Background

The N-methyl-D-aspartate (NMDA) subtype of glutamate receptor is blocked by ketamine, and this action likely contributes to ketamine's anesthetic and analgesic properties. Previous studies suggest that ketamine occludes the open channel by binding to a site located within the channel pore. This hypothesis was examined by investigating the effects of ketamine on single-channel currents from NMDA receptors.

Methods

The cell-attached and outside-out configurations of the patch clamp technique were used to study NMDA-activated currents recorded from cultured mouse hippocampal neurons.

Results

In cell-attached patches, NMDA evoked currents that had an apparent mean open time (tau o) of 3.26 ms. The probability of at least one channel being open (Po') was 0.058. The addition of ketamine (0.1 microM or 1 microM) to the pipette solution decreased Po' to 53% and 24% of control values, respectively. At 1 microM ketamine, this reduction was due to a decrease in both the frequency of channel opening and the mean open time (44% and 68% of control values, respectively). Ketamine did not influence channel conductance and no new components were required to fit the open- or closed-duration distributions. Ketamine (50 microM), applied outside the recording pipette, reduced the opening frequency of channels recorded in the cell attached configuration. This observation suggests that ketamine gained access to a binding site by diffusing across the hydrophobic cell membrane. In outside-out patches, ketamine potency was lower than that observed in cell-attached patches: 1 microM and 10 microM ketamine reduced Po' to 63% and 34% of control values, respectively, and this reduction was due primarily to a decrease in the frequency of channel opening with little change in mean open time.

Conclusions

These observations are consistent with a model whereby ketamine inhibits the NMDA receptor by two distinct mechanisms: (1) Ketamine blocks the open channel and thereby reduces channel mean open time, and (2) ketamine decreases the frequency of channel opening by an allosteric mechanism.

Glutamate is the major excitatory neurotransmitter in the mammalian central nervous system. Based on pharmacologic sensitivity, ion permeability, and channel kinetics, at least two major classes of ionotropic glutamate receptors have been identified: the N-methyl-D-aspartate (NMDA) receptors, and the non-NMDA or alpha-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid (AMPA) kainate receptors. [1] The NMDA receptor is a unique neurotransmitter receptor in that channel activation requires the binding of both glutamate and the coagonist glycine. [2] The channel is blocked in a voltage-dependent manner by physiologic concentrations of Mg2- and undergoes calcium-dependent and calcium-independent forms of inactivation. [3–5] Receptors for NMDA are thought to play a pivotal role in various physiologic functions, including synaptic plasticity and neuronal development. In addition, NMDA receptors may contribute to pathologic processes such as seizure, ischemia-related neuronal death, and several neurodegenerative diseases. [6]

Ketamine (2-O-chlorophenyl-2-methylamino cyclohexanone) blocks NMDA receptors, and this action is thought to contribute to ketamine's potent anesthetic and analgesic properties. [7–10] Electrophysiologic and binding studies have revealed that various other antagonists, including phencyclidine, memantine, amantadine, and MK-801, induce a “use-dependent” inhibition of the NMDA receptor. [11–15] The blockade is use dependent in that the rates of onset and the recovery from blockade are increased by applying agonist. These observations have been interpreted as indicating that the channel must be in the open state before ketamine can bind to or dissociates from a blocking site possibly located within the channel pore. This type of blockade is called “uncompetitive inhibition” because it is contingent on the binding of agonist to the receptor. [16] In contrast, a “noncompetitive” blocker binds independently of the presence of agonist.

Previous single-channel studies of the NMDA channel have provided insights into the mechanisms by which dissociative anesthetics block the NMDA receptor. Heuttner and Bean [13] examined the effects of MK-801 on NMDA-activated single-channel currents and found that MK-801 produced a dose-dependent decrease in the mean duration of channel opening. [13] Similarly, phencyclidine decreased the frequency and duration of NMDA channel openings. [17] These actions are consistent with a sequential open channel blocking mechanism that can be represented by scheme 1 Figure 1.

Figure 1. Scheme 1.

The rates of channel opening and closing are beta and alpha, the rates of association and dissociation of the blocker are k sub +B and k sub -B, and [A] and [C] are the concentrations of agonist and blocker, respectively. [18] Because this model assumes that the receptor exists in a limited number of discrete states, the rates governing the transitions between these states can be estimated from the time constants obtained from the dwell time distributions. According to scheme 1, the mean duration of channel opening is decreased by the blocking drug because the open (ion-conducting) state can be terminated by either channel closure (alpha) or transitions into the open, nonconducting state referred to as “blocked”(k sub +B [C]). Thus the mean single-channel open time (tauo) will be inversely related to the sum of the two transition rates leading from the open state; that is, tau sub = 1/(k sub +B [C]+ alpha). Accordingly, open-channel blockade is characterized by a decrease in the mean duration of the open state of the channel.

Scheme 1, however, cannot account for agonist-dependent recovery from channel blockade because the model assumes that the agonist remains bound until the blocker dissociates from the receptor. Thus the rate of recovery from blockade should be not be influenced by the extracellular concentration of agonist. To account for this inconsistency, a model of “drug trapping” was postulated, as summarized in scheme 2 Figure 2. Here the agonist can dissociate and the receptor close with ketamine bound to the receptor. Closing of the channel “traps” ketamine on the receptor, and the probability of dissociation of ketamine remains low until the channel is stimulated to reopen after agonist binding. [19] In its simplest form, this model proposes that ketamine does not directly influence channel “gating” or the transition rates between the open and closed states. The true affinity of the binding site remains unchanged even though the site is “guarded”(that is, access is restricted by the conformation of the receptor). Scheme 2, which is summarized here, is discussed in detail by MacDonald et al. [7]

Figure 2. Scheme 2.

Use-dependent blockade does not necessarily imply occlusion of the open pore or preferential binding to the open state. Receptors for NMDA can exist in several agonist-associated closed states, such as the desensitized form of the receptor. Preferential binding of ketamine to a agonist-associated closed state would also be use dependent (scheme 3 Figure 3). This scheme resembles the modified-receptor model proposed by Hille [20] that describes the effects of certain local anesthetics on voltage-activated sodium channels. In scheme 3, the normal channel is represented as making transitions between three states: closed, open, and desensitized. The NMDA channel resides in more that three states, but for our purpose the scheme has been simplified. The rates of entry into, and recovery from, a nonconducting desensitized state are indicated by k sub +D and k sub -D, respectively. The rates of association and disassociation of ketamine are assumed to be negligible when the channel is in a non-desensitized state.

Figure 3. Scheme 3.

Similar to scheme 2, ketamine can be trapped when the agonist is removed and the channel reverts to the non-ligand-bound state so that the onset and recovery from blockade will be agonist dependent. However, in contrast to scheme 2, scheme 3 predicts that ketamine will reduce the number of channels available for opening without influencing mean channel-open time.

The kinetic properties of the ketamine blockade of the NMDA receptor have not been examined previously at the single-channel level. It is plausible that the mechanism of ketamine block differs substantially from that of phencyclidine or MK-801: Ketamine's actions on the nicotinic acetylcholine receptor and non-NMDA (quisqualate)-type glutamate receptor in locust suggest that it interacts with both the open and closed conformational state of these channels. [21–23] Furthermore, the physical site(s) of binding may not be identical for all uncompetitive antagonists of the NMDA receptor. Molecular studies of the binding sites for phencyclidine and MK-801 indicate that the phencyclidine binding site appears to be confined to the putative M2 segment, whereas blockade by MK-801 requires an interplay between residues associated with the M2 and M3 regions of the protein. [24] Different binding sites and mechanisms of action might account for the diverse clinical properties of uncompetitive NMDA receptor antagonists and represent potential targets for the development of new pharmaceutical agents.

Here we report the effects of ketamine on single-channel currents activated by NMDA and examine the validity of the various kinetic schemes intended to describe the blockade. Our results suggest that there are two mechanisms of inhibition: an open-channel blockade and a closed-channel blockade that involves a site accessible from the hydrophobic membrane phase. Part of this work was published previously in abstract form.*

Dissociation and Culture of Hippocampal Neurons 

Animal protocols were approved by the University of Toronto Animal Care Committee. Cultures of fetal hippocampal neurons were prepared from Swiss white mice, as previously described. [21] Briefly, hippocampal neurons were dissected from fetal mice pups. Neurons were dissociated by using enzymatic digestion and trituration. Cells were plated on collagen-coated dishes, incubated at 37 degrees Celsius in 7% carbon dioxide, and maintained in culture for 10–21 days.

Solutions and Single-channel Current Recordings 

Before recording, cells were rinsed thoroughly with a standard extracellular recording solution containing 140 mM NaCl, 1.3 mM CaCl2, 5.4 mM KCl, 25 mM N'-2-hydroxy-ethylpiperazine-N'-2-ethanesulphonic acid, 33 mM glucose, and 300 nM tetrodotoxin. The solution was buffered to a pH of 7.4 with NaOH and the osmolality was adjusted to 297–300 mOsm. Glycine (1 micro Meter) was added to the solution because it is an essential co-agonist for gating of the NMDA receptor. [5] Patch electrodes were constructed from thin-walled borosilicate glass (1.5 mm outer diameter; World Precision Instruments) using a two-stage vertical puller (Narishige PP-83). Pipettes were coated near the tips with Sylgard 184 (Dow Corning, Midland, MI) and fire polished to a final resistance of 2–5 M Omega. The electrodes used for cell-attached recordings were filled with a solution containing 70 mM NaCl, 70 mM Na2SO4, 1 mM CaCl sub 2, 10 micro Meter NMDA, and 1 micro Meter glycine. Ketamine (0 micro Meter, 0.1 micro Meter, or 1.0 micro Meter) was added to the pipette solution. In a second series of cell-attached experiments, ketamine (50 micro Meter) was added to the bath solution. For outside-out patch recordings, the pipettes were filled with solutions containing 140 mM CsF, 10 mM N'-2-hydroxy-ethylpiperazine-N'-2-ethane-sulphonic acid, 11 mM ethylene glycol-O,O'-bis(2-aminoethyl)-N,N,N',N'-tetraacetic acid, and 1 mM CaCl2. The pH was adjusted to 7.4 with CsOH. Patches were voltage clamped using an Axopatch 1B or Axopatch 200 amplifier (Axon Instruments). For the outside-out patch recordings, drugs and agonist were applied through a perfusion barrel that was positioned close to the membrane patch. [2] Patches were exposed to NMDA for approximately 5–15 min before records were sampled for kinetic analysis. We assumed that channels were functioning under near steady-state conditions and the extent of glycine-insensitive desensitization had stabilized. [25,26] All experiments were conducted at room temperature (20–25 degrees Celsius).

Single-channel currents were filtered at 2 kHz and stored on VHS tape for off-line analysis. Data were digitized (T1–1 interface, Axon Instruments, Inc., Foster City, CA) and sampled every 100 micro s using pCLAMP software (Axon Instruments, Inc., Foster City, CA). Current records were selected for detailed kinetic analysis based on the stability of the recording and the number of channels present in the patch. Only patches that remained stable over a 5- to 45-min period were analyzed. It was assumed that patches contained at least as many channels as conductance levels observed and that each channel behaved independently. Only records with infrequent (< 10%) openings to a second open level were used for the analysis. However, most patches had at least two open levels.

The threshold method was used to detect transitions between the open and closed states. [27] The threshold was set at one half the maximal current amplitude, and all records were monitored visually to ensure stability of the baseline. To minimize the effect of false events (such as transitions due to random noise), the minimum event duration was set at 300 micro s. This strict criteria, as well as the limitations of the recording system, would inevitably result in “missed” or undetected events. For example, if the duration of channel closure was shorter than the rise time of the recording system or the specified minimum event duration, closed events would be missed. [27,28] This would result in an overestimation of channel open time. Conversely, missed brief openings would result in an overestimation of closed times. Methods to correct for missed events were not used. However, the effect of missed events on our interpretation of ketamine's effects was minimized by using identical analysis criteria for records obtained in the absence and presence of ketamine. Furthermore, all records were analyzed at least twice to ensure that a substantial number of brief events were not excluded from the analysis. Data were binned into open- and closed-interval histograms and plotted using the pSTAT program (pCLAMP 6.0, Axon Instruments, Inc., Foster City, CA). We observed that the duration of closed events spanned several orders of magnitude. To detect multiple kinetic components of channel closure, closed-interval histograms were plotted with logarithmically scaled bin widths. Open interval histograms were constructed with a linearly scaled bin width because no new components were evident when histograms were plotted on a linear-log scale. The time constants and relative areas were determined by fitting the sums of exponential functions to the plotted data using the method of maximum likelihood. [27] The number of exponential components was increased until the improved fit was not statistically significant according to the F-value (where the F-value compares the sum of the squared errors extracted from the fit for the various models). Critical values of F corresponding to P < 0.05 were obtained from standard tables. Results are presented as means +/- SEM unless otherwise stated.

The mean duration of channel opening and closure were extracted from the event lists. Unfortunately, the number of channels in the patch cannot be determined accurately if the channels take long sojourns in desensitized or inactivated states. [29] Our records reflect the steady-state activity of a population of receptors, and prolonged closings of individual channels cannot be distinguished. Because we could not determine accurately the number of channels present in each patch, we defined a modified open probability (Po') in which Po' is the probability that at least one channel is open. Po' was calculated as to/tiwhere towas the total dwell time at the open level and tiwas the time interval over which Po' is measured. The frequency of channel opening was determined by the number of open events divided by the duration of a specified time interval. A discrepancy between Po' and the open probability of a single channel (calculated by multiplying the frequency of channel opening and mean open time) is expected because Po' does not distinguish between single or double openings, whereas the mean open time was calculated from single open events.

Drugs and Other Chemicals 

Ketamine is an arylcyclohexylamine that contains an asymmetric carbon atom. It exists as two isomers: the (R) and (S) form, where the (S) isomer is the more potent general anesthetic and NMDA receptor antagonist. [30,31] The commercially available preparations of ketamine used in these experiments are racemic mixtures (Parke-Davis, Scarborough, Canada). All other agonists and compounds were purchased from Sigma Chemical Co. (St. Louis, MO) unless otherwise specified.

NMDA-activated Currents Recorded in the Cell-attached Configuration 

The addition of NMDA (10 micro Meter) to the pipette solution resulted in single-channel openings, as illustrated in Figure 4. With the membrane voltage clamped at the resting potential of the cell (pipette potential = 0 mV), currents were characterized by bursts of openings interrupted by brief and long channel closures. The mean current amplitude for seven patches was 4.33 +/- 0.32 pA. The current-voltage relation, measured at pipette holding potentials ranging from -30 to +30 mV, was linear with an estimated slope conductance of 43.2 +/- 9.4 pS.

Figure 4. Single-channel currents, evoked by applications of 10 micro Meter N-methyl-D-aspartate to cell-attached patches from hippocampal neurons, are shown. Each patch was voltage clamped at the resting membrane potential of the cell. Currents were recorded from two different cells, in the absence and presence of 0.1 micro Meter ketamine. Inward currents are represented by downward deflections.

Figure 4. Single-channel currents, evoked by applications of 10 micro Meter N-methyl-D-aspartate to cell-attached patches from hippocampal neurons, are shown. Each patch was voltage clamped at the resting membrane potential of the cell. Currents were recorded from two different cells, in the absence and presence of 0.1 micro Meter ketamine. Inward currents are represented by downward deflections.

Close modal

We observed a large variation in the frequency of channel opening among different patches (1.66 sup -1 to 55.5 s sup -1) However, in individual patches, the frequency of channel opening remained stable during the duration of each recording. Therefore we assumed that this variability resulted from differences in the number of channels present in each membrane patch. The probability of at least one channel being open (Po') was 0.058 +/- 0.015. The mean frequency of channel opening was 13.4 +/- 4.4 s sup -1, and the mean open time (tauo) was 3.26 +/- 0.25 ms (n = 13).

The duration of channel openings and closings were measured and event lists were used to construct open- and closed-duration histograms. Table 1summarizes the mean time intervals, time constants, and relative areas estimated from the dwell-time distributions.

Table 1. Distribution of Open and Closed Intervals, Cell Attached Patches with Ketamine in the Pipette 

Table 1. Distribution of Open and Closed Intervals, Cell Attached Patches with Ketamine in the Pipette 
Table 1. Distribution of Open and Closed Intervals, Cell Attached Patches with Ketamine in the Pipette 

Nowak and Wright [32] reported that membrane hyperpolarization caused a decrease in the probability of NMDA channel opening. To hyperpolarize the membrane patch, a positive holding potential of +30 mV was applied to the recording electrode. Conversely, the membrane was depolarized by applying a pipette potential of -30 mV. The value of Po' was 0.03 +/- 0.02 (n = 8) at a pipette potential of +30 mV and 0.07 +/- 0.04 (n = 6) at a pipette potential of -30 mV. The mean channel-open time also decreased with membrane hyperpolarization (V pipette =+30 mV, tauo= 2.85 +/- 0.48 ms; V pipette =-30 mV tauo= 9.97 +/- 2.00 ms).

Ketamine Modulation of Single-channel Currents from Cell-attached Patches 

Ketamine was shown previously to reversibly inhibit NMDA-evoked whole-cell currents recorded from hippocampal and striatal neurons. The median inhibitory concentration for blockade is reported to range from 0.43 to 10 micro Meter. [8,12,33] To determine if this inhibition by ketamine resulted from a decrease in channel conductance, frequency of channel opening or channel-open time, we studied, in the cell-attached configuration, the effects of ketamine on single NMDA receptors. We selected ketamine concentrations in the range of previously reported IC sub 50 (concentration that inhibits 50% of the maximal response) of whole-cell currents.

Inclusion of ketamine (0.1 micro Meter or 1.0 micro Meter) in the recording electrode produced a pronounced reduction in the frequency of channel opening but no change in single-channel conductance (Figure 5). We also tried to record single-channel events with 10 micro Meter ketamine in the pipette solution, but channel openings were so infrequent that kinetic analysis of these records was not possible.

Figure 5. Ketamine decreases the probability and frequency of N-methyl-D-aspartate receptor channel opening in cell-attached patches. (A) The steady-state probability of at least one channel being open (Po') was 0.058 +/- 0.015 (n = 13), 0.031 +/- 0.01 (n = 8), and 0.014 +/- 0.007 (n = 5), in the presence of 0 micro Meter, 0.1 micro Meter, and 1 micro Meter ketamine, respectively. (B) The frequency of channel opening (f) decreased from a control value of 13.4 +/- 4.4 s sup -1 to 9.1 +/- 4.11 s sup -1 and 5.95 +/- 2.45 s sup -1 in patches exposed to 0.1 micro Meter and 1.0 micro Meter ketamine, respectively (P < 0.05, by the Mann-Whitney test). (C) The current-voltage relations for single-channel currents recorded in the cell-attached and inside-out configurations are shown. Currents were recorded from the same cell before and after patch excision. Vpipetterefers to the pipette holding potential (the resting potential of the cell is not known, so current amplitude is shown relative to the various pipette potentials). The slope conductance, estimated from the regression lines fitted to the data points for the cell-attached and the inside-out patches were 51 pS and 45 pS, respectively. The resting membrane potential of the cell estimated from the reversal potential of currents recorded in the cell-attached configuration was -70 mV. In the inside-out configuration, currents reversed polarity at 0 mV.

Figure 5. Ketamine decreases the probability and frequency of N-methyl-D-aspartate receptor channel opening in cell-attached patches. (A) The steady-state probability of at least one channel being open (Po') was 0.058 +/- 0.015 (n = 13), 0.031 +/- 0.01 (n = 8), and 0.014 +/- 0.007 (n = 5), in the presence of 0 micro Meter, 0.1 micro Meter, and 1 micro Meter ketamine, respectively. (B) The frequency of channel opening (f) decreased from a control value of 13.4 +/- 4.4 s sup -1 to 9.1 +/- 4.11 s sup -1 and 5.95 +/- 2.45 s sup -1 in patches exposed to 0.1 micro Meter and 1.0 micro Meter ketamine, respectively (P < 0.05, by the Mann-Whitney test). (C) The current-voltage relations for single-channel currents recorded in the cell-attached and inside-out configurations are shown. Currents were recorded from the same cell before and after patch excision. Vpipetterefers to the pipette holding potential (the resting potential of the cell is not known, so current amplitude is shown relative to the various pipette potentials). The slope conductance, estimated from the regression lines fitted to the data points for the cell-attached and the inside-out patches were 51 pS and 45 pS, respectively. The resting membrane potential of the cell estimated from the reversal potential of currents recorded in the cell-attached configuration was -70 mV. In the inside-out configuration, currents reversed polarity at 0 mV.

Close modal

(Table 1) summarizes the distributions of open and closed intervals measured with ketamine (0 micro Meter, 0.1 micro Meter, or 1 micro Meter) in the pipette. Ketamine did not significantly influence channel conductance as the mean current amplitude of NMDA channels exposed to ketamine (0.1 micro Meter or 1 micro Meter) was 3.82 +/- 0.16 pA (n = 8) and 3.96 +/- 0.73 pA (n = 5), respectively. Ketamine (0.1 micro Meter or 1 micro Meter) reduced Po' to 53%(Po' = 0.031 +/- 0.010, n = 8) and 24%(Po' = 0.014 +/- 0.007, n = 5, P < 0.05, by the Mann-Whitney test) of control values. The frequency of channel opening (f) was also decreased to 68%(f = 9.1 +/- 4.11 s sup -1) and 44%(f = 5.9 +/- 24.4 s sup -1) of control values (P < 0.05, by the Mann-Whitney test). A significant decrease in channel-open time was observed at the higher concentration of ketamine. With 1 micro Meter ketamine, channel-open time was reduced to 68% of control. The difference in the fractional reduction of Po' compared with f with 1 micro Meter ketamine can be attributed to the decrease in channel-open time.

The concentration-dependent decrease in mean open time is consistent with a bimolecular interaction between ketamine and a site on the open channel, as illustrated in scheme 1. The relation between drug concentration (ketamine 0 micro Meter, 0.1 micro Meter, and 1 micro Meter) and channel-open time was used to approximate the forward rate constant (k sub +B) of ketamine blockade (Figure 6). The rate of association, calculated from the slope of the plot of 1/tau o, was 1.4 x 108m sup -1 [center dot] s sup -1.

Figure 6. (A) Histograms of the distribution of open times from a cell-attached patch exposed to 10 micro Meter N-methyl-D-aspartate, in the absence (tauo= 3.174 ms, n = 8,643 open events) or presence of 0.1 micro Meter ketamine (tauo= 2.71 ms, n = 1,144) or 1 micro Meter ketamine (tauo= 2.26 ms, n = 1,542). The distribution of open times was fitted with the sum of two exponential functions with the following time constants (and relative areas): control tau1= 1.05 ms, a1= 0.16 tau2= 3 ms, a2= 0.84; 0.1 micro Meter ketamine tau1= 0.65 ms, a1= 0.11, 72 = 2.69 ms, a2= 0.89; 1 micro Meter ketamine tau1= 0.20 ms, a1= 0.65, tau2= 1.8, a2= 0.35. (B) A linear plot of the inverse of the mean open time (1/tauo) as a function of the ketamine concentration is also shown. The rate of onset of the block, estimated from the slope of the regression line fitted to the data points, was 1.4 x 108m sup -1 [center dot] sec sup -1. The inverse of mean channel open time (1/tauo), measured in the absence of ketamine, was 306 s sup -1 (n = 13). Data points from control, 0.1 micro Meter ketamine, and 1 micro Meter ketamine are the average of 13, 8, and 5 patches, respectively.

Figure 6. (A) Histograms of the distribution of open times from a cell-attached patch exposed to 10 micro Meter N-methyl-D-aspartate, in the absence (tauo= 3.174 ms, n = 8,643 open events) or presence of 0.1 micro Meter ketamine (tauo= 2.71 ms, n = 1,144) or 1 micro Meter ketamine (tauo= 2.26 ms, n = 1,542). The distribution of open times was fitted with the sum of two exponential functions with the following time constants (and relative areas): control tau1= 1.05 ms, a1= 0.16 tau2= 3 ms, a2= 0.84; 0.1 micro Meter ketamine tau1= 0.65 ms, a1= 0.11, 72 = 2.69 ms, a2= 0.89; 1 micro Meter ketamine tau1= 0.20 ms, a1= 0.65, tau2= 1.8, a2= 0.35. (B) A linear plot of the inverse of the mean open time (1/tauo) as a function of the ketamine concentration is also shown. The rate of onset of the block, estimated from the slope of the regression line fitted to the data points, was 1.4 x 108m sup -1 [center dot] sec sup -1. The inverse of mean channel open time (1/tauo), measured in the absence of ketamine, was 306 s sup -1 (n = 13). Data points from control, 0.1 micro Meter ketamine, and 1 micro Meter ketamine are the average of 13, 8, and 5 patches, respectively.

Close modal

In the Cell-attached Configuration, Addition of Ketamine to the Bath Solution Inhibits NMDA Channels Recorded in the Cell-attached Configuration 

The uncharged form of ketamine is highly lipid soluble. [34] Thus it is plausible that ketamine can diffuse into the cell membrane and gain access to a hydrophobic binding site on the NMDA receptor. Therefore we tried to determined if ketamine, added to the bath solution, could inhibit NMDA receptors isolated in the tip of the recording electrode.

Stable single-channel currents were recorded with 10 micro Meter NMDA in the pipette solution. Adding 50 micro Meter ketamine to the bath solution caused a gradual, reversible inhibition of channel opening (Figure 7). The probability of channel opening was significantly reduced compared with control values (Po', control = 0.046 +/- 0.010, Po', ketamine = 0.051 +/- 0.003, n = 5, P < 0.05, by the Wilcoxon signed rank test).

Figure 7. Ketamine (50 micro Meter) added to the bath solution reversibly inhibited channels present in cell-attached patches. Single-channel currents were recorded from the same cell, in the absence and presence of ketamine. The durations of closed times are plotted as the distribution of log time versus the bin frequency. In the absence of ketamine, the distribution was fitted with the sum of three exponential functions, with time constants and relative areas of tau1= 0.32 ms, a1= 0.56, tau2= 5.3, a2= 0.13, tau3= 80.7 ms, and a3= 0.32, whereas tau1= 0.23 ms, a1= 0.53, tau2= 5.1 ms, a2= 0.14, tau3, = 838.93 ms, a3= 0.33 for currents recorded in the presence of ketamine. The mean closed times for the illustrated tracings were tauccontrol = 38.9 +/- 1.21 ms (n = 3662 events) and taucketamine = 412 +/- 36.86 (n = 390), whereas the mean channel-open times were tauocontrol = 2.82 +/- 0.48 ms (n = 3,539) and tauoketamine = 1.61 +/- 0.07 ms (n = 482).

Figure 7. Ketamine (50 micro Meter) added to the bath solution reversibly inhibited channels present in cell-attached patches. Single-channel currents were recorded from the same cell, in the absence and presence of ketamine. The durations of closed times are plotted as the distribution of log time versus the bin frequency. In the absence of ketamine, the distribution was fitted with the sum of three exponential functions, with time constants and relative areas of tau1= 0.32 ms, a1= 0.56, tau2= 5.3, a2= 0.13, tau3= 80.7 ms, and a3= 0.32, whereas tau1= 0.23 ms, a1= 0.53, tau2= 5.1 ms, a2= 0.14, tau3, = 838.93 ms, a3= 0.33 for currents recorded in the presence of ketamine. The mean closed times for the illustrated tracings were tauccontrol = 38.9 +/- 1.21 ms (n = 3662 events) and taucketamine = 412 +/- 36.86 (n = 390), whereas the mean channel-open times were tauocontrol = 2.82 +/- 0.48 ms (n = 3,539) and tauoketamine = 1.61 +/- 0.07 ms (n = 482).

Close modal

This inhibition was primarily characterized by a decrease in the frequency of channel opening, rather than by a reduction in the mean channel open time (Table 2). Three of the five patches remained stable for a prolonged period, and perfusion of the cell with the ketamine-free solution resulted in a gradual increase in the probability of channel opening (Po', recovery = 0.011 +/- 0. 002).

Table 2. Ketamine Added to the Bath Solution 

Table 2. Ketamine Added to the Bath Solution 
Table 2. Ketamine Added to the Bath Solution 

Ketamine Inhibition of NMDA Channels Recorded in the Outside-out Configuration 

A major advantage of the cell-attached recording method is that the plasma membrane is not disrupted after seal formation. This results in minimal disruption of cytosolic factors that might influence channel gating or drug sensitivity. [35] However, the pipette solution cannot be easily exchanged, so the same population of receptors cannot be examined before and after exposure to a known concentration of drug. To study the same receptor population in the absence and presence of ketamine, we examined the effects of ketamine on NMDA receptors in outside-out membrane patches (Figure 8).

Figure 8. N-methyl-D-aspartate-evoked inward currents from an outside-out patch voltage clamped at -60 mV. Sample traces of single-channel currents show ketamine inhibition and recovery from blockade. Traces were successively record from the same cell.

Figure 8. N-methyl-D-aspartate-evoked inward currents from an outside-out patch voltage clamped at -60 mV. Sample traces of single-channel currents show ketamine inhibition and recovery from blockade. Traces were successively record from the same cell.

Close modal

In patches held at -60 mV, applications of NMDA (10 micro Meter) evoked channel openings with a mean current amplitude of 2.69 +/- 0.11 pA (n = 9); the estimated single channel conductance was 44.8 pS. Application of ketamine (1 micro Meter or 10 micro Meter) did not change the single-channel current amplitude (2.47 +/- 0.13 pA n = 5, and 2.50 +/- 0.16 pA, n = 4, respectively), (Figure 9).

Figure 9. Currents were recorded from a single outside-out patch, in the absence and presence of ketamine. (A) Current traces are displayed at two different time scales (100 ms and 33.3 ms). The solid bars with arrows indicate the data selected for higher resolution. (B) Histograms of the current amplitude for all openings longer than 300 micro second, recorded in the absence and presence of ketamine (1 micro Meter), are shown. The distributions were fitted by a single Gaussian curve with mean values of 3.03 +/- 0.54 pA (n = 2,450 events) and 2.95 +/- 0.47 pA (n = 6,345 events), respectively.

Figure 9. Currents were recorded from a single outside-out patch, in the absence and presence of ketamine. (A) Current traces are displayed at two different time scales (100 ms and 33.3 ms). The solid bars with arrows indicate the data selected for higher resolution. (B) Histograms of the current amplitude for all openings longer than 300 micro second, recorded in the absence and presence of ketamine (1 micro Meter), are shown. The distributions were fitted by a single Gaussian curve with mean values of 3.03 +/- 0.54 pA (n = 2,450 events) and 2.95 +/- 0.47 pA (n = 6,345 events), respectively.

Close modal

As summarized in Table 3, ketamine (1 micro Meter or 10 micro Meter) reduced Po' to 63% and 34% of control values, respectively. For five patches, the steady-state probabilities of at least one channel being open, measured before and after exposure to ketamine (1 micro Meter), were 0.087 +/- 0.028 and 0.056 +/- 0.019 (P < 0.05, by the Wilcoxon signed rank test), respectively. Thus 1 micro Meter ketamine reduced Po' to 64% of control, and 10 micro Meter ketamine decreased Po' to 23% of control. These reductions in Po' were due primarily to a decrease in the frequency of channel opening, rather than to a reduction in the mean channel-open time (Table 3). Applications of the competitive NMDA receptor antagonist, DL-2-amino-5-phosphovaleric acid (10–100 micro Meter), reversibly inhibited residual channel openings (n = 3), indicating that the currents were indeed mediated by the NMDA receptor.

Table 3. Distribution of Open and Closed Intervals, Outside-out Patches 

Table 3. Distribution of Open and Closed Intervals, Outside-out Patches 
Table 3. Distribution of Open and Closed Intervals, Outside-out Patches 

Similar to the dwell-time histograms constructed for currents recorded in the cell-attached configuration, open-interval histograms were best fit with the sum of two exponential functions, whereas the closed-duration histograms were best fit with the sum of three exponential functions (Figure 10). A possible decrease in channel open time was apparent with the higher concentration of ketamine (10 micro Meter), but this effect was not significant (P = 0.437, by the Wilcoxon signed rank test).

Figure 10. Open- and closed-time distributions of N-methyl-D-aspartate receptors recorded in the outside-out configuration. (A) The open-duration histogram was fitted with the sum of two exponential functions that had fast and slow time constants (ms) and relative areas of tau1= 0.74 ms, a10.26, tau2= 3.6 ms, a2= 0.75. (B) In the presence of 1 micro Meter ketamine, tau1= 0.81 ms, a10.2, tau2= 4.1 ms and a2= 0.81. (C) The closed-duration histograms were best fit with three exponential functions. Under control conditions, tau1= 0.60 ms, a1= 0.6, tau2= 8.11 ms, a2= 0.16, tau3= 39.47 ms, a3= 0.25, whereas tau1= 0.8 ms, a1= 0.6, tau sub 2 = 7.6 ms, a2= 0.09, tau3= 65 ms and a3= 0.34 in the presence of ketamine (D).

Figure 10. Open- and closed-time distributions of N-methyl-D-aspartate receptors recorded in the outside-out configuration. (A) The open-duration histogram was fitted with the sum of two exponential functions that had fast and slow time constants (ms) and relative areas of tau1= 0.74 ms, a10.26, tau2= 3.6 ms, a2= 0.75. (B) In the presence of 1 micro Meter ketamine, tau1= 0.81 ms, a10.2, tau2= 4.1 ms and a2= 0.81. (C) The closed-duration histograms were best fit with three exponential functions. Under control conditions, tau1= 0.60 ms, a1= 0.6, tau2= 8.11 ms, a2= 0.16, tau3= 39.47 ms, a3= 0.25, whereas tau1= 0.8 ms, a1= 0.6, tau sub 2 = 7.6 ms, a2= 0.09, tau3= 65 ms and a3= 0.34 in the presence of ketamine (D).

Close modal

Single-channel studies were undertaken to identify the mechanisms underlying ketamine blockade of the NMDA receptor. In the absence of ketamine, channel conductance was similar to that previously reported for NMDA receptors in hippocampal neurons. [36,37] Furthermore, the duration of NMDA channel openings was within the range reported for native hippocampal receptors and cloned NR1a-NM epsilon 1 and NR1-NR2A subunits expressed in HEK 293 cells and Xenopus oocytes. [36–41] The effects of ketamine on single-channel NMDA currents are summarized here.

In cell-attached patches, the inclusion of ketamine (0.1 micro Meter or 1 micro Meter) in the pipette solution caused a concentration-dependent decrease in both the duration and frequency of NMDA channel openings. These actions are similar to the effects of phencyclidine and its analogs and are consistent with an open-channel mechanism of blockade. [17] However, the decreases in channel-open times induced by 0.1 micro Meter and 1 micro Meter ketamine (6% and 32%, respectively) were small relative to the reduction in open-channel frequency (32% and 56%, respectively). Furthermore, in cell-attached patches, externally applied ketamine (50 micro Meter) caused an 89% decrease in the frequency of channel opening without a significant change in channel-open time. Ketamine blockade of receptors recorded in the outside-out patch configuration was similarly characterized by a decrease in the frequency of open events with little change in open time.

Mechanisms of Ketamine Blockade 

As discussed previously, several mechanisms could account for ketamine blockade of the NMDA receptor. Schemes 1, 2, and 3 will now be considered in light of our observations. Scheme 1 is not consistent with the use-dependent recovery from block previously described for ketamine. [7] In addition, scheme 1 predicts that a new closed state with a mean duration of 1/kBshould be evident in the distribution of the closed times. We observed no change in the fast or intermediate components of the closed-time distributions and no new closed-time components. Thus we rejected scheme 1 as a possible model of ketamine blockade. Both scheme 2 and 3 are consistent with use-dependent recovery from blockade; however, scheme 2 predicts a reduction in mean open time, whereas scheme 3 does not.

In the cell-attached configuration with ketamine added to the recording electrode, there was a concentration-dependent decrease in channel-open time, consistent with scheme 2. The rate of association of ketamine with the open channel was estimated to be 14 x 108m sup -1 [center dot] sec sup -1. This value is considerably greater than that previously reported for ketamine inhibition of whole-cell currents (1.7 x 104m sup -1 [center dot] sec sup -1 and 9 x 104m sup -1 [center dot] sec sup -1). [7,33] However, the time course of inhibition of whole-cell currents depends not only on the association rate constant but also on the probability of channel opening (Po) where kon= 1/Po (alpha + k sub +B [B]). The maximum probability of NMDA channel opening is not known, and an overestimation of Po results in an underestimation of k sub +B. Thus molecular rate constants are more directly determined from single-channel experiments. Although our value of 1.4 x 108m sup -1 [center dot] sec sup -1 is presented only as an approximation, it suggests that the rate of ketamine binding is of the order of magnitude expected for a fast channel blocker. However, the observed decrease in channel-open time accounted, only in part, for the decrease in Po', suggesting that the channel could close with ketamine bound to the receptor. This closed-blocked state would reduce the number of channels available for activation and thereby reduce the frequency of channel opening. These observations are consistent with scheme 2.

In outside-out patches or cell-attached patches with ketamine added outside the recording electrode, a large reduction in opening frequency was observed with no significant change in mean open time (despite the presence of high concentrations of ketamine). Detailed single-channel studies on NMDA receptor function suggest that the receptor kinetics are characterized by several agonist-bound inactive states. [42–47] Any one of these states might preferentially enhance ketamine binding and lead to blocked states. This would reduce the number of channels available for opening but produce no change in mean open time. We observed that only the longest closed time was prolonged by ketamine, as expected from a decrease in the number of channels available for normal activation. The lack of change in the open-time constants and the fast and intermediate closed-time constants suggests that ketamine, acting through a membrane delimited pathway does not modulate channel gating.

Our data suggest that ketamine acts by at least two mechanisms: an open-channel block, as evidenced by the decrease in tauo, and a closed block characterized by a decrease in open frequency with no change in open time. It is not known if distinct sites mediate open and closed blockade. Furthermore, the kinetics of the two mechanisms appear to depend on the configuration of the membrane patch. There was an apparent decrease in drug sensitivity after patch excision. Factors that could contribute to changes in drug sensitivity include disruption of the cytoskeleton and alterations of cytosolic factors that influence the sensitivity of the receptor to antagonists. [35]

Open and closed mechanisms of blockade were previously proposed for ketamine inhibition of the nicotinic acetylcholine receptor. [21,22,47–50] Ketamine, as well as various other intravenous and volatile anesthetics, selectively bind to the closed nicotinic acetylcholine receptor receptor. [17,22] These anesthetics drive the receptor into a nonactivatable configuration that is similar to the agonist-induced desensitized state (for a review of this, see Forman and Miller [51]). Furthermore, binding studies indicate that the site that facilitates desensitization of the nicotinic acetylcholine receptor by [sup 3 H]phencyclidine is a hydrophobic region of the nicotinic acetylcholine receptor receptor that is not directly accessible from the aqueous phase. [52]

Ketamine Gained Access to a Binding Site through a Hydrophobic Pathway 

In addition to the mechanisms of ketamine blockade, the site(s) of antagonism are of interest. It was previously suggested that ketamine binds in the vicinity of the extracellular mouth of the NMDA channel pore. [7] However, we observed that ketamine added to the bath solution inhibited NMDA receptors recorded in the cell-attached configuration. It is unlikely that ketamine gained access to the extracellular mouth of the channel by diffusing across the high-resistance seal (> 10 G Omega) and accumulating in the tip of the recording electrode. Ketamine is formulated as a hydrochloride salt and is highly water soluble, with apKa (negative logarithm of the acid ionization constant) of 7.5. [34] Under physiologic conditions, a large fraction of the drug exists in the lipid-soluble form (44%). The octanol/buffer partition coefficient was reported to be 398 (pH 7.4).** Therefore the concentration of ketamine in the lipid phase is several orders of magnitude greater than in the aqueous phase. Our results suggest that ketamine gained access to a blocker site associated with the lipid membrane or the lipid-protein interface, as indicated in Figure 11. Indeed, the presence of a hydrophobic pathway for dissociated anesthetics was predicted from binding studies of the antagonist [sup 3 H]MK-801. [46] Javitt and Zukin [46] postulated that [sup 3 H]MK-801 gained access to the slow hydrophobic pathway, which was independent of channel gating and a fast hydrophilic pathway associated with channel opening.

Figure 11. Patch configurations used to investigate ketamine blockade of the N-methyl-D-aspartate (NMDA) receptor. For the cell-attached patches, NMDA was included in the pipette solution. Ketamine was included either in the pipette or added externally to the bath solution. In the outside-out patch configuration, both ketamine and NMDA were added to the external solution. Changes in NMDA receptor activity are schematically interpreted. Our results suggest that ketamine may interact with the receptor at two potentially distinct sites: one site located within the channel pore and a second site associated with a hydrophobic domain of the protein. The binding of the agonist to the receptor is assumed to modify the binding of ketamine to both sites. Binding of ketamine at the site associated with the channel pore would decrease channel-open time (scheme 2), whereas binding to the membrane-associated site does not require the channel to be in the open state (scheme 3).

Figure 11. Patch configurations used to investigate ketamine blockade of the N-methyl-D-aspartate (NMDA) receptor. For the cell-attached patches, NMDA was included in the pipette solution. Ketamine was included either in the pipette or added externally to the bath solution. In the outside-out patch configuration, both ketamine and NMDA were added to the external solution. Changes in NMDA receptor activity are schematically interpreted. Our results suggest that ketamine may interact with the receptor at two potentially distinct sites: one site located within the channel pore and a second site associated with a hydrophobic domain of the protein. The binding of the agonist to the receptor is assumed to modify the binding of ketamine to both sites. Binding of ketamine at the site associated with the channel pore would decrease channel-open time (scheme 2), whereas binding to the membrane-associated site does not require the channel to be in the open state (scheme 3).

Close modal

The site(s) of blockade by uncompetitive antagonists, including ketamine, have been investigated using cloned NMDA receptor subunits and site-directed mutagenesis. [53,54] Dissociative anesthetics are thought to bind to a site associated with the Mg2+ binding location on the putative transmembrane domain. Replacement of the conserved asparagine 598 residue in the M2 region by glutamine caused a decrease in the receptor sensitivity to MK-801 and Mg2+. [54] However, examination of a series of NMDA subunit heteromers revealed that not all combinations of NMDA receptor subunits were equally sensitive to dissociative anesthetics: Not all the dissociative anesthetics bind to an identical site on the NMDA receptor, and their effects were mediated by a site other than the Mg2+ binding domain. [53]

Our results are somewhat analogous to the actions of local anesthetics on voltage-activated sodium channels. [55] This family of compounds induce both a use-dependent or “phasic” block and “tonic” inhibition of sodium channels (for a review, see Butterworth and Strichartz [56]). Local anesthetics are thought to dissolve in the cytoplasm and reach the channel via the hydrophylic pathway. In addition, they may diffuse into the membrane and bind via a hydrophobic pathway. Clues to the location of binding sites of local anesthetics have been obtained by examining the effects of structurally diverse local anesthetics and their effects under different recording conditions (such as changes in membrane potential or pH). Similarly, the location of the ketamine-binding site(s) might be elucidated by examining the stereoselectivity and voltage sensitivity of the closed- and open-channel blockade by ketamine of NMDA receptors.

Clinical Implications 

Ketamine has unique clinical properties and is the anesthetic of choice in specific situations, such as in patients with a compromised hemodynamic status or asthma. [10] More recently, ketamine has been used in subanesthetic doses as a potent analgesic. In humans, inhibition of the NMDA receptor is thought to mediate ketamine's clinical properties. [57,58] During anesthesia, the peak plasma concentration in humans is approximately 8.5–9.5 micro Meter, whereas analgesia is associated with plasma concentrations of approximately 0.55 micro Meter. [10,59] In rats, the peak plasma concentration associated with anesthesia is less than 36 micro Meter. [60] Ketamine in the plasma exists in two states: a free unbound form and a form bound to plasma proteins. [61,62] The free fraction of ketamine determines the rate of diffusion to a site of action. In humans, as much as 47% of ketamine is bound to plasma proteins. [61] However, ketamine is highly lipid soluble and the brain-to-plasma ratio for ketamine is estimated to be 6.5:1, suggesting that ketamine preferentially accumulates in the brain. [63] Thus it is possible that, during anesthesia, the concentration of ketamine present at the NMDA receptor is considerably higher than the plasma concentration.

We observed that clinically relevant concentrations of ketamine induced both an open and closed blockade of the NMDA receptor. It is of interest to determine which of these mechanisms contribute to ketamine's clinical actions. The predominance of closed-channel blockade observed at low concentrations of ketamine suggests that ketamine's analgesic properties might result from closed-rather than open-channel blockade. In this regard it is notable that the anti-Parkinsonian drugs, memantine and amantadine, have no appreciable anesthetic or analgesic properties and inhibit the NMDA receptor by open-channel blockade. [6] It will be of interest to determine whether the two modes of blockade can be pharmacologically dissociated. [39,64–66]

Our results are consistent with a dual mechanism of ketamine blockade: occlusion of the open channel from the aqueous phase and closed-channel blockade from the membrane phase.

The authors thank Lydia Brandes and Claire Bartlett for technical assistance, Valerie Oxorn for graphic design, and Patricia Watson for reviewing the manuscript.

*Orser BA, MacDonald JF: Ketamine blocks NMDA channels via a hydrophobic pathway. (Abstract) Can J Anaesth 1992; 39:A12.

**Correspondence from Parke-Davis, Scarborough, Canada.

1.
Mori H, Mishina M: Review: Neurotransmitter receptors VIII. Structure and function of the NMDA receptor channel. Neuropharmacology 1995; 34:219-37.
2.
Johnson JW, Ascher P: Glycine potentiates the NMDA response in cultured mouse brain neurones. Nature 1987; 325:522-5.
3.
Legendre P, Rosenmund C, Westbrook GL: Inactivation of NMDA channels in cultured hippocampal neurons by intracellular calcium. J Neurosci 1993; 13:674-84.
4.
Turecek R, Vylicky L, Vlachova V, Vylicky L, G-protein modulation of glycine-resistant NMDA receptor desensitization in rat cultured hippocampal neurons. Eur J Neurosci 1995; 7:1826-30.
5.
McBain CJ, Mayer ML: N-methyl-D-aspartic acid receptor structure and function. Physiol Rev 1994; 74:723-60.
6.
Lipton SA, Rosenberg PA: Excitatory amino acids as a final common pathway for neurological disorders. N Engl J Med 1994; 330:613-22.
7.
MacDonald JF, Bartlett MC, Mody I, Pahapill P, Reynolds JN, Salter MW, Schneiderman JH, Pennefather PS: Action of ketamine, phencyclidine and MK-801 on NMDA receptor currents in cultured mouse hippocampal neurones. J Physiol 1991; 432:483-508.
8.
Mayer ML, Westbrook GL, Vyklicky L: Sites of antagonist action on N-methyl-D-aspartate acid receptors studied using fluctuation analysis and a rapid perfusion technique. J Neurophysiol 1988; 60:645-63.
9.
Irfune M, Shimizu, Nomoto M, Fukuda T: Ketamine-induced anesthesia involves the N-methyl-D-aspartate-channel complex in mice. Brain Res 1992; 596:1-9.
10.
Reich D, Silvay G: Ketamine: An update on the first twenty-five years of clinical experience. Can J Anaesth 1989; 36:186-97.
11.
Honey CR, Miljkovic Z, MacDonald JF: Ketamine and phencyclidine causes a voltage-dependent block of responses to L-aspartic acid. Neurosci Lett 1985; 61:135-9.
12.
MacDonald JF, Miljkovic Z, Pennefather P: Use-dependent block of excitatory amino acid currents in cultured neurones by ketamine. J Neurosci 1987; 58:251-66.
13.
Huettner J, Bean BP: Block of N-methyl-D-asparate-activated current by the anticonvulsant MK-801: Selective binding to open channels. Proc Natl Acad Sci USA 1988; 85:1301-11.
14.
Anis NS, Berry SC, Burton NR, Lodge D: The dissociative anaesthetics, ketamine and phencyclidine, selectively reduce excitation of central mammalian neurons by N-methyl-aspartate. Br J Pharmacol 1983; 79:565-75.
15.
Parsons CG, Gruner R, Rozental J, Millar J, Lodge D: Patch clamp studies on the kinetics and selectivity of N-methyl-D-aspartate receptor antagonism by memantine (1-amino-3,5-dimethyladamantan). Neuropharmacology 1993; 32:1337-50.
16.
Pennefather P, Quastel DMJ: Modification of dose-response curves by effector blockade and uncompetitive antagonism. Mol Pharmacol 1992; 22:369-80.
17.
Ramoa AS, Albuquerque EX: Phencyclidine and some of its analogues have distinct effects on NMDA receptors of rat hippocampal neurons. FEBS Lett 1988; 235:156-62.
18.
Adams PR: Drug blockade of open end-plate channels. J Physiol 1976; 260:531-52.
19.
MacDonald JF, Novak LM: Mechanisms of blockade of excitatory amino acid receptors channels. Trends Pharmacol Sci 1990; 11:167-72.
20.
Hille B: Local anaesthetics: Hydrophilic and hydrophobic pathways for the drug-receptor reaction. J Gen Physiol 1977; 69:497-515.
21.
Ashford MLJ, Boden P, Ramsey RL, Usherwood PNR: Enhancement of desensitization of quisqualate-type glutamate receptors by the dissociative anaesthetic ketamine. J Exp Biol 1989; 141:73-86.
22.
Aguayo LG, Witkop B, Albuquerque EX: Voltage- and time-dependent effects of phencyclidines on the endplate current arise from open and closed channel blockade. Proc Natl Acad Sci USA 1986; 83:3523-7.
23.
Wachtel RE, Wegrzynowicz ES: Kinetics of nicotinic acetylcholine ion channel in the presence of intravenous anaesthetics and induction agents. Br J Pharmacol 1992; 106:623-7.
24.
Ferrer-Montiel AV, Sun W, Montal M: Molecular design of the N-methyl-D-aspartate receptor binding site for phencyclidine and dizolcipine. Proc Natl Acad Sci USA 1995; 92:8021-5.
25.
MacDonald JF, Mody I, Salter MW: Regulation of NMDA receptors revealed by intracellular dialysis. J Physiol 1986; 414:17-23.
26.
Tong G, Jahr C: Regulation of glycine-insensitive desensitization of the NMDA receptor in outside-out patches. J Neurophysiol 1994; 72:754-61.
27.
Colquhoun D, Sigworth FJ: Fitting and statistical analysis of single-channel records, Single-channel Recording. Edited by Sackmann B, Neher E. New York, Plenum Press, 1995, pp 483-587.
28.
Hawkes AG, Jalali A, Colquhoun D: Asymptotic distributions of the apparent open times and shut times in a single channel record allowing for omission of brief events. Phil Trans R Soc Lond [Biol] 1992; 337:383-404.
29.
Hille B: Ionic channels of excitable membranes. Sunderland MA., Sinauer Associates, 1992, pp 390-422.
30.
Zeilhofer HU, Swandulla D, Geisslinger G, Brune K: Differential effects of ketamine enantiomers on NMDA receptor currents in cultured neurons. Eur J Pharmacol 1992; 213:155-8.
31.
White PF, Ham J, Way WL, Trevor AJ: Pharmacology of ketamine isomers in surgical patients. Anesthesiology 1980; 52:231-9.
32.
Nowak LM, Wright JM: Slow voltage-dependent changes in channel open-state probability underlies hysteresis of NMDA responses in Mg sup 2+ -free solutions. Neuron 1992; 8:181-7.
33.
Parsons GC, Quack G, Bresink I, Baran L, Przeganlinski E, Kostowski W, Krzascik P, Hartmann S, Danysz W: Comparison of the potency, kinetics and voltage-dependency of a series of uncompetitive NMDA receptor antagonists in vitro with anticonvulsive and motor impairment activity in vivo. Neuropharmacology 1995; 34:1239-58.
34.
Cohen ML, Trevor AJ: On the cerebral accumulation of ketamine and the relationship between metabolism of the drug and its pharmacological effects. J Pharmacol Exp Ther 1974; 189:351-8.
35.
Covarrubias M, Steinbach JH: Excision of membrane patches reduces the mean open time of nicotinic acetylcholine receptors. Pflugers Arch 1990; 416:385-92.
36.
McLaron JG, Sawyer D: Dependence of single channel properties of the N-methyl-D-asparate ion channel on sterioisomer agonist. Exp Brain Res 1993; 95:8-14.
37.
McLaron JG, Curry K: Single channel properties of the N-methyl-D-aspartate receptor channel using NMDA and NMDA agonists. Exp Brain Res 1990; 82:82-8.
38.
Lima-Landman MTR, Albuquerque EX: Ethanol potentiates and block NMDA-activated single-channel currents in rat hippocampal pyramidal cells. FEBS Lett 1989; 247:61-7.
39.
Parsons CG, Zong X, Lux HD: Whole cell and single channel analysis of the kinetics of glycine-sensitive N-methyl-D-aspartate receptor desensitization. Br J Pharmacol 1993; 109:213-21.
40.
Stern P, Cik M, Colquhoun D, Stephenson FA: Single channel properties of cloned NMDA receptors in human cell line: Comparison with results from Xenopus oocytes. J Physiol 1994; 476:391-7.
41.
Gibb AJ, Colquhoun D: Activation of N-methyl-D-aspartate receptors by L-glutamate in cells dissociated from adult rat hippocampus. J Physiol 1992; 456:143-79.
42.
Benveniste M, Clements J, Vyklicky L, Mayer ML: A kinetic analysis of the modulation of N-methyl-D-aspartic acid receptors by glycine in mouse cultured hippocampal neurons. J Physiol 1990; 428:333-7.
43.
Lin F, Stevens CF: A quantitative description of NMDA receptor-channel kinetic behavior. J Neurosci 1994; 14:2153-60.
44.
Colquhoun D, Hawkes AG: Desensitization of N-methyl-D-aspartate receptors: A problem of interpretation. Proc Natl Acad USA 1995; 92:10327-9.
45.
Zukin SR, Javitt DC: Phencyclidine receptor binding as a probe of NMDA receptor functioning: Implications for drug abuse research. NIDA Res Monogr 1993; 133:1-12.
46.
Javitt DC, Zukin SR: Biexponential kinetics of [3]MK-801 binding: Evidence for access to closed and open N-methyl-D-aspartate receptor channels. Mol Pharmacol 1989; 35:387-93.
47.
Javitt DC, Frusciante J, Zukin SR: Rat brain N-methyl-D-aspartate receptors require multiple molecules of agonist for activation. Mol Pharmacol 1990; 37:603-7.
48.
Scheller M, Bufler J, Schneck HJ, Franke C, Kochs E: Ketamine and its stereoisomers block the nicotinic acetylcholine receptor of mouse myotubules. Anesth Analag 1996; 82:S396.
49.
Wachtel RE: Ketamine decreases the open time of single-channel currents activated by acetylcholine. Anesthesiology 1988; 68:563-70.
50.
Heidmann T, Oswald RE, Changeux JP: Multiple sites of action for noncompetitive blockers on acetylcholine receptor rich membrane fragments from Torpedo marmorata. Biochemistry 1983; 22:3112-27.
51.
Forman SA, Miller K: Molecular sites of anesthetic action in postsynaptic nicotinic membranes. Trends Pharmacol Sci 1989; 10:447-52.
52.
Hertz JM, Johnson DA, Taylor P: Distance between the agonist and noncompetitive inhibitor sites on the nicotinic acetylcholine receptor. J Biochem Chem 1989; 264:12439-48.
53.
Yamakura T, Mori H, Masaki H, Shimoji K, Mishina M: Different sensitivities of NMDA receptor channel subtypes to non-competitive antagonists. NeuroReport 1993; 4:687-90.
54.
Mori H, Masaki H, Yamakura T, Mishina M: Identification by mutagenesis of a Mg sup 2+ -block site of the NMDA receptor channel. Nature 1992; 358:673-5.
55.
Lee-Son S, Wang GK, Concus A, Crill E, Strichartz G: Stereoselective inhibition of neuronal sodium channels by local anesthetics: Evidence for two sites of action? Anesthesiology 1992; 77:324-35.
56.
Butterworth JF, Strichartz GR: Molecular mechanisms of local anesthetics: A review. Anesthesiology 1990; 72:711-34.
57.
Johnson KM, Jones SM: Neuropharmacology of phencyclidine: Basic mechanisms and therapeutic potential. Ann Rev Pharmacol Toxicol 1990; 30:707-50.
58.
Hartig P, Valtysson J, Lindner K-J, Kristensen J, Karlsten R, Gustafsson LL, Persson J, Svensson JO, Oye I, Antoni G, Westerberg G, Langstrom B: Central nervous system effects of subdissociative doses of (S)-ketamine are related to plasma and brain concentrations measured with positron emission tomography in healthy volunteers. Clinical Pharmacol Ther 1995; 58:165-73.
59.
White PF, Ham J, Way WL, Trevor AJ: Pharmacology of ketamine isomers in surgical patients. Anesthesiology 1980; 52:231-9.
60.
Pedraz JL, Lanao JM, Dominguez-Gil: Interspecies pharmacokinetics of ketamine, Status of Ketamine in Anesthesiology. Edited by Domino EF. Ann Arbor, NPP Books, 1990, pp 285-95.
61.
Dayton PG, Stiller RL., Cook DR, Perel JM: The binding of ketamine to plasma proteins: Emphasis on human plasma. Eur J Clin Pharmacol 1983; 24:825-31.
62.
Wood M: Plasma binding and limitation of drug access to site of action. Anesthesiology 1991; 75:721-3.
63.
Cohen ML, Chan SL, Way WL, Trevor AJ: Distribution in the brain and metabolism of ketamine in the rat after intravenous administration. Anesthesiology 1973; 39:370-6.
64.
Kronhuber J, Quack G, Danysz W, Jellinger K, Danielczyk W, Gsell W, Riederer P: Therapeutic brain concentration of the NMDA receptor antagonist amantadine. Neuropharmacology 1995; 34:713-21.
65.
Kronhuber J, Quack G: Cerebrospinal fluid and serum concentrations of the N-methyl-D-aspartate (NMDA) receptor antagonist memantime in man. Neurosci Lett; 195:137-9.
66.
Fletcher JE, MacDonald JF: Haloperidol interacts with the strychnine-insensitive glycine site at the NMDA receptor in cultured mouse hippocampal neurones. Eur J Pharmacol 1993; 235:291-5.