Tipe Regulates Na: - Dependent Repetitive Firing in Drosophila Neurons
Tipe Regulates Na: - Dependent Repetitive Firing in Drosophila Neurons
1044-7431/02 $35.00
© 2002 Elsevier Science (USA)
402 All rights reserved.
tipE Regulates Neuronal Excitability 403
TABLE 2
Comparison of Electrical Properties in Wild-Type and tipE ⫺ Neurons at 2–3 DIV
Wild type 6.5 ⫾ 1.2 (35) 1.3 ⫾ 0.1 (43) ⫺38.0 ⫾ 2.0 (53) 22.3 ⫾ 1.3 (45)
tipE ⫺ 6.7 ⫾ 1.0 (33) 1.4 ⫾ 0.2 (46) ⫺37.0 ⫾ 2.1 (43) 19.0 ⫾ 1.2 (54)
studies comparisons included analysis of neurons in firing classes were readily apparent between tipE ⫺ and
tipE ⫺:tipE ⫹ cultures examined after heat shock and in wild-type neurons. In four independent experiments,
sibling cultures that were maintained continuously at half of the tipE ⫺:tipE ⫹ cultures were exposed to two 1-h
ambient temperature. heat shocks at 42 and 49 h after plating while the
To address the role of tipE in neuronal excitability, remaining cultures were not heat shocked (Fig. 5, top).
independent of development, tipE ⫺:tipE ⫹ cultures were To control for the affects of heat shock alone, wild-type
grown in the absence of heat shock for the first 2 days cultures prepared in parallel were exposed to the same
in vitro, by which time differences in the three major heat-shock regime. Cultures were coded and examined
blind with respect to genotype and heat-shock condi-
tions. Examples of firing properties in three different
neurons recorded from a wild-type (⫹HS), a tipE ⫺:tipE ⫹
(⫺HS), and a tipE ⫺:tipE ⫹ (⫹HS) culture are illustrated
in Fig. 5A. In the absence of heat shock, there were very
few multiple-spiking neurons in the tipE ⫺:tipE ⫹ cul-
tures, with the majority of excitable cells split between
the graded multiple-spiking and single-spiking firing
classes (Fig. 5B), similar to the distribution seen previ-
ously in the tipE ⫺ cultures. These data demonstrate that
the level of wild-type tipE product in the absence of
heat shock does not rescue the mutant firing phenotype
in transgenic cultures. However, the altered firing type
distribution was fully rescued within 24 h after heat
shock (⫹HS), with over 90% of the neurons in the
FIG. 4. Heat shock drives expression of wild-type tipE mRNA in a multiple-spiking firing class and the remainder classi-
tipE ⫺ background. (A) Schematic representation of the tipE cDNA fied as single spiking, similar to the distribution seen in
with the orientation of primer pair M1 (radioactively labeled) and M2. wild-type cultures (⫹HS) examined in parallel (Fig. 5B).
Positions of two RsaI restriction enzyme sites, R1* (eliminated in the The percentage of spontaneously firing neurons, low in
tipE mutant) and R2 are shown. RsaI digestion of the PCR product
generated by amplification using M1/M2 yields a labeled fragment of
the tipE ⫺:tipE ⫹ cultures in the absence of heat shock
85 bp in tipE ⫺ and 66 bp in wild type. (B) Autoradiogram of RsaI (⫺HS), was also rescued following heat shock (⫹HS)
digests of PCR products from wild-type, tipE ⫺, and tipE ⫺:tipE ⫹ (⫾HS) (Fig. 5C). Unexpectedly, the reduced action potential
cultures. Only the 66-bp product is amplified in RNA prepared from amplitude in the transgenic neurons was not rescued
wild-type neurons. Amplification of only the 85-bp product is ob- following heat shock (Fig. 5D).
served in RNA prepared from tipE ⫺ neurons. Heat shock does not
alter expression of these products in either the wild-type or the tipE ⫺
neurons. RNA prepared from tipE ⫺:tipE ⫹ transgenic embryo cultures Repolarization-Dependent Recovery of Sodium
expresses both the wild-type and the mutant product even in the
absence of heat shock. However, following heat shock there is a Currents and Excitability in Wild-Type, Mutant,
dramatic increase in the relative abundance of the wild-type (66 bp) and Transgenic Neurons
versus the mutant (85 bp) message. Cultures were heat shocked by
incubation in a 37°C, 5% CO 2 incubator for 1 h at 16, 33, and 40 h after Previous studies had demonstrated that the wild-
plating (see Experimental Methods for details). Total RNA was ex- type tipE gene product upregulates the amplitude and
tracted 2 h after the last heat shock. alters the kinetic properties of para sodium currents in a
tipE Regulates Neuronal Excitability 407
cortical neurons described in the animal and in disso- tipE Regulates Sodium-Dependent
ciated cell culture (Connors and Gutnick, 1990; Massen- Repetitive Firing
gill et al., 1997). The single-spiking neurons in the Dro-
Assessment of the firing properties in primary neu-
sophila cultures are similar to the recently described
rons from genetic mutants is a useful strategy for ex-
on-spiking neurons found in the rodent auditory cortex,
amining the role of specific genes in regulating neuro-
which fire only one or two spikes that occur within 10
nal excitability. Alterations in spontaneous activity of
ms of the onset of a maintained intracellular depolar-
neurons cultured from Hyperkinetic mutant embryos
ization (Metherate and Aramakis, 1999). The presence
(Yao and Wu, 1999) supported an early study indicating
of similar firing classes in the cultured Drosophila neu-
that this gene, encoding a K channel  subunit, is in-
rons and rodent cortical neurons suggests strong con- volved in regulation of neuronal firing properties
servation of the functional elements contributing to (Ikeda and Kaplan, 1970). Our analysis of tipE ⫺ neurons
CNS circuitry between these distantly related species. revealed reductions in repetitive firing, spontaneous
It should also be noted that the resting potentials of firing, action potential amplitude, peak sodium current
the Drosophila neurons reported in this study are more density, and sodium current recovery during repeated
depolarized than is standard for many mature mamma- activation, suggesting that these are linked to each other
lian neurons. However, hyperpolarizing shifts in mem- and to tipE. Rescue experiments, involving expression
brane potential, from ⫺40 to ⫺65 mV, have been re- of the wild-type tipE transgene in tipE ⫺ neurons, con-
ported during early development in some populations firmed that tipE is important in regulation of repetitive
of mammalian cortical neurons (Agmon et al., 1996; firing, spontaneous firing, and the rate of recovery of
Zhou and Hablitz, 1996). This suggests that the depo- sodium currents during repeated activation. Our data
larized resting potentials could be related to the rela- also demonstrate that induction of tipE ⫹ expression in
tively young age at which most of the recordings were transgenic neurons beginning at 2 days, after neurons
obtained, 2–3 days of the neuronal birth date. More have already established their firing properties, is suf-
negative resting potentials of ⫺55 mV have been re- ficient to rescue the mutant firing phenotypes. This
ported for Drosophila “giant neurons” examined at suggests that regulation of tipE may play a role, not
slightly later stages, between 2 and 5 days in culture only in establishment of neuronal firing phenotype, but
(Yao and Wu, 1999). In addition, we have observed also in modulation of firing properties in differentiated
more hyperpolarized resting potentials (⫺55 mV) when neurons.
recordings are done at 3– 4 days (unpublished data). The slower rate of recovery of sodium currents dur-
Despite the depolarized resting potentials, intracellu- ing repetitive activation in tipE ⫺ neurons predicts that a
lar (whole cell) recordings revealed spontaneous action diminished sodium current will be available for gener-
potentials in the absence of current injection in some ation of the second spike in an action potential train in
neurons. This does not seem likely to be injury-induced the mutant neurons. This could thus contribute to the
spiking as spontaneously active neurons were observed decrease in probability of mutant neurons firing repet-
at a similar frequency in extracellular (cell attached) itively during sustained depolarization. Concomitant
recordings. Previous studies from our lab have also rescue of sodium current recovery and repetitive firing,
demonstrated the presence of action potential mediated following induction of the tipE ⫹ transgene in tipE ⫺ neu-
spontaneous excitatory postsynaptic currents in many rons, suggests linkage between these two phenotypes.
of these cultured neurons, in which activity in the pre- The difference in the level of recovery of sodium cur-
synaptic neuron is clearly independent of technical rents during repolarization seen between wild-type and
artifacts that could be potentially associated with mutant neurons, though significant, was not large (ap-
whole-cell recording electrodes (Lee and O’Dowd, proximately 10%), and therefore it was not clear how
1999). Finally, recordings from neurons in the Drosoph- this property might influence repetitive firing rates.
ila embryonic nerve cord have revealed large spontane- However, analysis of the recovery of excitability as a
ously active currents, thought to underlie action poten- function of interstimulus interval in the different firing
tials, in neurons held at ⫺40 mV, that were rarely seen classes is consistent with the suggestion that reduced
in those held at more hyperpolarized potentials (Baines rate of recovery of sodium currents contributes to the
and Bate, 1998). Together these findings suggest that decrease in repetitive firing in mutant neurons. In sin-
young embryonic Drosophila neurons, both in vivo and gle-spiking neurons, an interstimulus interval was re-
in vitro, are excitable at relatively depolarized voltages. quired for recovery of the ability to fire a second action
412 Hodges et al.
potential. In addition, the duration of the interstimulus logues may identify novel pathways involved in regu-
interval necessary to fire a second full-sized action po- lation of sodium currents that can influence action po-
tential was significantly longer in single- versus multi- tential propagation in mammalian neurons.
ple-spiking transgenic neurons. Most of the spontaneously firing neurons in wild-
In Drosophila, as in mammals, the sodium channels type cultures were in the multiple-spiking class. Alter-
that underlie the whole-cell sodium currents are tran- ations that decrease the probability of firing a second
siently activated by a sustained depolarizing voltage spike in the mutant neurons in response to depolariza-
step and recovery from inactivation requires return to tion could also decrease the probability of firing
hyperpolarized potentials (O’Dowd and Aldrich, 1988). spontaneously. However, additional changes in the un-
A decrease in the rate of recovery from inactivation of derlying currents may contribute to the reduced spon-
the underlying sodium channels is one mechanism that taneous activity in the mutant neurons. For example, in
could contribute to the reduced recovery of sodium oocytes, coexpression of the wild-type tipE product in-
currents seen in the tipE ⫺ neurons. Studies in other fluenced both the density and the fast decay kinetics of
systems have clearly demonstrated a relationship be- the para sodium currents (Warmke et al., 1997). The fast
tween rate of recovery of sodium channels from inacti- kinetic properties of sodium currents were not assessed
vation and repetitive firing. In hippocampal pyramidal in the present study due to inadequate voltage-clamp in
neurons spikes in the dendrites are attenuated by high- excitable cells. Therefore, tipE might also affect fast
frequency stimulation and this is correlated with a rel- gating properties of sodium channels that could con-
atively slow rate of recovery of sodium channels from tribute to the altered firing phenotypes observed.
inactivation (Colbert et al., 1997; Jung et al., 1997). A The oocyte studies further suggested that tipE might
computational model supports the hypothesis that de- be functioning like sodium channel  subunits (1 and
layed recovery of sodium channels from inactivation 2) as these are known to influence both expression
can result in attenuation of action potentials (Migliore, levels and fast kinetic properties of mammalian sodium
1996). Additionally, hyperexcitability characterized by channels (Isom et al., 1994). A newly identified  sub-
elevated firing frequencies in spinal sensory neurons unit (3), cloned from human and rat, has been shown
following injury has been associated with the emer- to influence the rate of sodium current recovery from
gence of sodium currents that recover rapidly from inactivation (Morgan et al., 2000), similar to the role
inactivation (Cummins and Waxman, 1997; Cummins et suggested for tipE by the present study. Our single-cell
al., 2000). However, in the present study the majority of RT-PCR analyses demonstrate that tipE is coexpressed
the data on sodium currents were obtained from neu- with para in most cells, and coimmunoprecipitation in
rons that could not be well voltage-clamped. Therefore, Xenopus oocytes suggests that the two proteins can
we cannot rule out the possibility that a use-dependent physically associate (L. M. Hall and C. Ericsson, unpub-
change in space constant, rather than a change in the lished results). Taken together these data suggest that,
sodium channel inactivation properties, could contrib- although tipE has little amino acid sequence identity
ute to the observed decrease in recovery of the currents. with sodium channel  subunits, it could be functioning
For example, a failure to reach the same membrane as an auxiliary subunit important in regulating sodium
potential during the two sequential depolarizing steps channel function in wild-type Drosophila neurons. A
could cause a reduction in amplitude of the sodium prediction of this hypothesis is that wild-type neurons
current evoked by the second pulse. We do not believe that fire multiple spikes express more tipE than those
this was a factor since the latency and waveform of the that fire only single action potentials. A quantitative
currents, also influenced by space constant, did not analysis of gene and/or protein levels, not undertaken
vary significantly between the two steps (Fig. 6B). In in the present studies, would be necessary to address
addition, for this mechanism to account for the differ- this question.
ences seen between tipE ⫺ and wild-type neurons and The inability of wild-type tipE transgene expression
the rescue by tipE ⫹, it would necessitate invoking to rescue the reduced sodium current density and ac-
genotype-specific differences in the properties of use- tion potential amplitude in transgenic neurons was sur-
dependent alterations in space clamp. In either case, our prising. It is possible that these features, while related to
rescue studies clearly demonstrate that tipE is impor- each other, are not necessarily linked to tipE. However,
tant for regulating recovery of sodium currents from we cannot rule out the possibility that tipE plays a role
repeated activation and sodium-dependent repetitive in regulation of sodium current density and action po-
firing. Therefore, isolation of vertebrate tipE ortho- tential amplitude in primary neurons. For example,
tipE Regulates Neuronal Excitability 413
induction conditions or the timing of the assay could be stimulation due to the reduced sodium current density
suboptimal for detecting regulation mechanisms in- and depressed recovery of sodium currents during re-
volving coassembly of tipE products with para sodium peated stimulation. Furthermore, a rise in temperature
channels prior to membrane insertion. In either case, speeds up the gating kinetics of all channels and may
rescue of the repetitive firing phenotype in the absence result in potassium currents overwhelming the altered
of restoration of sodium current density and action sodium currents leading to an even more pronounced
potential amplitude demonstrates that these can be alteration at elevated temperatures. This could contrib-
functionally separated. ute to the temperature-induced paralysis. The reduced
sodium current density and altered repolarization-
dependent sodium current recovery in tipE ⫺ mutant
tipE ⴙ Is Not Necessary for Repetitive Spiking
neurons could also contribute to the enhanced sensitiv-
in All Drosophila Neurons
ity to temperature-induced action potential blockade
Repetitive firing, spontaneous activity, and fast re- previously reported in the para;tipE double mutants
covery of sodium currents from repeated activation in (Ganetzky, 1986).
some of the tipE ⫺ neurons demonstrate that tipE ⫹ ex-
pression is not necessary for manifestation of these
electrophysiological phenotypes in all cultured neu- EXPERIMENTAL METHODS
rons. It is possible that a tipE homologue, identified in a
recent analysis of the Drosophila genome (Littleton and Drosophila stocks and cell culture. Embryos were
Ganetzky, 2000), encodes a protein that substitutes for collected from Canton-S homozygous wild-type, tipE
the mutant tipE in cells that fire repetitively. Alterna- sepia (tipE ⫺), and w;tipE sepia flies transformed with a
tively, the tipE mutant used in this study, an EMS- wild-type tipE cDNA under control of the heat-shock
induced recessive mutation (Kulkarni and Padhye, promoter (tipE ⫺:tipE ⫹) (Feng et al., 1995b). Neurons
1982) resulting in a premature stop codon, may act as a were prepared from midgastrula-stage embryos and
hypomorph rather than a true null (Feng et al., 1995b). cultured in Drosophila defined medium 1 (DDM1) at
The tipE ⫺ neurons with apparently wild-type properties 22–25°C and 4 –5% CO 2, as previously described
could be due to residual function of the mutant protein. (O’Dowd, 1995). Cultures stained with anti-horseradish
This second possibility seems less likely because previ- peroxidase (HRP) antibodies were fixed in 4% parafor-
ous studies have shown that mutant tipE was not able to maldehyde for 30 min at room temperature followed by
rescue adult paralysis (Feng et al., 1995b). Additionally, a 1-h incubation with fluorescein-conjugated anti-HRP
mutant tipE cRNA expressed in Xenopus oocytes does antibodies (1:100; Organon Teknika). Coverslips were
not enhance para sodium current expression (M. Chopra mounted on glass slides. Images were acquired with a
and L. M. Hall, unpublished observations). Spot cooled CCD camera (Diagnostic Instruments)
mounted on a Nikon Optiphot microscope and pre-
pared for presentation in Adobe PhotoShop.
Functional Significance
Electrophysiological recordings. To minimize po-
In nap and para mutants, a temperature-dependent tential bias in selection of cells for analysis, whole-cell
blockade of action potential propagation in larval motor recordings from wild-type, tipE ⫺, and tipE ⫺:tipE ⫹ trans-
nerves has been associated with the temperature-sensi- gene neurons were performed blind with respect to
tive paralysis (Wu and Ganetzky, 1992). In contrast, it genotype and heat-shock treatment. Unpolished re-
was unclear why tipE ⫺ larvae exhibit normal extracel- cording pipettes had open pipette resistances of 2–5
lularly recorded action potential propagation in motor M⍀. For assessment of firing properties the internal
nerves both at the behaviorally permissive and at the pipette solution contained (in mM): potassium glu-
nonpermissive temperature (Ganetzky, 1986). Our data conate (120), NaCl (20), EGTA (1.1), CaCl 2 (0.1), MgCl 2
demonstrate that tipE ⫺ neurons are capable of generat- (2), Hepes (10), pH 7.2. Sodium currents were examined
ing a single action potential in response to a discrete using an internal solution containing (in mM): d-glu-
stimulus, consistent with the apparently normal com- conic acid (120), cesium hydroxide (120), NaCl (20),
pound action potential recorded in larval motor neu- CaCl 2 (0.1), MgCl 2 (2), EGTA (1.1), Hepes (10), pH 7.2.
rons. However, our findings suggest that action poten- The external solution contained (in mM): NaCl (140),
tial propagation in tipE ⫺ mutant nerves could be KCl (3), MgCl 2 (4), CaCl 2 (1), Hepes (5), pH 7.2. A 5-mV
compromised during high-frequency, repetitive nerve liquid junction potential has been subtracted from all
414 Hodges et al.
TABLE 3
Primer Pairs Used in RT-PCR Studies
membrane potentials noted in this report. Whole-cell 1995) using the primer pairs shown in Table 3. Ampli-
capacitance was determined by measuring the area un- fied products, visualized by inclusion of 2–5 ⫻ 10 5 dpm
der the capacitative transient current record obtained of 32P-end-labeled forward primers in the PCR, were
immediately after break into the cell. Data were col- separated by electrophoresis on 8 or 10% nondenatur-
lected and analyzed using a List EPC-7 patch-clamp ing polyacrylamide gels. The amount of product was
amplifier, a Dell computer, and pCLAMP software quantified by phosphorimager analysis (Molecular Dy-
(Axon Instruments). All recordings were performed at namics, Sunnyvale, CA).
room temperature. Identification of wild-type and mutant tipE PCR
Heat-shock induction of tipE ⴙ expression in trans- products was performed by RsaI restriction enzyme
genic neurons. Cultures were prepared from midgas- analysis of an aliquot of the PCR products using stan-
trula-stage embryos obtained from wild-type, tipE ⫺, dard procedures (Sambrook et al., 1989). In the devel-
and tipE ⫺:tipE ⫹ flies. For PCR analysis of gene expres- opmental study a single reverse transcription reaction
sion, half of the cultures from each genotype were heat was performed on each RNA sample for each time
shocked by transfer to a 37°C, 5% CO 2 incubator for 1 h point. This was divided into three equal aliquots in
at 16, 33, and 40 h after plating. The remainder of the which PCR products were amplified using primers spe-
time they were maintained at ambient temperature (22– cific for ribosomal protein 49 (rp49) (21 cycles) or para or
25°C). The sibling cultures were maintained continu- tipE (25 cycles). Cycle numbers were chosen to yield
ously in a 5% CO 2 incubator at ambient temperature for products within the linear range of amplification. To
42 h. Total RNA was extracted at 42 h (2 h after the last minimize differences in reaction conditions, primers of
heat shock) from both control and heat-shocked cul- similar size and specific activities were used. Phospho-
tures. For the electrophysiological studies, half of the imager optical density measurements for developmen-
cultures from each genotype were heat shocked by tally regulated PCR products were normalized to opti-
transfer to a 37°C, 5% CO 2 incubator for 1 h at 42 and cal density values obtained from PCR amplification of
49 h after plating. The remainder of the time they were rp49, a mRNA that is not developmentally regulated
maintained in a 5% CO 2 incubator at ambient temper- (O’Connell and Rosbash, 1984). Single-cell amplifica-
ature. The sibling cultures were maintained continu- tion of total RNA aspirated from neurons after electro-
ously at ambient temperature. All electrophysiological physiological recordings was performed as previously
recordings were done at 66 –74 h after plating. described (O’Dowd et al., 1995).
RT-PCR analysis of gene expression in cultured neu- Primer pairs. para: To amplify a single product
rons. Total RNA from cultured neurons was prepared common to all para transcripts the primer set para-
using Tri-Reagent (Molecular Research Center, Inc., ComF/R was used. To examine the distribution of para
Cincinnati, OH) according to a single-step method transcripts containing alternatively spliced exons a and
(Chomczynski and Sacchi, 1987). First-strand cDNA i, a primer pair (paraDP3/DP4) flanking these exons
was generated by random-primed reverse transcription was used. tipE: For developmental profiles and single-
of total RNA, and PCR amplification of the cDNA was cell experiments, PCR amplification of tipE mRNA was
performed as previously described (O’Dowd et al., performed using the primer pair tipEComF/R. To dif-
tipE Regulates Neuronal Excitability 415
ferentiate between wild-type and tipE mutant mRNA, Feng, G., Deak, P., Kasbekar, D. P., Gil, D. W., and Hall, L. M. (1995b).
primers tipEM1/M2 were used. rp49: rp49F/R primer Cytogenetic and molecular localization of tipE: A gene affecting
sodium channels in Drosophila melanogaster. Genetics 139: 1679 –
pair was used for PCR amplification of ribosomal pro- 1688.
tein transcripts. Sequences, nucleotide positions, and Ganetzky, B. (1986). Neurogenetic analysis of Drosophila mutations
GenBank accession numbers for all of the primer sets affecting sodium channels: Synergistic effects on viability and
used are detailed in Table 3. nerve conduction in double mutants involving tipE. J. Neurogenet. 3:
19 –31.
Huguenard, J. R., Hamill, O. P., and Prince, D. A. (1988). Develop-
mental changes in sodium conductances in rat neocortical neurons:
ACKNOWLEDGMENTS Appearance of a slowly inactivating component. J. Neurophysiol. 59:
778 –794.
This work was supported by NIH Grants NS27501 and NS01854 to Ikeda, K., and Kaplan, W. (1970). Patterned neural activity of a mutant
D.K.O’D., NIH Grant NS16204 and an American Cancer Society Drosophila melanogaster. Proc. Natl. Acad. Sci. USA 66: 765–772.
Scholar Award to L.M.H., and American Heart Association Postdoc- Isom, L. L., DeJongh, K. S., and Catterall, W. A. (1994). Auxiliary
toral Grant 95-98 to D.D.H. subunits of voltage-gated ion channels. Neuron 12: 1183–1194.
Jackson, F. R., Wilson, S. D., and Hall, L. M. (1986). The tip-E muta-
tions of Drosophila decrease saxitoxin binding and interact with
other mutations affecting nerve membrane excitability. J. Neuro-
REFERENCES genet. 3: 19 –31.
Jung, H. Y., Mickus, T., and Spruston, N. (1997). Prolonged sodium
Agmon, A., Hollrigel, G., and O’Dowd, D. K. (1996). Functional channel inactivation contributes to dendritic action potential atten-
GABAergic synaptic connections in neonatal mouse barrel cortex. uation in hippocampal pyramidal neurons. J. Neurosci. 17: 6639 –
J. Neurosci. 16: 4684 – 4695. 6646.
Aizenman, C. D., and Linden, D. J. (2000). Rapid, synaptically driven Kulkarni, S. J., and Padhye, A. (1982). Temperature-sensitive paralytic
increases in the intrinsic excitability of cerebellar deep nuclear mutations on the second and third chromosomes of Drosophila
neurons. Nat. Neurosci. 3: 109 –111. melanogaster. Genet. Res. 40: 191–199.
Armano, S., Rossi, P., Taglietti, V., and D’Angelo, E. (2000). Long-term Lee, D., and O’Dowd, D. K. (1999). Fast excitatory synaptic transmis-
potentiation of intrinsic excitability at the mossy fiber-granule cell sion mediated by nicotinic acetylcholine receptors in Drosophila
synapse of rat cerebellum. J. Neurosci. 20: 5208 –5216. neurons. J. Neurosci. 19: 5311–5321.
Baines, R. A., and Bate, M. (1998). Electrophysiological development Littleton, J. T., and Ganetzky, B. (2000). Ion channels and synaptic
of central neurons in the Drosophila embryo. J. Neurosci. 18: 4673– organization: Analysis of the Drosophila genome. Neuron 26: 35– 43.
4683. Loughney, K., Kreber, R., and Ganetzky, B. (1989). Molecular analysis
Barish, M. E. (1986). Differentiation of voltage-gated potassium cur- of the para locus, a sodium channel gene in Drosophila. Cell 58:
rent and modulation of excitability in cultured amphibian spinal 1143–1154.
neurones. J. Physiol. 375: 225–229. Massengill, J. L., Smith, M. A., Son, D. I., and O’Dowd, D. K. (1997).
Catterall, W. A. (2000). From ionic currents to molecular mechanisms: Differential expression of K4-AP currents and Kv3.1 potassium
The structure and function of voltage-gated sodium channels. Neu- channel transcripts in cortical neurons that develop distinct firing
ron 26: 13–25. phenotypes. J. Neurosci. 17: 3136 –3147.
Chomczynski, P., and Sacchi, N. (1987). Single-step method of RNA Metherate, R., and Aramakis, V. B. (1999). Intrinsic electrophysiology
isolation by acid guanidinium thiocyanate–phenol– chloroform ex- of neurons in thalamorecipient layers of developing rat auditory
traction. Anal. Biochem. 162: 156 –159. cortex. Dev. Brain. Res. 115: 131–144.
Colbert, C. M., Magee, J. C., Hoffman, D. A., and Johnston, D. (1997). Migliore, M. (1996). Modeling the attenuation and failure of action
Slow recovery from inactivation of Na ⫹ channels underlies the potentials in the dendrites of hippocampal neurons. Biophys. J. 71:
activity-dependent attenuation of dendritic action potentials in hip- 2394 – 403.
pocampal CA1 pyramidal neurons. J. Neurosci. 17: 6512– 6521. Morgan, K., Stevens, E., Shah, B., Cox, P., Kixon, A., Lee, K., Pinnock,
Connors, B. W., and Gutnick, M. J. (1990). Intrinsic firing patterns of R., Hughes, J., Richardson, P., Mizuguchi, K., and Jackson, A.
diverse neocortical neurons. Trends Neurosci. 13: 99 –104. (2000). 3: An additional auxiliary subunit of the voltage-sensitive
Cummins, T., Dib-Hajj, S., Black, J., and Waxman, S. (2000). Sodium sodium channel that modulates channel gating with distinct kinet-
channels and the molecular pathophysiology of pain. In Progress in ics. Proc. Natl. Acad. Sci. USA 97: 2308 –2313.
Brain Research, pp. 3–19. Elsevier, Amsterdam. O’Connell, P. O., and Rosbash, M. (1984). Sequence, structure, and
Cummins, T., and Waxman, S. (1997). Downregulation of tetrodo- codon preference of the Drosophila ribosomal protein 49 gene.
toxin-resistant sodium currents and upregulation of rapidly Nucleic Acids Res. 12: 5495–5513.
repriming tetrodotoxin-sensitive sodium current in small spinal O’Dowd, D. K. (1995). Voltage-gated currents and firing properties of
sensory neurons after nerve injury. J. Neurosci. 17: 3503–3514. embryonic Drosophila neurons grown in a chemically defined me-
Desai, N. S., Rutherford, L. C., and Turrigiano, G. G. (1999). Plasticity dium. J. Neurobiol. 27: 113–126.
in the intrinsic excitability of cortical pyramidal neurons. Nat. Neu- O’Dowd, D. K., and Aldrich, R. W. (1988). Voltage-clamp analysis of
rosci. 2: 515–520. sodium channels in wildtype and mutant Drosophila neurons.
Feng, G., Deák, P., Chopra, M., and Hall, L. M. (1995a). Cloning and J. Neurosci. 8: 3633–3643.
functional analysis of tipE, a novel membrane protein that enhances O’Dowd, D. K., Gee, J. R., and Smith, M. A. (1995). Sodium current
Drosophila para sodium channel function. Cell 82: 1001–1011. density correlates with expression of specific alternatively spliced
416 Hodges et al.
sodium channel mRNAs in single neurons. J. Neurosci. 15: 4005– potentials associated with the Shaker complex locus of Drosophila.
4012. Proc. Natl. Acad. Sci. USA 78: 6548 – 6552.
O’Dowd, D. K., Germeraad, S., and Aldrich, R. W. (1989). Alterations Turrigiano, G., LeMasson, G., and Marder, E. (1995). Selective regu-
in the expression and gating of Drosophila sodium channels by lation of current densities underlies spontaneous changes in the
mutations in the para gene. Neuron 2: 1301–1311. activity of cultured neurons. J. Neurosci. 15: 3640 –3652.
O’Dowd, D. K., Ribera, A. B., and Spitzer, N. C. (1988). Development Warmke, J., Reenan, R., Wang, P., Qian, S., Arena, J., Wang, J.,
of voltage-dependent calcium, sodium and potassium currents in Wunderler, D., Liu, K., Kaczorowski, G., Wan der Ploeg, L.,
Xenopus spinal neurons. J. Neurosci. 8: 792– 805. Ganetzky, B., and Cohen, C. (1997). Functional expression of Dro-
Reenan, R. A., Hanrahan, C. J., and Ganetzky, B. (2000). The mle napts sophila para sodium channels: Modulation by the membrane pro-
RNA helicase mutation in Drosophila results in a splicing catastro-
tein tipE and toxin pharmacology. J. Gen. Physiol. 110: 119 –133.
phe of the para Na ⫹ channel transcript in a region of RNA editing.
Wu, C.-F., and Ganetzky, B. (1992). Neurogenetic studies of ion chan-
Neuron 25: 139 –149.
nels in Drosophila. In Ion Channels, pp. 261–314. Plenum, New York.
Sambrook, J. M., Fritsch, E. F., and Maniatis, T. (1989). Molecular
Yao, W. D., and Wu, C. F. (1999). Auxiliary hyperkinetic beta subunit
Cloning: A Laboratory Manual. Cold Spring Harbor Laboratory Press,
Cold Spring Harbor, New York. of K ⫹ channels: Regulation of firing properties and K ⫹ currents in
Spitzer, N. C. (1991). A developmental handshake: Neuronal control Drosophila neurons. J. Neurophysiol. 81: 2472–2484.
of ionic currents and their control of neuronal differentiation. Zhao, M.-L., and Wu, C.-F. (1997). Alterations in frequency coding
J. Neurobiol. 22: 659 – 673. and activity dependence of excitability in cultured neurons of Dro-
Spitzer, N. C., Gu, X., and Olson, E. (1994). Action potentials, calcium sophila memory mutants. J. Neurosci. 17: 2187–2199.
transients and the control of differentiation of excitable cells. Curr. Zhou, F.-M., and Hablitz, J. J. (1996). Postnatal development of mem-
Opin. Neurobiol. 4: 70 –77. brane properties of layer I neurons in rat neocortex. J. Neurosci. 16:
Tanouye, M. A., Ferrus, A., and Fujita, S. C. (1981). Abnormal action 1131–1139.