and, T. Hökfelt. Coexistence of peptides and classical neurotransmitters. Trends Neurosci. 6: 325–333, 1983.
 477. Lynch, G., and P. Schubert. The use of in vitro brain slices for multidisciplinary studies of synaptic function. Annu. Rev. Neurosci. 3: 1–22, 1980.
 478. MacDonald, J. F., A. V. Porietis, and J. M. Wojtowicz. l‐Aspartic acid induces a region of negative slope conductance in the current‐voltage relationship of cultured spinal cord neurons. Brain Res. 237: 248–253, 1982.
 479. MacDonald, J. F., and J. M. Wojtowicz. Two conductance mechanisms activated by applications of l‐glutamic, l‐aspartic, dl‐homocysteic, N‐methyl‐D‐aspartic, and DL‐kainic acids to cultured mammalian central neurones. Can. J. Physiol. Pharmacol. 58: 1393–1397, 1980.
 480. MacDonald, J. F., and J. M. Wojtowicz. The effects of L‐glutamate and its analogues upon the membrane conductance of central murine neurones in culture. Can. J. Physiol. Pharmacol. 60: 282–296, 1982.
 481. Macdonald, R. L., and P. G. Nelson. Specific‐opiate‐induced depression of transmitter release from dorsal root ganglion cells in culture. Science 199: 1449–1451, 1978.
 482. Macdonald, R. L., and L. M. Nowak. Substance P and somatostatin actions on spinal cord neurons in primary dissociated cell culture. In: Neurosecretion and Brain Peptides: Implications for Brain Function and Neurological Disease, edited by J. B. Martin, S. Reichlin, and K. L. Bick. New York: Raven, 1981, p. 159–173.
 483. Macdonald, R. L., and A. B. Young. Pharmacology of GABA‐mediated inhibition of spinal cord neurons in vivo and in primary dissociated cell culture. Mol. Cell. Biochem. 38: 147–162, 1981.
 484. MacMillan, S. J., and G. Clarke. Opioid peptides have differential actions on sub‐populations of arcuate neurones. Life Sci. 33: 529–532, 1983.
 485. Madison, D. V., and R. A. Nicoll. Noradrenaline blocks accommodation of pyramidal cell discharge in the hippocampus. Nature London 299: 636–638, 1982.
 486. Magistretti, P. J., J. H. Morrison, W. J. Shoemaker, V. Sapin, and F. E. Bloom. Vasoactive intestinal polypeptide induces glycogenolysis in mouse cortical slices: a possible regulatory mechanism for the local control of energy metabolism. Proc. Natl. Acad. Sci. USA 78: 6535–6539, 1981.
 487. Magistretti, P. J., and M. Schorderet. VIP and noradrenaline act synergistically to increase cyclic AMP in cerebral cortex. Nature London 308: 280–282, 1984.
 488. Marciani, M. G., P. Stanzione, E. Cherubini, and G. Bernardi. Action mechanisms of γ‐aminobutyric acid (GABA) and glycine on rat cortical neurons. Neurosci. Lett. 18: 169–172, 1980.
 489. Marshall, K. C. Catecholamines and their actions in the spinal cord. In: Handbook of the Spinal Cord. Pharmacology, edited by R. A. Davidoff. New York; Dekker, 1983, vol. 1, p. 275–328.
 490. Marshall, K. C., and I. Engberg. Reversal potential for noradrenaline‐induced hyperpolarization of spinal motoneurons. Science 205: 422–424, 1979.
 491. Marshall, K. C., R. Y. K. Pun, W. J. Hendelman, and P. G. Nelson. A coeruleo‐spinal system in culture. Science 213: 355–357, 1981.
 492. Martin, D. L. Carrier‐mediated transport and removal of GABA from synaptic regions. In: GABA in Nervous System Function, edited by E. Roberts, T. N. Chase, and D. B. Tower. New York: Raven, 1976, p. 347–386.
 493. Masukawa, L. M., L. S. Benardo, and D. A. Prince. Variations in electrophysiological properties of hippocampal neurons in different subfields. Brain Res. 242: 341–344, 1982.
 494. Masukawa, L. M., and D. A. Prince. Enkephalin inhibition of inhibitory input to CA1 and CA3 pyramidal neurons in the hippocampus. Brain Res. 249: 271–280, 1982.
 495. Masukawa, L. M., and D. A. Prince. Synaptic control of excitability in isolated dendrites of hippocampal neurons. J. Neurosci. 4: 217–227, 1984.
 496. Mathers, D. A., and J. L. Barker. Chemically induced ion channels in nerve cell membranes. Int. Rev. Neurobiol. 23: 1–34, 1982.
 497. Matsuzaki, Y., Y. Hamasaki, and S. I. Said. Vasoactive intestinal peptide: a possible transmitter of nonadrenergic relaxation of guinea pig airways. Science 210: 1252–1253, 1980.
 498. McCaman, R. E., D. G. McKenna, and J. K. Ono. A pressure system for intracellular and extracellular ejections of picoliter volumes. Brain Res. 136: 141–147, 1977.
 499. McCance, I., and J. W. Phillis. Cholinergic mechanisms in the cerebellar cortex. Int. J. Neuropharmacol. 7: 447–462, 1968.
 500. McGinty, J. F., S. J. Henriksen, A. Goldstein, L. Terenius, and F. E. Bloom. Dynorphin is contained within hippocampal mossy fibers: immunochemical alterations after kainic acid administration and colchicine‐induced neurotoxicity. Proc. Natl. Acad. Sci. USA 80: 589–593, 1983.
 501. McKenzie, S. G., and H. P. Bar. On the mechanism of adenyl cyclase inhibition by adenosine. Can. J. Physiol. Pharmacol. 51: 190–196, 1973.
 502. McKinney, M., J. T. Coyle, and J. C. Hedreen. Topographic analysis of the innervation of the rat neocortex and hippocampus by the basal forebrain cholinergic system. J. Comp. Neurol. 217: 103–121, 1983.
 503. McLennan, H. On the nature of the receptors for various excitatory amino acids in the mammalian central nervous system. In: Glutamate as a Neurotransmitter, edited by G. Di Chiara and G. L. Gessa. New York: Raven, 1981, p. 253–262. (Adv. Biochem. Psychopharmacol. Ser., vol. 27.)
 504. Mellgren, S. I., and B. Srebro. Changes in acetylcholinesterase and distribution of degenerating fibers in the hippocampal region after septal lesions in the rat. Brain Res. 52: 19–36, 1973.
 505. Minota, S., and K. Koketsu. Effects of adrenaline on the action potential of sympathetic ganglion cells in bullfrogs. Jpn. J. Physiol. 27: 353–366, 1977.
 506. Mizobe, F., V. Kosousek, D. M. Dean, and B. G. Livett. Pharmacological characterization of adrenal paraneurons: substance P and somatostatin as inhibitory modulators of the nicotinic response. Brain Res. 178: 555–566, 1979.
 507. Monaghan, D. T., V. R. Holets, D. W. Toy, and C. W. Cotman. Anatomical distributions of four pharmacologically distinct 3H‐l‐glutamate binding sites. Nature London 306: 176–179, 1983.
 508. Moore, R. Y. Monoamine neurons innervating the hippocampal formation and septum: organization and response to injury. In: The Hippocampus, edited by R. L. Isaacson and K. H. Pribram. New York: Plenum, 1975, vol. 5, p. 215–237.
 509. Moore, R. Y., and F. E. Bloom. Central catecholamine neuron systems: anatomy and physiology of the norepinephrine and epinephrine systems. Annu. Rev. Neurosci. 2: 113–168, 1979.
 510. Morel, N., and F. M. Meunier. Simultaneous release of acetylcholine and ATP from stimulation cholinergic synaptosomes. J. Neurochem. 36: 1766–1773, 1981.
 511. Morita, K., R. A. North, and T. Tokimasa. The calcium‐activated potassium conductance in guinea‐pig myenteric neurones. J. Physiol. London 329: 341–354, 1982.
 512. Morita, K., R. A. North, and T. Tokimasa. Muscarinic agonists inactivate potassium conductance of guinea‐pig myenteric neurones. J. Physiol. London 333: 125–139, 1982.
 513. Morris, R., T. E. Salt, M. V. Sofroniew, and R. G. Hill. Actions of microiontophoretically applied oxytocin, vasopressin and neurophysin in the rat caudal medulla. Neurosci. Lett. 18: 163–168, 1980.
 514. Morrison, J. H., R. Benoit, P. J. Magistretti, and F. E. Bloom. Immunohistochemical distribution of pro‐somatostatin‐related peptides in cerebral cortex. Brain Res. 262: 344–351, 1983.
 515. Morrison, J. H., R. Benoit, P. J. Magistretti, N. Ling, and F. E. Bloom. Immunohistochemical distribution of prosomatostatin‐related peptides in hippocampus. Neurosci. Lett. 34: 137–142, 1982.
 516. Morrison, J. H., P. J. Magistretti, R. Benoit, and F. E. Bloom. The distribution and morphological characteristics of the intracortical VIP‐positive cell: an immunohistochemical analysis. Brain Res. 292: 269–282, 1984.
 517. Morrison, J. H., M. E. Molliver, R. Grzanna, and J. T. Coyle. The intra‐cortical trajectory of the coeruleo‐cortical projection in the rat: a tangentially organized cortical afferent. Neuroscience 6: 139–158, 1981.
 518. Mosco, S., G. Lynch, and C. Cotman. The distribution of septal projections to the hippocampus of the rat. J. Comp. Neurol. 152: 163–174, 1973.
 519. Moss, R. L., R. E. J. Dyball, and B. A. Cross. Excitation of antidromically identified neurosecretory cells of the paraventricular nucleus by oxytocin applied iontophoretically. Exp. Neurol. 34: 95–102, 1972.
 520. Mountcastle, V. B. Central nervous mechanisms in mechanoreceptive sensibility. In: Handbook of Physiology. The Nervous System, edited by J. M. Brookhart and V. B. Mountcastle. Bethesda, MD: Am. Physiol. Soc., 1984, sect. 1, vol. III, pt. 2, chapt. 18, p. 789–878.
 521. Mudge, A. W., S. E. Leeman, and G. D. Fischbach. Enkephalin inhibits release of substance P from sensory neurons in culture and decreases action potential duration. Proc. Natl. Acad. Sci. USA 76: 526–530, 1979.
 522. Mueller, A. L., B. J. Hoffer, and T. V. Dunwiddie. Noradrenergic responses in rat hippocampus: evidence for mediation by α and β receptors in the in vitro slice. Brain Res. 214: 113–126, 1981.
 523. Mueller, A. L., K. L. Kirk, B. J. Hoffer, and T. V. Dunwiddie. Noradrenergic responses in rat hippocampus: electrophysiological actions of direct‐ and indirect‐acting sympathomimetics in the in vitro slice. J. Pharm. Exp. Ther. 223: 599–605, 1982.
 524. Mueller, A. L., M. R. Palmer, B. J. Hoffer, and T. V. Dunwiddie. Hippocampal noradrenergic responses in vivo and in vitro: characterization of alpha and beta components. Naunyn‐Schmiedebergs Arch. Pharmacol. 318: 259–266, 1982.
 525. Mühlethaler, M., S. Charpak, and J. J. Dreifuss. Contrasting effects of neurohypophysial peptides on pyramidal and non‐pyramidal neurones in the rat hippocampus. Brain Res. 308: 97–107, 1984.
 526. Mühlethaler, M., J. J. Dreifuss, and B. H. Gähwiler. Vasopressin excites hippocampal neurones. Nature London 296: 749–751, 1982.
 527. Mühlethaler, M., W. H. Sawyer, M. M. Manning, and J. J. Dreifuss. Characterization of a uterine‐type oxytocin receptor in the rat hippocampus. Proc. Natl. Acad. Sci. USA 80: 6713–6717, 1983.
 528. Müller, J. E., E. Straus, and R. S. Yalow. Cholecystokinin and its COOH‐terminal octapeptide in the pig brain. Proc. Natl. Acad. Sci. USA 74: 3035–3037, 1977.
 529. Murase, K., V. Nedelikov, and M. Randic. The actions of neuropeptides on dorsal horn neurons in the rat spinal cord slice preparation: an intracellular study. Brain Res. 234: 170–176, 1982.
 530. Nadi, N. S., W. J. McBride, and M. H. Aprison. Distribution of several amino acids in regions of the cerebellum of the rat. J. Neurochem. 28: 453–455, 1977.
 531. Nakai, Y., and S. Takaori. Influence of norepinephrine‐containing neurons derived from the locus coeruleus on lateral geniculate neuronal activities of cats. Brain Res. 71: 47–60, 1974.
 532. Nakajima, S., and K. Takahashi. Post‐tetanic hyperpolarisation and electrogenic Na pump in stretch receptor neurone of crayfish. J. Physiol. London 187: 105–127, 1967.
 533. Neher, E., B. Sakmann, and J. H. Steinbach. The extracellular patch clamp: a method for resolving currents through individual open channels in biological membranes. Pfluegers Arch. 375: 219–228, 1978.
 534. Neher, E., and C. F. Stevens. Conductance fluctuations and ionic pores in membranes. Annu. Rev. Biophys. Bioeng. 6: 345–381, 1977.
 535. Nelson, P. G., and M. Lieberman (editors). Excitable Cells in Tissue Culture. New York: Plenum, 1981.
 536. Neuman, R. S., and C. W. Harley. Long‐lasting potentiation of the dentate gyrus population spike by norepinephrine. Brain Res. 273: 162–165, 1983.
 537. Newberry, N. R., and R. A. Nicoll. A bicuculline‐resistant inhibitory post‐synaptic potential in rat hippocampal pyramidal cells in vitro. J. Physiol. London 348: 239–254, 1984.
 538. Newberry, N. R., and R. A. Nicoll. Comparison of the action of baclofen with γ‐aminobutyric acid on rat hippocampal pyramidal cells in vitro. J. Physiol. London 360: 161–185, 1985.
 539. Nicoll, R. A. Presynaptic action of barbiturates in the frog spinal cord. Proc. Natl. Acad. Sci. USA 72: 1460–1463, 1975.
 540. Nicoll, R. A. The action of thyrotropin‐releasing hormone, substance P and related peptides on frog spinal motoneurons. J. Pharmacol. Exp. Ther. 207: 817–824, 1978.
 541. Nicoll, R. A. Responses of central neurons to opiates and opioid peptides. In: Regulatory Peptides: From Molecular Biology to Function, edited by E. Costa and M. Trabucchi. New York: Raven, 1982, p. 337–346.
 542. Nicoll, R. A., and B. E. Alger. Presynaptic inhibition: transmitter and ionic mechanisms. Int. Rev. Neurobiol. 21: 217–258, 1979.
 543. Nicoll, R. A., B. E. Alger, and C. E. Jahr. Enkephalin blocks inhibitory pathways in the vertebrate CNS. Nature London 287: 22–25, 1980.
 544. Nicoll, R. A., B. E. Alger, and C. E. Jahr. Peptides as putative excitatory neurotransmitters: carnosine, enkephalin, substance P and TRH. Proc. R. Soc. London Ser. B 210: 133–149, 1980.
 545. Nicoll, R. A., and J. L. Barker. Excitation of supraoptic neurosecretory cells by angiotensin II. Nature London New Biol. 233: 172–173, 1971.
 546. Nicoll, R. A., and J. L. Barker. The pharmacology of recurrent inhibition in the supraoptic neurosecretory system. Brain Res. 35: 501–511, 1971.
 547. Nicoll, R. A., J. C. Eccles, T. Oshima, and F. Rubia. Prolongation of hippocampal inhibitory postsynaptic potentials by barbiturates. Nature London 258: 625–627, 1975.
 548. Nicoll, R. A., C. Schenker, and S. E. Leeman. Substance P as a transmitter candidate. Annu. Rev. Neurosci. 3: 227–268, 1980.
 549. Nicoll, R. A., G. R. Siggins, N. Ling, F. E. Bloom, and R. Guillemin. Neuronal actions of endorphin and enkephalin among brain regions: a comparative microiontophoretic study. Proc. Natl. Acad. Sci. USA 74: 2584–2588, 1977.
 550. Nilsson, G., T. Hökfelt, and B. Pernow. Distribution of substance P‐like immunoreactivity in the rat central nervous system as revealed by immunohistochemistry. Med. Biol. 52: 424–427, 1974
 551. Nishi, S. Cellular pharmacology of ganglionic transmission. In: Advances in General and Cellular Pharmacology, edited by T. Narahashi and C. P. Bianchi. New York: Plenum, 1976, vol. 1, p. 179–245.
 552. Nishi, S., H. Soeda, and K. Koketsu. Studies on sympathetic B and C neurons and patterns of preganglionic innervation. J. Cell. Comp. Physiol. 66: 19–32, 1965.
 553. Nishi, S., H. Soeda, and K. Koketsu. Unusual nature of ganglionic slow EPSP studied by a voltage‐clamp method. Life Sci. 8: 33–42, 1969.
 554. Nistri, A., and A. Constanti. Pharmacological characterization of different types of GABA and glutamate receptors in vertebrates and invertebrates. Prog. Neurobiol. 13: 117–235, 1979.
 555. North, R. A. Opiates, opioid peptides and single neurones. Life Sci. 24: 1527–1546, 1979.
 556. North, R. A., and T. Tokimasa. Muscarinic synaptic potentials in guinea‐pig myenteric plexus neurones. J. Physiol. London 333: 151–156, 1982.
 557. North, R. A., and M. Yoshimura. The actions of noradrenaline on neurones of the rat substantia gelatinosa in vitro. J. Physiol. London 349: 43–55, 1984.
 558. Nowak, L., P. Bregestovski, P. Ascher, A. Herbet, and A. Prochiantz. Magnesium gates glutamate‐activated channels in mouse central neurones. Nature London 307: 462–465, 1984.
 559. Nowak, L. M., and R. L. Macdonald. Substance P decreases a potassium conductance of spinal cord neurons in cell culture. Brain Res. 214: 416–423, 1981.
 560. Nowak, L. M., and R. L. Macdonald. Substance P: ionic basis for depolarizing responses of mouse spinal cord neurons in cell culture. J. Neurosci. 2: 1119–1128, 1982.
 561. Nowak, L. M., and R. L. Macdonald. Muscarine‐sensitive voltage‐dependent potassium current in cultured murine spinal cord neurons. Neurosci. Lett. 35: 85–91, 1983.
 562. Oakley, B., and R. Schafer. Experimental Neurobiology: A Laboratory Manual. Ann Arbor: Univ. of Michigan Press, 1978.
 563. Obata, K., M. Ito, R. Ochi, and N. Sato. Pharmacological properties of the postsynaptic inhibition by Purkinje cell axons and the action of γ‐aminobutyric acid on deiters neurones. Exp. Brain Res. 4: 43–57, 1967.
 564. Obata, K., K. Takeda, and H. Shinozaki. Further study on pharmacological properties of the cerebellar‐induced inhibition of deiters neurones. Exp. Brain Res. 11: 327–342, 1970.
 565. Oertel, W. H., A. M. Graybiel, E. Mugnaini, R. P. Elde, D. E. Schmechel, and I. J. Kopin. Coexistence of glutamate decarboxylase immunoreactivity and somatostatin‐like immuno‐reactivity in neurones of nucleus reticularis thalami of the cat. Soc. Neurosci. Abstr. 7: 3 1981.
 566. Oertel, W. H., A. M. Graybiel, E. Mugnaini, R. P. Elde, D. E. Schmechel, and I. J. Kopin. Coexistence of glutamic‐acid decarboxylase‐ and somatostatin‐like immunoreactivity in neurons of the feline nucleus reticularis thalami. J. Neurosci. 3: 1322–1332, 1983.
 567. Ogata, N. Substance P causes direct depolarisation of neurones of guinea pig interpeduncular nucleus in vitro. Nature London 277: 480–481, 1979.
 568. Okada, Y., and M. Saito. Inhibitory action of adenosine, 5‐HT (serotonin) and GABA (γ‐aminobutyric acid) on the postsynaptic potential (PSP) of slices from olfactory cortex and superior colliculus in correlation to the level of cyclic AMP. Brain Res. 160: 368–371, 1979.
 569. Okada, Y., and C. Shimada. Intrahippocampal distribution of GABA and GAD activity in the guinea pig: microassay method for the determination of GAD activity. In: GABA in Nervous System Function, edited by E. Roberts, T. N. Chase, and D. B. Tower. New York: Raven, 1976, p. 223–228.
 570. Olpe, H.‐R., V. J. Balcar, H. Bittiger, H. Rink, and P. Sieber. Central actions of somatostatin. Eur. J. Pharmacol. 63: 127–133, 1980.
 571. Olpe, H.‐R., and V. Baltzer. Vasopressin activates noradrenergic neurones in the rat locus coeruleus: a microiontophoretic investigation. Eur. J. Pharmacol. 73: 377–378, 1981.
 572. Olpe, H.‐R., M. Baudry, L. Fagni, and G. Lynch. The blocking action of baclofen on excitatory transmission in the rat hippocampal slice. J. Neurosci. 2: 698–703, 1982.
 573. Olpe, H.‐R., A. Glatt, J. Laszlo, and A. Schellenberg. Some electrophysiological and pharmacological properties of the cortical noradrenergic projection of the locus coeruleus in the rat. Brain Res. 186: 9–19, 1980.
 574. Olsen, R. W., D. Greenlee, P. Van Ness, and M. K. Ticku. Studies of the gamma‐aminobutyric acid receptor/ionophore proteins in mammalian brain. In: Amino Acids as Chemical Transmitters, edited by F. Fonnum. New York: Plenum, 1978, p. 467–486.
 575. Olsen, R. W., and F. L. Lundberg. Convulsant and anticonvulsant drug binding sites related to GABA‐regulated chloride ion channels. In: GABA and Benzodiazepine Receptors, edited by E. Costa, G. Di Chiara, and G. L. Gessa. New York: Raven, 1981, p. 93–102.
 576. Oomura, Y., M. Otita, H. Kita, S. Ishibashi, and T. Okajima. Hypothalamic neuron response to glucose, phlorizin and cholecystokinin. In: Iontophoresis and Transmitter Mechanisms in the Mammalian Central Nervous System, edited by R. W. Ryall and J. S. Kelly. Amsterdam: Elsevier/North‐Holland Biomed., 1978, p. 120–123.
 577. Otmakhov, N. A., and A. G. Bragin. Effects of norepinephrine and serotonin upon spontaneous activity and responses to mossy fiber stimulation of CA3 neurons in hippocampal slices. Brain Res. 253: 173–183, 1982.
 578. Otsuka, M., and S. Konishi. GABA in the spinal cord. In: GABA in Nervous System Function, edited by E. Roberts, T. N. Chase, and D. B. Tower. New York: Raven, 1976, p. 197–202.
 579. Otsuka, M., S. Konishi, M. Yanagisawa, A. Tsunoo, and H. Akagi. Role of substance P as a sensory transmitter in spinal cord and sympathetic ganglia. In: Substance P in the Nervous System, edited by G. Lawrenson. London: Pitman, 1982, p. 13–34. (Ciba Found. Symp. No. 91.)
 580. Ozawa, S., and Y. Okada. Adenosine inhibits synaptic transmission in the guinea pig hippocampus in vitro. Neurosci. Lett. 4, Suppl.: S52, 1980.
 581. Padjen, A. L. Effects of somatostatin on frog spinal cord. Soc. Neurosci. Abstr. 3: 411 1977.
 582. Padjen, A. L., and P. A. Smith. Possible role of divalent cations in amino acid responses of frog spinal cord. In: Amino Acid Neurotransmitters, edited by F. V. DeFeudis and P. Mandel. New York: Raven, 1981, p. 271–280.
 583. Palay, S. L., and, V. Chan‐Palay. Cerebellar Cortex: Cytology and Organization. New York: Springer, 1974.
 584. Palmer, M. R., R. Freedman, and T. V. Dunwiddie. Interactions of a neuroleptic drug (Fluphenazine) with catecholamines in hippocampus. Psychopharmacology 76: 122–129, 1982.
 585. Palmer, M. R., S. M. Wuerthele, and B. J. Hoffer. Physical and physiological characteristics of micropressure ejection of drugs from multibarreled pipettes. Neuropharmacology 19: 931–938, 1980.
 586. Pasternak, G. W., R. Goodman, and S. H. Snyder. An endogenous morphine‐like factor in mammalian brain. Life Sci. 16: 1765–1769, 1975.
 587. Patel, Y. C., and S. Reichlin. Somatostatin in hypothalamus, extra hypothalamic brain and peripheral tissues of the rat. Endocrinology 102: 523–530, 1978.
 588. Paterson, S. J., L. E. Robson, and H. W. Kosterlitz. Classification of opioid receptors. Br. Med. Bull. 39: 31–36, 1983.
 589. Patterson, P. H. Environmental determination of autonomic neurotransmitter functions. Annu. Rev. Neurosci. 1: 1–17, 1978.
 590. Paull, W. K., J. Schöler, A. Arimura, C. A. Meyers, J. K. Chang, D. Chang, and M. Shimizu. Immunocytochemical localization of CRF in the ovine hypothalamus. Peptides 3: 183–191, 1982.
 591. Pedersen, C. A., J. A. Ascher, Y. L. Monroe, and A. J. Prangle, Jr. Oxytocin induces maternal behavior in virgin female rats. Science 216: 648–650, 1982.
 592. Pellmar, T. C., and D. O. Carpenter. Voltage‐dependent calcium current induced by serotonin. Nature London 277: 483–484, 1979.
 593. Pellmar, T. C., and D. O. Carpenter. Serotonin induces a voltage‐sensitive calcium current in neurons of Aplysia californica. J. Neurophysiol. 44: 423–439, 1980.
 594. Pepper, C. M., and G. Henderson. Opiates and opioid peptides hyperpolarize locus coeruleus neurons in vitro. Science 209: 394–396, 1980.
 595. Petrusz, P., M. Sar, G. H. Grossman, and J. S. Kizer. Synaptic terminals with somatostatin‐like immunoreactivity in the rat brain. Brain Res. 137: 181–187, 1977.
 596. Phillips, M. I., and D. Felix. Specific angiotensin II receptive neurons in the cat subfornical organ. Brain Res. 109: 531–540, 1976.
 597. Phillis, J. W. The role of calcium in the central effects of biogenic amines. Life Sci. 14: 1189–1201, 1974.
 598. Phillis, J. W. Responses of central neurons to neuronally localized peptides. In: Aminergic and Peptidergic Receptors, edited by E. S. Vizi and M. Wollemann. New York: Pergamon, 1980, p. 151–168.
 599. Phillis, J. W., and J. P. Edstrom. Effects of adenosine analogs on rat cerebral cortical neurons. Life Sci. 19: 1041–1054, 1976.
 600. Phillis, J. W., J. P. Edstrom, G. K. Kostopoulos, and J. R. Kirkpatrick. Effects of adenosine and adenine nucleotides on synaptic transmission in the cerebral cortex. Can. J. Physiol. Pharmacol. 57: 1289–1312, 1979.
 601. Phillis, J. W., and J. R. Kirkpatrick. The actions of adenosine and various nucleosides and nucleotides on the isolated toad spinal cord. Gen. Pharmacol. 9: 239–247, 1978.
 602. Phillis, J. W., and J. R. Kirkpatrick. Actions of various gastrointestinal peptides on the isolated amphibian spinal cord. Can. J. Physiol. Pharmacol. 57: 887–899, 1979.
 603. Phillis, J. W., and J. R. Kirkpatrick. The actions of motilin, luteinizing hormone releasing hormone, cholecystokinin, somatostatin, vasoactive intestinal polypeptide and other peptides on rat cerebral cortical neurons. Can. J. Physiol. Pharmacol. 58: 612–623, 1980.
 604. Phillis, J. W., J. R. Kirkpatrick, and S. I. Said. Vasoactive intestinal polypeptide excitation of central neurons. Can. J. Physiol. Pharmacol. 56: 337–340, 1978.
 605. Phillis, J. W., and G. P. Kostopoulos. Activation of a noradrenergic pathway from the brain stem to rat cerebral cortex. Gen. Pharmacol. 8: 379–384, 1977.
 606. Phillis, J. W., and J. J. Limacher. Substance P excitation of cerebral cortical Betz cells. Brain Res. 69: 158–163, 1974.
 607. Phillis, J. W., A. K. Tebecis, and D. H. York. Depression of spinal motoneurones by noradrenaline, 5‐hydroxytryptamine and histamine. Eur. J. Pharmacol. 4: 471–475, 1968.
 608. Phillis, J. W., and P. H. Wu. The role of adenosine and its nucleotides in central synaptic transmission. Prog. Neurobiol. 16: 187–239, 1981.
 609. Phillis, J. W., and P. H. Wu. Catecholamines and the sodium pump in excitable cells. Prog. Neurobiol. 17: 141–184, 1981.
 610. Pinget, M., E. Straus, and R. S. Yalow. Release of cholcystokinin peptides from a synaptosome enriched fraction of rat cerebral cortex. Life Sci. 25: 339–342, 1979.
 611. Piper, P. J., S. I. Said, and J. R. Vane. Effects on smooth muscle preparations of unidentified vasodilator peptide from intestine and lung. Nature London 225: 1144–1146, 1970.
 612. Pittman, Q. J., and G. R. Siggins. Somatostatin hyperpolarizes hippocampal pyramidal cells in vitro. Brain Res. 221: 402–408, 1981.
 613. Plotsky, P. M., and W. Vale. Hemorrhage‐induced secretion of corticotropin‐releasing factor‐like immunoreactivity into the rat hypophysial portal circulation and its inhibition by glucocorticoids. Endocrinology 114: 164–169, 1984.
 614. Porter, K. R. Motility in cells. Cold Spring Harbor Conf. Cell Proliferation 3: 1–28, 1976.
 615. Potashner, S. J. Baclofen: effects on amino acid release and metabolism in slices of guinea pig cerebral cortex. J. Neurosci. 32: 103–109, 1979.
 616. Proctor, W. R., and T. V. Dunwiddie. Adenosine inhibits calcium spikes in hippocampal pyramidal neurons in vitro. Neurosci. Lett. 35: 197–201, 1983.
 617. Pull, I., and H. McIlwain. Adenine derivates as neurohumoral agents in the brain. The quantities liberated on excitation of superfused cerebral tissues. Biochem. J. 130: 979–981, 1972.
 618. Pun, R. Y. K. Studies on synaptic transmission in spinal cord cultures: a comparison of postsynaptic actions of classical neurotransmitters with the peptides. Peptides 3: 249–257, 1982.
 619. Purves, R. D. Microelectrode Methods for Intracellular Recording and Ionophoresis. London: Academic, 1981.
 620. Quik, M., L. L. Iversen, and S. R. Bloom. Effect of vasoactive intestinal peptide (VIP) and other peptides on cAMP accumulation in rat brain. Biochem. Pharmacol. 27: 2209–2213, 1978.
 621. Rall, W. Theory of physiological properties of dendrites. Ann. NY Acad. Sci. 96: 1071–1092, 1962.
 622. Ramón y Cajal, S. The Structure of Ammon's Horn, translated by L. M. Kraft. Springfield, IL: Thomas, 1968.
 623. Ramsay, D. J., L. C. Keil, M. C. Sharpe, and J. Shinsako. Angiotensin II infusion increases vasopressin, ACTH, and 11‐hydroxycorticosteroid secretion. Am. J. Physiol. 234 (Regulatory Integrative Comp. Physiol. 3): R66–R71, 1978.
 624. Ranck, J. B., Jr. Studies on single neurons in dorsal hippocampal formation and septum in unrestrained rats. I. Behavioral correlates and firing repertoires. Exp. Neurol. 41: 461–531, 1973.
 625. Randic, M., and V. Miletic. Effect of substance P in cat dorsal horn neurones activated by noxious stimuli. Brain Res. 128: 164–169, 1977.
 626. Randic, M., and V. Miletic. Depressant actions of methionine‐enkephalin and somatostatin in cat dorsal horn neurones activated by noxious stimuli. Brain Res. 152: 196–202, 1978.
 627. Rapp, P. E., and M. J. Berridge. Oscillations in calicum‐cycle AMP control loops from the basis of pacemaker activity and other high frequency biological rhythms. J. Theor. Biol. 66: 497–525, 1977.
 628. Reddington, M., K. S. Lee, and P. Schubert. An A1‐adenosine receptor, characterized by [3H]cyclohexyladenosine binding, mediates the depression of evoked potentials in a rat hippocampal slice preparation. Neurosci. Lett. 28: 275–279, 1982.
 629. Rehfeld, J. F. Immunochemical studies on cholecystokinin. II. Distribution and molecular heterogeneity in the central nervous system and small intestine of man and hog. J. Biol. Chem. 253: 4022–4030, 1978.
 630. Rehfeld, J. F., N. Goltermann, L.‐I. Larsson, P. M. Emson, and C. M. Lee. Gastrin and cholecystokinin in central and peripheral neurons. Federation Proc. 38: 2325–2329, 1979.
 631. Reichlin, S. Peptides in neuroendocrine regulation. In: Peptides: Integrators of Cell and Tissue Function, edited by F. E. Bloom. New York: Raven, 1980, p. 235–250. (Soc. Gen. Physiologists Ser., vol. 35.)
 632. Renaud, L. P., H. W. Blume, Q. J. Pittman, Y. Lamour, and A. T. Tan. Thyrotropin‐releasing hormone selectivity depresses glutamate excitation of cerebral cortical neurons. Science 205: 1275–1277, 1979.
 633. Renaud, L. P., J. B. Martin, and P. Brazeau. Depressant action of TRH, LH‐RH and somatostatin on activity of central neurones. Nature London 255: 233–235, 1975.
 634. Ribeiro, J. A. ATP; related nucleotides and adenosine on neurotransmission. Life Sci. 22: 1373–1380, 1978.
 635. Rivier, J., J. Spiess, M. Thorner, and W. Vale. Characterisation of a growth hormone‐releasing factor from a human pancreatic islet tumour. Nature London 300: 276–278, 1982.
 636. Rivier, J., J. Spiess, and W. Vale. Characterization of rat hypothalamic corticotropin‐releasing factor. Proc. Natl. Acad. Sci. USA 80: 4851–4855, 1983.
 637. Roberts, E. New directions in GABA research. I. Immunocytochemical studies of GABA neurons. In: GABA‐Neurotransmitters, edited by P. Krogsgaard‐Larsen, J. Scheel‐Krüger, and H. Kofod. New York: Academic, 1979, p. 28–45.
 638. Roberts, E. Roles of GABA in neurons in information processing in the vertebrate CNS. In: Neuronal Information Transfer, edited by A. Karlin, V. M. Tennyson, and H. J. Vogel. New York: Academic, 1978, p. 213–239.
 639. Roberts, M. H. T., and D. M. Wright. Functional identification of units in the rat dorsal horn responding to a Substance P analogue. J. Physiol. London 281: 33–34, 1978.
 640. Robinson, J. H., and S. A. Deadwyler. Kainic acid produces depolarisation of CA3 pyramidal cells in the in vitro hippocampal slice. Brain Res. 221: 117–127, 1981.
 641. Robinson, J. H., and S. A. Deadwyler. Intracellular correlates of morphine excitation in the hippocampal slice preparation. Brain Res. 224: 375–387, 1981.
 642. Robison, G. A., R. W. Butcher, and E. W. Sutherland. Cyclic AMP. New York: Academic, 1971.
 643. Rodan, G. A., L. A. Bourrel, and L. A. Norton. DNA synthesis in cartilage cells is stimulated by oscillating electric fields. Science 199: 690–692, 1978.
 644. Rogawski, M. A. Cholecystokinin octapeptide: effects on the excitability of cultured spinal neurons. Peptides 3: 545–551, 1982.
 645. Rogawski, M. A., and G. K. Aghajanian. Activation of lateral geniculate neurons by norepinephrine: mediation by an α‐adrenergic receptor. Brain Res. 182: 345–359, 1980.
 646. Rogawski, M. A., and G. K. Aghajanian. Modulation of lateral geniculate neurone excitability by noradrenaline microionotophoresis or locus coeruleus stimulation. Nature London 287: 731–734, 1980.
 647. Rogawski, M. A., and G. K. Aghajanian. Norepinephrine and serotonin: opposite effects on the activity of lateral geniculate neurons evoked by optic pathway stimulation. Exp. Neurol. 69: 678–694, 1980.
 648. Rogawski, M. A., and J. L. Barker (editors). Neurotransmitter Actions in the Vertebrate Nervous System. New York: Plenum, in press.
 649. Rose, A. M., T. Hattori, and H. C. Fibiger. Analysis of the septo‐hippocampal pathway by light and electron microscopic autoradiography. Brain Res. 108: 170–174, 1976.
 650. Rose, G., K. Pang, M. Palmer, and R. Freedman. Differential effects of phencyclidine upon hippocampal complex‐spike and theta neurons. Neurosci. Lett. 45: 141–146, 1984.
 651. Rose, G., and P. Schubert. Release and transfer of [3H] adenosine derivatives in the cholinergic septal system. Brain Res. 121: 353–357, 1977.
 652. Rosenfeld, M. G., J.‐J. Mermod, S. G. Amara, L. W. Swanson, P. E. Sawchenko, J. Rivier, W. W. Vale, and R. M. Evans. Production of a novel neuropeptide encoded by the calcitonin gene via tissue‐specific RNA processing. Nature London 304: 129–135, 1983.
 653. Rudomin, P. Presynaptic control of information transmission in the vertebrate spinal cord. In: Regulatory Mechanisms of Synaptic Transmission, edited by R. Tapia and C. W. Cotman. New York: Plenum, 1981, p. 297–330.
 654. Ryall, R. W. Amino acid receptors in CNS. I. GABA and glycine in spinal cord. In: Handbook of Psychopharmacology. Amino Acid Neurotransmitters, edited by L. L. Iversen, S. D. Iversen, and S. H. Snyder. New York: Plenum, 1975, vol. 4, p. 83–128.
 655. Ryall, R. W. Modulation of cholinergic transmission by substance P. In: Substance P in the Nervous System, edited by G. Lawrenson. London: Pitman, 1982, p. 267–280. (Ciba Found. Symp. No. 91.)
 656. Ryan, G. P., J. C. Hackman, C. J. Wohlberg, and R. A. Davidoff. Catecholamine effects on frog dorsal root terminals. Neurosci. Lett. 36: 63–68, 1983.
 657. Saffran, M., and A. V. Schally. The release of corticotropin by anterior pituitary tissue in vitro. Can. J. Biochem. Physiol. 33: 408–415, 1955.
 658. Saffran, M., and A. V. Schally. Corticotropin‐releasing factor: isolation and chemical properties. Ann. NY Acad. Sci. 297: 395–404, 1977.
 659. Said, S. I. Vasoactive intestinal polypeptide (VIP): isolation, distribution, biological actions, structure‐function relations and possible functions. In: Gastrointestinal Hormones, edited by G. B. Jerzy Glass. New York: Raven, 1980, p. 245–273.
 660. Said, S. I., and V. Mutt. Isolation from porcine intestinal wall of a vasoactive octacosapeptide related to secretin and to glucagon. Eur. J. Biochem. 28: 199–204, 1972.
 661. Sakai, M., H. Sakai, and C. D. Woody. Intracellular staining of cortical neurons by pressure microinjection of horseradish peroxidase and recovery by core biopsy. Exp. Neurol. 58: 138–144, 1978.
 662. Sakmann, B., and E. Neher (editors). Single‐Channel Recording. New York: Plenum, 1983.
 663. Salmoiraghi, G. C., and F. Weight. Micromethods in neuropharmacology: an approach to the study of anesthetics. Anesthesiology 28: 54–64, 1967.
 664. Salt, T. E., G. J. De Vries, R. E. Rodriguez, P. M. B. Cahusac, R. Morris, and R. G. Hill. Evaluation of (D‐Pro2, D‐Trp7,9)‐substance P as an antagonist of substance P responses in the rat central nervous system. Neurosci. Lett. 30: 291–295, 1982.
 665. Salt, T. E., and R. G. Hill. Vasoactive intestinal polypeptide (VIP) applied by microiontophoresis excites single neurones in the trigeminal nucleus caudalis of the rat. Neuropeptides 1: 403–408, 1981.
 666. Salt, T. E., and R. G. Hill. The effects of C‐terminal fragments of cholecystokinin on the firing of single neurones in the caudal trigeminal nucleus of the rat. Neuropeptides 2: 301–306, 1982.
 667. Salt, T. E., and R. G. Hill. Excitation of single sensory neurones in the rat caudal trigeminal nucleus by iontophoretically applied adenosine 5′‐triphosphate. Neurosci. Lett. 35: 53–57, 1983.
 668. Sandoval, M. E., and C. W. Cotman. Evaluation of glutamate as a neurotransmitter of cerebellar parallel fibers. Neuroscience 3: 199–206, 1978.
 669. Sastry, B. R. Morphine and met‐enkephalin effects on sural A‐delta afferent terminal excitability. Eur. J. Pharmacol. 50: 269–273, 1978.
 670. Sastry, B. R. Substance P effects on spinal nociceptive neurones. Life Sci. 24: 2169–2177, 1979.
 671. Satoh, M., A. Akaike, and H. Takagi. Excitation by morphine and enkephalin of single neurons of nucleus reticularis paragigantocellularis in the rat: a probable mechanism of analgesic action of opioids. Brain Res. 169: 406–410, 1979.
 672. Sattin, A., and T. W. Rall. The effect of adenosine and adenine nucleotides on the cyclic AMP content of guinea pig cerebral cortex slices. Mol. Pharmacol. 6: 13–23, 1970.
 673. Sawada, S., S. Takada, and C. Yamamoto. Electrical activity recorded from thin sections of the bed nucleus of the stria terminalis, and the effects of neurotensin. Brain Res. 188: 578–581, 1980.
 674. Sawchenko, P. E., L. W. Swanson, and W. W. Vale. Co‐expression of corticotropin‐releasing factor and vasopressin immunoreactivity in parvocellular neurosecretory neurons of the adrenalectomized rat. Proc. Natl. Acad. Sci. USA 81: 1883–1887, 1984.
 675. Sawyer, W. H., Z. Grzonka, and M. Manning. Neurohypophysial peptides. Design of tissue specific agonists and antagonists. Mol. Cell. Endocr. 22: 117–134, 1981.
 676. Scatton, B., H. Simon, M. Le Moal, and S. Bischoff. Origin of dopaminergic innervation of the rat hippocampal formation. Neurosci. Lett. 18: 125–131, 1980.
 677. Scholfield, C. N. Depression of evoked potentials in brain slices by adenosine compounds. Br. J. Pharmacol. 63: 239–244, 1978.
 678. Schon, F., and L. L. Iversen. The use of autoradiographic techniques for the identification and mapping of transmitter‐specific neurones in the brain. Life Sci. 15: 157–175, 1974.
 679. Schubert, P., W. Komp, and G. W. Kreutzberg. Correlation of 5′‐nucleotidase activity and selective transneuronal transfer of adenosine in the hippocampus. Brain Res. 168: 419–424, 1979.
 680. Schubert, P., K. Lee, and G. W. Kreutzberg. Neuronal release of adenosine derivatives and modulation of signal processing in the CNS. In: Progress in Brain Research. Chemical Transmission in the Brain, edited by R. M. Buijs, P. Pevet, and D. F. Swaab. Amsterdam: Elsevier, 1982, vol. 55, p. 225–238.
 681. Schubert, P., K. Lee, M. West, S. Deadwyler, and G. Lynch. Stimulation‐dependent release of 3H‐adenosine derivatives from central axon terminals to target neurones. Nature London 260: 541–542, 1976.
 682. Schubert, P., and U. Mitzdorf. Analysis and quantitative evaluation of the depressive effect of adenosine on evoked potentials in hippocampal slices. Brain Res. 172: 186–190, 1979.
 683. Schubert, P., M. Reddington, and G. W. Kreutzberg. On the possible role of adenosine as a modulatory messenger in the hippocampus and other regions of the CNS. In: Progress in Brain Research. Development of Chemical Specificity of Neurons, edited by M. Cuénod. Amsterdam: Elsevier, 1979, vol. 51, p. 149–165.
 684. Schwartzkroin, P. A., and M. Slawsky. Probable calcium spikes in hippocampal neurons. Brain Res. 135: 157–161, 1977.
 685. Schwartzkroin, P. A., and C. E. Stafstrom. Effects of EGTA on the calcium‐activated after hyperpolarization in hippocampal CA3 pyramidal cells. Science 210: 1125–1126, 1980.
 686. Segal, M. Lithium and the monoamine neurotransmitters in the rat hippocampus. Nature London 250: 71–73, 1974.
 687. Segal, M. Morphine and enkephalin interactions with putative neurotransmitters in rat hippocampus. Neuropharmacology 16: 587–592, 1977.
 688. Segal, M. The acetylcholine receptor in the rat hippocampus: nicotinic, muscarinic or both? Neuropharmacology 17: 619–623, 1978.
 689. Segal, M. The action of serotonin in the rat hippocampal slice preparation. J. Physiol. London 303: 423–439, 1980.
 690. Segal, M. Histamine produces a Ca2+‐sensitive depolarization of hippocampal pyramidal cells in vitro. Neurosci. Lett. 19: 67–71, 1980.
 691. Segal, M. The action of norepinephrine in the rat hippocampus: intracellular studies in the slice preparation. Brain Res. 206: 107–128, 1981.
 692. Segal, M. Intracellular analysis of a postsynaptic action of adenosine in the rat hippocampus. Eur. J. Pharmacol. 79: 193–199, 1982.
 693. Segal, M. Multiple actions of acetylcholine at a muscarinic receptor studied in the rat hippocampal slice. Brain Res. 246: 77–87, 1982.
 694. Segal, M. Norepinephrine modulates reactivity of hippocampal cells to chemical stimulation in vitro. Exp. Neurol. 77: 86–93, 1982.
 695. Segal, M., and J. L. Barker. Rat hippocampal neurons in culture: properties of GABA‐activated Cl− ion conductance. J. Neurophysiol. 51: 500–515, 1984.
 696. Segal, M., and J. L. Barker. Rat hippocampal neurons in culture: voltage‐clamp analysis of inhibitory synaptic connections. J. Neurophysiol. 52: 469–487, 1984.
 697. Segal, M., and F. E. Bloom. The action of norepinephrine in the rat hippocampus. I. Ionotophoretic studies. Brain Res. 72: 79–97, 1974.
 698. Segal, M., and F. E. Bloom. The action of norepinephrine in the rat hippocampus. II. Activation of the input pathway. Brain Res. 72: 99–114, 1974.
 699. Segal, M., and F. E. Bloom. The action of norepinephrine in the rat hippocampus. III. Hippocampal cellular responses to locus coeruleus stimulation in the awake rat. Brain Res. 107: 499–511, 1976.
 700. Segal, M., and F. E. Bloom. The action of norepinephrine in the rat hippocampus. IV. The effects of locus coeruleus stimulation on evoked hippocampal unit activity. Brain Res. 107: 513–525, 1976.
 701. Segal, M., V. Greenberger, and R. Hofstein. Cyclic AMP‐generating systems in rat hippocampal slices. Brain Res. 213: 351–364, 1981.
 702. Segal, M., and S. Landis. Afferents to the hippocampus of the rat studied with the method of retrograde transport of horseradish peroxidase. Brain Res. 78: 1–15, 1974.
 703. Segal, M., D. B. Sagie, and A. Mayevsky. Metabolic changes induced in rat hippocampal slices by norepinephrine. Brain Res. 202: 387–399, 1980.
 704. Sejnowski, T. J. Peptidergic synaptic transmission in sympathetic ganglia. Federation Proc. 41: 2923–2928, 1982.
 705. Shain, W., and D. O. Carpenter. Mechanisms of synaptic modulation. Int. Rev. Neurobiol. 22: 205–250, 1981.
 706. Shepherd, G. M. Neurobiology. New York: Oxford Univ. Press, 1983, p. 120–139.
 707. Shute, C. C. D., and P. R. Lewis. Cholinesterase‐containing systems of the brain of the rat. Nature London 199: 1160–1164, 1963.
 708. Siegelbaum, S. A., and Tsien, R. W. Modulation of gated ion channels as a mode of transmitter action. Trends Neurosci. 6: 307–313, 1983.
 709. Siggins, G. R. The electrophysiological effects of cyclic nucleotides on excitable tissue. In: Cyclic Nucleotides: Mechanisms of Action, edited by H. Cramer and J. Schultz. London: Wiley, 1977, p. 317–336.
 710. Siggins, G. R. Electrophysiological assessment of mononucleotides and nucleosides as first and second messengers in the nervous system. In: Neuronal Information Transfer, edited by A. Karlin, V. M. Tennyson, and H. J. Vogel. New York: Academic, 1978, p. 339–355.
 711. Siggins, G. R. Electrophysiological role of dopamine in striatum: excitatory or inhibitory? In: Psychopharmacology: A Generation of Progress, edited by M. A. Lipton, A. DiMascio, and K. F. Killam. New York: Raven, 1978, p. 143–157.
 712. Siggins, G. R. Catecholamines and endorphins as neurotransmitters and neuromodulators. In: Regulatory Mechanisms of Synaptic Transmission, edited by R. Tapia and C. W. Cotman. New York: Plenum, 1981, p. 1–27.
 713. Siggins, G. R. Cyclic nucleotides: regulation of cellular excitability. In: Handbook of Experimental Pharmacology. Cyclic Nucleotides, edited by J. A. Nathanson and J. W. Kebabian. Berlin: Springer‐Verlag, 1982, vol. 58, pt. 2, p. 305–346.
 714. Siggins, G. R., and F. E. Bloom. Modulation of unit activity by chemically coded neurons. In: Brain Mechanisms of Perceptual Awareness and Purposeful Behavior, edited by O. Pompeiano and C. Ajmone‐Marsan, 1981, New York: Raven, p. 431–448.
 715. Siggins, G. R., and E. D. French. Pro‐somatostatin‐related peptides alter the discharge rate of rat cortical and hippocampal neurons in vivo: an iontophoretic study. Soc. Neurosci. Abstr. 10: 810 1984.
 716. Siggins, G. R., and D. L. Gruol. Norepinephrine (NE) and isoproterenol (IP) do not consistently hyperpolarize CA1 pyramidal neurons (PNs) of the rat hippocampus slice in vitro. Soc. Neurosci. Abstr. 7: 725 1981.
 717. Siggins, G. R., D. Gruol, J. Aldenhoff, and Q. Pittman. Electrophysiological actions of corticotropin‐releasing factor in the central nervous system. Federation Proc. 44: 237–242, 1985.
 718. Siggins, G. R., D. Gruol, A. Padjen, and D. Forman. Purine and pyrimidine mononucleotides depolarise neurones of explanted amphibian sympathetic ganglia. Nature London 270: 263–265, 1977.
 719. Siggins, G., D. Gruol, A. Padjen, and D. Forman. Action of purine and pyrimidine mononucleotides on central and cultured sympathetic neurons. In: Iontophoresis and Transmitter Mechanisms in the Mammalian Central Nervous System, edited by R. W. Ryall and J. S. Kelly. Amsterdam: Elsevier/North‐Holland Biomed., 1978, p. 453–455.
 720. Siggins, G. R., and S. J. Henriksen. Analogues of cyclic adenosine monophosphate: correlation of inhibition of Purkinje neurons with protein kinase activation. Science 189: 559–561, 1975.
 721. Siggins, G. R., S. J. Henriksen, and F. E. Bloom. Iontophoresis of Li+ antagonizes noradrenergic synaptic inhibition of rat cerebellar Purkinje cells. Proc. Natl. Acad. Sci. USA 76: 3015–3018, 1979.
 722. Siggins, G. R., S. J. Henriksen, and S. C. Landis. Electrophysiology of Purkinje neurons in the weaver mouse: iontophoresis of neurotransmitters and cyclic nucleotides, and stimulation of the nucleus locus coeruleus. Brain Res. 114: 53–69, 1976.
 723. Siggins, G. R., B. J. Hoffer, and F. E. Bloom. Studies on norepinephrine‐containing afferents to Purkinje cells of rat cerebellum. III. Evidence for mediation of norepinephrine effects by cyclic 3′,5′‐adenosine monophosphate. Brain Res. 25: 535–553, 1971.
 724. Siggins, G. R., B. J. Hoffer, F. E. Bloom, and U. Ungerstedt. Cytochemical and electrophysiological studies of dopamine in the caudate nucleus. In: The Basal Ganglia, edited by M. D. Yahr. New York: Raven, 1976, p. 227–248.
 725. Siggins, G. R., B. J. Hoffer, A. P. Oliver, and F. E. Bloom. Activation of a central noradrenergic projection to cerebellum. Nature London 233: 481–483, 1971.
 726. Siggins, G. R., B. J. Hoffer, and U. Ungerstedt. Electrophysiological evidence for involvement of cyclic adenosine monophosphate in dopamine responses of caudate neurons. Life Sci. 15: 779–792, 1974.
 727. Siggins, G. R., J. F. McGinty, J. H. Morrison, Q. J. Pittman, W. Zieglgänsberger, P. J. Magistretti, and D. L. Gruol. The role of neuropeptides in the hippocampal formation. In: Regulatory Peptides: From Molecular Biology to Function, edited by E. Costa and M. Trabucchi. New York: Raven, 1982, p. 413–422.
 728. Siggins, G. R., A. P. Oliver, B. J. Hoffer, and F. E. Bloom. Cyclic adenosine monophosphate and norepinephrine: effects on transmembrane properties of cerebellar Purkinje cells. Science 192–194, 1971.
 729. Siggins, G. R., and P. Schubert. Adenosine depression of hippocampal neurons in vitro: an intracellular study of dose‐dependent actions on synaptic and membrane potentials. Neurosci. Lett. 23: 55–60, 1981.
 730. Siggins, G. R., and, W. Zieglgänsberger. Morphine and opioid peptides reduce inhibitory synaptic potentials in hippocampal pyramidal cells in vitro without alteration of membrane potential. Proc. Natl. Acad. Sci. USA 78: 5235–5239, 1981.
 731. Simonnet, G., B. Bioulac, F. Rodriguez, and J. D. Vincent. Evidence of a direct action of angiotensin II on neurones in the septum and in the medial preoptic area. Pharmacol. Biochem. Behav. 13: 359–363, 1980.
 732. Simonnet, G., and M. F. Giorguieff‐Chesselet. Stimulating effect of angiotensin II on the spontaneous release of newly synthetized [3H]dopamine in rat striatal slices. Neurosci. Lett. 15: 153–158, 1979.
 733. Simonnet, G., M. F. Giorguieff‐Chesselet, A. Carayon, B. Bioulac, F. Cesselin, J. Glowinski, and J. D. Vincent. Angiotensine II et systéme nigro‐néostriatal. J. Physiol. Paris 77: 71–79, 1981.
 734. Simonnet, G., F. Rodriguez, F. Fumoux, P. Czernichow, and J. D. Vincent. Vasopressin release and drinking induced by intracranial injection of angiotensin II in monkey. Am. J. Physiol. 237 ( Regulatory Integrative Comp. Physiol. 6): R20–R25, 1979.
 735. Sirett, N. E., A. S. McLean, J. J. Bray, and J. I. Hubbard. Distribution of angiotensin II receptors in rat brain. Brain Res. 122: 299–312, 1977.
 736. Skirboll, L. R., A. A. Grace, D. W. Hommer, J. Rehfeld, M. Goldstein, T. Hökfelt, and B. S. Bunney. Peptide‐monoamine coexistence: studies of the actions of cholecystokinin‐like peptide on the electrical activity of midbrain dopamine neurons. Neuroscience 6: 2111–2124, 1981.
 737. Smith, C. M. The release of acetylcholine from rabbit hippocampus. Br. J. Pharmacol. 45: 172P, 1972.
 738. Smith, T. G., Jr., J. L. Barker, B. M. Smith, and T. R. Colburn. Voltage clamp techniques applied to cultured skeletal muscle and spinal neurons. In: Excitable Cells in Tissue Culture, edited by P. G. Nelson and M. Lieberman. New York: Plenum, 1981, p. 111–136.
 739. Snodgrass, S. R., W. F. White, B. Biales, and M. Dichter. Biochemical correlates of GABA function in rat cortical neurons in culture. Brain Res. 190: 123–138, 1980.
 740. Sofroniew, M. V. Projections from vasopressin, oxytocin, and neurophysin neurons to neural targets in the rat and human. J. Histochem. Cytochem. 28: 475–478, 1980.
 741. Sofroniew, M. V., and A. Weindl. Central nervous system distribution of vasopressin, oxytocin and neurophysin. In: Endogenous Peptides and Learning and Memory Processes, edited by J. L. Martinez, Jr., R. A. Jensen, R. B. Messing, H. Rigter, and J. L. McGangh. New York: Academic, 1981, p. 327–369.
 742. Sokoloff, L., M. Reivich, C. Kennedy, M. H. Des Rosiers, C. S. Patlak, K. D. Pettigrew, O. Sakurada, and M. Shinohara. The [14C]deoxyglucose method for the measurement of local cerebral glucose utilization: theory, procedure and normal values in the conscious and anesthetized albino rat. J. Neurochem. 28: 897–916, 1977.
 743. Somogyi, P., A. D. Smith, M. G. Nunzi, A. Gorio, H. Takagi, and J. Y. Wu. Glutamate decarboxylase immunoreactivity in the hippocampus of the cat: distribution of immunoreactive synaptic terminals with special reference to the axon initial segment of pyramidal neurons. J. Neurosci. 3: 1450–1468, 1983.
 744. Stallcup, W. B., and J. Patrick. Substance P enhances cholinergic receptor desensitization in a clonal nerve cell line. Proc. Natl. Acad. Sci. USA 77: 634–638, 1980.
 745. Stämpfli, R. A new method for measuring membrane potentials with external electrodes. Experientia 10: 508–509, 1954.
 746. Stanzione, P., and, W. Zieglgänsberger. Action of neurotensin on spinal cord neurons in the rat. Brain Res. 268: 111–118, 1983.
 747. Starke, K., H. D. Taube, and E. Borowski. Presynaptic receptor systems in catecholaminergic transmission. Biochem. Pharmacol. 26: 259–268, 1977.
 748. Steinbach, J. H. Activation of nicotinic acetylcholine receptors. In: The Cell Surface and Neuronal Function, edited by C. W. Cotman, G. Poste, and G. L. Nicolson. New York: North‐Holland, 1980, p. 119–156.
 749. Stengaard‐Pedersen, K., and L.‐I. Larsson. Comparative immunocytochemical localization of putative opioid ligand in the central nervous system. Histochemistry 73: 89–114, 1981.
 750. Stengaard‐Pedersen, K., and L.‐I. Larsson. Localization and opiate receptor binding of enkephalin, CCK and ACTH/β‐endorphin in the rat central nervous system. Peptides 2, Suppl. 1: 3–19, 1981.
 751. Stevens, C. F. Study of membrane permeability changes by fluctuation analysis. Nature London 270: 391–396, 1977.
 752. Stevens, C. F. Ionic channels in neuromembranes: methods for studying their properties. In: Molluscan Nerve Cells: From Biophysics to Behavior, edited by J. Koester and J. H. Byrne. New York: Cold Spring Harbor, 1980, p. 11–31.
 753. Stone, T. W. Glutamate as the neurotransmitter of cerebellar granule cells in the rat: electrophysiological evidence. Br. J. Pharmacol. 616: 291–296, 1979.
 754. Stone, T. W. Physiological roles for adenosine and adenosine 5′‐triphosphate in the nervous system. Neuroscience 6: 523–555, 1981.
 755. Stone, T. W., and M. N. Perkins. Adenine nucleotides effects on rat cortical neurones. Brain Res. 229: 241–245, 1981.
 756. Stone, T. W., and D. A. Taylor. Microiontophoretic studies of the effects of cyclic nucleotides on excitability of neurones in the rat cerebral cortex. J. Physiol. London 266: 523–543, 1977.
 757. Storm‐Mathisen, J. Distribution of the components of the GABA system in neuronal tissue: cerebellum and hippocampus—effects of axotomy. In: GABA in Nervous System Function, edited by E. Roberts, T. N. Chase, and D. B. Tower. New York: Raven, 1976, p. 149–168.
 758. Storm‐Mathisen, J. Localization of transmitter candidates in the brain: the hippocampal formation as a model. Prog. Neurobiol. 8: 119–181, 1977.
 759. Storm‐Mathisen, J., A. K. Leknes, A. T. Bore, J. L. P. Vaaland, Edminson, F.‐M., S. Haug, and O. P. Ottersen. First visualization of glutamate and GABA in neurones by immunocytochemistry. Nature London 301: 517–520, 1983.
 760. Study, R. E., and J. L. Barker. Diazepam and (—)‐pentobarbital: fluctuation analysis reveals different mechanisms for potentiation of γ‐aminobutyric acid responses in cultured central neurons. Proc. Natl. Acad. Sci. USA 78: 7180–7184, 1981.
 761. Su, C., J. A. Bevan, and G. Burnstock. [3H]adenosine triphosphate: release during stimulation of enteric nerves. Science 173: 337–339, 1971.
 762. Sulakhe, P. V., and J. W. Phillis. The release of 3H‐adenosine and its derivatives from cat sensorimotor cortex. Life Sci. 17: 551–556, 1975.
 763. Sutcliffe, J. G., R. J. Milner, T. M. Shinnick, and F. E. Bloom. Identifying the protein products of brain‐specific genes with antibodies to chemically synthesized peptides. Cell 33: 671–682, 1983.
 764. Sutherland, E. W., T. W. Rall, and T. Menon. Adenyl cyclase. I. Distribution, preparation and properties. J. Biol. Chem. 237: 1220–1227, 1962.
 765. Sutton, R. E., G. F. Koob, M. Le Moal, J. Rivier, and W. Vale. Corticotropin relaxing factor produces behavioral action in rats. Nature London 297: 331–333, 1982.
 766. Suzue, T., and T. Jessell. Opiate analgesics and endorphins inhibit rat dorsal root potential in vitro. Neurosci. Lett. 16: 161–166, 1980.
 767. Suzue, T., N. Yanaihara, and M. Otsuka. Actions of vasopressin, gastrin releasing peptide and other peptides on neurons of newborn rat spinal cord in vitro. Neurosci. Lett. 26: 137–142, 1981.
 768. Swanson, L. W., and B. K. Hartman. The central adrenergic system. An immunofluorescence study of the location of cell bodies and their efferent connections in the rat utilizing dopamine‐β‐hydroxylase as a marker. J. Comp. Neurol. 163: 467–506, 1975.
 769. Swanson, L. W., P. E. Sawchenko, J. Rivier, and W. W. Vale. Organization of ovine corticotropin‐releasing factor immunoreactive cells and fibers in the rat brain: an immunohistochemical study. Neuroendocrinology 36: 165–186, 1983.
 770. Swanson, L. W., T. J. Teyler, and R. F. Thompson. Neurosciences Research Program Bulletin. Cambridge, MA: MIT Press, 1982, vol. 20.
 771. Takeuchi, A. Junctional transmission. I. Postsynaptic mechanisms. In: Handbook of Physiology. The Nervous System, edited by J. M. Brookhart and V. B. Mountcastle. Bethesda, MD: Am. Physiol. Soc., 1977, sect. 1, vol. I, pt. 1, chapt. 9, p. 295–328.
 772. Tanaka, C., C. Inagaki, and H. Fujiwara. Labeled noradrenaline release from rat cerebral cortex following electrical stimulation of locus coeruleus. Brain Res. 106: 384–389, 1976.
 773. Tanaka, S., and A. Tsujimoto. Somatostatin facilitates the serotonin release from rat cerebral cortex, hippocampus and hypothalamus slices. Brain Res. 208: 219–222, 1981.
 774. Tatemoto, K., M. Carlquist, and V. Mutt. Neuropeptide Y—a novel brain peptide with structural similarities to peptide YY and pancreatic polypeptide. Nature London 296: 659–660, 1982.
 775. Taxi, T., and, J. Storm‐Mathisen. Uptake of d‐aspartate and l‐glutamate in excitatory axon terminals in hippocampus: autoradiographic and biochemical comparison with γ‐amino‐butyrate and other amino acids in normal rats and in rats with lesions. Neuroscience 11: 79–100, 1984.
 776. Taylor, D. A., and T. W. Stone. The action of adenosine on noradrenergic neuronal inhibition induced by stimulation of locus coeruleus. Brain Res. 183: 367–376, 1980.
 777. Teyler, T. J. Brain slice preparation: hippocampus. Brain Res. Bull. 5: 391–403, 1980.
 778. Thalmann, R. H. Reversal properties of an EGTA‐resistant late hyperpolarization that follows synaptic stimulation of hippocampal neurons. Neurosci. Lett. 46: 103–108, 1984.
 779. Thalmann, R. H., and G. F. Ayala. A late increase in potassium conductance follows synaptic stimulation of granule neurons of the dentate gyrus. Neurosci. Lett. 29: 243–248, 1982.
 780. Thalmann, R. H., E. J. Peck, and G. F. Ayala. Biphasic response of hippocampal pyramidal neurons to GABA. Neurosci. Lett. 21: 319–324, 1981.
 781. Thomas, R. C. Electrogenic sodium pump in nerve and muscle cells. Physiol. Rev. 52: 563–594, 1972.
 782. Tielen, A. M., F. H. Lopes da Silva, W. J. Mollevanger, and F. H. de Jonge. Differential effects of enkephalin within hippocampal areas. Exp. Brain Res. 44: 343–346, 1981.
 783. Timmermans, P. B., and P. A. van Zwieten. α2‐Adrenoreceptors: classification, localization, mechanisms, and targets for drugs. J. Med. Chem. 25: 1389–1401, 1982.
 784. Tokimasa, T., K. Morita, and A. North. Opiates and clonidine prolong calcium‐dependent after‐hyperpolarisations. Nature London 294: 162–163, 1981.
 785. Tsujimoto, A., and T. Shokichi. Stimulatory effect of somatostatin on norepinephrine release from rat brain cortex slices. Life Sci. 28: 903–910, 1981.
 786. Tsunoo, A., S. Konishi, and M. Otsuka. Substance P as an excitatory transmitter of primary afferent neurons in guinea‐pig sympathetic ganglia. Neuroscience 7: 2025–2037, 1982.
 787. Ulbricht, W. Ionic channels and gating currents in excitable membranes. Annu. Rev. Biophys. Bioeng. 6: 7–31, 1977.
 788. Ungerstedt, U. Stereotaxic mapping of the monoamine pathways in the rat brain. Acta Physiol. Scand. Suppl. 367: 1–48, 1971.
 789. U'Pritchard, D. C., T. D. Reisine, S. T. Masón, H. C. Fibiger, and H. I. Yamamura. Modulation of rat brain α‐ and β‐adrenergic receptor populations by lesion of the dorsal noradrenergic bundle. Brain Res. 187: 143–154, 1980.
 790. U'Prichard, D. C., and S. H. Snyder. Distinct α‐noradrenergic receptors differentiated by binding and physiological relationships. Life Sci. 24: 79–88, 1979.
 791. Urca, G., H. Frenk, J. C. Liebeskind, and A. N. Taylor. Morphine and enkephalin: analgesic and epileptic properties. Science 197: 83–86, 1977.
 792. Vale, W., J. Spiess, C. Rivier, and J. Rivier. Characterization of a 41‐residue ovine hypothalamic peptide that stimulates secretion of corticotropin and β‐endorphin. Science 213: 1394–1397, 1981.
 793. Valentino, R. J., and R. Dingledine. Presynaptic inhibitory effect of acetylcholine in the hippocampus. J. Neurosci. 1: 784–792, 1981.
 794. Valentino, R. J., and R. Dingledine. Pharmacological characterization of opioid effects in the rat hippocampal slice. J. Pharmacol. Exp. Ther. 223: 502–509, 1982.
 795. Valentino, R. J., S. L. Foote, and, G. Aston‐Jones. Corticotropin‐releasing factor activates noradrenergic neurons of the locus coeruleus. Brain Res. 270: 363–367, 1983.
 796. Van Calker, D., M. Muller, and B. Hamprecht. Adenosine inhibits the accumulation of cyclic AMP in cultured brain cells. Nature London 276: 839–841, 1978.
 797. Vanderhaeghen, J. J., F. Lofstra, J. Demay, and C. Gilles. Immunochemical localization of cholecystokinin‐ and gastrin‐like peptides in the brain and hypophysis of the rat. Proc. Natl. Acad. Sci. USA 77: 1190–1194, 1980.
 798. Vanderhaeghen, J. J., J. C. Signeau, and W. Gepts. New peptide in the vertebrate CNS reacting with gastrin antibodies. Nature London 257: 604–605, 1975.
 799. VanderMaelen, C. P., and G. K. Aghajanian. Intracellular studies showing modulation of facial motoneurone excitability by serotonin. Nature London 287: 346–347, 1980.
 800. Vincent, J.‐D., and J. L. Barker. Substance P: evidence for diverse roles in neuronal function from cultured mouse spinal neurons. Science 205: 1409–1412, 1979.
 801. Vincent, S. R., T. Hökfelt, I. Christensson, and L. Terenius. Dynorphin‐immunoreactive neurons in the central nervous system of the rat. Neurosci. Lett. 33: 185–190, 1982.
 802. Vincent, S. R., O. Johannson, L. Skirbolland, and, T. Hökfelt. Coexistence of somatostatin‐like and avian pancreatic polypeptide‐like immunoreactivities in striatal neurons which are selectively stained for NADPH‐diaphorase activity. Adv. Biol. Psychiatry 33: 453–462, 1982.
 803. Walker, J. M., H. C. Moises, D. H. Coy, G. Baldrighi, and H. Akil. Nonopiate effects of dynorphin and des‐Tyrdynorphin. Science 218: 1136–1138, 1982.
 804. Wall, P. D., and M. Fitzgerald. If substance P fails to fulfill the criteria as a neurotransmitter in somatosensory afferents, what might be its function? In: Substance P in the Nervous System, edited by G. Lawrenson. London: Pitman, 1982, p. 249–266. (Ciba Found. Symp. No. 91.)
 805. Walter, D. S., and D. Eccleston. Increase of noradrenaline metabolism following electrical stimulation of the locus coeruleus in the rat. J. Neurochem. 21: 281–289, 1973.
 806. Waszcak, B. L., and J. R. Walters. Dopamine modulation of the effects of γ‐aminobutyric acid on substantia nigra pars reticulata neurons. Science 220: 218–221, 1983.
 807. Watkins, J. C. Pharmacology of excitatory amino acid receptors. In: Glutamate: Transmitter in the Central Nervous System, edited by P. J. Roberts, J. Storm‐Mathisen, and G. A. R. Johnston. New York: Wiley, 1981, p. 1–24.
 808. Watkins, J. C., J. Davies, R. H. Evans, A. A. Francis, and A. W. Jones. Pharmacology of receptors for excitatory amino acids. In: Glutamate as a Neurotransmitter, edited by G. Di Chiara and G. L. Gessa. New York: Raven, 1981, p. 263–273. (Adv. Biochem. Psychopharmacol. Ser., vol. 27.)
 809. Watkins, J. C., and R. H. Evans. Excitatory amino acid transmitters. Annu. Rev. Pharmacol. Toxicol. 21: 165–204, 1981.
 810. Wayner, M. J., T. Ono, and D. Nolley. Effects of angiotensin II on central neurons. Pharmacol. Biochem. Behav. 1: 679–691, 1973.
 811. Weight, F. F. Synaptic potentials resulting from conductance decreases. In: Synaptic Transmission and Neuronal Interaction, edited by M. V. L. Bennett. New York: Raven, 1974, p. 141–152.
 812. Weight, F. F., and A. Padjen. Slow synaptic inhibition in sympathetic ganglion cells: evidence for synaptic inactivation of sodium conductance. Brain Res. 55: 219–224, 1973.
 813. Weight, F. F., and G. C. Salmoiraghi. Motoneurone depression by norepinephrine. Nature London 213: 1229–1230, 1967.
 814. Weight, F. F., and J. Votava. Slow synaptic excitation in sympathetic ganglion cells: evidence for synaptic inactivation of potassium conductance. Science 170: 755–758, 1970.
 815. Werman, R. Criteria for identification of a central nervous system transmitter. Comp. Biochem. Physiol. 18: 745–766, 1966.
 816. Werman, R., R. A. Davidoff, and M. H. Aprison. Inhibitory action of glycine on spinal neurons in the cat. J. Neurophysiol. 31: 81–95, 1968.
 817. Werz, M. A., and R. L. Macdonald. Heterogeneous sensitivity of cultured dorsal root ganglion neurones to opioid peptides selective for μ‐ and δ‐opiate receptors. Nature London 299: 730–733, 1982.
 818. Werz, M. A., and R. L. Macdonald. Opiate alkaloids antagonize postsynaptic glycine and GABA responses: correlation with convulsant action. Brain Res. 236: 107–119, 1982.
 819. Werz, M. A., and R. L. Macdonald. Opioid peptides selective for mu‐ and delta‐opiate receptors reduce calcium‐dependent action potential duration by increasing potassium conductance. Neurosci. Lett. 42: 173–178, 1983.
 820. Werz, M. A., and R. L. Macdonald. Opioid peptides with differential affinity for mu and delta receptors decrease sensory neuron calcium‐dependent action potentials. J. Pharmacol. Exp. Ther. 227: 394–402, 1983.
 821. Werz, M. A., and R. L. Macdonald. Dynorphin decreases a voltage‐dependent Ca conductance in mouse dorsal root ganglion neurons in cell culture. Neurosci. Lett. 46: 185–190, 1984.
 822. White, S. R., and R. S. Neuman. Facilitation of spinal motoneurone excitability by 5‐hydroxytryptamine and noradrenaline. Brain Res. 188: 119–127, 1980.
 823. White, W. F., M. A. Dichter, and S. R. Snodgrass. Benzodiazepine binding and interactions with the GABA receptor complex in living cultures of rat cerebral cortex. Brain Res. 215: 162–176, 1981.
 824. White, W. F., J. V. Nadler, and C. W. Cotman. The effect of acidic amino acid antagonists on synaptic transmission in the hippocampal formation in vitro. Brain Res. 164: 177–194, 1979.
 825. White, W. F., S. R. Snodgrass, and M. Dichter. Identification of GABA neurons in rat cortical cultures by GABA uptake autoradiography. Brain Res. 180: 139–152, 1980.
 826. Wikberg, J. E. S. The pharmacological classification of adrenergic α1 and α2 receptors and their mechanisms of action. Acta Physiol. Scand. Suppl. 106: 1–99, 1979.
 827. Williams, J. T., T. M. Egan, and R. A. North. Enkephalin opens potassium channels on mammalian central neurones. Nature London 299: 74–77, 1982.
 828. Williams, J., G. Henderson, and A. North. Locus coeruleus neurons. In: Brain Slices, edited by R. Dingledine. New York: Plenum, 1984, p. 297–311.
 829. Williams, J. T., and R. A. North. Inhibition of firing of myenteric neurones by somatostatin. Brain Res. 155: 165–168, 1978.
 830. Williams, J. T., and R. A. North. Vasoactive intestinal polypeptide excites neurones of the myenteric plexus. Brain Res. 175: 174–177, 1979.
 831. Williams, J., and, W. Zieglgänsberger. Mature spinal ganglion cells are not sensitive to opiate receptor mediated actions. Neurosci. Let. 21: 211–216, 1981.
 832. Wilson, C. J., P. M. Groves, and E. Fifkova. Monoaminergic synapses, including dendro‐dendritic synapses in the rat substantia nigra. Exp. Brain Res. 30: 161–174, 1977.
 833. Wilson, W. A., and, M. Monahan‐Goldner. Voltage clamping with a single microelectrode. J. Neurobiol. 6: 411–422, 1975.
 834. Wojtowicz, J. M., M. Gysen, and J. F. MacDonald. Multiple reversal potentials for responses to l‐glutamic acid. Brain Res. 213: 195–200, 1981.
 835. Wojtowicz, J. M., K. C. Marshall, and W. J. Hendelman. Electrophysiological and pharmacological studies of the inhibitory projection from the cerebellar cortex to the deep cerebellar nuclei in tissue culture. Neuroscience 3: 607–618, 1978.
 836. Wong, R. K. S., D. A. Prince, and A. I. Basbaum. Intradendritic recordings from hippocampal neurons. Proc. Natl. Acad. Sci. USA 78: 986–990, 1979.
 837. Woodward, D. J., B. J. Hoffer, and J. Altman. Physiological and pharmacological properties of Purkinje cells in rat cerebellum degranulated by postnatal X‐irradiation. J. Neurobiol. 5: 283–304, 1974.
 838. Woodward, D. J., B. J. Hoffer, G. R. Siggins, and F. E. Bloom. The ontogenetic development of synaptic junctions, synaptic activation and responsiveness to neurotransmitter substances in rat cerebellar Purkinje cells. Brain Res. 34: 73–97, 1971.
 839. Woodward, D. J., B. J. Hoffer, G. R. Siggins, and A. P. Oliver. Inhibition of Purkinje cells in the frog cerebellum. II. Evidence for GABA as the inhibitory transmitter. Brain Res. 33: 91–100, 1971.
 840. Woodward, D. J., H. C. Moises, B. D. Waterhouse, B. J. Hoffer, and R. Freedman. Modulatory actions of norepinephrine in the central nervous system. Federation Proc. 38: 2109–2116, 1979.
 841. Woody, C. D., B. E. Swartz, and E. Gruen. Effects of acetylcholine and cyclic GMP on input resistance of cortical neurons in awake cats. Brain Res. 158: 373–395, 1978.
 842. Yamamoto, C., and N. Kawai. Presynaptic action of acetylcholine in thin sections from the guinea pig dentate gyrus in vitro. Exp. Neurol. 19: 176–187, 1967.
 843. Yanagisawa, M., S. Konishi, T. Suzue, and M. Otsuka. Effects of substance P and capsaicin on isolated spinal cord of new‐born rat. In: Neuropeptides and Neural Transmission, edited by C. Ajmon‐Marsan and W. Z. Traczyk. New York: Raven, 1980, p. 43–49.
 844. Yarbrough, G. G. TRH potentiates excitatory actions of acetylcholine on cerebral cortical neurones. Nature London 263: 523–524, 1976.
 845. Yarbrough, G. G., N. Lake, and J. W. Phillis. Calcium antagonism and its effect on the inhibitory actions of biogenic amines on cerebral cortical neurones. Brain Res. 67: 77–88, 1974.
 846. York, D. H. Amine receptors in CNS. II. Dopamine. In: Handbook of Psychopharmacology. Biogenic Amine Receptors, edited by L. L. Iversen, S. D. Iversen, and S. H. Snyder. New York: Plenum, 1975, vol. 6, p. 23–61.
 847. Yoshimura, M., and R. A. North. Substantia gelatinosa neurones in vitro hyperpolarised by enkephalin. Nature London 305: 529–530, 1983.
 848. Young, A. B., M. L. Oster‐Granite, R. M. Herndon, and S. H. Snyder. Glutamic acid: selective depletion by viral‐induced granule cell loss in hamster cerebellum. Brain Res. 73: 1–13, 1974.
 849. Zaczek, R., K. Koller, R. Cotter, D. Heller, and J. T. Coyle. N‐acetylaspartylglutamate: an endogenous peptide with high affinity for a brain “glutamate” receptor. Proc. Natl. Acad. Sci. USA 80: 1116–1119, 1983.
 850. Zieglgänsberger, W. Actions of amino acids, amines and neuropeptides on target cells in the mammalian central nervous system. In: Progress in Brain Research. Chemical Transmission in the Brain, edited by R. M. Buijs, P. Pevet, and D. F. Swaab. Amsterdam: Elsevier, 1982, vol. 55, p. 297–320.
 851. Zieglgänsberger, W., and J. Bayerl. The activity by opiates in the spinal cord of cat. Brain Res. 115: 111–128, 1976.
 852. Zieglgänsberger, W., E. D. French, G. R. Siggins, and F. E. Bloom. Opioid peptides may excite hippocampal pyramidal neurones by inhibiting adjacent inhibitory interneurons. Science 205: 415–417, 1979.
 853. Zieglgänsberger, W., and J. Fry. Actions of enkephalin on cortical and striatal neurones of naive and morphine/tolerant dependent rats. In: Opiates and Endogenous Opioid Peptides, edited by H. W. Kosterlitz. Amsterdam: Elsevier/North‐Holland, 1976, p. 213–238.
 854. Zieglgänsberger, W., and C. H. Reiter. A cholinergic mechanism in the spinal cord of cats. Neuropharmacology 13: 519–527, 1974.
 855. Zieglgänsberger, W., and B. Sutor. Responses of substantia gelatinosa neurones to putative neurotransmitters in an in vitro preparation of the adult rat spinal cord. Brain Res. 279: 316–320, 1983.
 856. Zieglgänsberger, W., and I. F. Tulloch. Effects of substance P on neurones in the dorsal horn of the spinal cord of the cat. Brain Res. 166: 273–282, 1979.
 857. Zieglgänsberger, W., and I. F. Tulloch. The effects of methionine‐ and leucine‐enkephalin on spinal neurons of the cat. Brain Res. 167: 53–64, 1979.

Contact Editor

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

* Required Field

How to Cite

George R. Siggins, Donna L. Gruol. Mechanisms of Transmitter Action in the Vertebrate Central Nervous System. Compr Physiol 2011, Supplement 4: Handbook of Physiology, The Nervous System, Intrinsic Regulatory Systems of the Brain: 1-114. First published in print 1986. doi: 10.1002/cphy.cp010401