You are on page 1of 29

ARTICLES

T-type Ca2+ channels, SK2 channels and SERCAs gate


© 2008 Nature Publishing Group http://www.nature.com/natureneuroscience

sleep-related oscillations in thalamic dendrites


Lucius Cueni1, Marco Canepari1, Rafael Luján2, Yann Emmenegger3, Masahiko Watanabe4, Chris T Bond5,
Paul Franken3, John P Adelman5 & Anita Lüthi1,6

T-type Ca21 channels (T channels) underlie rhythmic burst discharges during neuronal oscillations that are typical during sleep.
However, the Ca21-dependent effectors that are selectively regulated by T currents remain unknown. We found that, in dendrites
of nucleus reticularis thalami (nRt), intracellular Ca21 concentration increases were dominated by Ca21 influx through T channels
and shaped rhythmic bursting via competition between Ca21-dependent small-conductance (SK)-type K1 channels and Ca21
uptake pumps. Oscillatory bursting was initiated via selective activation of dendritically located SK2 channels, whereas Ca21
sequestration by sarco/endoplasmic reticulum Ca21-ATPases (SERCAs) and cumulative T channel inactivation dampened
oscillations. Sk2–/– (also known as Kcnn2) mice lacked cellular oscillations, showed a greater than threefold reduction in low-
frequency rhythms in the electroencephalogram of non–rapid-eye-movement sleep and had disrupted sleep. Thus, the interplay
of T channels, SK2 channels and SERCAs in nRt dendrites comprises a specialized Ca21 signaling triad to regulate oscillatory
dynamics related to sleep.

Neurons in the thalamocortical system cooperate to produce synchro- corticothalamic and thalamocortical excitatory input and by reciprocal
nized, rhythmic network activity, which underlies the slow waves that connections between nRt cells1,13.
are characteristic of sleep electroencephalograms (EEGs)1,2. Rhythmo- Here we report a highly specialized Ca2+ signaling network in nRt
genesis is accompanied by low-threshold burst discharges in thalamic dendrites, in which Ca2+ influx through T channels has a dominant
neurons2,3, which are carried by the three members of the low voltage– role. Ca2+ influx activates competing targets, SK2 channels and
activated Cav3 Ca2+ channel family, also called T channels4. Although SERCAs to generate and regulate the strength of nRt oscillations. In
Ca2+ ions entering through T channels are the electrical charge carriers Sk2–/– mice, non–rapid-eye-movement sleep (NREMS) EEG power
underlying low-threshold bursts, the associated intracellular Ca2+ density was markedly reduced in the delta and spindle frequencies, and
concentration ([Ca2+]i) dynamics, the intracellular Ca2+ signaling sleep was fragmented. Our findings suggest that Ca2+ influx through
and their role in sleep physiology remain largely unknown. T channels acts in concert with SK2 channels and SERCAs to influence
To investigate Ca2+-dependent mechanisms that govern oscillatory characteristic frequency bands of NREMS.
rhythms important for sleep, we focused on the nRt, a thin inhibitory
network interposed between thalamocortical projection neurons and RESULTS
the cortex that is crucial for information transfer and arousal control5,6. Selective coupling between T and SK currents in nRt cells
Prominent forms of rhythmic bursting in the nRt accompany the major In Cs+-based whole-cell patch-clamp recordings, we evoked rapid
types of low-frequency EEG oscillations: in particular, delta oscillations inward currents following 0.5-s step hyperpolarizations (by –40 mV
(1–4 Hz), spindle waves (10–15 Hz) and slow oscillations (o1 Hz), from a holding potential of –55 to –60 mV) that were reduced by the
with low-threshold bursts showing peculiarly long durations and high T channel blocker mibefradil (10–50 mM, from –494 ± 101 pA to –72 ±
numbers of superimposed action potentials1,2,5–7. T channels in nRt are 34 pA, n ¼ 7, P o 0.02; Fig. 1a)4. The cells showed T currents with
composed of Cav3.2 and Cav3.3 subunits8, and are heavily expressed properties that fulfilled previously established criteria for acceptable
along the somatodendritic axis9. In nRt neurons, bursts are typically voltage control in nRt cells (Supplementary Methods online). In
followed by an afterhyperpolarization (AHP) generated by small- K+-based electrodes permitting K+ flow, the response recorded after
conductance Ca2+-activated SK-type K+ currents10–12. The dynamics a hyperpolarizing command (–40 mV, 125 ms) at –67 to –62 mV was
and synchrony of these endogenous oscillatory activities are shaped by biphasic; a small outward current (B10–200 pA) typically followed the

1Division of Pharmacology and Neurobiology, Biozentrum, University of Basel, Klingelbergstr. 70, CH-4056 Basel, Switzerland. 2Departamento de Ciencias Médicas,
Crib-Facultad De Medicina, Campus Biosanitario C/ Almansa 14, Universidad de Castilla-La Mancha, 02006 Albacete, Spain. 3Center for Integrative Genomics, Génopode
Building, Quartier UNIL-Sorge, University of Lausanne, CH-1015 Lausanne-Dorigny, Switzerland. 4Department of Anatomy, Hokkaido University School of Medicine, Kita-
15 Nishi-7 Kita-ku Sapporo 060-8638, Japan. 5Vollum Institute, Oregon Health & Science University, 3181 SW Sam Jackson Park Road, Portland, Oregon 97239, USA.
6Department of Cell Biology and Morphology, Faculty of Biology and Medicine, Rue du Bugnon 9, University of Lausanne, CH-1005 Lausanne, Switzerland. Correspondence

should be addressed to A.L. (anita.luthi@unil.ch).


Received 25 March; accepted 16 April; published online 18 May 2008; doi:10.1038/nn.2124

NATURE NEUROSCIENCE VOLUME 11 [ NUMBER 6 [ JUNE 2008 683


ARTICLES

Figure 1 SK currents are selectively activated by


a b Apa-sens current
c T currents to control nRt cell oscillations.
Apa-sens current (a) Membrane current responses elicited via a
K+ control 1 = 28 ms 100 pA
hyperpolarizing voltage command (from –60
500 ms
+Apa (100 nM) to –100 mV, 125 ms) in K+-based solutions
2 = 329 ms 14.2 ms (K+ control) and in the presence of apamin (+Apa,
+Mibe 100 pA
200 pA 100 nM). The last 15 ms of the holding current
100 ms during hyperpolarization are also visible. Inset,
25 ms
Cs+ control 2+ T current, evoked with a Cs+-based patch solution
T-type Ca current
(Cs+ control) and after mibefradil addition (+Mibe,
d Apa-sens current
e f 50 mM, bath applied for 10 min). (b) Apamin-
© 2008 Nature Publishing Group http://www.nature.com/natureneuroscience

(5 mM BAPTA) Apa-sens current (TTX) Apa-sens current (Mibe)


Apa-sens current (Ctrl) sensitive current (Apa-sens), obtained from traces
in a. The bi-exponential fit to the trace (dark line)
T-type Ca2+
current (Mibe)
described by t1 and t2 had a correlation coeffici-
2+
ent of 0.91 ± 0.01 (n ¼ 5 cells). (c) Overlay of
T-type Ca current 2+
T-type Ca current the T current (in apamin) and the apamin-sensitive
2+
T-type Ca current
100 pA 100 pA 100 pA current from cell in a. (d) Apamin-sensitive
current with a BAPTA-containing recording
100 ms 100 ms 100 ms
pipette, superimposed on the T current from the
same cell. (e) Apamin-sensitive currents in control
(Ctrl) and in TTX, superimposed with the T current
g h in TTX and apamin. (f) The apamin-sensitive
40 –77 mV Ctrl
current in mibefradil (50 mM) and the T current
T current decay slope

before mibefradil. A small inward current


(pA ms–1)

remained in the presence of mibefradil (inset).


+TTX
20 ** ** (g) Decay slope of T current in K+-based solution
(n ¼ 6), in apamin (Apa, n ¼ 6) and with
+Apa,TTX 10 mV intracellular Cs+ (n ¼ 13), presented as means ±
0 s.e.m. **P o 0.01. (h) Discharge patterns of
K+ +Apa Cs+ 200 ms
an nRt cell after brief negative current injections
(–100 pA) in control conditions, TTX (0.5 mM) and
apamin (+Apa, 100 nM). Dotted lines indicate
resting membrane potential.

T current, which was abolished by apamin (100 nM), a selective SK reduced decay slope (from 29.8 ± 6.7 pA ms–1 to 12.9 ± 3.4 pA ms–1,
channel blocker (Fig. 1a). The apamin-sensitive current, obtained by n ¼ 6, P o 0.01) to values that were found with Cs+-based electrodes
subtracting currents before and after apamin application, was 349 ± (13.5 ± 2.5 pA ms–1, n ¼ 13, P 4 0.05; Fig. 1g). Thus, apamin-sensitive
59 pA (n ¼ 5; Fig. 1b) and decayed with a bi-exponential time course. SK channels underlie the outward K+ current following the T currents
The fast component had a time constant of t1 ¼ 30.1 ± 2.6 ms (n ¼ 5), and the repolarizing effect of the SK current accelerated the decay of the
whereas the slow component decayed in t2 ¼ 834 ± 227 ms (n ¼ 5) and T current.
contributed 17.3 ± 4.2% to the total current amplitude. The latency Cells of nRt are well known for their propensity to generate
from the peak of the T current to the peak of the apamin-sensitive oscillatory bursting5,6. In whole-cell (Fig. 1h) or extracellular recordings
current was 14.1 ± 0.3 ms (n ¼ 5; Fig. 1c), suggesting rapid SK channel (Supplementary Fig. 1 online), nRt cells presented dampened oscilla-
activation. Thus, large SK currents were expressed in nRt cells and the tory, low-threshold burst discharges around 4–10 Hz, typically of 2–10
voltage-clamp approach appeared to be suitable for characterizing the action potentials at 150–250 Hz. TTX (0.5 mM) blocked action
Ca2+ signals leading to their activation (Supplementary Methods). potentials, isolating the low-threshold Ca2+ spike, but only marginally
We examined whether Ca2+ entry through T channels was required reduced the number of low-threshold spikes per oscillation (control, 3.0
for SK current activation by testing the effects of Ca2+ chelators and of ± 0.3 bursts; TTX, 2.6 ± 0.2 bursts; n ¼ 6, P o 0.05; Fig. 1h).
the T channel blocker mibefradil. No apamin-sensitive current could be Subsequent apamin application (100 nM) blocked oscillations, unmask-
elicited with the rapid Ca2+ chelator 1,2-bis(2-aminophenoxy)ethane- ing a plateau potential (n ¼ 6; Fig. 1h). Conversely, the SK channel
N,N,N¢,N¢-tetraacetic acid (BAPTA, 1–5 mM, included in patch pip- agonist, 1-ethyl-2-benzimidazolinone (1-EBIO, 0.1 mM), which
ette) (5.9 ± 3.2 pA, n ¼ 9, P o 0.001 compared with BAPTA-free increases the apparent Ca2+ sensitivity of SK channels, but does not
conditions; Fig. 1d), although T currents remained unaltered (current alter their maximal activation14, potentiated oscillatory activity (Sup-
amplitudes, 426 ± 49 pA, P 4 0.05). Similar results were obtained plementary Fig. 1). These pharmacological experiments suggested that
when the slow Ca2+ chelator EGTA (5 mM) was included (16 ± 13 pA, the coupling between low-threshold Ca2+ bursts and SK channels was
n ¼ 10, P o 0.05 compared with recordings in 0.1 mM EGTA). the central event underlying the oscillatory activity in nRt neurons.
Furthermore, apamin-sensitive currents persisted in the presence of
tetrodotoxin (TTX, 0.5 mM; 366 ± 34 pA versus 423 ± 54 pA, n ¼ 3, T currents dominate D[Ca2+]i in dendrites
P 4 0.05; Fig. 1e), a Na+ channel blocker, but were abolished in To investigate the Ca2+ dynamics underlying the observed functional
mibefradil (50 mM, apamin-sensitive current amplitude –11 ± 4 pA, coupling between T currents and SK currents, we visualized burst-
n ¼ 4, P o 0.002; Fig. 1f). induced changes in the intracellular free Ca2+ concentration (D[Ca2+]i)
We explored the role of rapid SK-current activation by quantifying in cells filled with the Ca2+ dye magfura-2 (3 mM). This low-affinity
the decay phase of T currents. Because these show a biphasic inward- Ca2+ indicator permits real-time imaging of the D[Ca2+]i in the
outward waveform before apamin and a monophasic decay in apamin, micromolar range15 and quantification of the free [Ca2+]i reached
we measured current decay slope (Supplementary Methods). Apamin during both single and oscillatory low-threshold bursts. We focused on

684 VOLUME 11 [ NUMBER 6 [ JUNE 2008 NATURE NEUROSCIENCE


ARTICLES

signals in nRt dendrites, as dendritic T currents are essential for larger Ca2+ influx during a low-threshold burst or differential intra-
low-threshold bursting9,16 (Fig. 2a). The D[Ca2+]i evoked by a low- cellular buffering. To distinguish between these possibilities, we mea-
threshold burst could be measured up to B150–200 mm from the sured the total Ca2+ entering during a low-threshold burst with the
soma, with no obvious major differences in the amplitude and kinetics high-affinity Ca2+ indicator bis-fura-2 (1 mM). The change in the
of the signals at different dendritic sites (Supplementary Methods and concentration of dye-bound Ca2+ (D[DCa2+]i, see Supplementary
Fig. 2b). Therefore, we determined the D[Ca2+]i from fluorescence Methods)18 was 135 ± 22 mM (n ¼ 4) for a low-threshold burst and
signals averaged over the entire imaged dendrite. The D[Ca2+]i asso- 1.35 ± 0.18 mM (n ¼ 4) per action potential (Fig. 2h). The large [Ca2+]i
ciated with the low-threshold burst reached peak levels of 713 ± 71 nM increase in nRt cell dendrites during a low-threshold burst was thus
(n ¼ 6 dendrites from six different nRt neurons) and decayed with a dominated by Ca2+ influx through T currents, with an action potential
time constant of 56 ± 9 ms (fitted for n ¼ 5 dendrites). The delay crowning the burst contributing B1%, and an equal buffer capacity of
© 2008 Nature Publishing Group http://www.nature.com/natureneuroscience

between the D[Ca2+]i and the AHP, measured as a peak-to-peak B200 for both Ca2+ entry resulting from burst and tonic discharge.
latency, was 37 ± 6 ms (range of 21–46 ms, sampled at 500 frames
s–1 in n ¼ 4 experiments; Fig. 2c). Taking into account that the AHP SK2 channels mediate the SK currents in nRt cells
time-to-peak is the convolution of T current decay, SK current The selective and rapid coupling of T type Ca2+ to SK channels requires
activation and the charging of cell capacitance (t ¼ B12–22 ms), that SK channels be expressed at sites where T channels dominate Ca2+
this value is consistent with the peak-to-peak latency between T and SK entry. mRNAs encoding SK1 and SK2 are expressed in nRt neurons19.
currents (B14 ms; Fig. 1c) and confirms a
rapid gating of SK channels by Ca2+ entry
during a low-threshold burst. a b c 46 ms
0.25 µM
Both low voltage– and high voltage–acti-
vated (HVA) Ca2+ currents activate SK cur- 100 ms
17
rents in the nRt . Apamin-sensitive currents 50 ms

(194 ± 24 pA, n ¼ 11), evoked following


depolarizing voltage steps (from –67 to
–37 mV for 10–125 ms), were reduced by 20 mV
the P/Q- and N-type Ca2+ channel blocker o-
conotoxin MVIIC (o-CTXMVIIC, B1 mM, –82 mV

18.7 ± 6.4% of control amplitude, n ¼ 5, P o 50 µm

0.02), but not by the L-type Ca2+ channel d Ctrl e Ctrl f 1.0
+Mibe +HVA blockers
blocker nifedipine (10 mM, 88.9 ± 7.1% of the (50 µM)
control amplitude, n ¼ 9, P 4 0.05). The 0.25 µM 0.5
T channel blocker mibefradil (50 mM) abol-

∆ [Ca2+ ]i (µM)
*
ished low-threshold bursts and reduced the 0
Ctrl Mibe
burst-evoked D[Ca2+]i by 90.2 ± 3.6% (to 1.0
87 ± 57 nM, n ¼ 3, P o 0.05; Fig. 2d). 20 mV
20 mV
However, a cocktail of HVA channel blockers 0.5
–85 mV –82 mV
(10 mM nifedipine, B1 mM o-CTXMVIIC, –82 mV –82 mV
40 ms
n ¼ 7, supplemented with 100 nM SNX-482, 0
Ctrl HVA
an R type channel blocker, n ¼ 4 cells)
2+ g h blockers
preserved low-threshold bursts and D[Ca ]i
0.25 µM
(D[Ca2+]i ¼ 738 ± 133 nM in control, 673 ± 100 µM
120 nM in cocktail, n ¼ 7, P 4 0.05,
Fig. 2e,f). Therefore, Ca2+ entry through T
channels was principally important for gating
SK channels during a low-threshold burst.
Does burst-induced Ca2+ entry represent a –69 mV –53 mV –71 mV
–55 mV

unique source for [Ca2+]i in nRt dendrites?


We compared the D[Ca2+]i associated with 20 mV 20 mV

the low-threshold burst to that observed from 100 ms 100 ms


tonic action-potential discharge, which was 2+
induced by depolarizing neurons from a rest- Figure 2 T channels dominate dendritic [Ca ]i increases during low-threshold bursts. (a) Reconstruction
of a magfura-2–filled nRt cell (3 mM). Fluorescent signals were averaged over boxed areas. (b) D[Ca2+]
ing membrane potential between –55 and (upper traces) elicited by low-threshold burst (lower trace) acquired at 125 Hz in the areas indicated i
–45 mV. Tonic discharge rates reached max- in a. (c) Overlay of low-threshold burst and D[Ca2+]i acquired at 500 Hz. Same cell as in a and b.
imal frequencies around 150–220 Hz, similar (d) Effects of mibefradil (+Mibe, 50 mM, applied 10 min) on burst-evoked D[Ca2+]i. (e) Effects of HVA
to those reached during burst discharges channel blockers (10 mM nifedipine, 1 mM o-CTXMVIIC, 100 nM SNX-482), applied 10 min, on burst-
2+
(150–250 Hz). Under these conditions, the evoked D[Ca ]i. (f) Pooled data for mibefradil (n ¼ 3) and HVA channel blockers (n ¼ 7), presented as
2+
D[Ca ]i associated with the tonic discharges means ± s.e.m. *P o 0.05. (g) [Ca2+]i increases for another cell produced by a low-threshold burst (left
traces) and by tonic action potentials (evoked by +100 pA injection) are shown. The onset of D[Ca2+]i is
was markedly smaller (15.9 ± 5.4 nM per delayed with respect to burst onset (dotted line) as a result of magfura-2’s low Ca2+ affinity. (h) Same
action potential, n ¼ 5, P o 0.01; Fig. 2g) experiment as in g, for a cell filled with bis-fura-2 (1 mM). Note the lack of a burst-associated AHP, in
than that associated with low-threshold contrast to g. For this high-affinity Ca2+ dye, the transient starts at the onset of the low-threshold burst
bursts. This difference could be because of a (dotted line).

NATURE NEUROSCIENCE VOLUME 11 [ NUMBER 6 [ JUNE 2008 685


ARTICLES

Figure 3 SK currents in nRt are carried by SK2-containing SK channels.


a (a) Membrane voltage responses to negative current injections (–100 pA,
–/–
Sk2 LT burst 400 ms, left) and apamin-sensitive (Apa-sens) currents (right) in nRt cells of
Apa-sens current
Sk2 –/– mice. Note that a low-threshold (LT) burst and a T current were clearly
–66 mV
present. (b) Same as a for a cell derived from a Sk1–/– mouse. (c) Pooled data
of 4–33 cells, presented as means ± s.e.m. WT, wild type. **P o 0.005,
10 mV
100 pA ***P o 0.001.
200 ms
100 ms
T-type Ca2+
current
b nRt (Fig. 4a), predominantly in the neuropil surrounding the spindle-
Sk1–/–
© 2008 Nature Publishing Group http://www.nature.com/natureneuroscience

Apa-sens current shaped cell bodies (Fig. 4b), but found none in Sk2–/– animals
(Supplementary Fig. 2 online). Using pre-embedding immunogold
–64 mV electron microscopy, we observed only a few immunogold particles in
cell bodies, and most of these were associated with the rough endo-
10 mV 100 pA
T-type Ca2+ plasmic reticulum; we found almost no particles at the somatic plasma
200 ms 100 ms
current membrane (0.87 ± 0.32 immunogold particles per mm2) (Fig. 4c). Most
SK2 immunogold particles were located in dendrites along the extra-
c Apa-sens
current, T current–
Apa-sens
current, depolarization
T current decay
–1 synaptic plasma membrane of dendritic shafts (10.87 ± 2.11 immuno-
slope (pA ms )
activated (pA) activated (pA) gold particles per mm2, P o 0.01 compared with somatic density; Fig.
300 40
400
4d–g). Only background levels of immunogold labeling were detected in
200 the Sk2–/– animals (Supplementary Fig. 2).
20
200 ** Notably, we often observed immunoparticles close to excitatory
100
*** *** synapses (146 immunoparticles within 200 nm of the edge of the
0
–/– –/–
0 0 postsynaptic density of 50 excitatory profiles), but never close to
WT Sk1 Sk2 WT Sk1–/– Sk2 –/– WT Sk1 –/– Sk2 –/–
inhibitory synapses (0 immunoparticles for 25 inhibitory profiles)
(Fig. 4e–g), suggesting that Ca2+ entry during glutamatergic synaptic
To determine which channel isoforms carry the SK current, we transmission may gate SK2 channels. A large portion of total labeling
examined Sk1–/– (also known as Kcnn1) and Sk2–/– mice20. Cellular (1,181 immunoparticles out of 1,737; 68%) was also found to be
morphology and basic electrophysiological properties were similar in associated with intracellular membranes (Fig. 4d), potentially reflect-
cells from Sk1–/–, Sk2–/– and wild-type Sk2+/+ littermate animals (Sup- ing protein trafficking. Thus, SK2 channel density was highest in
plementary Note online). In Sk2–/– animals, however, the oscillatory dendrites, where Ca2+ influx occurred almost exclusively through
discharge was replaced by a single, slowly decaying depolarization (27/ T channels during a low-threshold burst.
27 cells tested; Fig. 3a). Sk1–/– cells showed unaltered discharge behavior
compared with wild-type littermates in all 39 cells studied (Fig. 3b).
Sk2–/–, but not Sk1–/–, nRt neurons lacked an apamin-sensitive current Sk2 +/+
a b
following both T-type and HVA Ca2+ current activation (Fig. 3a–c).
Moreover, the slope of the T current decay in Sk2–/– neurons (11.6 ± 1.2
pA ms–1, n ¼ 20) was similar to that obtained in apamin (P 4 0.05), but
smaller than that in wild type (24.9 ± 3.2 pA ms–1; n ¼ 15; P o 0.05)
and in Sk1–/– cells (25.0 ± 2.9 pA ms–1; n ¼ 33, P o 0.01; Fig. 3c).
Finally, application of 1-EBIO failed to induce an outward current
following the T current in Sk2–/– cells (Supplementary Fig. 1).
The low-threshold bursting properties of thalamocortical neurons in
the ventrobasal thalamic nucleus, adjacent to the nRt, including sag
potentials, burst amplitudes and burst discharge frequencies, appeared c d
to be unaltered in the Sk2–/– mice (n ¼ 4) compared with wild-type
animals (n ¼ 5, data not shown). Thus, in the thalamic network, the
lack of SK2 channels selectively compromised oscillatory bursting in
nRt neurons.

Selective expression of SK2 channels in nRt dendrites


Consistent with previous observations at the mRNA level19, we detected
immunoreactivity to SK2 throughout the dorsoventral extension of the
e f g
Figure 4 SK2 channel subunits are selectively expressed in nRt dendrites.
(a–b) Immunoreactivity for SK2 protein in nRt-containing sections of wild-
type animals (Sk2 +/+) at the light-microscopic level. Scale bars, 1 mm in a
and 0.1 mm in b. IC, internal capsule; Th, thalamus. (c–g) Immunoreactivity
for SK2 at the ultrastructural level. Arrows, immunogold particle on somatic
and dendritic membranes; crossed arrows, gold labeling on intracellular
membranes. B, bouton; Den, dendrite; ER, endoplasmic reticulum; IB,
inhibitory bouton; N, nucleus. Scale bars, 0.2 mm.

686 VOLUME 11 [ NUMBER 6 [ JUNE 2008 NATURE NEUROSCIENCE


ARTICLES

a a Ctrl +CPA (10 µM) +Apa (100 nM)

100 pA
50 µm 0.25 µM 100 ms

b 300 CPA (10 µM)


c CPA
Ctrl

Apa-sens current (pA)


250
20 mV 100 pA
200
© 2008 Nature Publishing Group http://www.nature.com/natureneuroscience

200 ms
150 CPA + Apa
–70 mV CPA
100
200 ms
50 Ctrl
0
–200 0 200 400
Time (s)
d e
Peak amplitude (µM)

b 1
0.8 CPA (10 µM) Sk2 –/–

Apa-sens current (pA)


0.6 250 Ctrl +CPA +Apa
0.25 µM
50 ms 0.4 *** 200
2
0.2 150 200 pA
3 Sk1–/– 200 ms
100
0.0
Burst 1 Burst 2 Burst 3 50
Ctrl +CPA100 pA
50 ms
c d
Area under Ca transient

0
Rise slope (µM s–1)

120 200 pA
–200 0 200 400 600
(0.25 × 10–3 µM s)

4
* Time (s)
Decay  (ms)

2+

80 Burst 1
2 f Sk2
–/–
Ctrl g 250
40 **

Apa-sens current
in CPA (% of ctrl)
200
0 Burst 2 +CPA 200 pA
** Burst 3 200 ms 150 **
0
Burst 1 Burst 2 Burst 3 0 10 20 30 Sk1–/–
Ctrl 100
Ampl of LT burst (mV)
50
Figure 5 Repetitive low-threshold bursting is accompanied by a decrease in +CPA 200 pA 0
the amplitude of D[Ca2+]i. (a) Fluorescent signals during the generation of 200 ms WT Sk2 –/– Sk1–/–
repetitive LT bursting. D[Ca2+]i was acquired at 125 Hz in a dendritic
segment shown in the inset (boxed area). (b) Overlay of the first three [Ca2+]i Figure 6 SERCAs antagonize SK2 current activation by T-type Ca2+, but
transients shown in a, labeled as 1, 2 and 3. Means ± s.e.m. of the peak not by HVA Ca2+, currents. (a) Current responses after membrane
D[Ca2+]i reached in five cells for bursts 1 and 2, and in two cells for burst 3 hyperpolarization (–40 mV, 125 ms) in control (Ctrl), CPA (10 mM) and
are shown in the histogram. *** denotes P o 0.001. (c) Rising slope (filled apamin (+Apa, 100 nM). (b) Time course of apamin-sensitive current before
bars) and decay time constant (open bars) of D[Ca2+]i in three successive and during CPA application (n ¼ 7). Average current before CPA is also
bursts, presented as means ± s.e.m. **P o 0.005 and *P o 0.02. shown (dotted line). (c) Apamin-sensitive currents before (Ctrl) and after CPA
(d) Linear relation between the LT burst amplitude and the area underneath application (10 mM), with superimposed black lines showing the bi-
the Ca2+ transient. The LT burst amplitude was determined from the burst exponential fits (time constants t1 ¼ 33 ms and t2 ¼ 429 ms in Ctrl, and t1
threshold to the peak of the T current–induced depolarization. Data points ¼ 41 ms and t2 ¼ 946 ms in CPA). The slow current component contributed
were pooled for all bursts from n ¼ 5 cells. The squared correlation 32.6% to the total current in CPA and 19.6% in control. Current traces used
coefficient for the linear regression was 0.834 and P o 0.0001. for digital subtraction are shown overlaid in inset. Scale bars, 100 pA,
200 ms. (d) Time course of apamin-sensitive currents evoked by depolari-
zation (30 mV, 125 ms) before and during CPA application (n ¼ 7). Average
current before CPA is shown (dotted line). (e) As in a, with cells from Sk2 –/–
Regulation of T-to-SK2 channel coupling and Sk1–/– mice. (f) Apamin-sensitive currents from cells recorded in e.
Are there mechanisms that regulate SK2 channel activation by Dotted lines denote 0 pA. (g) CPA-induced changes in apamin-sensitive
T channels and the dampening of oscillations? We assessed whether currents relative to baseline in wild-type (WT, n ¼ 9), Sk2 –/– (n ¼ 5) and
variable T-to-SK2 channel coupling contributed to a diversification of Sk1–/– (n ¼ 10) cells, presented as means ± s.e.m. In Sk2 –/–, the current
nRt discharge patterns7,10 by quantifying Ca2+ signals during repetitive obtained as apamin-sensitive before CPA was o1 pA. **P o 0.01.
low-threshold bursts in magfura-2–filled cells. Up to three bursting
cycles of a dampened oscillation were imaged (Fig. 5a), each of which oscillatory cycles (P o 0.0001; Fig. 5d), and repeated voltage gating
was accompanied by rapid elevations of the [Ca2+]i that largely decayed of T currents at frequencies comparable to those of dampened
(o10% of the peak) before the generation of the next transient oscillations resulted in cumulative current inactivation (Supplemen-
(n ¼ 5 cells). The D[Ca2+]i values showed a marked decrement in tary Fig. 3 online). Consequently, diminished D[Ca2+]i produced
amplitude from one oscillatory cycle to the next (n ¼ 5, P o 0.001; smaller SK2 currents (Supplementary Figs. 1 and 3). Taken together,
Fig. 5b), suggesting that fewer T channels open in successive oscillatory use-dependent inactivation mechanisms limited activation of
cycles (Supplementary Movie 1 online). Furthermore, the rising slope T currents during repetitive oscillations, thereby leading to
of successive D[Ca2+]i became shallower, while decay time constants fading D[Ca2+]i, attenuated SK2 channel activation and oscilla-
increased (Fig. 5c), indicating a temporal blurring in the synchrony of tory dampening.
T channel activation. Finally, the area beneath the D[Ca2+]i signals The gating of SK currents may be potentiated by Ca2+-induced Ca2+
correlated linearly with low-threshold burst amplitude throughout release21–24. To assess whether depleting intracellular Ca2+ stores

NATURE NEUROSCIENCE VOLUME 11 [ NUMBER 6 [ JUNE 2008 687


ARTICLES

Ctrl +CPA (10 µM) Sk2–/– mice, but potentiated the outward cur-
a b Control 6 rents in nRt neurons of Sk1–/– mice (Fig. 6e),
–64 mV –64 mV
5 CPA (10 µM) and revealed a slowly decaying apamin-sensi-

Number of LT bursts
4
tive current component (Fig. 6f,g). Thus,
SERCAs selectively clear the Ca2+ flowing
3
through T channels and antagonize SK2 chan-
–79 mV –80 mV 2 nel activation. These data suggested that T
1 channels, SK2 channels and SERCAs are func-
0
tionally colocalized, such that SK2 channels
© 2008 Nature Publishing Group http://www.nature.com/natureneuroscience

–87 mV –87 mV and SERCA compete for available Ca2+ enter-


–90 –80 –70 –60
Membrane potential (mV)
ing through T channels.
20 mV
We evaluated whether such competitive
200 ms
interaction has a role in rhythmic nRt dis-
c *** d charges. In control nRt cells, the number of
* low-threshold bursts showed a bell-shaped
–67 mV dependence on the resting membrane poten-
Number of LT bursts

2
tial preceding a hyperpolarizing current injec-
tion (–100 pA, 400 ms; Fig. 7a,b). CPA
1 (10 mM) induced a prolongation of the
AHPs following the low-threshold bursts
–68 mV (n ¼ 9; Fig. 7a) and lengthened the time
0
10 mV
spent bursting. Increased bursting occurred
–90 –80 –70 –60 –50 at the peak of the bell-shaped curve
400 ms
Membrane potential (mV)
(Fig. 7b,c), indicating that SERCA regulates
the dynamic range of bursting. The membrane
Figure 7 SERCAs modulate the strength of nRt oscillations. (a) Discharge patterns of an nRt cell before
(Ctrl) and after CPA application, evoked by negative current injection (–100 pA). The cell was held at potential responses to the hyperpolarizing
different membrane potentials, which are indicated to the left of each trace. (b) The number of LT burst current steps remained unaltered (control,
discharges in the cell presented in a, plotted against holding potential. (c) As in b, showing pooled data –22.8 ± 3.3 mV and –24.4 ± 2.2 mV from
for eight cells (means ± s.e.m.). Membrane potentials were binned into units of 2.5 mV. *P o 0.05. –72 and –77 mV; CPA, –24.2 ± 2.3 mV and
(d) Effect of CPA (10 mM, bottom trace) on oscillatory discharges of an nRt cell recorded in –26.4 ± 2.1 mV; n ¼ 9, P 4 0.05). The effects
perforated-patch mode.
of CPA were also tested on cells using perfo-
rated patch-clamp recordings that preserve
weakened T-to-SK2 channel coupling in nRt, we bath-applied the intracellular Ca2+ homeostasis. In this configuration, nRt neurons
SERCA inhibitor cyclopiazonic acid (CPA, 10 mM). Contrary to showed more bursts, as well as more variable burst discharge patterns
expectation, the amplitude of the outward SK2 current following T (Fig. 7d). The CPA effects were evident as a marked prolongation of the
currents was markedly enhanced (from 14 ± 30 pA to 91 ± 33 pA, n ¼ bursting pattern (n ¼ 5 different nRt cells).
10, P o 0.001; Fig. 6a). Digital subtraction in a subgroup of seven cells We reproduced basic aspects of nRt oscillatory dynamics and
confirmed that the apamin-sensitive current increased in the presence their regulation by SERCA in a computational model of a single-
of CPA (n ¼ 7, P o 0.02; Fig. 6b). Thapsigargin (4 mM in the patch compartment cell incorporating previously described phenomenolo-
pipette), a less-selective inhibitor of SERCAs, also produced a current gical models of T currents, SK currents and SERCAs (Supplementary
augmentation (51 ± 12 pA to 153 ± 33 pA, n ¼ 7, P o 0.03). In the Fig. 4 online). This model supported the conclusion that we had
presence of CPA, apamin-sensitive currents showed a bi-exponential identified the three major interacting partners controlling the dynamics
decay that was similar to that of controls (P 4 0.4 for both time of nRt cell oscillations.
constants), but the slow component now contributed 29.7 ± 7.6% of
the total current (n ¼ 7, P o 0.05 compared with control; Fig. 6c). The Altered NREM sleep in Sk2–/– mice
amplitude of the T current was largely preserved (B10% decrease; Oscillatory burst discharges in nRt in vivo accompany characteristic
control, –510 ± 88 pA; CPA, –456 ± 80 pA; n ¼ 7, P o 0.05) and the slow waves found in EEGs during NREMS1,2. To determine whether the
decay time remained unaltered compared with cells exposed to apamin cellular mechanisms identified here are relevant for sleep oscillations, we
only (tdecay ¼ 31.3 ± 3.1 ms, n ¼ 7, P 4 0.05). Blocking SERCAs thus studied sleep EEGs in Sk2–/– mice20. EEG spectral profiles between
potentiated SK2 current amplitudes and a slowly decaying component, 0–35 Hz were examined in freely behaving Sk2–/– (n ¼ 6) and wild-type
consistent with rapid Ca2+ sequestration in endoplasmic reticulum, (n ¼ 7) mice during NREMS, REMS and while the mice were awake.
rather than Ca2+ release from endoplasmic reticulum, regulating T-to- Low-frequency, high-amplitude oscillations dominate the NREMS
SK2 channel coupling. EEG, and the thalamic contributions to these are well established1,2.
The effects of SERCA antagonism on SK2 currents could reflect The REMS EEG is instead dominated by theta (5–10 Hz) oscillations,
increased [Ca2+]i levels or SK2 channel modifications25, rather than an partially of hippocampal origin26. Various frequency components con-
effect on T-to-SK2 channel coupling. We tested whether CPA affected tribute to the waking EEG, including theta activity. In Sk2 –/– mice, the
SK2 currents activated via Ca2+ entry through HVA channels. CPA EEGs of all arousal states showed a reduction in power (Fig. 8a),
produced a small, but nonsignificant, reduction of the apamin-sensitive consistent with the expression of SK2 in numerous brain regions19.
SK2 currents evoked by depolarization (n ¼ 7, P ¼ 0.055; Fig. 6d and However, the lack of SK2 channels compromised the NREMS EEG to
Supplementary Methods). Moreover, CPA (10 mM) had no effect on the greatest extent. We observed an almost fourfold decrease in the delta
currents activated after a hyperpolarizing command in nRt neurons of (1–4 Hz) frequency range and a more than threefold reduction in the

688 VOLUME 11 [ NUMBER 6 [ JUNE 2008 NATURE NEUROSCIENCE


ARTICLES

Figure 8 Lack of Sk2 greatly impacts the EEG


and fragmentation of NREMS. (a) Examples of a 150 µV Sk2
–/–
Sk2
+/+ 700
NREMS REMS
c
EEG
EEG and EMG traces for Sk2 –/– (left) and Sk2 +/+ 300 µV 175

Sigma peak (%)


600
(right) mice at the NREMS-to-waking (upper) and 100 µV
EMG *
REMS-to-waking (lower) transitions. Despite a 150

Sigma peak (%)


NREMS Waking NREMS Waking
4s 500
pronounced reduction in EEG amplitude, NREMS
125
preserved its characteristic EEG signature (see 400
REMS Waking REMS Waking
inset at twofold amplification). Scale bars apply to 100
300
all traces (excepting the inset). Examples were b
taken from two simultaneously recorded mice at –/– +/+ 200

µV per 0.25 Hz
10 Sk2 Sk2
1 h after lights on. (b) Spectral analysis of the +/+
Sk2
© 2008 Nature Publishing Group http://www.nature.com/natureneuroscience

EEG power density (µV per 0.25 Hz)


30
100
EEG quantified the reduction in EEG activity in 5 Sk2
–/–

Sk2 –/– (upper left) versus Sk2 +/+ (right) mice. 25


–3 –2 –1 0 1 2

2
Genotype differences (lower left) were most 0 Time (min)
0 10 20 30 20
pronounced during NREMS at frequencies below –/–
Sk2 versus Sk2
+/+
15 Hz. Horizontal bars mark frequency bins with 1
15
NREMS episodes
d
Fold decrease
Brief
significant genotype differences (post hoc t-tests,

2
2 awakenings Short Long
10
P o 0.05). Color coding of bars matches that of

No. episodes per


NREM sleep
REM sleep 60 * 60

h of NREMS
the EEG spectra. (c) Time course of EEG activity 3 Waking
5 *
in the spindle frequency band (that is, sigma 40 40
4
power; 10–15 Hz) at NREM-REMS transitions. 0
20 20
0 10 20 30 0 10 20 30
The prevailing level of sigma power during
EEG frequency [Hz] 0 0
NREMS (dashed horizontal lines) and its surge
before REMS onset (see inset; peak values
expressed as % of the prevailing level) were reduced. Sigma power was expressed as the percentage of values during REMS when spindle oscillations are
absent (see Supplementary Fig. 5 for details). Red bar denotes 4-s epochs with significant genotype differences in sigma power; asterisk indicates significant
difference in sigma peak power (P o 0.05). (d) Sleep fragmentation, measured as the number of brief awakenings (r16 s) and the number of short NREMS
episodes (r1 min; both expressed per h of NREMS), was increased in Sk2–/– mice. Asterisks mark significant genotype differences (P o 0.05). Histogram
data are presented as means ± s.e.m.

sleep spindle (10–15 Hz) band (Fig. 8b). The reduced contribution contributes to the amplification of low-frequency thalamic oscillations
of slow oscillations to the NREMS EEG persisted even after taking into large-scale EEG waves.
the differences in overall EEG amplitude into account (Supple-
mentary Fig. 5 online). For waking and REMS, a pronounced T channels in nRt dendrites
reduction was observed in the 10-Hz range (Fig. 8b). This decrease T current–dependent Ca2+ signals account for the overwhelming
was primarily a result of a slowing of EEG theta oscillations (Supple- majority of D[Ca2+]i in nRt dendrites during the low-threshold burst
mentary Fig. 5), consistent with SK2 channel involvement in hippo- mode of action-potential discharge, with negligible contributions from
campal discharges at theta frequencies27. In addition, Sk2–/– mice HVA currents. Moreover, tonic action-potential discharge evoked
showed a diminished surge of sleep spindle activity that is characteristic minor [Ca2+]i signals, showing that low-threshold bursting represents
of the transition from NREMS to REMS (Fig. 8c and Supple- a unique Ca2+ source for nRt dendrites. T currents are responsible for
mentary Fig. 5)28,29. the majority of Ca2+ entry in the dendrites of some types of cerebellar31
Behaviorally, Sk2–/– mice showed greater NREMS fragmen- and olfactory bulb neurons32 and in invertebrate heart interneurons33,
tation, such as more-frequent brief awakenings from NREMS and a but not in bursting thalamocortical neurons34,35. Thus, the nRt may
higher number of short NREMS periods (Fig. 8d), while leaving the belong to a small group of neuronal types with dendrites that are
time spent in NREMS unaltered (Supplementary Table 1 online). These specialized for handling Ca2+ entry through T channels and in
are behavioral signs that are indicative of decreased sleep depth, compartmentalizing targets for these, including Ca2+-dependent ion
consistent with the reduction of EEG delta activity30. This suppression channels and Ca2+ uptake sites, but also sites of vesicular release32,33.
of prominent low-frequency components of the NREMS EEG, Cells in the nRt express mRNAs for Cav3.2 and Cav3.3 isoforms8, and
accompanied by sleep disturbances, suggested that SK2 channel T currents show distinct kinetic properties according to their sub-
activity contributed to generating some of the physiological hallmarks cellular localization, characterized by rapidly inactivating somatic and
of NREMS. more slowly decaying dendritic currents9. How the peculiar inactiva-
tion and recovery characteristics of the corresponding channels36,37
DISCUSSION shape Ca2+ entry and SK2 current kinetics remains to be determined.
We report here that, in the dendrites of nRt neurons, the T currents In addition, synaptic activity during network oscillations, in
that are essential for low-threshold burst generation are components of particular Ca2+ influx through NMDA receptors, may further con-
a specialized Ca2+ signaling triad that also includes SK2 channels and tribute to D[Ca2+]i in nRt dendrites and regulate synaptically located
SERCAs. We found that the selective interaction between T and SK2 channels38.
SK2 channels is determined by their spatial coexpression in a substantial Assuming a single-channel conductance of 1 pS and a dendritic
portion of nRt dendrites and underlies the robust oscillatory properties diameter of 5–8 mm, a T-type Ca2+ conductance density of B0.6–1.0
of nRt cells. The lack of SK2 channels was accompanied by a mS cm–2 would be required to achieve the [Ca2+]i quantified here,
marked decrement in the NREMS EEG frequency bands that correlated which is consistent with a computational study16. Combined with
with rhythmic burst discharges of thalamic neurons, including those their extended arborizations, electrical connectivity via gap junctions
from the nRt1,2,5,6. Consistent with weakened NREM sleep power, and long-range reciprocal interactions5, nRt dendrites appear to
Sk2–/– mice showed fragmented sleep patterns. We propose that nRt be uniquely equipped to form a plexus of vigorously oscillating
dendrites are endowed with a dendritic Ca2+ signaling complex that membrane surfaces.

NATURE NEUROSCIENCE VOLUME 11 [ NUMBER 6 [ JUNE 2008 689


ARTICLES

SK2 channels in nRt dendrites Role of SK2 channels in low-frequency EEG waves of NREMS
The latency to SK current activation was B14 ms from the peak Burst discharges in nRt are most prominent in vivo during periods of
of the T current, indicating a close-to-maximal rate of exposure of EEG synchronization, such as during NREMS, while tonic firing
SK2 channels to the Ca2+ entering through T channels14,39. Moreover, predominates during waking and during REMS1,2,5. In particular, we
the peak [Ca2+]i levels reached in the dendrites (B0.7 mM) are close to note that EEG power was reduced most strongly during NREMS in
what is needed to maximally activate SK2 channels14,40. These estimates Sk2–/– mice, suggesting that SK2 channels and rhythmic cellular dis-
underscore the idea that the thin dendrites of nRt cells are functionally charges are important in boosting the network activities underlying low-
specialized to ensure rapid and selective Ca2+-activated K+ signaling frequency sleep oscillations. Such a role is consistent with the pacemak-
via colocalization with a high density of Ca2+ sources, thereby supplant- ing function of nRt for thalamocortical oscillations, which is based on its
© 2008 Nature Publishing Group http://www.nature.com/natureneuroscience

ing the need for integrating the channels into a nanodomain endogenous cellular rhythms and arises out of the widespread synaptic
protein complex. projections of nRt into thalamus and its strong cortical innervation5,6.
SK2 channel–dependent AHPs are permissive for repetitive low- At present, however, we cannot exclude other explanations, as EEG
threshold burst generation because they allow T channels to recover rhythms result from interplay between thalamic and cortical net-
from inactivation. These AHPs also accelerate low-threshold burst works1,2. Indeed, in the absence of SK2 channels, low-frequency
termination, as abolishing them pharmacologically or genetically oscillations in the EEG persist, although they are weakened, which is
reduced the T current decay slope and led to slowly decaying plateau consistent with the nRt not being the only site of rhythm generation1,2.
potentials. Such a role is consistent with previous descriptions of SK Which aspects of nRt function in thalamocortical rhythm generation
currents in the context of terminating plateau potentials22,41. Further- could relate to the major weakening of low-frequency bands in the
more, decremental activation of SK2 currents also accompanies, and NREM EEG of Sk2 –/– mice? The thalamic counterpart of delta EEG
contributes to, the cessation of endogenous nRt oscillations. Such waves has been principally attributed to thalamocortical neurons, which
dampening was previously attributed to slowly activating, Ca2+- oscillate independently of SK channels1,2. However, clock-like dis-
dependent cationic currents11,12,42, but the extent to which these charges of nRt cells have been described11,12,46, and cortex may facilitate
additional currents were instrumental for oscillatory dampening and burst generation of thalamocortical neurons at delta frequencies
what their biophysical and molecular properties were remained open through nRt activation1, thus leaving room for SK2 channel-dependent
questions. We found minor temporal summation of [Ca2+]i during nRt oscillations to contribute to thalamic discharges in the delta band.
repetitive low-threshold bursts, and thus little support for a scenario of The spindle frequency band relies on reciprocal synaptic interactions
an accruing activation of Ca2+-dependent cationic currents, although between thalamocortical projection neurons and nRt neurons, and on
localized submembrane Ca2+ signaling events cannot be excluded. the synaptically triggered recruitment of low-threshold bursts in both
Instead, T-type Ca2+ currents diminished during repetitive activation, cell types1,5,6. The weakened sleep spindle generation in Sk2–/– mice
produced smaller [Ca2+]i elevations and failed to generate rapid out- suggests that these intrathalamic interactions are strongly attenuated,
ward SK2 currents. For these reasons, we propose that a mechanism possibly because of failures in nRt low-threshold bursting caused by
involving T-type Ca2+ and SK2 channel coupling may, for the most improper membrane repolarization. Thus, SK2 channels help enable
part, account for dampened nRt oscillations. oscillatory bursting triggered through glutamatergic input, eventually
boosting oscillatory interaction in the nRt-thalamic network. Alto-
Role of SERCAs in regulating T-to-SK2 channel coupling gether, although the nRt is clearly implicated in some NREMS oscilla-
SERCAs ensure the appropriate filling of the endoplasmic reticulum tions1,6, a substantial experimental effort in recording in vivo will be
with Ca2+, which can become available to regulate ion channels via needed to elucidate the detailed role of SK2 channels in nRt dendrites in
Ca2+-induced Ca2+ release. However, the possibility that SERCAs the neuronal networks underlying NREMS EEG oscillations.
actively control electrical rhythms via their sequestrating function has The genetic loss of K+ channels has been linked to changes in the
not been described, with the exception of cardiac cells43. Notably, sleep-wake cycle of mammals47,48 and is accompanied by increased
SERCAs act specifically on T current–dependent Ca2+ entry, suggesting wakefulness and diminished power in the sleep EEG. Moreover,
that they are colocalized with T-type Ca2+ and SK2 channels, but not patients with Morvan’s syndrome, a rare autoimmune disorder with
with HVA Ca2+ channels. SERCAs may be rapid to compete with SK2 pronounced insomnia, produce autoantibodies against K+ channels49.
channels for available Ca2+ ions, as they bind Ca2+ with an affinity of The EEG alterations resulting from the lack of these ion channels have
0.27–0.4 mM44 and, in cerebellar Purkinje neurons at 20–24 1C, remove not yet been linked to distinct neuronal oscillatory mechanisms. To the
Ca2+ at a rate of up to B0.6 mM s–1 for [Ca2+]i in the low micromolar best of our knowledge, our study is the first to start establishing a
range45. In support of this model, SK2 currents are potentiated in total correlation between the molecular bases of endogenous neuronal
amplitude and in the slowly decaying current component when oscillations and EEG waves. We propose that determining the roles
SERCAs are blocked. Moreover, interburst AHPs are lengthened and of SK2 channels in thalamocortical networks could point out targets for
the number of bursts is increased in CPA. Thus, it is likely that SERCAs improving NREMS continuity and/or depth in disorders in which
shorten the exposure of SK2 channels to Ca2+ during AHP, thereby NREMS quality is impacted, such as in primary insomnia or insomnias
limiting T channel recovery and promoting oscillatory dampening. associated with psychiatric or neurological disorders.
However, SERCAs may act rapidly enough to antagonize SK2 channel–
induced deactivation of T channels, thereby effectively potentiating
T channel activation and SK2 channel exposure to [Ca2+]i. More Ca2+ METHODS
Electrophysiological recordings. Horizontal slices 300-mm thick were prepared
imaging experiments will be required to understand the detailed
from 17–23-d-old wild-type, Sk1–/– or Sk2–/– mice and their corresponding
consequences of SERCA activity on dendritic [Ca2+]i signals and wild-type littermates, as described previously35 and approved by the
on SK2 channel function. Physiologically controlled SERCA Veterinäramt of the Canton Basel-Stadt. Extracellular and patch-clamp record-
activity in nRt cells could represent a method of regulating ings were obtained according to well-established procedures at 33.5–35 1C. For
intrinsic oscillatory strength and associated Ca2+ uptake into the information on recording solutions, data acquisition, analysis and control
endoplasmic reticulum. experiments, see Supplementary Methods.

690 VOLUME 11 [ NUMBER 6 [ JUNE 2008 NATURE NEUROSCIENCE


ARTICLES

Genotyping. Homozygous Sk2–/– and Sk1–/– mice were obtained from crossings 3. Contreras, D. The role of T-channels in the generation of thalamocortical rhythms. CNS
of heterozygotic pairs, bred to a wild-type C57Bl/6J background and genotyped Neurol. Disord. Drug Targets 5, 571–585 (2006).
4. Perez-Reyes, E. Molecular physiology of low voltage–activated T-type calcium channels.
as described20. Physiol. Rev. 83, 117–161 (2003).
5. Pinault, D. The thalamic reticular nucleus: structure, function and concept. Brain Res.
Fluorescent Ca2+ imaging. For Ca2+ imaging experiments, the patch pipette Brain Res. Rev. 46, 1–31 (2004).
solution was supplemented with magfura-2 (3 mM, with the Kd determined 6. Fuentealba, P. & Steriade, M. The reticular nucleus revisited: intrinsic and network
experimentally, see Supplementary Methods) or bis-fura-2 (1 mM, Kd ¼ properties of a thalamic pacemaker. Prog. Neurobiol. 75, 125–141 (2005).
0.525 mM), but without EGTA, unless otherwise specified. Fluorescence was 7. Domich, L., Oakson, G. & Steriade, M. Thalamic burst patterns in the naturally sleeping
cat: a comparison between cortically projecting and reticularis neurones. J. Physiol.
excited with a 150 W ultra-stable Xenon Arc lamp (Cairn Research) at 387 ± (Lond.) 379, 429–449 (1986).
6 nm, detected with a NeuroCCD-SM camera (RedShirt) at 510 ± 45 nm and 8. Talley, E.M. et al. Differential distribution of three members of a gene family encoding
sampled typically at 125 or 500 frames s–1 to resolve the signal kinetics. The
© 2008 Nature Publishing Group http://www.nature.com/natureneuroscience

low voltage–activated (T-type) calcium channels. J. Neurosci. 19, 1895–1911 (1999).


field was 125  125 mm (80  80 pixels), giving a pixel size of B1.5  1.5 mm. 9. Joksovic, P.M., Bayliss, D.A. & Todorovic, S.M. Different kinetic properties of two T-type
Ca2+ currents of rat reticular thalamic neurones and their modulation by enflurane.
Bleaching was taken into account by subtracting trials without negative current J. Physiol. (Lond.) 566, 125–142 (2005).
injection. The brightest dendrite was selected for recordings. Data analysis was 10. Avanzini, G., de Curtis, M., Panzica, F. & Spreafico, R. Intrinsic properties of nucleus
carried out on D[Ca2+]i signals obtained by averaging fluorescence over the reticularis thalami neurones of the rat studied in vitro. J. Physiol. (Lond.) 416, 111–122
entire visible dendrite. For details on data analysis and dye calibration, see (1989).
11. Bal, T. & McCormick, D.A. Mechanisms of oscillatory activity in guinea-pig nucleus
Supplementary Methods. reticularis thalami in vitro: a mammalian pacemaker. J. Physiol. (Lond.) 468, 669–691
(1993).
Immunohistochemistry and pre-embedding immunoelectron microscopy. 12. Blethyn, K.L., Hughes, S.W., Toth, T.I., Cope, D.W. & Crunelli, V. Neuronal basis of the
Sections were treated for light and electron microscopy immunolabeling as slow (o1 Hz) oscillation in neurons of the nucleus reticularis thalami in vitro.
described previously50. These experiments were approved by the Institutional J. Neurosci. 26, 2474–2486 (2006).
13. Huguenard, J.R. & McCormick, D.A. Thalamic synchrony and dynamic regulation of
Animal Care and Use Committee of the University of Castilla-La Mancha.
global forebrain oscillations. Trends Neurosci. 30, 350–356 (2007).
Affinity-purified rabbit and guinea pig antibodies to SK2 were raised against 14. Pedarzani, P. et al. Control of electrical activity in central neurons by modulating the
amino acid residues 536–574 of the mouse SK2 (accession number NM080465) gating of small conductance Ca2+-activated K+ channels. J. Biol. Chem. 276,
and diluted at 1–2 mg ml–1. 9762–9769 (2001).
15. Ogden, D., Khodakhah, K., Carter, T., Thomas, M. & Capiod, T. Analogue computation of
EEG monitoring and analyses. We used adult (2.5–4 months old), female transient changes of intracellular free Ca2+ concentration with the low affinity Ca2+
indicator furaptra during whole-cell patch-clamp recording. Pflugers Arch. 429,
Sk2+/+ (n ¼ 7) and Sk2–/– (n ¼ 6) mice for EEG recordings. Mice were 587–591 (1995).
equipped with EEG and electromyogram (EMG) electrodes according to 16. Destexhe, A., Contreras, D., Steriade, M., Sejnowski, T.J. & Huguenard, J.R. In vivo,
standard procedures (see Supplementary Methods) approved by the Veterinary in vitro and computational analysis of dendritic calcium currents in thalamic reticular
Office of the Canton de Vaud. Continuous EEG and EMG recordings were neurons. J. Neurosci. 16, 169–185 (1996).
17. Debarbieux, F., Brunton, J. & Charpak, S. Effect of bicuculline on thalamic activity:
obtained for 24 (n ¼ 5) or 48 h (n ¼ 8). Offline, the behavioral states waking, a direct blockade of IAHP in reticularis neurons. J. Neurophysiol. 79, 2911–2918
NREMS and REMS were determined by visual inspection of the EEG and EMG (1998).
signals for consecutive 4-s intervals. The spectral content of the EEG was 18. Canepari, M., Auger, C. & Ogden, D. Ca2+ ion permeability and single-channel properties
estimated using a discrete Fourier transformation routine. of the metabotropic slow EPSC of rat Purkinje neurons. J. Neurosci. 24, 3563–3573
(2004).
19. Stocker, M. & Pedarzani, P. Differential distribution of three Ca2+-activated K+ channel
Computational modeling. For details of the model, see Supplementary Figure subunits, SK1, SK2 and SK3, in the adult rat central nervous system. Mol. Cell.
4. Equations were solved in Mathematica V.5.0 (Wolfram Research) using Neurosci. 15, 476–493 (2000).
Runge-Kutta integration. 20. Bond, C.T. et al. Small conductance Ca2+-activated K+ channel knockout mice reveal the
identity of calcium-dependent afterhyperpolarization currents. J. Neurosci. 24,
Statistical analysis. Two-tailed paired and unpaired t-tests were used for within 5301–5306 (2004).
and between group comparisons, respectively. For multiple comparisons 21. Stocker, M. Ca2+-activated K+ channels: molecular determinants and function of the SK
family. Nat. Rev. Neurosci. 5, 758–770 (2004).
between data obtained from wild-type, Sk1–/– and Sk2–/– mice, Dunnett’s T3 22. Bond, C.T., Maylie, J. & Adelman, J.P. SK channels in excitability, pacemaking and
post hoc test was used after significance was reached in a one-way analysis of synaptic integration. Curr. Opin. Neurobiol. 15, 305–311 (2005).
variance with factor ‘genotype’ and homogeneity of variances was tested with 23. Cui, G., Okamoto, T. & Morikawa, H. Spontaneous opening of T-type Ca2+ channels
Levene’s test. P o 0.05 was considered statistically significant. Analyses were contributes to the irregular firing of dopamine neurons in neonatal rats. J. Neurosci. 24,
11079–11087 (2004).
carried out using the Statistical Package for the Social Sciences V. 14. (SPSS). 24. Richter, T.A., Kolaj, M. & Renaud, L.P. Low voltage–activated Ca2+ channels are coupled
See Supplementary Figure 5 and Supplementary Table 1 for further details. to Ca2+-induced Ca2+ release in rat thalamic midline neurons. J. Neurosci. 25,
All data are presented as means ± s.e.m. 8267–8271 (2005).
25. Bildl, W. et al. Protein kinase CK2 is coassembled with small conductance Ca2+-
Note: Supplementary information is available on the Nature Neuroscience website. activated K+ channels and regulates channel gating. Neuron 43, 847–858 (2004).
26. Buzsáki, G. Theta oscillations in the hippocampus. Neuron 33, 325–340 (2002).
27. Kramár, E.A. et al. A novel mechanism for the facilitation of theta-induced long-term
ACKNOWLEDGMENTS
potentiation by brain-derived neurotrophic factor. J. Neurosci. 24, 5151–5161 (2004).
We thank M. Tafti, K. Vogt and N. Wanaverbecq for constructive input and 28. Franken, P., Malafosse, A. & Tafti, M. Genetic variation in EEG activity during sleep in
A. Reisch for carrying out preliminary electrophysiological experiments. We inbred mice. Am. J. Physiol. 275, R1127–R1137 (1998).
thank L. Acsády, A. Destexhe, B. Gähwiler, U. Gerber, C. Kopp and D. Ulrich for 29. Gottesmann, C. The transition from slow-wave sleep to paradoxical sleep: evolving facts
stimulating discussions and helpful comments on the manuscript. This work was and concepts of the neurophysiological processes underlying the intermediate stage of
supported by grants from the Swiss National Science Foundation (A.L.), the US sleep. Neurosci. Biobehav. Rev. 20, 367–387 (1996).
National Institutes of Health (P.F. and J.P.A.) and the Spanish Ministry of 30. Franken, P., Malafosse, A. & Tafti, M. Genetic determinants of sleep regulation in inbred
Education and Science (R.L.). mice. Sleep 22, 155–169 (1999).
31. Diana, M.A. et al. T-type and L-type Ca2+ conductances define and encode the bimodal
firing pattern of vestibulocerebellar unipolar brush cells. J. Neurosci. 27, 3823–3838
Published online at http://www.nature.com/natureneuroscience
(2007).
Reprints and permissions information is available online at http://npg.nature.com/ 32. Egger, V., Svoboda, K. & Mainen, Z.F. Mechanisms of lateral inhibition in the olfactory
reprintsandpermissions bulb: efficiency and modulation of spike-evoked calcium influx into granule cells.
J. Neurosci. 23, 7551–7558 (2003).
33. Ivanov, A.I. & Calabrese, R.L. Intracellular Ca2+ dynamics during spontaneous and
1. Steriade, M. Grouping of brain rhythms in corticothalamic systems. Neuroscience 137, evoked activity of leech heart interneurons: low-threshold Ca currents and graded
1087–1106 (2006). synaptic transmission. J. Neurosci. 20, 4930–4943 (2000).
2. Crunelli, V., Cope, D.W. & Hughes, S.W. Thalamic T-type Ca2+ channels and NREM sleep. 34. Munsch, T., Budde, T. & Pape, H.C. Voltage-activated intracellular calcium transients in
Cell Calcium 40, 175–190 (2006). thalamic relay cells and interneurons. Neuroreport 8, 2411–2418 (1997).

NATURE NEUROSCIENCE VOLUME 11 [ NUMBER 6 [ JUNE 2008 691


ARTICLES

35. Kuisle, M. et al. Functional stabilization of weakened thalamic pacemaker channel 44. Lytton, J., Westlin, M., Burk, S.E., Shull, G.E. & MacLennan, D.H. Functional compar-
regulation in rat absence epilepsy. J. Physiol. (Lond.) 575, 83–100 (2006). isons between isoforms of the sarcoplasmic or endoplasmic reticulum family of calcium
36. Frazier, C.J. et al. Gating kinetics of the a1I T-type calcium channel. J. Gen. Physiol. pumps. J. Biol. Chem. 267, 14483–14489 (1992).
118, 457–470 (2001). 45. Fierro, L., DiPolo, R. & Llanò, I. Intracellular calcium clearance in Purkinje cell somata
37. Uebachs, M., Schaub, C., Perez-Reyes, E. & Beck, H. T-type Ca2+ channels encode prior from rat cerebellar slices. J. Physiol. (Lond.) 510, 499–512 (1998).
neuronal activity as modulated recovery rates. J. Physiol. (Lond.) 571, 519–536 46. Amzica, F., Nunez, A. & Steriade, M. Delta frequency (1–4 Hz) oscillations of
(2006). perigeniculate thalamic neurons and their modulation by light. Neuroscience 51,
38. Ngo-Anh, T.J. et al. SK channels and NMDA receptors form a Ca2+-mediated feedback 285–294 (1992).
loop in dendritic spines. Nat. Neurosci. 8, 642–649 (2005). 47. Vyazovskiy, V.V. et al. Sleep EEG in mice that are deficient in the potassium channel
39. Marrion, N.V. & Tavalin, S.J. Selective activation of Ca2+-activated K+ channels by subunit K.v.3.2. Brain Res. 947, 204–211 (2002).
colocalized Ca2+ channels in hippocampal neurons. Nature 395, 900–905 (1998). 48. Joho, R.H., Marks, G.A. & Espinosa, F. Kv3 potassium channels control the duration of
40. Kohler, M. et al. Small-conductance, calcium-activated potassium channels from different arousal states by distinct stochastic and clock-like mechanisms. Eur. J.
© 2008 Nature Publishing Group http://www.nature.com/natureneuroscience

mammalian brain. Science 273, 1709–1714 (1996). Neurosci. 23, 1567–1574 (2006).
41. Cai, X. et al. Unique roles of SK and Kv4.2 potassium channels in dendritic integration. 49. Liguori, R. et al. Morvan’s syndrome: peripheral and central nervous system and cardiac
Neuron 44, 351–364 (2004). involvement with antibodies to voltage-gated potassium channels. Brain 124,
42. Pape, H.C., Munsch, T. & Budde, T. Novel vistas of calcium-mediated signaling in the 2417–2426 (2001).
thalamus. Pflugers Arch. 448, 131–138 (2004). 50. Luján, R., Nusser, Z., Roberts, J.D., Shigemoto, R. & Somogyi, P. Perisynaptic
43. Misquitta, C.M., Mack, D.P. & Grover, A.K. Sarco/endoplasmic reticulum Ca2+ location of metabotropic glutamate receptors mGluR1 and mGluR5 on dendrites
(SERCA)-pumps: link to heart beats and calcium waves. Cell Calcium 25, 277–290 and dendritic spines in the rat hippocampus. Eur. J. Neurosci. 8, 1488–1500
(1999). (1996).

692 VOLUME 11 [ NUMBER 6 [ JUNE 2008 NATURE NEUROSCIENCE


Cueni et al. Supplementary Information 1

Supplementary Information
T–type Ca2+ channels, SK2 channels, and SERCAs gate sleep–related

oscillations in thalamic dendrites

Lucius Cueni, Marco Canepari, Rafael Luján, Yann Emmenegger, Masahiko Watanabe, Chris
T. Bond, Paul Franken, John P. Adelman and Anita Lüthi

1) Supplementary Methods
Electrophysiological recordings. Patch–clamp recordings were obtained according to well–
established procedures at 33.5 – 35 °C. The bath was constantly perfused with fresh medium
at a rate of 2.5 – 4 mlmin–1 that contained (in mM): 131 NaCl; 2.5 KCl; 1.25 NaH2PO4; 1.2
MgCl2; 2 CaCl2; 26 NaHCO3, 18 dextrose, 1.7 L(+)–ascorbic acid. The caudal portion of the
nRt was localized before pipette positioning using a low–power (10 x) objective, and a high–
power water immersion objective (40 x) and near–infrared differential interference contrast
optics were used for visualizing cells. Patch pipettes were pulled from borosilicate glass
tubing (TW150F–4, outer diameter 1.5 mm, World Precision Instruments) on a vertical two–
step puller (PP–83, Narishige) and filled with the following solution (in mM): 130 KMeSO4,
10 KCl, 10 HEPES, 0.1 mM EGTA, 2 MgCl2, 2 K–ATP, 0.2 Na–GTP, 10 phosphocreatine,
adjusted to 290 mOsm with sucrose, pH 7.25. The low concentration of EGTA adds to the
intracellular Ca2+ buffering capacity by > 500 and hence clamps steady–state Ca2+ levels to
low values, while not affecting [Ca2+]i transients generated by low-threshold bursts (see
Supplementary Methods on “Fluorescent Ca2+ imaging”). To isolate T–currents, Cs+–based
intracellular solutions were used, in which KMeSO4 was replaced by CsGluconate and KCl
by CsCl. Electrodes had resistances of 2.1 – 3.8 MΩ and yielded series resistances in the
range between 10 – 20 MΩ. Series resistance was constantly monitored throughout the
experiments and increases > 20% were not accepted. For MeSO4– and gluconate–based
pipettes, a liquid junction potential of –17 mV and –10 mV was taken into account,
respectively. Data from voltage– and current–clamp recordings were collected through an
Axopatch 200B amplifier (Molecular Devices), filtered at 2 kHz and acquired at 5 kHz using
pClamp 9.2 software (Molecular Devices). T–currents were evoked by hyperpolarizing
voltage commands (from –60 to –100 mV, 500 ms). Cells included in this analysis showed T–
currents with properties that fulfilled previously established criteria for acceptable voltage
control in intact nRt cells, including a smooth activation and decay (τdecay = 28.1 ± 3.2 ms at –
55 to –60 mV at 33.5 – 35 °C, n = 12) and a steady–state inactivation curve with a V50 of –

1
Cueni et al. Supplementary Information 2

79.3 ± 0.3 mV and a slope of 5.0 ± 0.2 mV (n = 12), close to values published previously for
acutely dissociated cells1. In K+–based recordings, neurons were generally clamped around
their resting membrane potentials (~–67 to –80 mV) and SK currents were evoked via hyper–
(–40 mV, 125 ms) or depolarizing voltage steps (+30 mV, 125 ms) from holding potentials
between –62 and –67 mV. T–current decay slope was determined by a linear fit to the 15 ms
time interval after current peak, during which decay is linear. Dampened oscillations were
obtained after brief membrane hyperpolarization (–100 pA, 400 ms), similar to findings in
vivo and in vitro2, 3. The discharge frequency decreased from ~10 Hz when current pulses
were injected at –60 mV, to ~ 4 – 6 Hz around –78 mV. Note that the properties of action
potentials and associated rapid AHPs may be distorted due to the electronic design of the
Axopatch 200B amplifier. Perforated patch–clamp recordings were achieved via including
gramicidin at 2.8 µM in the prefiltered patch pipette solution, which was then sonicated for 30
s. Gramicidin was prepared freshly in a 2.8 mM stock solution in dimethylsulfoxide. The
pipette tip was initially filled with gramicidin–free solution by brief immersion and backfilled
with gramicidin–containing patch pipette solution. Minimal pressure was applied to the patch
pipette only while crossing the surface of the bath and before cell contact. The cell–attached
configuration, with a seal resistance > 0.7 GΩ was obtained by applying negative pressure to
the patch pipette. Perforation was assessed in voltage–clamp by monitoring current responses
to 10 mV hyperpolarizing steps. When current transients reached values > 90 pA, the
recording configuration was switched to current–clamp and the experiment was started. Data
were analyzed off–line using pClamp 9.2 and Igor Pro V.5.0.5 software and are indicated as
means ± standard error (SEM).
SK currents were recorded at a constant holding voltage following a hyperpolarizing
step to gate T–currents. Apamin subtraction was carried out by digitally subtracting averaged
traces (typically 2 – 5 sweeps) obtained following bath application of 100 nM apamin. We
assessed the quality and stability of voltage–clamp control in these recordings with K+–based
electrodes in several ways. First, we tested whether apamin–sensitive currents, obtained by
digital subtraction, remained stable for the period required to perform application of
pharmacological substances via the bath or through the recording pipette (ca. 10 min).
Apamin–sensitive currents were unchanged after this time (109.2 ± 7.9 % of control
amplitude, n = 4, p > 0.4). Second, we took care to only carry out the subtraction when all
recording conditions, such as input resistance, series resistance and capacitive currents
remained unaltered before and after apamin application. We noted that some small inward T–
currents (< 15 % of total current) remained which were due to a slight, apamin–induced

2
Cueni et al. Supplementary Information 3

increase in the peak of the T–current. This likely resulted from apamin–induced decelerated
decay of the T–current (see e.g. Fig. 1a, b). However, these currents were small (< ~ 50 pA)
and decayed rapidly compared to the time course of the apamin–sensitive outward current.
These remaining currents will thus lead to an overestimation of the latency between peak T–
and SK currents, an error which does not affect the interpretation of the tight temporal
coupling between the two currents. Third, the decay time constant of the T–currents was 32.3
± 1.2 ms (n = 6) in apamin, close to the value obtained with Cs+–based electrodes (p > 0.05).
This suggests that, in the cells selected, voltage–clamp was comparable to the situation in
which K+ currents were blocked. Finally, we noted that the peak–to–peak latency between T–
currents and apamin–sensitive currents showed little variability in the 5 cells included in the
analysis (14.1 ± 0.3 ms, range 12.9 – 14.8 ms), further pointing to a stable time course of T–
currents and reproducible voltage–clamp across experiments. From these tests, we concluded
that our experimental conditions were such that voltage–clamp of T–currents at –62 to –67
mV was stable enough to allow for reliable activation of apamin–sensitive currents. However,
we also noted that our T–current amplitudes were comparatively small, which indicated that
we clamped only a portion of the whole–cell currents4, 5, likely those contained in soma and
proximal dendrites.
CPA produced a small but non–significant reduction of the apamin–sensitive SK2
currents evoked by depolarization (Fig. 6d). We obtained similar results when the
depolarizing voltage step was shortened to 60 ms to reduce Ca2+ influx (data not shown),
showing that limiting the duration of Ca2+ entry did not alter the polarity of CPA actions.

Fluorescent Ca2+ imaging. For experiments with magfura–2, fluorescence signals were
converted into changes of free Ca2+ concentration defined as Δ[Ca2+]i = Kd * (Fmin – F)/(F –
Fmax) where F is the fluorescence intensity (after correction for the slice auto–fluorescence)
and Fmin and Fmax are the fluorescence intensities at 0 and at saturating Ca2+, respectively.
Bis–fura–2–mediated signals were converted into dye–bound Ca2+ defined as Δ[DCa2+]i =
1mM * (Fmin – F)/(Fmin – Fmax)6. In the conversions of fluorescence signals either into Δ[Ca2+]i
or Δ[DCa2+]i, Fmin was approximated with the initial resting fluorescence, whereas F – Fmax
and Fmin – Fmax were approximated with F and Fmin respectively. For magfura–2 experiments
at 125 Hz acquisition rate (single trials), peak–to–peak noise for the brightest pixels was ~ 1
µM. Thus, to discriminate Δ[Ca2+]i signals in the order of 100 nM, fluorescence was averaged
over several trials (typically 4 – 9) and over dendritic segments of ~50 µm. Within this length,
inspection of smaller dendritic subsegments of 10 µm each revealed no non–uniformity of
signal amplitude within the experimental error.

3
Cueni et al. Supplementary Information 4

At 0.1 mM EGTA, Δ[Ca2+]i measured with magfura–2 was undistinguishable from that
without EGTA (Δ[Ca2+]i = 775 ± 82 nM, τ = 46 ± 8 ms, n = 4, p > 0.05).
Ca2+ signals were evaluated after calibration of dye affinities with solutions containing
known free [Ca2+] using established procedures. The Kd of magfura–2 with Mg2+ at pH 7.3
and 34 ºC, measured using EGTA–buffered solutions ([Ca2+] ~ 1 – 30 µM ), was 25 ± 5 µM
(standard deviation from least–squares interpolation). This estimate was similar to that
reported7, but slightly smaller than in studies8, 9 carried out at 22 – 24 ºC. The Kd of bis–fura–
2 with Mg2+, not relevant for the estimate of [DCa2+]i, was considered to be that reported by
Molecular Probes (0.525 µM). The buffer capacity of the cell, defined as K = [BCa2+]/[Ca2+]
where [BCa2+] is the transient Ca2+ bound to the endogenous cell buffer, was estimated as
Δ[DCa2+]i/Δ[Ca2+]i. This estimate is based on the approximation that 3 mM magfura–2 (buffer
capacity ~ 120) does not significantly alter the physiological Δ[Ca2+] and that in the presence
of 1 mM bis–fura–2 (buffer capacity ~ 1900) all the Ca2+ that enters the cell binds to the dye.

Immunohistochemistry and electron microscopy. Three P21 mice were deeply


anaesthetised by intraperitoneal injection of ketamine–xylazine 1 : 1 (0.1 mlkg–1 body weight)
and perfused through the ascending aorta for 13 – 18 min, first with 0.9 % saline for 1 min
followed by freshly prepared ice–cold fixative containing 4 % paraformaldehyde, 0.05 %
glutaraldehyde and ~ 0.2 % picric acid made up in 0.1 M phosphate buffer (PB, pH 7.4). After
perfusion, brains were removed from the skull and immersed in the same fixative for 2 h.
Tissue blocks containing the nRt were dissected and washed thoroughly in 0.1 M PB for
several hours. Coronal 60 µm – thick sections were then cut on a Vibratome (Leica VT1000)
and collected in 0.1 M PB. For light microscopy, free–floating sections were incubated in 10
% normal goat serum (NGS, Vector Laboratories, USA) diluted in Tris–buffered saline (TBS)
for 1 h. Sections were then incubated for 48 h in a solution of a primary antibody against SK2
at a final protein concentration of 1–2 µgml–1 each, diluted in TBS containing 1 % NGS. After
washes in TBS, the sections were incubated for 2 h in biotinylated goat anti–rabbit or goat
anti–guinea pig IgGs (Vector Laboratories) diluted 1 : 50 in TBS containing 1 % NGS. They
were then transferred into avidin–biotin–peroxidase complex (ABC kit, Vector Laboratories)
diluted 1 : 100 and left for 2 h at room temperature. Peroxidase enzyme activity was revealed
using 3,3´–diaminobenzidine tetrahydrochloride (DAB; 0.05 % in TB, pH 7.4) as the
chromogen and 0.01 % H2O2 as substrate. Finally, the sections were air–dried and
coverslipped prior to observation with a photomicroscope (DMRS, Leica) equipped with
differential interference contrast optics. Sections from Sk2–/– mice were incubated for the
same times in the same vials. For ultrastructural analysis, the silver–enhanced immunogold
4
Cueni et al. Supplementary Information 5

technique was used and immunogold particles along the plasma membrane and at intracellular
sites of morphologically identifiable somata, dendritic shafts and axon terminals were
assessed. We also measured the density of immunoparticles in particle/effective membrane
area in EM pictures (in particleμm–2) over plasma membrane compartments and statistically
compared these to the non–specific labelling densities (also given in particleμm–2).
Background labelling, assessed by determining particle density over nucleus, mitochondria
and myelin, was 0.05 ± 0.01 immunogoldμm–2.

EEG monitoring and analysis. Mice were kept individually in polycarbonate cages (31 x 18
x 18 cm) with food and water available ad libitum, and maintained on a 12 h light – 12 h dark
cycle (lights–on at 9:00 AM) at an ambient temperature of 24.5–25.5 °C. Body (Sk2+/+: 19.4 ±
0.7 g; Sk2–/–: 21.2 ± 0.6 g) and brain (Sk2+/+: 452 ± 8 mg; Sk2–/–: 449 ± 13 mg) weight did not
differ between genotypes. Age at time–of–recording was 17 weeks for 4 of the Sk2–/– mice; all
others were 11–weeks old. EEG and electromyogram (EMG) electrodes were implanted under
deep anaesthesia with a mixture of ketamine and xylazine (i.p., 75 and 10 mgkg–1,
respectively, at a volume of 8 μlg–1). Two gold–plated miniature screws (diameter 1.1 mm)
served as EEG electrodes and were screwed into the cranium over the right cerebral
hemisphere, in a fronto–parietal position10. Four additional anchor screws were implanted;
one over the right hemisphere and three over the left hemisphere. Two semi–rigid gold wires
served as EMG electrodes and were inserted between two neck muscles. The EEG and EMG
electrodes were soldered to a connector and the anchor screws were cemented to the skull.
Four to 8 days of recovery from surgery were allowed before animals were connected to the
recording leads. A minimum of 6 adaptation days (or 10 including recovery from surgery)
were scheduled before data collection. EEG and EMG signals were recorded continuously for
24 h (n = 5) or 48 h (n = 8) under undisturbed baseline conditions. The analogous signals
were digitized at 2 kHz and subsequently stored at 200 Hz on hard disc. The EEG was
subjected to a discrete–Fourier transformation yielding power spectra (range: 0.25 – 90 Hz,
resolution: 0.25 Hz, window function: hamming) for consecutive 4 – s epochs. Hardware
(EMBLA™) and software (Somnologica–3™) were purchased from Medcare/Flaga (Island).
Based on the EEG and EMG signals, the animal’s behavior was classified as REMS, NREMS,
or wakefulness, for consecutive 4 – s epochs according to standard criteria10. States were
scored by visual inspection of the EEG and EMG signals displayed on a PC monitor. Four–
second epochs containing EEG artifacts were marked, so they could be excluded from EEG
spectral analyses. For each state, an EEG spectral profile was constructed by averaging all 4 –

5
Cueni et al. Supplementary Information 6

s epochs scored as that state. Spectral changes at the NREM–to–REMS transition, calculation
of theta peak frequency in the REMS and waking EEG and the fragmentation of NREMS
were calculated as described previously10-12.

Pharmacological agents. Apamin, 1–EBIO and Thapsigargin were obtained from Tocris,
CPA, ω–conotoxinMVIIC and SNX–482 from Alomone Labs, TTX from Latoxan and
magfura–2, bis–fura–2 from Molecular Probes. EGTA, BAPTA, gramicidin D, nifedipine and
standard salts for electrophysiological solutions were purchased from Sigma–Aldrich or
Merck, L(+)–Ascorbic acid from VWR Prolabo and KMeSO4 from ICN Biomedicals.
Mibefradil was a kind gift of F. Hoffmann–La Roche Ltd, Basel Switzerland. Drugs (1–EBIO,
mibefradil, apamin, TTX, ω–CTXMVIIC, nifedipine, CPA) were maintained in 200 – 1000–
fold concentrated stock concentrations and applied to the bath at the concentrations indicated,
whereas SNX–482 was dissolved directly at final concentration in the bath solution. In some
experiments, 1–EBIO, apamin and ω–CTXMVIIC were applied focally through a local
puffing pipette attached to a picospritzer (World Precision Instruments). Thapsigargin was
included at 4 µM in the patch pipette through back–filling, as described for gramicidin.

2) Supplementary Note

Electrophysiological properties of Sk1–/– and Sk2–/– and littermate control animals

Values for passive input resistance, measured with brief 10 mV hyperpolarizing voltage
steps, were 250 ± 16 MΩ (n = 41) in Sk1–/– animals, and 207 ± 16 MΩ in Sk2–/– animals (n =
25), not significantly different from a randomly selected group of wild-type littermate controls
(225 ± 16 MΩ, n = 33, p > 0.05). Moreover, cells showed unaltered amplitudes of resting
membrane potential (for Sk1–/–: –79.8 ± 1.5 mV, n = 40; for Sk2–/–: –77.6 ± 1.6 mV, n = 27; in
control: –77.9 ± 1.6 mV, n = 39; p > 0.05), similar amplitudes of T–type Ca2+ currents at a
test potential of –62 to 67 mV (for Sk1–/–: –481 ± 48 pA, n = 33; for Sk2–/–: –429 ± 34 pA, n
=20; for wild–type: –478 ± 53 pA, n = 15, p > 0.05), overlapping steady–state inactivation
curves, and similar time courses of recovery from inactivation (data not shown).

6
Cueni et al. Supplementary Information 7

3) Supplementary Figures and Table

Supplementary Fig. 1
1–EBIO enhances the apparent Ca2+ affinity of native SK channels13, and is expected to
potentiate channel activation. To test the effects of this compound on nRt oscillations,
extracellular recordings were obtained in interface–style recording chambers with low–
resistance (< 1 MΩ) tungsten electrodes (Frederick Haer) and band–pass filtered between 0.3
kHz and 10 kHz using an extracellular amplifier (Warner Instruments) from slices perfused
with (in mM): 131 NaCl, 2.5 KCl, 1.25 NaH2PO4, 26 NaHCO3, 2 CaCl2, 1.2 MgCl2, 18
dextrose, 1.7 L(+)–ascorbic acid. A single–unit was identified by series of spontaneous tonic
or bursts of action potentials. Bursts showed an accelerando–decelerando pattern in the action
potential discharge frequency14. Single electric shocks (50 – 200 μA, 1 ms) were applied at
0.05 – 0.1 Hz via bipolar stimulation electrodes (Frederick Haer) placed in the internal
capsule adjacent to the nRt. These stimuli typically silenced active units and then elicited
repetitive burst discharges. The number of bursts was determined by the average of the bursts
after 5 – 10 successive stimuli. Tonic action potential frequency was determined by counting
action potentials in the first second after cessation of burst discharge.
Single units were identified in extracellular recordings by a stable action potential
amplitude (20 – 100 μV) above baseline (~2 – 5 μV) and repetitive, high–frequency (220 –
500 Hz), burst discharges. An electric shock to the internal capsule transformed a tonically
discharging into a repetitively bursting unit for 2 – 8 s, before tonic action potential discharge
was resumed (Supplementary Fig. 1a, b). Local application of 1–EBIO (~ 0.1 mM)
provoked an almost three–fold increase in the number of low-threshold bursts
(Supplementary Fig. 1c) and a prolongation of the interburst intervals in the initial portion of
the oscillation (Supplementary Fig. 1a), while the frequency of the tonic discharge remained
unaffected (Supplementary Fig. 1d). Interestingly, the prolongation of the interburst
intervals was significant for all but the first of 6 intervals measured (first 6 intervals in
control: 402 ± 40 ms, 338 ± 34 ms, 325 ± 25 ms, 296 ± 26 ms, 295 ± 17 ms, 277 ± 18 ms; in
1–EBIO: 482 ± 53 ms, 398 ± 36 ms, 344 ± 33 ms, 316 ± 33 ms, 320 ± 24 ms, 297 ± 21 ms; p
< 0.03 for 2nd to 6th interval, p = 0.067 for 1st interval), suggesting that, except for the first
interburst interval, the SK channel gating by bursts is normally submaximal.. Bursting was
abolished by local application of apamin (100 nM) (data not shown). In whole–cell recordings
(see main manuscript), 1–EBIO markedly enhanced SK currents in wild–type animals, while

7
Cueni et al. Supplementary Information 8

not having any effect on T–type Ca2+ current–dependent outward currents in Sk2–/– cells
(Supplementary Fig. 1e, f).

Suppl. Fig. 1.
Effects of 1–EBIO, an SK channel enhancer, on nRt cells derived from wild-type and
Sk2–/– mice. (a) Extracellular recordings of a tonically discharging nRt cell, which is
transformed into bursting by electrically stimulating synaptic inputs (200 μA, 0.1 ms, arrow),
before (Control) and after local application of 1–EBIO (~0.1 mM). Inset shows a single burst
at an expanded time scale. The cell generates a serious of burst discharges, before resuming
tonic firing. In 1–EBIO, the number of burst discharges was increased. (b) Expanded portions
of recordings numbered 1 and 2 in A. Double–headed arrows denote the interburst interval.
(c) Average number of bursts discharged per stimulation, before and after 1–EBIO
application. Data are presented as means ± SEM of 10 units. * denotes p < 0.05. (d) Average
number of action potentials discharged in the first second after resuming tonic firing. Data are
presented as means ± SEM of 7 units. (e) Bath–application of 1–EBIO (0.1 mM) onto
voltage–clamped nRt cells from wild–type (WT) and Sk2–/– mice. 1–EBIO promotes the
generation of an outward afterhyperpolarization current (AHP–current) that follows the T–
current in wild-type, but not in Sk2–/– cells. Cells were hyperpolarized to –120 mV for 125 ms,
before being repolarized to –62 mV. Dotted lines denote steady–state holding current at –62
mV. (f) Time course of 1–EBIO effects on AHP–current in recordings from wild-type (black
circles) and Sk2–/– cells (white circles). Dotted lines represent the average of the last 5
responses before the start of 1–EBIO application. Note that the current values in Sk2–/–
animals are negative at the time point at which an AHP–current is generated in wild-type
cells, due to the slower decay of T–currents in Sk2–/– cells. 1–EBIO increased the current in
wild-type (p < 0.005), but not in Sk2–/– cells (p > 0.05). Data are presented as means ± SEM of
5 cells per genotype.

8
Cueni et al. Supplementary Information 9

Supplementary Fig. 2

Suppl. Fig. 2.
Absence of SK2 immunolabeling in Sk2–/– tissue. At both the light microscopic (left) and
electronmicroscopic (right) level, no immunostaining could be detected. The single
immunoparticle present in the figure on the right is located on the myelin of a myelinated
axon. Ctx, Cortex, Th, Thalamus, IC, internal capsule, B, bouton, Den, Dendrite. Scale bars:
Left, 1 mm; Right, 0.5 μm.

9
Cueni et al. Supplementary Information 10

Supplementary Fig. 3

To assess for use–dependent decrease in T–current activation, a voltage–protocol that


involves repeated depolarization–hyperpolarization cycles was used. This protocol mimics the
rapid membrane potential changes occurring during dampened oscillations (Supplementary
Fig. 3a) and allows to measure full T–current amplitudes at each cycle. Rapid inward currents
were generated after each depolarizing upstroke (asterisks), which markedly diminished in
amplitude in subsequent cycles, and reached steady–state levels after about three cycles
(Supplementary Fig. 3b). These inward currents were T–currents since they were largely
blocked by mibefradil (50 μM, Supplementary Fig. 3c). Averaged values of four
experiments involving five depolarization–hyperpolarization cycles yielded a significant
decrease of current response after 3 cycles (to 45.6 ± 10 % of control, p < 0.05) and after 5
cycles (to 30 ± 8% after 5 cycles, p < 0.05 compared to third cycle). This was accompanied by
a decrease in the decay slope of the T–current (Supplementary Fig. 3d). Taken together,
whole–cell T–current in nRt cells showed use–dependent cumulative inactivation,
accompanied by a decrement in SK channel recruitment. These gating properties are
consistent with the observed accruing decrement of Ca2+ signals during dampened
oscillations.

Suppl. Fig. 3

10
Cueni et al. Supplementary Information 11

T–current in nRt cells shows use–dependent cumulative inactivation. (a) Voltage–clamp


protocol involving five hyperpolarizing–depolarizing cycles, approaching the membrane
potential events during a dampened oscillation. Depolarized phases were maintained for 80
ms to allow for full decay of T–current amplitudes. In this way, current amplitudes could be
measured directly. (b) Representative current responses to the protocol illustrated in a. Note
the rapid generation of an inward current at the onset of each depolarizing cycle (*), reflecting
activation of T–current, and its decrement in current amplitude from one cycle to the next.
The T–current activated in the first cycle is followed by an outward current, which strongly
weakens, reflecting weakened SK2 channel recruitment. (c) Overlay of current responses
before (Ctrl, thin lines) and after Mibefradil (Mibe, thick lines) (50 μM) application. (d)
Normalized T–current responses and decay slope of T-current plotted against cycle number
elicited by the protocol shown in a. Values are significantly smaller starting with the third
cycle (p < 0.05).

11
Cueni et al. Supplementary Information 12

Supplementary Fig. 4
Previous computational models of nRt cells generated dampened oscillations within a
single–compartment, containing Hodgkin–Huxley models of voltage–gated channels and two
Ca2+–activated conductances, a K+ and a cation conductance15. Following this model, a
single–compartment model incorporating T–channels, Ca2+–dependent K+ channels, and
sequestration mechanisms was used. Passive properties were implemented as described15,
with

CmdV / dt = –gL*(V – EL) – IT – I SK

where V is the membrane potential, Cm = 1 μFcm-2, EL = – 78 mV, and gL = 0.05


mScm–2, IT is the T–current, and ISK the Ca2+–dependent K+ current.

T–channels were modelled according to

IT = gca * m2 * h * (V – Eca)

with the activation parameter m(V,t) described by

dm / dt = – [1 / τm (V)] * [m – m∞ (V)]

and the inactivation parameter h(V,t)

dh / dt = – [1 / τh (V)] * [h – h∞ (V)]

The voltage dependence of m, h, and the time constants followed

m∞(V) = 1 / [1 + exp [(– V + 52) / 7.4]]

h∞(V) = 1 / [1 + exp [( V + 80) / 5]]

τm (V) = 0.44 + 0.15 / [exp ( (V + 27) / 10) + exp ( –(V + 102) / 15)]

τh (V) = 22.7 + 0.27 / [exp ( (V + 48) / 4) + exp ( –(V + 407) / 50)]

The maximum conductance gca was 1.75 mScm-2, and ECa = 100 mV the reversal
potential.

Similarly, ISK was described by

ISK = gSK * mSK * hSK * (V – Eca)

with the gating parameter mSK described by

12
Cueni et al. Supplementary Information 13

dmSK / dt = – [1/τmSK([Ca2+]i)] * [mSK – mSK,∞([Ca2+]i)]

The steady–state activation parameter mSK,∞([Ca2+]i) was described by a dependence on


the fourth power (n = 4) of [Ca2+]i and α was set to 0.4 * 10^12 ms–1mM–4 and β to 0.025 ms–
1
, according to

mSK,∞ ([Ca2+]i)] = α[Ca2+]in / (α[Ca2+]in + β)

τmSK ([Ca2+]i)] = 1 / (α[Ca2+]in + β)

In this manner, half–activation of the K+ channels occurred at ~0.5 µM [Ca2+]i. The


maximal K+ conductance was set to 3 mScm–2.

Ca2+ sequestration mechanisms were modelled as described15.

d [Ca2+]i / dt = – KT * [Ca2+]i / [[Ca2+]i + Kd]

Two such mechanisms were implemented, the first with Michaelis–Menten constants of
KT = 10–4 mMms–1 and Kd = 10–4 mM and the second with 5–fold smaller KT and Kd.
[Ca2+]i(t) was then modelled according to the sum sequestration and influx, with the latter one
defined as

d [Ca2+]i / dt = – [k /2Fd] * IT

where k = 0.1, F is the Faraday constant and d = 1 μm.


Initial conditions were a resting membrane potential of –90 mV, a [Ca2+]i of 100 nM,
and the gating parameters for the T–current were set to m (t = 0) = 0.0 and h (t = 0) = 1.0.
Slow recovery from inactivation was introduced by requiring that the time evolution of
the inactivation parameter h(t) is governed by τh (V) when h(t) > h∞(V) but by τslow = 2 s when
h(t) < h∞(V). In this manner, the voltage–dependence of inactivation described previously16
was approximated. Blocking SERCA was modelled by removing the second Ca2+
sequestration mechanism.
In this model, we reproduced on–going oscillatory bursting in the absence of the
cationic conductance. Removing Ca2+ sequestration from this model fully abolished the
oscillations, because Ca2+ was not cleared and the cells became tonically hyperpolarized (data
not shown). This is not observed experimentally when blocking SERCA selectively, and
indicates that more than one sequestration mechanism controls Ca2+ removal after a low-
threshold burst in a real cell. To take this into account, we implemented two sequestration
mechanisms with a 5–fold difference in affinity and kinetics. This, by itself, did not affect the

13
Cueni et al. Supplementary Information 14

on–going oscillations (left traces). We then implemented a formalism following values


reported16 to phenomenologically take slow recovery from inactivation into account. This led
to a marked dampening that involved the generation of submaximal low-threshold bursts and
Ca2+ signals (middle traces). Within this extended model, we studied the role of [Ca2+]i
handling. Removal of the low–affinity sequestration only generated a pattern in which
oscillations occurred, albeit with a weakened dampening (right traces). This reduction was
accompanied by an enhanced number of [Ca2+]i transients, and a small decrease in the
amplitude of the first two Ca2+ signals. This illustrates that the recovery from inactivation
strongly shapes the temporal evolution of oscillatory dampening, while the sequestration of
Ca2+ finely modulates the interaction between T–channels and rapidly activated K+ channels.

Suppl. Fig. 4.
A computational model of nRt oscillations. This model reproduces on–going oscillations
when T–channels were coupled to a Ca2+–activated K+ conductance (left). The cell contained
two Ca2+ sequestration mechanisms. Time course of [Ca2+]i is shown below. When slow
recovery was introduced (middle), oscillations were dampened. Removal of a slow Ca2+
sequestration, mimicking SERCA blockade, attenuated the dampening (right).

14
Cueni et al. Supplementary Information 15

Supplementary Figure 5a, 5b and Supplementary Table 1


Large differences in EEG power density were observed between genotypes (Fig. 8).
To verify whether, besides difference in absolute values, the relative contribution of the
various frequencies to the EEG was affected by genotype, relative spectra were calculated by
expressing the power density in each frequency bin as a percentage of total EEG power over
the entire frequency range (excluding the 45 – 55 Hz range) for that state (Supplementary
Fig. 5). During NREMS, the relative contribution of delta activity to the EEG was decreased,
and of a wide range of fast frequencies, including beta (18 – 25 Hz) and gamma (35 – 60 Hz),
increased. Differences in the relative REMS spectra were most prominent in theta (5 – 10 Hz)
and beta frequency ranges. Gamma activity was not affected (not shown). Differences in the
theta range were due to a significant (p < 0.05, t–test, asterisk) 0.3 Hz slowing of theta peak
frequency (see inset) during this state in Sk2–/– mice. Changes in the relative waking EEG
were limited to a pronounced decrease at 10 Hz that resulted from a (non–significant) slowing
in theta oscillations also in this state (see inset of Supplementary Fig. 5a). Theta peak
frequency was determined by selecting the frequency bin with the highest power density
within the theta range (5 – 10Hz) within individual mice (mean ± SEM; bi–directional). A
contour plot shows the spectral composition of the EEG at frequencies up to 90 Hz at
NREMS to REMS transitions (Supplementary Fig. 5b).
Despite pronounced differences in NREMS fragmentation (Fig. 8), overall time–
spent–asleep was not significantly affected by genotype (Supplementary Table 1).
Calculated over 24 h, Sk2–/– mice (n = 4) slept somewhat less (–30min) compared to Sk2+/+ (n
= 5), due largely to a 43 min deficit in sleep time during the 12 h dark (D) or active period.

15
Cueni et al. Supplementary Information 16

Suppl. Fig. 5a.


Mean relative EEG spectra for NREMS, REMS, and waking during baseline. Relative
EEG spectra (upper panels) were calculated for each state by expressing power density in
each frequency bin as a percentage of total power over all frequency bins within each state.
Genotype differences (lower panels) were calculated as log2–transformed Sk2–/– / Sk2+/+
ratios. Red bars mark frequency bins that significantly differed between genotype (post–hoc,
t–tests, p < 0.05).

16
Cueni et al. Supplementary Information 17

Suppl. Fig. 5b.


Contour plot of the changes in the EEG spectral composition at the transition from
NREMS (–3 to 0 min) to REMS (0 to + 2min) for Sk2+/+ and Sk2–/– mice. ‘Heat map’ was
constructed by aligning and averaging spectra from all transitions selected during the 24– or
48 h baseline recordings, first within and then among mice. Power density within 0.25 Hz
bins was expressed as a percentage of the mean power density for that bin over 4 – s epochs
scored as NREMS in the 1st 2 min (i.e., –3 to –1 min) of the transition to visualize relative
spectral changes. Contour lines connect levels of similar relative power in 8 color – coded 20
% increments. White dashed lines at time = 0 indicate the time between the last 4 – s epoch
scored as NREMS and the 1st 4 – s epoch scored as REMS. EEG changes start during the 1
min prior to REMS onset and entail a marked increase in EEG activity in the spindle
frequency range (10 – 15 Hz; see also Fig. 8d) shortly followed by an increase in theta
activity (5 – 10 Hz). Maximum spindle activity is reached around –25 s, while maximum theta
activity is reached at the transition. After the transition, spectral values reach their typical
REMS levels with, below 35 Hz, no other activity than theta and above 35 Hz, including the
gamma band (35 – 60 Hz) activity that exceeds high frequency EEG activity in NREMS up to
3 – fold. Genotype differences (Sk2+/+, n = 7; Sk2–/– , n = 6) concern the less prominent surge
in spindle activity prior to REMS onset (see Fig. 8d) and a smaller relative increase in
gamma/high–frequency activity and smaller relative decreases in delta and spindle activity
(dark blue areas) during REMS in Sk2–/– mice.
17
Cueni et al. Supplementary Information 18

Genotype Waking [h] NREMS [h] REMS [min]


12h L–period Sk2–/– 4.97 ± 0.15 6.16 ± 0.20 52.2 ± 6.3
Sk2+/+ 5.16 ± 0.13 5.85 ± 0.11 59.3 ± 3.1
12h D–period Sk2–/– 10.22 ± 0.17 1.60 ± 0.15 10.3 ± 1.6
Sk2+/+ 9.50 ± 0.26 2.23 ± 0.23 16.0 ± 2.1
L–D difference Sk2–/– 5.20 ± 0.13* 4.47 ± 0.15* 43.8 ± 2.9
Sk2+/+ 4.34 ± 0.21 3.61 ± 0.17 43.4 ± 3.0
24 h Sk2–/– 15.19 ± 0.28 7.78 ± 0.28 62.5 ± 7.9
Sk2+/+ 14.67 ± 0.31 8.08 ± 0.29 75.3 ± 3.3

Suppl. Table 1
Sleep and wake times of Sk2+/+ and Sk2–/– mice. Despite pronounced differences in NREMS
fragmentation (Fig. 8) overall time-spent-asleep was not significantly affected by genotype.
Calculated over 24 h, Sk2–/– mice (n = 4) slept somewhat less (–30 min) compared to Sk2+/+ (n
= 5), due largely to a 43 min deficit in sleep time during the 12 h dark (D) or active period.
The (non–significant) decrease in NREMS time in the D-period combined with the (non–
significant) increase during the light period, let to a significantly increased L–D difference in
NREMS time in Sk2–/– mice (2-way ANOVA interaction between factors ‘Genotype’ and
‘LD-period’, with repeated measures for ‘LD-period’: Waking p = 0.0064; NREMS p =
0.0032). *Mark significant genotype differences (p < 0.01; post-hoc t-test). Analyses were
based on 2 baseline recordings and represent means ± SEM.

References

1. Huguenard, J.R. & Prince, D.A. A novel T-type current underlies prolonged Ca2+-
dependent burst firing in GABAergic neurons of rat thalamic reticular nucleus. J Neurosci 12,
3804-3817 (1992).
2. Avanzini, G., de Curtis, M., Panzica, F. & Spreafico, R. Intrinsic properties of nucleus
reticularis thalami neurones of the rat studied in vitro. J Physiol 416, 111-122 (1989).
3. Steriade, M., Contreras, D., Curro Dossi, R. & Nunez, A. The slow (< 1 Hz)
oscillation in reticular thalamic and thalamocortical neurons: scenario of sleep rhythm
generation in interacting thalamic and neocortical networks. J Neurosci 13, 3284-3299
(1993).
4. Joksovic, P.M., Bayliss, D.A. & Todorovic, S.M. Different kinetic properties of two
T-type Ca2+ currents of rat reticular thalamic neurones and their modulation by enflurane. J
Physiol 566, 125-142 (2005).
5. Sun, Q.Q., Huguenard, J.R. & Prince, D.A. Neuropeptide Y receptors differentially
modulate G-protein-activated inwardly rectifying K+ channels and high-voltage-activated
Ca2+ channels in rat thalamic neurons. J Physiol 531, 67-79 (2001).
6. Canepari, M., Auger, C. & Ogden, D. Ca2+ ion permeability and single-channel
properties of the metabotropic slow EPSC of rat Purkinje neurons. J Neurosci 24, 3563-3573
(2004).
7. Hyrc, K.L., Bownik, J.M. & Goldberg, M.P. Ionic selectivity of low-affinity
ratiometric calcium indicators: mag-Fura-2, Fura-2FF and BTC. Cell Calcium 27, 75-86
(2000).
8. Ogden, D., Khodakhah, K., Carter, T., Thomas, M. & Capiod, T. Analogue
computation of transient changes of intracellular free Ca2+ concentration with the low affinity
Ca2+ indicator furaptra during whole-cell patch-clamp recording. Pflügers Arch 429, 587-591
(1995).

18
Cueni et al. Supplementary Information 19

9. Naraghi, M. T-jump study of calcium binding kinetics of calcium chelators. Cell


Calcium 22, 255-268 (1997).
10. Franken, P., Malafosse, A. & Tafti, M. Genetic variation in EEG activity during sleep
in inbred mice. Am J Physiol 275, R1127-R1137 (1998).
11. Franken, P., Malafosse, A. & Tafti, M. Genetic determinants of sleep regulation in
inbred mice. Sleep 22, 155-169 (1999).
12. Franken, P., et al. NPAS2 as a transcriptional regulator of non-rapid eye movement
sleep: genotype and sex interactions. Proc Natl Acad Sci U S A 103, 7118-7123 (2006).
13. Pedarzani, P., et al. Control of electrical activity in central neurons by modulating the
gating of small conductance Ca2+-activated K+ channels. J Biol Chem 276, 9762-9769 (2001).
14. Domich, L., Oakson, G. & Steriade, M. Thalamic burst patterns in the naturally
sleeping cat: a comparison between cortically projecting and reticularis neurones. J Physiol
379, 429-449 (1986).
15. Destexhe, A., Contreras, D., Sejnowski, T.J. & Steriade, M. A model of spindle
rhythmicity in the isolated thalamic reticular nucleus. J Neurophysiol 72, 803-818 (1994).
16. Frazier, C.J., et al. Gating kinetics of the α1I T-type calcium channel. J Gen Physiol
118, 457-470 (2001).

19

You might also like