Cultured lobster olfactory receptor neurons (ORNs) were surveyed for their odor sensitivity with whole-cell, voltage-clamp recording. The nature of the adequate stimuli, the degree of tuning (response spectra) of the cells, the threshold of sensitivity and the dual polarity of the odor-evoked currents are consistent with chemosensitivity in the cultured ORNs being olfactory. The ability of odors to evoke currents in cultured ORNs that lack processes suggests that lobster ORNs can be induced in vitro to insert all the elements of the transduction cascade in the soma, including those that might normally be confined to processes. This should greatly facilitate analysis of olfactory transduction in these cells.

The small size and thin, elongated morphology of olfactory receptor neurons (ORNs) has made the study of olfactory transduction difficult. The advent of patch-clamp recording ameliorated this situation and facilitated progress towards understanding olfactory transduction (reviews: Anholt, 1991; Firestein, 1991). Central to this effort has been the ability to dissociate ORNs from their surrounding epithelium in order to study them directly or in sustained primary culture. Dissociated ORNs are not only accessible for patching, they often assume a more compact form than their counterparts in situ that allows a reasonable space-clamp and facilitates diffusion of membrane-impermeant probes from the electrode to the site of transduction. While transduction is thought to occur in the cilia (outer dendrites, in invertebrates) of the ORNs (Kurahashi, 1989; Firestein et al. 1990; Lowe and Gold, 1991), at least some elements of the transduction cascade are not confined to the cilia. Specifically, cyclic-AMP-gated cation channels that are the effectors in the transduction cascade in amphibian ORNs also occur on the dendrite and soma of the cells, although in lower densities than on the cilia (Firestein et al. 1991; Zufall et al. 1991a). Indeed, this variability in density was exploited to obtain favorable channel density for recording (Firestein et al. 1991; Zufall et al. 1991a).

Previously, we reported preliminary evidence that cultured lobster ORNs respond to odors regardless of whether the cells had sprouted processes (Fadool et al. 1991). This observation raises the interesting possibility that, in vitro, all elements of the transduction cascade may be expressed and inserted into the soma of ORNs that fail to sprout processes. Given the ease of patching the soma compared to the extremely thin cilia (outer dendrites) and the ability to culture ORNs from vertebrates (Calof and Chikaraishi, 1991; Coon et al. 1989; Ronnett et al. 1991; Pixley and Pun, 1990) and other invertebrates (Stengl et al. 1989; Zufall et al. 1991b), such a phenomenon could be of general use for studying olfactory transduction.

Without the normal polarity of the cell, however, it must be established that applied ‘odors’ are activating what would otherwise be ciliary (dendritic) chemoreceptors. The need to establish the adequacy of odor stimuli is particularly important when studying ORNs from aquatic animals such as fish and lobsters. Adequate olfactory stimuli for many aquatic animals are the blood-borne components of prey, compounds such as amino acids, amines and nucleotides (review: Carr, 1990). Such compounds could be expected to activate cells as neurotransmitters or neuromodulators. In order to establish the usefulness of cultured lobster ORNs for analysis of transduction mechanisms, we have attempted to obtain functional evidence that cultured lobster ORNs with no processes or varying numbers of processes are morphs of the same type of cell and that the odorevoked properties of the cultured cells reflect those of lobster ORNs in situ.

Tissue culture

The distinct clusters of the ORNs were dissected from the aesthetasc (olfactory) sensilla on the lateral antennular filament (olfactory organ) of adult specimens of the Caribbean spiny lobster, Panulirus argus Latreille. The clusters were enzymatically dissociated, and the resulting cells sustained in primary culture as described previously (Fadool et al. 1991). Briefly, the isolated clusters were incubated for 50min at 80revsmin−1 on an orbital shaker in a 0.2 µm filter-sterilized solution of 2.5mg papain and 12mg L-cysteine in 10ml of Panulirus saline (PS) containing 1% penicillin, streptomycin sulfate and amphotericin B (Gibco). Proteolytic digestion was stopped by replacing the enzyme solution with low-glucose L-15 medium supplemented with L-glutamine, dextrose, fetal calf serum and BME vitamins. Cells were immediately plated on poly-D-lysine-coated glass coverslips. Cells were maintained at saturation humidity in a modular incubator chamber (Billups-Rothenberg) at 24°C. Neurite outgrowth in individual cells was recorded on a TL Panasonic 6050 time-lapse video cassette recorder. Images were later captured and subsequently analyzed using Image 1 analysis software.

Electrophysiology

Voltage- and odor-activated currents were recorded in the whole-cell configuration with an integrating patch-clamp amplifier (Dagan 3900). The analog signal was filtered at 5 kHz and digitally sampled every 4ms. Data acquisition and subsequent storage and analysis of the digitized records were carried out using pCLAMP software (Axon Instruments). Cells were viewed at 40X magnification with Hoffman optics. Patch electrodes, pulled from 1.8mm o.d. borosilicate glass, were fire-polished to a tip diameter of approximately 1.0µm (bubble number 4.8; Mittman et al. 1987). High-resistance seals (8.0–14 GD) were formed by applying gentle suction to the lumen of the pipette upon contact with the cell. In all experiments, cells were voltage-clamped at a holding membrane potential of −60mV. Membrane resistance changes were determined by injecting current sufficient to elicit 30mV, 300ms hyperpolarizing voltage steps into the cells from the holding potential. Each cell was photographed to allow correlation of soma size, length of process and cell morphology with physiological properties.

Odor stimulation

Odors were ‘spritzed’ on the cells for 120ms from a seven-barrel (six barrels were used) glass micropipette (Frederick Haer) coupled to a pressurized valve system (Picospritzer, General Valve). In most trials, fluorescein was used as an indicator in one barrel of the pipette, which was varied from one experiment to another, to position the tip of the pipette relative to the cell and to ensure that the delivered odorant completely surrounded the cell and its associated processes. The magnitude of the response to odors was found to be independent of which of the six barrels contained the odor. Dilution of the odor between the pipette and the cell surface, an average distance of two cell diameters, was estimated to be approximately 9%, based on the calculated K+ permeability method of Firestein and Werblin (1989). Odor concentrations are reported as the pipette concentration and are not corrected for this dilution.

The odors used were solutions of (1) an equimolar mixture that included (10−3 mol l−1) betaine, glycine, lactic acid, taurine and trimethylamine oxide, referred to as S-1; (2) an aqueous extract of TetraMarin, a commercially available fish food, prepared as described earlier (Schmiedel-Jakob et al. 1990) and diluted 1000-fold, referred to as TET; and (3) single substances known to be effective odors for the lobster, which included (10−3 mol l−1) adenosine monophosphate (AMP), arginine, ascorbic acid, betaine, cysteine, glycine, histamine, proline, taurine and trimethylamine oxide (TMAO). All odorant solutions were prepared daily in modified L15 medium and applied at the stated concentration unless otherwise noted.

The number of different odors that stimulated a given cell (the response spectrum) was quantified using the breadth of responsiveness metric of Smith and Travers (1979). Here, the breadth of responsiveness (H) is defined as:
formula
where K is a proportionality constant, N is the number of odors tested, pi=| pA | the absolute current (pA) elicited from the Nth odor and expressed as a proportion of total pA elicited from all odors.

Solutions

Panulirus saline (PS) consisted of (in mmoll−1) 458 NaCl, 13.4 KCl, 9.8 MgCl2, 13.6 CaCl2, 13.6 Na2SO4, 3 Hepes and 2 glucose; pH7.4. Modified L15 medium consisted of 50ml of Liebowitz L15 stock, 50ml of 1.6 times normal concentration of PS, 0.6g of dextrose, 0.026g of L-glutamine, and 0.01% gentamicin. The patch electrode solution consisted of (mmol l−1) 30 NaCl, 11 EGTA, 10 Hepes, 1 CaCl2, 180 potassium acetate and 696 glucose; pH7.0. All salts were obtained from Sigma.

Morphology: neurite outgrowth

The cultures consisted largely of the five morphological types of small (8–16 µm diameter soma) ‘neuron-like’ cells, described previously (Fadool et al. 1991): (1) soma only, (2) soma with bud, (3) unipolar, (4) bipolar or (5) multiprocess, four of which were used in the present study (Fig. 1A). The processes ranged from 3 to 160 µm long. Each of the four morphological types was present as early as 2h post-plating. Initially, the predominant form was ‘soma only’, but the proportion of each morphological type changed over time; the relative proportion of cells lacking processes decreased, with a concomitant increase in the proportion of cells with processes (Fig. 1B). To distinguish whether the change in the relative proportion of the morphological types reflected selective loss of cells lacking processes or process proliferation, or both, 218 ‘soma only’ cells were followed individually with digital time-lapse imaging for three consecutive days starting 2h after plating. 42 (20%) of the cells died within the observation period, indicating that a selective loss of ‘soma only’ contributed to the change in the relative proportion of the cell types. While many of the 176 cells that persisted failed to sprout processes, 61 of the cells sprouted processes throughout the observation period, becoming uni-, bi- and, eventually, multipolar. The latter finding supports our contention that the four morphological types of ‘neuron-like’ cells are morphs of a single type of cell.

Fig. 1.

(A) Light micrographs of four morphs of cultured lobster ORNs observed under Hoffman modulation contrast optics. (i) Soma only, (ii) unipolar, (iii) bipolar, (iv) multiprocess. Magnification 780X. (B) Changes in the proportion of the four morphs over 9 days in culture. Each data point represents the incidence of that morph in 10 random fields of view, expressed as a percentage of all four types.

Fig. 1.

(A) Light micrographs of four morphs of cultured lobster ORNs observed under Hoffman modulation contrast optics. (i) Soma only, (ii) unipolar, (iii) bipolar, (iv) multiprocess. Magnification 780X. (B) Changes in the proportion of the four morphs over 9 days in culture. Each data point represents the incidence of that morph in 10 random fields of view, expressed as a percentage of all four types.

Physiology

Electrical properties

That the four types of ‘neuron-like’ cells were morphs of a single type of cell was supported by the finding that 268 cells tested for their current–voltage relationship had similar voltage-activated properties. The total membrane current evoked in a typical cell by depolarizing voltage steps consisted of a transient inward current (≤570 µs duration) that activated around −30mV, followed by a much larger, prolonged outward current that activated around −20mV and persisted with little decay throughout the 15ms duration of the pulse (Fig. 2A). The magnitude of the inward and outward currents was independent of the number of processes on the cell (ANOVA, P=0.47 for inward currents and P=0.71 for outward currents) (Fig. 2B). A subsample of 12 cells, including at least one cell of each morphological type, had a mean input resistance (RN) at rest of 1.1±0.2 GD and a membrane time constant (t0) of 67.3±11.3ms. No detectable equalizing time constant could be measured in any of the morphs, including morphs with multiple processes.

Fig. 2.

Total voltage-activated currents of cultured lobster ORNs. (A) Representative current–voltage relationship of one cell. Inset: macroscopic currents (upper traces) evoked by depolarizing voltage-steps (lower traces) when the cell was held at −60mV and stepped to +30mV in 10mV episodes. Records are not leak-subtracted. (B) Plot of the maximum amplitude (mean ± S.E.M.) of the inward (hatched bars) and outward (filled bars) currents of 268 cells grouped according to number of processes (N=68 soma only, 76 unipolar, 69 bipolar, 55 multiprocess).

Fig. 2.

Total voltage-activated currents of cultured lobster ORNs. (A) Representative current–voltage relationship of one cell. Inset: macroscopic currents (upper traces) evoked by depolarizing voltage-steps (lower traces) when the cell was held at −60mV and stepped to +30mV in 10mV episodes. Records are not leak-subtracted. (B) Plot of the maximum amplitude (mean ± S.E.M.) of the inward (hatched bars) and outward (filled bars) currents of 268 cells grouped according to number of processes (N=68 soma only, 76 unipolar, 69 bipolar, 55 multiprocess).

Physiology

Response to odors

472 cells (including the cells tested for their electrical properties) were tested for their ability to generate a current in response to stimulation with up to five different odors. The odor arrays usually, but not always, included the complex mixture, TET. 64% of the cells responded when they could be tested with at least one odor. This percentage increased to 89% when the cells could be tested with at least three different odors (N=182). Odors evoked a transient current that rose to a maximum over several hundred milliseconds and subsequently declined more slowly to rest (Fig. 3). The current could be of either polarity, depending on the cell and the odor tested and different odors could evoke currents of opposite polarity in the same cell (Fig. 3). In cells that could be tested with at least three different odors (N=182), the odors that were tested evoked only inward currents in 48 cells (26%), only outward currents in 58 cells (32%) and currents of both polarities in 56 (31%) cells. The remaining 11% of the cells did not respond to any of the odors tested.

Fig. 3.

Whole-cell, voltage-clamp recording from a cultured lobster ORN in response to spritzing the cell (arrow) with 10−3 mol l−1 proline (top trace), 10−3 mol l−1 taurine (second trace), TET (third trace) and culture medium only (bottom trace). Holding potential, −60mV for all traces.

Fig. 3.

Whole-cell, voltage-clamp recording from a cultured lobster ORN in response to spritzing the cell (arrow) with 10−3 mol l−1 proline (top trace), 10−3 mol l−1 taurine (second trace), TET (third trace) and culture medium only (bottom trace). Holding potential, −60mV for all traces.

Odor-evoked currents of both polarities were associated with an increase in membrane conductance, as indicated by a decrease in input resistance, when hyperpolarizing voltage steps were injected into the cells prior to and during odor-stimulation (Fig. 4). The input resistance (RN) decreased significantly under odor stimulation from a mean of 1.1±0.1 to 0.7±0.1 GD for the inward current (N=8) and from a mean of 1.3±0.6 to 0.7±0.1 GD for the outward current (N=4) (paired t-test, P<0.05). Concomitantly, the membrane time constant (t0), decreased significantly during odor stimulation from 59.6±12.6 to 21.6±4.1ms for the inward current (N=14) and from 82.5±20.7 to 34.6±9.5ms for the outward current (N=7) (paired t-test, P<0.05).

Fig. 4.

Whole-cell, voltage-clamp recordings from a cultured lobster ORN showing increased conductance associated with the odor-stimulated inward current. (A) Neither current (upper trace) nor change in conductance (lower trace) was evoked by stimulating (arrow) with culture medium control. (B) Inward current (upper trace) associated with an increase in conductance (lower trace) was evoked by stimulating with TET. Holding potential, −60mV. Solid bars denote 30mV, 300ms hyperpolarizing voltage pulses used to monitor membrane conductance.

Fig. 4.

Whole-cell, voltage-clamp recordings from a cultured lobster ORN showing increased conductance associated with the odor-stimulated inward current. (A) Neither current (upper trace) nor change in conductance (lower trace) was evoked by stimulating (arrow) with culture medium control. (B) Inward current (upper trace) associated with an increase in conductance (lower trace) was evoked by stimulating with TET. Holding potential, −60mV. Solid bars denote 30mV, 300ms hyperpolarizing voltage pulses used to monitor membrane conductance.

The latency to the onset of the odor-evoked currents, measured from the activation of the spritzer, ranged from less than 20ms to more than 1s, but typically was less than 100ms (Fig. 5). Overall, the mean latency to onset for the inward current, 186.2±31.3ms (N=100), was significantly longer than that for the outward current, 81.0±10.6ms (N=121) (t’ statistic, P<0.05). To determine whether this difference was caused by variations between cells, we performed a paired comparison of the latency in 19 cells that supported odor-evoked currents of both polarities. The mean latency in these cells for the inward current was not significantly longer than that for the outward current (paired t-test, P<0.05).

Fig. 5.

Plot of the distribution of the latency of odor-evoked currents in cultured lobster ORNs. Hatched bars, inward currents (N=100); filled bars, outward currents (N=121). Arrows denote the mean latency to onset of the inward (open arrow) and outward (filled arrow) currents, respectively.

Fig. 5.

Plot of the distribution of the latency of odor-evoked currents in cultured lobster ORNs. Hatched bars, inward currents (N=100); filled bars, outward currents (N=121). Arrows denote the mean latency to onset of the inward (open arrow) and outward (filled arrow) currents, respectively.

The peak amplitude of odor-evoked currents of both polarities increased with the concentration of the odor and saturated over 3–4 orders of magnitude (Fig. 6). The mean slope of the concentration–response function in the steepest region of the curve was 8.3±1.7pA per decade (N=3) for the inward current and 1.3±0.2pA per decade (N=7) for the outward current. Thresholds were less than 10−8 mol l−1, the lowest concentration tested, for currents of either polarity.

Fig. 6.

(A) Whole-cell, voltage-clamp recordings of outward currents in a cultured lobster ORN evoked by stimulating the cell (arrow) with decreasing concentrations of proline. Holding potential, −60mV. (B) Plot of the peak inward current (mean ± S.E.M.) elicited in three cells by taurine at the concentrations shown. (C) Plot of the peak outward current (mean ± S.E.M.) elicited in three other cells by proline at the concentrations shown. Current magnitudes in B and C are normalized to that evoked by 10−3 mol l−1 odor.

Fig. 6.

(A) Whole-cell, voltage-clamp recordings of outward currents in a cultured lobster ORN evoked by stimulating the cell (arrow) with decreasing concentrations of proline. Holding potential, −60mV. (B) Plot of the peak inward current (mean ± S.E.M.) elicited in three cells by taurine at the concentrations shown. (C) Plot of the peak outward current (mean ± S.E.M.) elicited in three other cells by proline at the concentrations shown. Current magnitudes in B and C are normalized to that evoked by 10−3 mol l−1 odor.

The peak amplitude of currents of both polarities evoked by a single concentration (10−3 mol l−1) of odor ranged, typically, from 5 to 85pA (Fig. 7A). The average magnitude of the inward current (39.2±3.0pA) was significantly greater than that of the outward current (20.1±1.4pA) measured across all cells and odors (N=386; t’ statistic, P≤0.05). The polarity and the magnitude of the currents evoked by odors were independent of the cell morphology for six different odors. The results for two of the six odors, proline ( N=111 cells) and taurine (N=102 cells), are shown in Fig. 7B. The peak amplitude of the odor-evoked currents of either polarity was also independent of the length of the process in cells bearing processes (N=55, correlation analysis, r>3.86) (Fig. 8A) and the diameter of the soma in cells lacking processes (N=60, correlation analysis, r>3.86) (Fig. 8B).

Fig. 7.

Plots of the peak amplitude of odor-evoked currents in cultured lobster ORNs as a function of cell morphology. (A) Distribution of the peak amplitude of the currents evoked in 386 cells by single odors tested at a common concentration (10−3 mol l−1). Hatched bars, inward currents; filled bars, outward currents. Arrows denote the mean amplitude of the inward (open arrow) and outward (filled arrow) currents. (B) Plots of the peak amplitude of currents of both polarities evoked by stimulation with 10−3 mol l−1 proline (N=111) and 10−3 mol l−1 taurine (N=102) grouped according to morphology. Inward currents are denoted as negative, outward as positive.

Fig. 7.

Plots of the peak amplitude of odor-evoked currents in cultured lobster ORNs as a function of cell morphology. (A) Distribution of the peak amplitude of the currents evoked in 386 cells by single odors tested at a common concentration (10−3 mol l−1). Hatched bars, inward currents; filled bars, outward currents. Arrows denote the mean amplitude of the inward (open arrow) and outward (filled arrow) currents. (B) Plots of the peak amplitude of currents of both polarities evoked by stimulation with 10−3 mol l−1 proline (N=111) and 10−3 mol l−1 taurine (N=102) grouped according to morphology. Inward currents are denoted as negative, outward as positive.

Fig. 8.

Plots of the peak amplitude of odor-evoked currents in cultured lobster ORNs as a function of length of the longest process (if more than one) for 55 cells (A) or the diameter of the soma in 60 cells lacking processes (B). The data set combines currents evoked by 10−3 mol l−1 proline, arginine, taurine, TMAO and betaine as odors (1 cell, 1 odor). Inward currents in both plots are denoted as negative, outward as positive.

Fig. 8.

Plots of the peak amplitude of odor-evoked currents in cultured lobster ORNs as a function of length of the longest process (if more than one) for 55 cells (A) or the diameter of the soma in 60 cells lacking processes (B). The data set combines currents evoked by 10−3 mol l−1 proline, arginine, taurine, TMAO and betaine as odors (1 cell, 1 odor). Inward currents in both plots are denoted as negative, outward as positive.

Single odors activated 14–77% of the cells (Table 1). The stimulatory effectiveness was betaine > histamine > glycine > proline > taurine > AMP > TMAO > ascorbate > arginine > cysteine. An equimolar mixture of five compounds (S-1: betaine, taurine, glycine, TMAO, lactate) ranked intermediate in stimulatory effectiveness, as did the complex mixture (TET), although the concentration of TET relative to that of the pure compounds is unknown. Culture medium itself, tested as a control, stimulated none of the cells tested, even though it contained some of the same amino acids as above as well as L-glutamine. Some odors selectively activated either inward or outward currents (Fig. 9). Taurine elicited inward currents twice as frequently as outward currents, while proline showed the opposite tendency. Betaine and ascorbic acid elicited strictly outward currents, although the number of cells tested in each of these instances was not large.

Table 1.

The percentage of cultured lobster olfactory receptor cells responding to individual odors, a mixture of five odors, and a complex odor

The percentage of cultured lobster olfactory receptor cells responding to individual odors, a mixture of five odors, and a complex odor
The percentage of cultured lobster olfactory receptor cells responding to individual odors, a mixture of five odors, and a complex odor
Fig. 9.

Plot of the percentage of cultured lobster ORNs that respond to various odors with inward and outward currents. All odors tested at 10−3 mol l−1, with the exception of TET, which was tested at 1000-fold dilution of stock. TMAO, trimethylamine oxide. AMP, adenosine monophosphate.

Fig. 9.

Plot of the percentage of cultured lobster ORNs that respond to various odors with inward and outward currents. All odors tested at 10−3 mol l−1, with the exception of TET, which was tested at 1000-fold dilution of stock. TMAO, trimethylamine oxide. AMP, adenosine monophosphate.

In a separate experiment, 53 cells that were determined to be odor-responsive by their ability to respond to TET were presented sequentially with the same five odors in order to determine their response spectra or ‘tuning’. The cells varied in the number of odors they responded to, as well as in the magnitude and the polarity of the response to any one odor (Fig. 10A). The mean breadth of responsiveness (H) for the sample population was 0.31±0.04 (Fig. 10B). Repeating this analysis using fewer odors or the same number of different odors yielded measures of H between 0.2 and 0.3 (data not shown).

Fig. 10.

(A) Graph of the response profiles of 10 cultured lobster ORNs to five odors tested at 10−3 mol l−1. Each horizontal line depicts the responses of a different cell; the height and direction of the bars on each line indicate the amplitude and the polarity, respectively, of the odor-evoked current (pA). Outward currents are denoted as positive (upward bars); inward currents as negative (downward bars). (B) Plot of the breadth of responsiveness of 53 cells. Arrow denotes the mean H value. TMAO, trimethylamine oxide.

Fig. 10.

(A) Graph of the response profiles of 10 cultured lobster ORNs to five odors tested at 10−3 mol l−1. Each horizontal line depicts the responses of a different cell; the height and direction of the bars on each line indicate the amplitude and the polarity, respectively, of the odor-evoked current (pA). Outward currents are denoted as positive (upward bars); inward currents as negative (downward bars). (B) Plot of the breadth of responsiveness of 53 cells. Arrow denotes the mean H value. TMAO, trimethylamine oxide.

The fact that the cells could be observed to sprout processes in culture, together with their common active electrical properties, strengthens our earlier contention that the various types of ‘neuron-like’ cells in the culture are morphs of one type of cell, the ORN (Fadool et al. 1991). This finding is consistent with the highly enriched source of the cultured cells; the somata were harvested from the lumen of the olfactory organ, which is literally filled with grape-like clusters of the somata of the ORNs (Grünert and Ache, 1988).

Although most neurons, as well as non-neuronal cells, are chemically irritable (Lerner et al. 1990), it is reasonable to interpret the responses of the cultured ORNs as olfactory. Most importantly, the cultured ORNs showed a diversity of response profiles (Fig. 10). Postsynaptic or neuromodulatory receptors, in contrast, would be expected to be much more homogeneously distributed across the population of cells and show either the same or, if possessing multiple receptors, a restricted set of response profiles across cells. For example, a histamine-gated chloride channel implicated in modulating the output of lobster ORNs in situ occurs on the soma of more than 95% of the cells (McClintock and Ache, 1989a). Second, the average breadth of responsiveness of the cultured ORNs (H=0.31) approximated that of lobster ORNs determined in situ with a similar, but not identical, array of compounds (H=0.2, Derby et al. 1984) and did not differ significantly from that of lobster ORNs determined in situ with the same array of compounds (H=0.35, Student’s t-test, P<0.05, pooled S.D.; M. Wachowiak, unpublished data). Third, as in their counterparts in situ, odors evoked both inward and outward currents in the cultured ORNs, and different compounds could evoke currents of opposite polarity in the same cell (McClintock and Ache, 1989b). Odors that tended frequently to elicit currents of one polarity in the cultured cells, e.g. proline and outward currents, also did so in lobster ORNs in situ (Schmiedel-Jakob et al. 1990; Michel et al. 1991). That currents of opposite polarity were associated with an increase in membrane permeability argues, as occurs in situ (Schmiedel-Jakob et al. 1990; Michel et al. 1991), that the odors were not up- and down-regulating a single conductance but, rather, were regulating two distinct conductances through two parallel transduction pathways. Finally, the threshold and dynamic range of the cultured ORNs matched those of lobster ORNs in situ in response to the same or similar types of odor molecules (Schmiedel-Jakob et al. 1989; W. C. Michel, unpublished data). We conclude that most and possibly all, of the chemosensitivity expressed by cultured lobster ORNs is sensory, i.e. is mediated by the same receptors and transduction pathways found in the outer dendrites of lobster ORNs in situ, and is not the result of activating synaptic, neuromodulatory or other receptors that may occur on the soma of lobster ORNs.

The inability of the culture medium to elicit responses from the cells when applied as a control stimulus, even though it contains amino acids shown to be odors (cysteine, glycine, arginine) and others (e.g. glutamine) that are potential odors, is not necessarily surprising. The cells were maintained and tested in the same medium and would be expected to adapt to continuous background stimulation, as do their counterparts in vivo (Borroni and Atema, 1988). However, adaptation to continuous background stimulation does shift the apparent threshold of the cells in vivo so that recording in medium might have lowered the overall sensitivity to those test stimuli present in the medium (cysteine, glycine, arginine). Indeed, the smallest percentage of cells tested responded to cysteine and arginine; in contrast, glycine was among the more effective odors tested (Table 1). The remaining test odors, including those used for threshold determinations, were not components of the culture medium.

The overall magnitudes of the current evoked by odors in cultured cells could differ appreciably from those in situ as a result of possible differences in the number of channels expressed in culture. It is interesting, therefore, that the average magnitudes of the odorevoked currents in the cultured ORNs (39.2±3.0pA, inward; 20.1±1.4pA outward) were within an order of magnitude of those evoked in lobster ORNs in situ. A 10-fold greater concentration of TET than that used in the present study elicited an average inward current of 25.4±5.1pA, while the same concentration of proline as that used in the present study elicited an average outward current of 4.5±0.6pA in lobster ORNs in situ (Michel and Ache, 1992). Odor-evoked currents of dendritic origin in situ would be electrotonically attenuated in the soma (where they were recorded), so the actual magnitude of the current in the dendrite would presumably be closer to that found in the cultured cells. Our results are consistent with the possibility that the cultured cells express a full complement of odor-activated channels.

The average amplitude of the odor-evoked inward currents in lobster ORNs is an order of magnitude smaller than those reported for salamander or rat ORNs (Firestein et al. 1991; Kurahashi, 1989; Lowe and Gold, 1991; Pixley and Pun, 1990), but is consistent with our still limited understanding of unitary currents that may underlie the macroscopic current. Inositol 1,4,5-trisphosphate, Ins(1,4,5)P3, the suspected excitatory second messenger, activated 1–3 channels in most (49/56) cell-free patches taken from the plasma membrane of the cultured cells (Fadool and Ache, 1992). On the assumption that these Ins(1,4,5)P3-activated channels are transductory, there are an estimated 330–1990 channels per cell, given a pipette tip diameter of 0.5–1.0 µm and a total cell surface area of 380 µm2 (Fadool et al. 1991); or 0.9–5.3 channels µm−2 membrane. Taking the smaller of the two Ins(1,4,5)P3-activated unitary conductances (2pA, Fadool and Ache, 1992), a macroscopic current of average magnitude (39pA; Fig. 7) would require activation of as few as 20 channels, or only 1–6% of the available channels. Such a small fraction of recruited channels is consistent with the presumed sensory function of the cells (e.g. Firestein et al. 1991).

The ability of the cultured cells to respond to odors did not correlate with the number or even the presence of processes, implying that the receptor sites for odors were not necessarily confined to the processes. At least in vertebrate ORNs, odor receptors are thought to be selectively localized to the cilia. Deciliating the frog olfactory epithelium with Triton X-100, for example, completely abolishes the electro-olfactogram (Adamek et al. 1984). Lowe and Gold (1991) confirmed this hypothesis by focally stimulating the cilia of dissociated salamander ORNs. Whether somatic sensitivity to odors in cultured lobster ORNs is induced or is typical of mature lobster ORNs in situ is unclear. Previous attempts to record responses to odors in freshly isolated somata have been consistently unsuccessful (e.g. Anderson and Ache, 1985), suggesting that somatic sensitivity to odors in cultured ORNs may have been induced by placing the cells in culture. This conclusion would be consistent with the observation that molluscan neurons raised in suspension culture to prevent process formation insert postsynaptic receptors otherwise normally confined to processes in the soma (Wong et al. 1981). Nevertheless, Hatt (1990) found odor-gated channels on the soma of freshly dissociated crayfish chemosensory neurons that had the same structure–activity profiles and sensitivity as the intact cells. Ligandgated histamine receptors on the soma of lobster ORNs resist the same enzymatic protocol used in the present study (McClintock and Ache, 1989a). Therefore, we cannot eliminate the possibility that the receptors in question normally occur on the somata of mature lobster ORNs, but are destroyed by the enzymatic treatment required to dissociate the cells and fail to recover in the subsequent 4–6h over which in situ recordings are usually made. It is known that enzymatic digestion used to prepare cells for patchclamping alters neurotransmitter responsiveness in invertebrate neurons (Oyama et al. 1990) and eliminates odor-responsiveness in amphibian ORNs (Firestein and Werblin, 1989).

The latency to response was extremely brief, as short as 20ms, with the majority of the responses occurring within 125ms of valve (spritzer) activation. These values fall below the lower limit of the range of latencies reported for dissociated salamander ORNs, the other system for which comparable measurements are available (175–600ms; Firestein et al. 1990). Both systems lack a mucus barrier that would impede stimulus access to the receptor sites, so the latencies presumably reflect the actual time course of transduction. Breer et al. (1990) recently reported that levels of two second messengers in rat and insect ORNs, cyclic adenosine monophosphate and inositol 1,4,5-trisphosphate, respectively, peak in response to odor stimulation within 25–50ms. Thus, while the latency of the response of cultured lobster ORNs is considerably shorter than that of salamander ORNs, it is consistent with second-messenger-mediated transduction.

Thus, we show that lobster ORNs not only survive in culture but express the odor sensitivity and selectivity of their counterparts in situ, including the ability of odors to excite as well as to inhibit the cells. As the cultured cells are morphologically and, presumably, electrotonically (as shown by the absence of a measurable equalizing time constant) more compact than their counterparts in situ, often consisting of only a spherical soma, the ability to characterize the transduction pathways and analyze potential interactions between them should be greatly facilitated in vitro.

This investigation was supported by ONR N0014-90-J-1566, NIMH NRSA 1F31MH10124-01A1. The authors would like to thank Ms Leslie Van Ekeris for technical assistance, Ms Lynn Milstead and Mr James Netherton for assistance with the illustrations and Drs Paul J. Linser and James M. Fadool for technical discussions.

Adamek
,
G. D.
,
Gesteland
,
R. C.
,
Mair
,
R. G.
and
Oakley
,
B.
(
1984
).
Transduction physiology of olfactory receptor cilia
.
Brain. Res
.
310
,
87
97
.
Anderson
,
P. A. V.
and
Ache
,
B. W.
(
1985
).
Voltage- and current-clamp recordings of the receptor potential in olfactory receptor cells in situ
.
Brain Res
.
338
,
273
280
.
Anholt
,
R. R. H.
(
1991
).
Odor recognition and olfactory transduction: the new frontier
.
Chem. Sens
.
16
,
421
427
.
Borroni
,
P. F.
and
Atema
,
J.
(
1988
).
Adaptation in chemoreceptor cells. I. Self-adapting backgrounds determine threshold and cause parallel shift of response function
.
J. comp. Physiol. A
164
,
67
74
.
Breer
,
H.
,
Boekhoff
,
I.
and
Tareilus
,
E.
(
1990
).
Rapid kinetics of second messenger formation in olfactory transduction
.
Nature
344
,
65
68
.
Calof
,
A. L.
and
Chikaraishi
,
D. M.
(
1991
).
Analysis of neurogenesis in a mammalian neuroepithelium: proliferation and differentiation of an olfactory neuron precursor in vitro
.
Neuron
3
,
115
127
.
Carr
,
W. E. S.
(
1990
).
Chemical signaling systems in lower organisms: A prelude to the evolution of chemical communication in the nervous system
. In
Evolution of the First Nervous Systems
(ed.
P. A. V.
Anderson
), pp.
81
94
.
New York
:
Plenum Press
.
Coon
,
H. G.
,
Curcio
,
F.
,
Sakaguchi
,
K.
,
Brand
,
M. L.
and
Swerdlow
,
R. D.
(
1989
).
Cell cultures of neuroblasts from rat olfactory epithelium that show odorant responses
.
Proc. natn. Acad. Sci. U. S. A
.
86
,
1703
1707
.
Derby
,
C. D.
,
Hamilton
,
K. A.
and
Ache
,
B. W.
(
1984
).
Processing of olfactory information at three neuronal levels in the spiny lobster
.
Brain. Res
.
300
,
311
319
.
Fadool
,
D. A.
and
Ache
,
B. W.
(
1992
).
Plasma membrane inositol 1,4,5-trisphosphate-activated channels mediate signal transduction in lobster olfactory receptor neurons
.
Neuron
9
,
1
20
.
Fadool
,
D. F.
,
Michel
,
W. C.
and
Ache
,
B. W.
(
1991
).
Sustained primary culture of lobster (Panulirus argus) olfactory receptor neurons
.
Tissue & Cell
23
,
719
732
.
Firestein
,
S.
(
1991
).
A noseful of odor receptors
.
Trends Neurosci
.
14
,
270
272
.
Firestein
,
S.
,
Shepherd
,
G. M.
and
Werblin
,
S. F.
(
1990
).
Time course of the membrane current underlying sensory transduction in salamander olfactory receptor neurones
.
J. Physiol., Lond
.
430
,
135
158
.
Firestein
,
S.
and
Werblin
,
F.
(
1989
).
Odor-induced membrane currents in vertebrate-olfactory receptor neurons
.
Science
244
,
79
82
.
Firestein
,
S.
,
Zufall
,
F.
and
Shepherd
,
G. M.
(
1991
).
Single odor-sensitive channels in olfactory receptor neurons are also gated by cyclic nucleotides
.
J. Neurosci
.
11
,
3565
3572
.
Grünert
,
U.
and
Ache
,
B. W.
(
1988
).
Ultrastructure of the aesthetasc (olfactory) sensilla of the spiny lobster, Panulirus argus
.
Cell Tissue Res
.
251
,
95
103
.
Hatt
,
H.
(
1990
).
Electrophysiological studies on chemoreceptive neurons in crustacea
. In
Frontiers in Crustacean Neurobiology
(ed.
K.
Wiese
,
W.-D.
Krenz
,
J.
Tautz
,
H.
Reichert
and
B.
Mulloney
), pp.
41
45
. Switzerland: Birkhäuser Verlag.
Kurahashi
,
T.
(
1989
).
Activation by odorants of cation-selective conductance in the olfactory receptor cell isolated from the newt
.
J. Physiol., Lond
.
419
,
177
192
.
Lerner
,
M. R.
,
Gyorgyi
,
T. K.
,
Reagan
,
J.
,
Roby-Shemkovitz
,
A.
,
Rybczynski
,
R.
and
Vogt
,
R. G.
(
1990
).
Peripheral events in moth olfaction
.
Chem. Senses
15
,
191
198
.
Lowe
,
G.
and
Gold
,
G. H.
(
1991
).
The spatial distributions of odorant sensitivity and odorant-induced currents in salamander olfactory receptor cells
.
J. Physiol., Lond
.
442
,
147
168
.
McClintock
,
T. S.
and
Ache
,
B. W.
(
1989a
).
Histamine directly gates a chloride channel in lobster olfactory receptor neurons
.
Proc. natn. Acad. Sci. U. S. A
.
86
,
8137
8141
.
McClintock
,
T. S.
and
Ache
,
B. W.
(
1989b
).
Hyperpolarizing receptor potentials in lobster olfactory receptor cells: implications for transduction and mixture suppression
.
Chem. Senses
14
,
637
647
.
Michel
,
W. C.
and
Ache
,
B. W.
(
1992
).
Cyclic nucleotides mediate an odor-evoked potassium conductance in lobster olfactory receptor cells
.
J. Neurosci
.
12
,
3979
3984
.
Michel
,
W. C.
,
McClintock
,
T. S.
and
Ache
,
B. W.
(
1991
).
Inhibition of lobster olfactory receptor cells by an odor-activated potassium conductance
.
J. Neurophysiol
.
65
,
446
453
.
Mittman
,
S. C.
,
Flaming
,
D. G.
,
Copenhagen
,
D. R.
and
Belgum
,
J. H.
(
1987
).
Bubble pressure measurement of micropipet tip outer diameter
.
J. Neurosci. Meth
.
22
,
161
166
.
Oyama
,
Y.
,
Hori
,
N.
,
Allen
,
C. N.
and
Carpenter
,
D. O.
(
1990
).
Influences of trypsin and collagenase on acetylcholine responses of physically isolated single neurons of Aplysia californica
.
Cell. molec. Neurobiol
.
10
,
193
205
.
Pixley
,
S. K.
and
Pun
,
R. Y. K.
(
1990
).
Cultured rat olfactory neurons are excitable and respond to odors
.
Devl Brain. Res
.
53
,
125
130
.
Ronnett
,
G. V.
,
Hester
,
L. D.
and
Synder
,
S. H.
(
1991
).
Primary culture of neonatal rat olfactory neurons
.
J. Neurosci
.
11
,
1243
1255
.
Schmiedel-Jakob
,
I.
,
Michel
,
W. C.
,
Anderson
,
P. A. V.
and
Ache
,
B. W.
(
1990
).
Whole cell recording from lobster olfactory receptor cells: multiple ionic bases for the receptor potential
.
Chem. Senses
15
,
397
405
.
Smith
,
D. V.
and
Travers
,
J. B.
(
1979
).
A metric for the breadth of tuning of gustatory neurons
.
Chem. Sens. Flavour
4
,
215
229
.
Stengl
,
M.
,
Zufall
,
F.
,
Dudel
,
J.
and
Hildebrand
,
J. G.
(
1989
).
Patch clamp analysis of male Manduca sexta olfactory receptor neurons in primary cell culture
.
Soc. Neurosci. Abstr
.
15
,
751
.
Wong
,
R. G.
,
Hadley
,
R. D.
,
Kater
,
S. B.
and
Hauser
,
G. C.
(
1981
).
Neurite outgrowth in molluscan organ and cell cultures: The role of conditioning factor(s)
.
J. Neurosci
.
1
,
1008
1021
.
Zufall
,
F.
,
Firestein
,
S.
and
Shepherd
,
G. M.
(
1991a
).
Analysis of single cyclic nucleotide-gated channels in olfactory receptor cells
.
J. Neurosci
.
11
,
3573
3580
.
Zufall
,
F.
,
Stengl
,
M.
,
Franke
,
C.
,
Hildebrand
,
J. G.
and
Hatt
,
H.
(
1991b
).
Ionic currents of cultured olfactory receptor neurons from antennae of male Manducasexta
.
J. Neurosci
.
11
,
956
965
.