Skip to main content

Main menu

  • HOME
  • CONTENT
    • Early Release
    • Featured
    • Current Issue
    • Issue Archive
    • Collections
    • Podcast
  • ALERTS
  • FOR AUTHORS
    • Information for Authors
    • Fees
    • Journal Clubs
    • eLetters
    • Submit
  • EDITORIAL BOARD
  • ABOUT
    • Overview
    • Advertise
    • For the Media
    • Rights and Permissions
    • Privacy Policy
    • Feedback
  • SUBSCRIBE

User menu

  • Log in
  • My Cart

Search

  • Advanced search
Journal of Neuroscience
  • Log in
  • My Cart
Journal of Neuroscience

Advanced Search

Submit a Manuscript
  • HOME
  • CONTENT
    • Early Release
    • Featured
    • Current Issue
    • Issue Archive
    • Collections
    • Podcast
  • ALERTS
  • FOR AUTHORS
    • Information for Authors
    • Fees
    • Journal Clubs
    • eLetters
    • Submit
  • EDITORIAL BOARD
  • ABOUT
    • Overview
    • Advertise
    • For the Media
    • Rights and Permissions
    • Privacy Policy
    • Feedback
  • SUBSCRIBE
PreviousNext
Articles

Requirement for Tyrosine Phosphatase during Serotonergic Neuromodulation by Protein Kinase C

Stefano Catarsi and Pierre Drapeau
Journal of Neuroscience 1 August 1997, 17 (15) 5792-5797; DOI: https://doi.org/10.1523/JNEUROSCI.17-15-05792.1997
Stefano Catarsi
1Centre for Research in Neuroscience, McGill University, and Montreal General Hospital Research Institute, Montreal, Quebec, Canada H3G 1A4
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Pierre Drapeau
1Centre for Research in Neuroscience, McGill University, and Montreal General Hospital Research Institute, Montreal, Quebec, Canada H3G 1A4
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • Article
  • Figures & Data
  • Info & Metrics
  • eLetters
  • PDF
Loading

Abstract

Tyrosine kinases and phosphatases are abundant in the nervous system, where they signal cellular differentiation, mediate the responses to growth factors, and direct neurite outgrowth during development. Tyrosine phosphorylation can also alter ion channel activity, but its physiological significance remains unclear. In an identified leech mechanosensory neuron, the ubiquitous neuromodulator serotonin increases the activity of a cation channel by activating protein kinase C (PKC), resulting in membrane depolarization and modulation of the receptive field properties. We observed that the effects on isolated neurons and channels were blocked by inhibiting tyrosine phosphatases. Serotonergic stimulation of PKC thus activates a tyrosine phosphatase activity associated with the channels, which reverses their constitutive inhibition by tyrosine phosphorylation, representing a novel form of neuromodulation.

  • single channel
  • serotonin
  • protein kinase C
  • tyrosine phosphorylation
  • identified neuron
  • leech

Regulation of ion channel activity by protein phosphorylation is a common mechanism of neuromodulation of transmitter-activated and voltage-gated ion channels (Kaczmarek and Levitan, 1987). The best characterized actions are those of the serine/threonine kinases activated by calcium or cyclic nucleotides, whose transient effects are terminated by dephosphorylation by protein phosphatases. More recently, tyrosine phosphorylation has been observed for various ligand-gated channels (Hopfield et al., 1988; Wang and Salter, 1994; Moss et al., 1995; Valenzuela et al., 1995) and voltage-gated channels (Huang et al., 1993; Wilson and Kaczmarek, 1993;Lev et al., 1995; Holmes et al., 1996a,b; Jonas et al., 1996) and can alter channel activity and synaptic transmission (Llinas et al., 1997), but the physiological significance of this process is unproven. To examine the physiological role of tyrosine phosphorylation in modulating ion channels, we have been characterizing the properties of channels in identified neurons that are modulated by serotonin [5-hydroxytryptamine (5-HT)].

5-HT is a modulator of neurons in a wide variety of species (Peroutka, 1993; Martin and Humphrey, 1994). We have shown that 5-HT modulates the receptive field properties of identified mechanosensory neurons of the leech (Mar and Drapeau, 1996). When these cells are isolated in culture, 5-HT retains its neuromodulatory effect, permitting a detailed analysis of the signal transduction pathway. In pressure-sensitive (P) neurons, 5-HT binding to a 5-HT2 receptor activates protein kinase C (PKC) (Sanchez-Armass et al., 1991), as in mammalian neurons (Peroutka, 1993; Martin and Humphrey, 1994), which then increases the activity of a cation channel (Drapeau, 1990; Catarsi and Drapeau, 1992,1993). The channels are also activated when tyrosine phosphorylation is reversed by exposing intact P cells to inhibitors of tyrosine kinases or by treating isolated membrane patches with a catalytically active tyrosine phosphatase (Aniksztejn et al., 1997).

These observations suggest a constitutive suppression of channel activity by tyrosine phosphorylation. We tested for an interaction between PKC and tyrosine dephosphorylation during serotonergic modulation by treating cultured P cells or isolated membrane patches with vanadate, an inhibitor of tyrosine phosphatases (Swarup et al., 1982), before activating the channels with 5-HT, phorbol myristate acetate (PMA) (a membrane-permeant activator of PKC), or directly with PKC. In all cases the effects of the modulators were blocked, indicating that reversal of constitutive tyrosine phosphorylation underlies neuromodulation of the cation channels.

MATERIALS AND METHODS

P cells were isolated from Hirudo medicinalis and cultured as described previously (Dietzel et al., 1986). Because 5-HT activates both chloride and cation conductances in the P cell (Sanchez-Armass et al., 1991), the latter was isolated during intracellular recordings by blocking potassium channels with 10 mm diaminopyridine and by replacing external chloride with sulfate in the recording solution (Catarsi et al., 1995), which contained 130 mm Na2SO4, 4 mm K2SO4, 5 mmMgSO4, 10 mm 3,4-diaminopyridine, 10 mm glucose, and 10 mm HEPES, pH 7.4, adjusted to 330 mOsm. In some experiments, chloride was replaced with gluconate, which gave similar responses to 5-HT and enhanced cell survival. Intracellular microelectrodes were filled with 4 mpotassium acetate and had resistances of 40–50 MΩ. 5-HT was applied by pressure ejection (15 psi) from a pipette with a tip opening of 5 μm. Sodium pervanadate was prepared freshly by mixing 1 part 500 mm H2O2 with 50 parts 10 mm sodium orthovanadate dissolved in saline; the solution was incubated for 15 min at room temperature to reduce H2O2 (Wallace, 1995). The mixture was then added at a final concentration of 100 μm 1 hr before the recordings were started.

Cell-attached recordings of cation channels were obtained as described previously (Drapeau, 1990; Catarsi and Drapeau, 1993). Briefly, cell-attached patches were formed using 5–10 MΩ electrodes filled with the same solution used to bathe the cells: 155 mmNaCl, 5 mm KCl, 1 mm MgCl2, 1 mm CaCl2, 10 mm glucose, and 10 mm HEPES, pH 7.4, adjusted to 330 mOsm. PMA or genistein (LC Services Corp.) were diluted 2000- or 1000-fold, respectively, from a stock solution made in DMSO to yield a final concentration of 0.5 or 30 μm, respectively. This diluted DMSO concentration had no effect on the cellular properties. Recordings were filtered at 1 kHz (−3 dB), digitized at 10 kHz, and stored for later analysis using pClamp software (Axon Instruments, Foster City, CA).

Inside-out recordings were obtained as described previously (Catarsi and Drapeau, 1992, 1993) and differed from the procedure described above for cell-attached patches in that 1.5 mm EGTA was added to the solutions to reduce the free calcium concentration and facilitate patch excision. Rat brain PKC (Terochem Laboratories) was added at a final concentration of 1 pU/μl in the presence of 1 mm ATP and 0.1 μm PMA. Truncated CD45 T-cell protein tyrosine phosphatase (PTPase) was a gift from Dr. E. H. Fischer (University of Washington, Seattle, WA) and was added at a final concentration of 50 nm.

RESULTS

As a measure of its macroscopic effect, 5-HT (1 mm) was applied briefly (for 100 msec) onto P cells in culture while the membrane potential was recorded with an intracellular microelectrode. As shown in Figure1A,B, this resulted in a prolonged depolarization (lasting several seconds) of the P cells (peak depolarization of 3.1 mV ± 0.5; n = 14). When the P cells were exposed to sodium pervanadate (100 μm), a membrane-permeant form of vanadate, before testing the response to 5-HT, this resulted in a significant suppression of the depolarization (0.4 mV ± 0.2; n = 12; p < 0.001 by ANOVA) (Fig. 1A,B). Pervanadate applied alone did not affect the physiological properties of the P cells (Aniksztejn et al., 1997).

Fig. 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 1.

Effect of pervanadate on P cell depolarization by 5-HT. A, The net voltage response (relative to the resting potential of −50 mV) to brief application of 5-HT was measured in cultured P cells using an intracellular electrode in either the absence (Normal) or presence of pervanadate. B, Histogram summarizing the voltage responses recorded in control P cells and in P cells treated with pervanadate. The number of experiments (n) is indicated for each group, and the error bars represent the SEM.

Our previous work has shown that the depolarizing response to 5-HT is attributable to activation of cation channels by PKC, which increases the frequency of channel openings without affecting the mean open time or current amplitude (Drapeau, 1990; Catarsi and Drapeau, 1992). The channels are easily recognized during patch-clamp recordings because of their characteristic properties, which include spontaneous (frequency of ∼1 Hz), brief (duration of ∼1 msec), and large channel openings (conductance of 60 pS) at the resting membrane potential (−50 mV) in the absence of 5-HT. We therefore examined whether pervanadate could inhibit the activation of cation channels. In cell-attached recordings, activation of endogenous PKC by PMA (0.5 μm) increased channel activity (3.6 ± 0.8-fold increase; n = 14); the mean open time and current amplitude were unaffected (Fig.2A,C), as reported previously (Drapeau, 1990; Catarsi and Drapeau, 1993). Because the patches had more than one channel, we could not determine accurately the close times, which were expected to be reduced to account for the increased rate of channel openings that we observed. When the P cells were exposed to pervanadate before application of PMA, the latter resulted in a small stimulation of channel activity (1.6 ± 0.3-fold increase; n = 14) (Fig. 2B,C). This effect was significantly lower than in the absence of pervanadate (p = 0.02), and the mean open time and current amplitude were unaffected. Application of pervanadate alone had no effect (p = 0.2) on channel activity (1.2 ± 0.2; n = 7) (Fig. 2C).

Fig. 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 2.

Effect of pervanadate on PMA activation of cation channels in cell-attached patches. A, Current traces (20 sec) of cation channel activity at the resting potential in a cell-attached patch before (left) and after (right) application of 0.5 μm PMA (indicated by the arrow). Note that PMA increased the number of channel openings. The current amplitude, probability of channel opening (Po ), and mean open times (MOTs) were −2.1 pA, 0.0006 and 0.89 msec before, and −1.9 pA, 0.0034 and 0.89 msec after addition of PMA. B, Current recordings (20 sec) in the presence of 100 μm pervanadate before (left) and after (right) the application of PMA. Note that in contrast to the traces depicted inA, PMA no longer increased channel activity. The values for the current amplitude, Po, and MOT were −2.8 pA, 0.0013 and 0.75 msec before, and −2.8 pA, 0.0009 and 0.70 msec after addition of PMA. C, Histogram summarizing the ratio of Po after relative to before (Po/Pocontrol) treatments with PMA (left), PMA in pervanadate (middle), and pervanadate alone (right). The dashed line indicates the basal activity level.

To examine whether inhibition of tyrosine phosphatases could suppress PKC modulation of cation channels isolated from P cells, we excised membrane patches in the inside-out configuration. When patches were held at a potential of 50 mV (i.e., −50 mV in the pipette) and PKC was added to the bathing solution, channel activity increased (4.6 ± 0.8-fold increase; n = 10); the mean open time and current amplitude were unaffected (Fig.3A,C), as reported previously (Catarsi and Drapeau, 1992). When the patches were exposed to orthovanadate before the addition of PKC (Fig. 3B,C), no increase in channel activity was observed (1.1 ± 0.3-fold increase; n= 8); this result was significantly different from the one observed with PKC in the absence of orthovanadate (p < 0.001). Exposure of patches to orthovanadate alone (0.9 ± 0.3-fold effect; n = 8) (Fig. 3C) or exposure to orthovanadate after pretreating P cells with PMA (1.1 ± 0.2; n = 5) (Fig.4B,C) had no significant effect on channel activity.

Fig. 3.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 3.

Effect of orthovanadate on PKC activation of cation channels in inside-out patches. A, Current traces (20 sec) of cation channel activity at 50 mV (i.e., −50 mV in the pipette) in an inside-out patch before (left) and after (right) application of PKC (indicated by thearrow). Note that PKC increased the number of channel openings. The values for current amplitude,Po, and MOT were 2.2 pA, 0.0025 and 1.46 msec before, and 1.9 pA, 0.0126 and 1.17 msec after addition of PMA. B, Current recordings (20 sec) in the presence of 100 μm orthovanadate before (left) and after (right) the application of PMA. Note that in contrast to the traces depicted in A, PKC no longer increased channel activity. The values for current amplitude,Po, and MOT were 2.6 pA, 0.0077 and 1.71 msec before, and 2.6 pA, 0.0078 and 1.71 msec after addition of PMA. C, Histogram summarizing the ratio ofPo after relative to before (Po/Po control) treatments with PKC (left), PKC in orthovanadate (middle), and orthovanadate alone (right).

Fig. 4.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 4.

Top.A, Current traces (20 sec) of cation channel activity at 50 mV in an inside-out patch excised from a P cell treated with 0.5 μm PMA before (left) and after (right) application of 50 nm PTPase (indicated by the arrow). Note that no variation of channel activity was observed. The values for current amplitude, Po, and MOT were 3.1 pA, 0.1157 and 1.36 msec before, and 3.2 pA, 0.0880 and 1.83 msec after addition of PTPase. B, Current traces (20 sec) of cation channel activity at 50 mV in an inside-out patch excised from a P cell treated with 0.5 μm PMA before (left) and after (right) application of 100 μm orthovanadate (indicated by thearrow). Note that no variation in channel activity was observed. The values for current amplitude,Po, and MOT were 3.4 pA, 0.0596 and 1.52 msec before, and 3.4 pA, 0.0741 and 1.77 msec after addition of orthovanadate. C. Histogram summarizing the ratio ofPo after relative to before (Po/Po control) treatments with PTPase (left) or orthovanadate (middle) in PMA, and PMA in 30 μmgenistein (right).

We next examined whether PKC and tyrosine phosphatase activities had additive effects or acted sequentially, i.e., with PKC activating a tyrosine phosphatase. As shown previously (Aniksztejn et al., 1997), exposing inside-out patches to a recombinant, catalytically active fragment of the CD45 T-cell PTPase (Zander et al., 1991) resulted in a 10.3 ± 3.6-fold increase in channel activity. In contrast, inside-out patches excised from P cells treated with PMA were not activated further by the addition of PTPase (1.4 ± 0.6-fold effect; n = 5) (Fig. 4A,C). This effect was significantly lower than that obtained with PTPase alone (p < 0.04), suggesting a lack of additive effects. As an alternative, we pretreated P cells with the tyrosine kinase inhibitor genistein (Akiyama et al., 1987), because this should result in a progressive dephosphorylation by endogenous tyrosine phosphatases, previously shown to increase channel activity (Aniksztejn et al., 1997). After treatment with genistein, PMA failed to increase channel activity (Fig. 4C) (1.0 ± 0.1-fold effect;n = 4), a result significantly different from that obtained with PMA alone (p < 0.05) and consistent with a lack of additive effects. Taken together, these results suggest that PKC activates a tyrosine phosphatase activity sequentially, rather than having independent, additive effects on the channels.

DISCUSSION

Our results show that the stimulation of cation channels by 5-HT, or by its intracellular mediator PKC, is blocked when tyrosine phosphatases are inhibited, suggesting that reversal of constitutive tyrosine phosphorylation increases the activity of these channels (Fig.5). Because this is also true for channels in isolated membrane patches, the simplest interpretation is that a tyrosine phosphatase is closely associated with the channels in the patch and is activated by PKC, as has been shown for some tyrosine phosphatases (Zor et al., 1993; Kiyomoto et al., 1994). This conclusion is supported by the lack of evidence for additive effects of PKC and tyrosine phosphatase. An alternative possibility is that PKC may inhibit a tyrosine kinase, but this is unlikely because we have shown that inhibitors of tyrosine kinases have no effect on the responses to 5-HT (Ching et al., 1993) and PMA (Catarsi and Drapeau, 1993). No “run-up” of activity was observed with inside-out patches in the absence of ATP (Catarsi and Drapeau, 1992; Aniksztejn et al., 1997), indicating that the phosphatase activity is inactive at rest. Neuromodulation by 5-HT is thus likely attributable to PKC activation of an otherwise inactive tyrosine phosphatase that then dephosphorylates the cation channels with which it is intimately associated. Interestingly, the mode of gating of a cation channel inAplysia bag cells is switched to a higher activity pattern by a tyrosine phosphatase that is regulated by a serine/threonine kinase within membrane patches, suggesting a close association between the phosphatase and the channel (Wilson and Kaczmarek, 1993). Our results suggest that the regulation of channel-associated tyrosine phosphatases may be a final step in the signaling cascades of serine/threonine kinases during neuromodulation. In contrast to signaling by transient serine/threonine phosphorylation of target molecules, transient tyrosine dephosphorylation may signal neuromodulation.

Fig. 5.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 5.

A model for neuromodulation by tyrosine dephosphorylation. Under resting conditions, cation channels are constitutively tyrosine phosphorylated. 5-HT stimulates PKC, which in turn activates a protein tyrosine phosphatase (PTPase) closely associated with the channels, resulting in their dephosphorylation and increased activity.

Tyrosine phosphorylation is emerging as an important signal not only for early events during cellular differentiation (Cantley et al., 1991;Greenwald and Rubin, 1992) but also for neuronal plasticity in the mature nervous system. Tyrosine kinases prevent rundown of the NMDA class of glutamate receptors (Wang and Salter, 1994), which are the major tyrosine phosphorylated proteins in postsynaptic densities in the mammalian brain (Moon et al., 1994). Furthermore, pharmacological inhibition (O’Dell et al., 1991) or genetic deletion of tyrosine kinases (Grant et al., 1992) alters long-term potentiation in the hippocampus, a process requiring activation of NMDA receptors. In contrast to tyrosine kinases, the roles of tyrosine phosphatases are not well defined (Fischer et al., 1991), but they are expected to be crucial for regulating the phosphorylation of ion channels during neuronal signaling and development and may be important for signal transduction in general.

Footnotes

  • This work was supported by a Medical Research Council (MRC) of Canada Fellowship to S.C. and by a Fonds de la Recherche en Santé du Québec Senior Research Scholarship and Medical Research Council grant to P.D. We thank P. V. Nguyen for his critical reading of this manuscript.

    Correspondence should be addressed to Dr. Pierre Drapeau, Department of Neurology, Montreal General Hospital, 1650 Cedar Avenue, Montreal, Quebec, Canada H3G 1A4.

REFERENCES

  1. ↵
    1. Akiyama T,
    2. Ishida J,
    3. Nakagawa S,
    4. Ogawara H,
    5. Watanabe S,
    6. Itoh N,
    7. Shibuya M,
    8. Fukami Y
    (1987) Genistein, a specific inhibitor of tyrosine-specific protein kinases. J Biol Chem 262:5592–5595.
    OpenUrlAbstract/FREE Full Text
  2. ↵
    1. Aniksztejn L,
    2. Catarsi S,
    3. Drapeau P
    (1997) Channel modulation by tyrosine phosphorylation in an identified leech neuron. J Physiol (Lond) 498:135–142.
    OpenUrlCrossRefPubMed
  3. ↵
    1. Cantley LC,
    2. Auger KR,
    3. Carpenter C,
    4. Duckworth B,
    5. Graziani A,
    6. Kapeller R,
    7. Soltoff S
    (1991) Oncogenes and signal transduction. Cell 64:281–302.
    OpenUrlCrossRefPubMed
  4. ↵
    1. Catarsi S,
    2. Drapeau P
    (1992) Loss of extrasynaptic channel modulation by protein kinase C underlies the selection of serotonin responses in an identified leech neuron. Neuron 8:275–281.
    OpenUrlCrossRefPubMed
  5. ↵
    1. Catarsi S,
    2. Drapeau P
    (1993) Tyrosine kinase-dependent selection of transmitter responses induced by neural contact. Nature 363:353–355.
    OpenUrlCrossRefPubMed
  6. ↵
    1. Catarsi S,
    2. Ching S,
    3. Merz DC,
    4. Drapeau P
    (1995) Tyrosine phosphorylation during synapse formation between identified leech neurons. J Physiol (Lond) 485:775–786.
    OpenUrlPubMed
  7. ↵
    1. Ching S,
    2. Catarsi S,
    3. Drapeau P
    (1993) Selection of transmitter responses at sites of neurite contact during synapse formation between identified leech neurons. J Physiol (Lond) 468:425–439.
    OpenUrlPubMed
  8. ↵
    1. Dietzel ID,
    2. Drapeau P,
    3. Nicholls JG
    (1986) Voltage dependence of 5-hydroxytryptamine release at a synapse between identified leech neurons in culture. J Physiol (Lond) 372:191–205.
    OpenUrlCrossRefPubMed
  9. ↵
    1. Drapeau P
    (1990) Loss of channel modulation by transmitter and protein kinase C during innervation of an identified leech neuron. Neuron 4:875–882.
    OpenUrlCrossRefPubMed
  10. ↵
    1. Fischer EH,
    2. Charbonneau H,
    3. Tonks NK
    (1991) Protein tyrosine phosphatases: a diverse family of intracellular and transmembrane enzymes. Science 253:401–406.
    OpenUrlAbstract/FREE Full Text
  11. ↵
    1. Grant SGN,
    2. O’Dell TJ,
    3. Karl KA,
    4. Stein PL,
    5. Soriano SP,
    6. Kandel ER
    (1992) Impaired long-term potentiation, spatial learning, and hippocampal development in fyn mutant mice. Nature 258:1903–1910.
    OpenUrl
  12. ↵
    1. Greenwald I,
    2. Rubin GM
    (1992) Making a difference: the role of cell-cell interactions in establishing separate identities for equivalent cells. Cell 68:271–281.
    OpenUrlCrossRefPubMed
  13. ↵
    1. Holmes TC,
    2. Fadool DA,
    3. Levitan IB
    (1996a) Tyrosine phosphorylation of the Kv1.3 potassium channel. J Neurosci 16:1581–1590.
    OpenUrlAbstract/FREE Full Text
  14. ↵
    1. Holmes TC,
    2. Fadool DA,
    3. Ren R,
    4. Levitan IB
    (1996b) Association of Src tyrosine kinase with a human potassium channel mediated by SH3 domain. Science 274:2089–2091.
    OpenUrlAbstract/FREE Full Text
  15. ↵
    1. Hopfield JF,
    2. Tank DW,
    3. Greengard P,
    4. Huganir RL
    (1988) Functional modulation of the nicotinic acetylcholine receptor by tyrosine phosphorylation. Nature 336:677–680.
    OpenUrlCrossRefPubMed
  16. ↵
    1. Huang X-Y,
    2. Morielli AD,
    3. Peralta EG
    (1993) Tyrosine kinase-dependent suppression of a potassium channel by the G protein-coupled m1 muscarinic acetylcholine receptor. Cell 75:1145–1156.
    OpenUrlCrossRefPubMed
  17. ↵
    1. Jonas EA,
    2. Knox RJ,
    3. Kaczmarek LK,
    4. Schwartz JH,
    5. Solomon DH
    (1996) Insulin receptor in Aplysia neurons: characterization, molecular cloning, and modulation of ion currents. J Neurosci 16:1645–1658.
    OpenUrlAbstract/FREE Full Text
  18. ↵
    1. Kaczmarek LK,
    2. Levitan IB
    (1987) Neuromodulation: the biochemical control of neuronal excitability. (Oxford UP, New York).
  19. ↵
    1. Kiyomoto H,
    2. Fouqueray B,
    3. Abboud HE,
    4. Choudhury GG
    (1994) Phorbol 12-myristate 13-acetic acid inhibits PTP1B activity in human mesangial cells: a possible mechanism of enhanced tyrosine phosphorylation. FEBS Lett 353:217–220.
    OpenUrlCrossRefPubMed
  20. ↵
    1. Lev S,
    2. Moreno H,
    3. Martinez R,
    4. Canoll P,
    5. Peles E,
    6. Musacchio JM,
    7. Plowman GD,
    8. Rudy B,
    9. Schlessinger J
    (1995) Protein tyrosine kinase PYK2 involved in Ca2+-induced regulation of ion channel and MAP kinase functions. Nature 376:737–745.
    OpenUrlCrossRefPubMed
  21. ↵
    1. Llinas R,
    2. Moreno H,
    3. Sugimori M,
    4. Mohammadi M,
    5. Schlessinger J
    (1997) Differential pre- and postsynaptic modulation of chemical transmission in the squid giant synapse by tyrosine phosphorylation. Proc Natl Acad Sci USA 94:1990–1994.
    OpenUrlAbstract/FREE Full Text
  22. ↵
    1. Mar A,
    2. Drapeau P
    (1996) Modulation of conduction block in leech mechanosensory neurons. J Neurosci 16:4335–4343.
    OpenUrlAbstract/FREE Full Text
  23. ↵
    1. Martin GR,
    2. Humphrey PPA
    (1994) Receptors for 5-hydroxytryptamine: current perspectives on classification and nomenclature. Neuropharmacology 33:261–273.
    OpenUrlCrossRefPubMed
  24. ↵
    1. Moon IS,
    2. Apperson ML,
    3. Kennedy MB
    (1994) The major tyrosine-phosphorylated protein in the postsynaptic density fraction is N-methyl-d-aspartate receptor subunit 2B. Proc Natl Acad Sci USA 91:3954–3958.
    OpenUrlAbstract/FREE Full Text
  25. ↵
    1. Moss SJ,
    2. Gorrie GH,
    3. Amato A,
    4. Smart TG
    (1995) Modulation of GABAA receptors by tyrosine phosphorylation. Nature 377:344–348.
    OpenUrlCrossRefPubMed
  26. ↵
    1. O’Dell TJ,
    2. Kandel ER,
    3. Grant SGN
    (1991) Long-term potentiation in the hippocampus is blocked by tyrosine kinase inhibitors. Nature 353:558–560.
    OpenUrlCrossRefPubMed
  27. ↵
    1. Peroutka SJ
    (1993) 5-hydroxytryptamine receptors. J Neurochem 60:408–416.
    OpenUrlCrossRefPubMed
  28. ↵
    1. Sanchez-Armass S,
    2. Merz DC,
    3. Drapeau P
    (1991) Distinct receptors, second messengers and conductances underlying the dual responses to serotonin in an identified leech neuron. J Exp Biol 155:531–547.
    OpenUrlAbstract/FREE Full Text
  29. ↵
    1. Swarup G,
    2. Cohen S,
    3. Garbers DL
    (1982) Inhibition of membrane phosphotyrosyl-protein phosphatase activity by vanadate. Biochem Biophys Res Commun 107:1104–1109.
    OpenUrlCrossRefPubMed
  30. ↵
    1. Valenzuela CF,
    2. Machu TK,
    3. McKernan RM,
    4. Whiting P,
    5. VanRenterghem BB,
    6. McManaman JL,
    7. Brozowski SJ,
    8. Smith GB,
    9. Olsen RW,
    10. Harris RA
    (1995) Tyrosine kinase phosphorylation of GABAA receptors. Mol Brain Res 31:165–172.
    OpenUrlCrossRefPubMed
  31. ↵
    1. Wallace BG
    (1995) Regulation of interaction of nicotinic acetylcholine receptors with the cytoskeleton by agrin-activated protein tyrosine kinase. J Cell Biol 128:1121–1129.
    OpenUrlAbstract/FREE Full Text
  32. ↵
    1. Wang YT,
    2. Salter MW
    (1994) Regulation of NMDA receptors by tyrosine kinases and phosphatases. Nature 369:233–235.
    OpenUrlCrossRefPubMed
  33. ↵
    1. Wilson GF,
    2. Kaczmarek LK
    (1993) Mode-switching of a voltage-gated cation channel is mediated by a protein kinase A-regulated tyrosine phosphatase. Nature 366:433–438.
    OpenUrlCrossRefPubMed
  34. ↵
    1. Zander NF,
    2. Lorenzen JA,
    3. Cool DE,
    4. Tonks NK,
    5. Daum G,
    6. Krebs EG,
    7. Fischer EH
    (1991) Purification and characterisation of a human recombinant T-cell protein-tyrosine-phosphatase from a bacculovirus expression system. Biochemistry 30:6964–6970.
    OpenUrlCrossRefPubMed
  35. ↵
    1. Zor U,
    2. Ferber E,
    3. Gergely P,
    4. Szucs K,
    5. Dombradi V,
    6. Goldman R
    (1993) Reactive oxygen species mediate phorbol ester-regulated tyrosine phosphorylation and phospholipase A2 activation: potentiation by vanadate. Biochem J 295:879–888.
    OpenUrlAbstract/FREE Full Text
Back to top

In this issue

The Journal of Neuroscience: 17 (15)
Journal of Neuroscience
Vol. 17, Issue 15
1 Aug 1997
  • Table of Contents
  • Index by author
Email

Thank you for sharing this Journal of Neuroscience article.

NOTE: We request your email address only to inform the recipient that it was you who recommended this article, and that it is not junk mail. We do not retain these email addresses.

Enter multiple addresses on separate lines or separate them with commas.
Requirement for Tyrosine Phosphatase during Serotonergic Neuromodulation by Protein Kinase C
(Your Name) has forwarded a page to you from Journal of Neuroscience
(Your Name) thought you would be interested in this article in Journal of Neuroscience.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Print
View Full Page PDF
Citation Tools
Requirement for Tyrosine Phosphatase during Serotonergic Neuromodulation by Protein Kinase C
Stefano Catarsi, Pierre Drapeau
Journal of Neuroscience 1 August 1997, 17 (15) 5792-5797; DOI: 10.1523/JNEUROSCI.17-15-05792.1997

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Respond to this article
Request Permissions
Share
Requirement for Tyrosine Phosphatase during Serotonergic Neuromodulation by Protein Kinase C
Stefano Catarsi, Pierre Drapeau
Journal of Neuroscience 1 August 1997, 17 (15) 5792-5797; DOI: 10.1523/JNEUROSCI.17-15-05792.1997
Reddit logo Twitter logo Facebook logo Mendeley logo
  • Tweet Widget
  • Facebook Like
  • Google Plus One

Jump to section

  • Article
    • Abstract
    • MATERIALS AND METHODS
    • RESULTS
    • DISCUSSION
    • Footnotes
    • REFERENCES
  • Figures & Data
  • Info & Metrics
  • eLetters
  • PDF

Keywords

  • single channel
  • serotonin
  • protein kinase C
  • tyrosine phosphorylation
  • identified neuron
  • leech

Responses to this article

Respond to this article

Jump to comment:

No eLetters have been published for this article.

Related Articles

Cited By...

More in this TOC Section

  • Choice Behavior Guided by Learned, But Not Innate, Taste Aversion Recruits the Orbitofrontal Cortex
  • Maturation of Spontaneous Firing Properties after Hearing Onset in Rat Auditory Nerve Fibers: Spontaneous Rates, Refractoriness, and Interfiber Correlations
  • Insulin Treatment Prevents Neuroinflammation and Neuronal Injury with Restored Neurobehavioral Function in Models of HIV/AIDS Neurodegeneration
Show more Articles
  • Home
  • Alerts
  • Visit Society for Neuroscience on Facebook
  • Follow Society for Neuroscience on Twitter
  • Follow Society for Neuroscience on LinkedIn
  • Visit Society for Neuroscience on Youtube
  • Follow our RSS feeds

Content

  • Early Release
  • Current Issue
  • Issue Archive
  • Collections

Information

  • For Authors
  • For Advertisers
  • For the Media
  • For Subscribers

About

  • About the Journal
  • Editorial Board
  • Privacy Policy
  • Contact
(JNeurosci logo)
(SfN logo)

Copyright © 2023 by the Society for Neuroscience.
JNeurosci Online ISSN: 1529-2401

The ideas and opinions expressed in JNeurosci do not necessarily reflect those of SfN or the JNeurosci Editorial Board. Publication of an advertisement or other product mention in JNeurosci should not be construed as an endorsement of the manufacturer’s claims. SfN does not assume any responsibility for any injury and/or damage to persons or property arising from or related to any use of any material contained in JNeurosci.