, 2010a) A more recent study provides evidence that disruptions

, 2010a). A more recent study provides evidence that disruptions in microglia function result in delayed maturation of hippocampal synaptic circuits (Paolicelli et al.,

2011). Moreover, data from these studies suggest that microglia may be phagocytosing dendritic spines. These intriguing studies raise several interesting and important questions. The precise function of microglia at synaptic sites, the molecular mechanism(s) underlying microglia-mediated synaptic engulfment, and the long term consequence(s) of disrupting microglia function on synaptic circuits remain a mystery. A candidate mechanism by which microglia could be Dabrafenib ic50 interacting with developing synapses is the classical complement cascade. Complement cascade components C1q and C3 localize to immature synapses and are necessary for the developmental pruning of retinogeniculate synapses (Stevens et al., 2007 and Stephan et al.,

2012). While provocative, the mechanism by which complement mediates synaptic pruning has remained completely unknown. Complement components function in the immune system by binding and targeting unwanted cells and cellular debris for rapid elimination through several different pathways. Among the many mechanisms by which complement may mediate synaptic pruning is phagocytosis, which makes microglia, the resident CNS phagocyte, a candidate. Veliparib Given the questions that have now emerged regarding the role of microglia

at CNS synapses, we sought to address precisely how microglia are interacting with developing synaptic circuits and determine the long-term consequences of disrupting microglia function on neural circuit development. In the current study, we demonstrate that microglia engulf presynaptic retinal inputs undergoing synaptic pruning in the postnatal brain and determine that this process is regulated by neuronal activity. Furthermore, we identify signaling through a phagocytic receptor, complement receptor 3 (CR3/CD11b-CD18/Mac-1), expressed on the surface of microglia and its ligand, complement component C3, localized to synaptically enriched regions, below as a key molecular mechanism underlying engulfment of developing synapses. Importantly, disruption of CR3/C3 signaling was specific to microglia in the CNS and resulted in sustained deficits in brain wiring. Taken together, these observations provide a role for microglia in the healthy, developing brain and provide a cellular and molecular mechanism by which microglia are physically interacting with synaptic elements. To investigate the functional role of microglia in developmental synaptic remodeling, we used the mouse retinogeniculate system, a classic model for studying activity-dependent developmental synaptic pruning (Feller, 1999, Huberman et al., 2008 and Shatz and Kirkwood, 1984).

Time courses of activity were analyzed by two-way repeated-measur

Time courses of activity were analyzed by two-way repeated-measures ANOVA; sleep quantities Selleck NLG919 were compared by one-way ANOVA. Arousal thresholds were tested by subjecting flies to mechanical stimuli generated by vibration motors (Precision Microdrives, model 310-113) ( van Alphen et al., 2013). Stimuli were delivered for 5 s and separated by 1 hr. Female flies were sleep deprived by the mechanical SNAP method while housed in TriKinetics DAM systems (Shaw et al., 2002). A cumulative sleep loss plot was calculated

for each individual by comparing the percentage of sleep lost during overnight sleep deprivation to the immediately preceding unperturbed night. Individual sleep rebound was quantified hourly for 24 hr by dividing the cumulative amount of sleep regained by the total amount of sleep lost

during deprivation. Individual flies were excluded from rebound analysis if sleep deprivation was less than 70% effective or if flies lost less than 60 min of sleep. Statistical significance was assessed buy Y-27632 by two-way repeated-measures ANOVA. Males were used for circadian analyses in order to prevent interference from developing embryos and larvae over a weeklong experiment. Flies were housed individually in 65 mm glass tubes containing 4% sucrose and 2% agar medium. Locomotor activity was measured in TriKinetics DAM systems for 7 days in constant darkness. χ2 periodogram analysis was completed for each individual fly with the ActogramJ plugin (Benjamin Schmid and Taishi Yoshii, University of Würzburg) for ImageJ (National Institutes of Health). Individual female flies were trained and analyzed in 50 mm chambers perfused with air-odor mixtures as previously described (Claridge-Chang et al., 2009). Baseline preference for 3-octanol (OCT) versus

4-methylcyclohexanol Megestrol Acetate (MCH) was measured by tracking a fly’s movements for 2 min. Flies then received two training cycles, which each consisted of two epochs in random order: a 1 min presentation of OCT without shock and a 1 min presentation of MCH with twelve 60-VDC electric shocks. Flies were allowed to recover for 5 min and then reanalyzed for 2 min. Learning is reported as the percentage change in time spent in MCH before and after training (Figure S2). To gain access for whole-cell patch-clamp recordings in vivo, we removed a small piece of cuticle from the head of female 104y-GAL4;UAS-CD8-GFP flies and targeted the recording electrodes visually to the fluorescent somata of dorsal FB neurons. Control recordings from olfactory PNs were obtained by sampling unlabeled antennal lobe neurons. PNs were distinguished from local neurons by their characteristic electrophysiological properties ( Wilson and Laurent, 2005). Borosilicate glass electrodes (7–13 MΩ) were filled with internal solution containing 140 mM potassium aspartate, 10 mM HEPES, 1 mM KCl, 4 mM Mg-ATP, 0.5 mM Na3GTP, 1 mM EGTA (pH 7.

This finding is consistent with observations that the intact coch

This finding is consistent with observations that the intact cochlear partition is roughly three orders of magnitude stiffer than an individual outer hair cell (He and Dallos, 1999; Olson and Mountain, 1991). A portion of the active process remained after photoinactivation of regions in which gain occurred, however, for damping increased even further after anoxia (Figures 4A and 4B). Although

the residual active process might reflect incomplete blockage of prestin, we ascertained that repeated exposure to UV light in the presence of 4-azidosalicylate did not further diminish the response (Figure 3D). Understanding the degree to which photoinactivation eliminates electromotility selleck kinase inhibitor in vivo could yield a more quantitative assessment of prestin’s contribution to amplification. Obtaining a clearer picture of the intact

active process would also necessitate an appreciation of the specificity with which photoinactivation affects electromotility; are other cellular processes affected? If, for example, active hair-bundle motility were entirely spared the effects of photoinactivation, then bundle forces might account for the balance of amplification. It is possible, however, that photoinactivating prestin affected hair-bundle forces as well. Changes in an outer hair cell’s membrane potential can elicit hair-bundle deflections (Jia and He, 2005). Moreover, hair bundles can Selleck Target Selective Inhibitor Library be displaced by somatic length changes of outer hair cells through the mechanical coupling in an intact organ of Corti. Although adaptation in hair bundles restores the set point of nonlinear amplification even for large static deflections (Martin et al., 2003), photoinactivation might force all prestin molecules into a conformation normally elicited only by extreme depolarization. In this circumstance, active hair-bundle motility could be compromised. Finally, salicylate might affect other aspects of hair-cell physiology. Although we failed to detect photolabeling of other proteins during our biochemical investigation, it remains possible that photoinactivation

modifies proteins in addition to prestin. Cell press These results demonstrate that an active process overcomes viscous damping to locally amplify the cochlear traveling wave and that this locally accrued gain accumulates spatially up to the wave’s peak. The results further indicate that prestin plays a crucial role in establishing this gain. It remains an open question, however, how this active process is locally tuned to yield a tonotopic map of amplification. Despite its critical contribution, there is no evidence yet that somatic motility exhibits resonance. Under physiological conditions, sound-evoked receptor potentials in mammalian outer hair cells modulate a resting potential of −40 mV by less than 10 mV at moderate stimulus levels (Johnson et al., 2011; Kössl and Russell, 1992).

If tuft excitatory synaptic input is isolated from the axon, and

If tuft excitatory synaptic input is isolated from the axon, and only weakly directly drives apical dendritic trunk spike initiation, how can it influence neuronal output? We demonstrate that the apical dendritic tuft can interact in

a nonlinear manner with more proximal integration zones (i.e., the axon and nexus), and that KV channels control this process. Previous findings have revealed that axosomatic and dendritic trunk integration compartments can synergistically interact in L5B pyramidal neurons (Larkum et al., 1999, Larkum et al., 2004 and Williams, 2005). Here, we show that apical dendritic KV channels tightly control this selleck inhibitor interaction. Indeed, KV channels also governed the interaction between subthreshold tuft excitatory input and trunk spike generation, with the pharmacological blockade of KV channels leading to the generation

of large amplitude long-duration plateau potentials in the apical dendritic arbor. As apical dendritic KV channels exhibit highly sensitive voltage- and time-dependent inactivation properties, widespread depolarization of the apical dendritic MAPK inhibitor tuft, as may occur during behavior (Xu et al., 2012), will lead to voltage-dependent KV channel inactivation, that is augmented by the local recruitment of NMDA-receptor- and Na+-channel-mediated supralinearities. These effects will powerfully influence the interaction between axosomatic and apical dendritic of trunk integration compartments. Consistent with this, we find that tuft Ca2+ signals during behavior are enhanced by pharmacological blockade of KV channels. These observations support a multilayer integration scheme for L5B pyramidal neurons, which we tested experimentally using triple recordings. Top-down, apical tuft excitatory input was found to decisively

influence the rate and pattern of AP firing evoked via coincident activation of axosomatic and apical dendritic trunk integration compartments by controlling the duration of apical dendritic plateau potentials. Therefore, trunk spikes are not a binary output mode of apical dendritic integration, but rather are analog signals, the duration of which is tuned by apical dendritic tuft excitatory input. Our direct demonstration that apical dendritic tuft input drives high-frequency AP burst firing suggests that top-down influences such as attention, expectation, and action command (Gilbert and Sigman, 2007, Gregoriou et al., 2009, Hupe et al., 1998 and Xu et al.

Extrapolating from ∼10% connectivity to ∼2000 total glomeruli in

Extrapolating from ∼10% connectivity to ∼2000 total glomeruli in the mouse (Soucy et al., 2009), one can estimate that each PCx neuron connects with ∼200 glomeruli. The number of possible 200-glomerulus combinations CT99021 mouse is >10500, which will be massively undersampled by the PCx population. Correspondingly, PCx firing was reliably triggered by MOB patterns with only 3–4 sites, suggesting cortical cells are not “grandmother” neurons with highly

specific input requirements. Instead, undersampling appears to be balanced by low-stringency coincidence detection requiring activity in a relatively small fraction of connected MOB glomeruli (Apicella et al., 2010 and Franks and Isaacson, 2006). Our results are qualitatively consistent with recent monosynaptic tracing of PCx input (Miyamichi et al., 2011), although we find substantially greater convergence of M/T input. Electrophysiological circuit mapping, besides revealing the functional strength of synaptic contacts, may allow detection of a greater proportion of MOB inputs. Our experiments treat glomeruli as elementary processing units. In vivo imaging supports this assumption for presynaptic OR input (Wachowiak et al., 2004). Postsynaptically, each glomerulus contacts ∼50–75 GABA receptor inhibition M/T neurons (Haberly, 1991), whose activity depends strongly on presynaptic OR input (Tan et al., 2010). However,

all such “sister” M/Ts do not necessarily respond identically (Dhawale et al., 2010, Egaña et al., 2005, Fantana et al., 2008 and Tan et al., 2010). Our data do not address whether sister M/Ts converge onto like cortical targets, although the small size of synaptic inputs suggested this was unlikely. Odor responses of second-order neurons are influenced by lateral interactions between glomeruli in both rat MOB (Fantana et al., 2008) and Drosophila antennal lobe ( Olsen et al., 2007, Olsen and Wilson, 2008 and Shang et al., 2007). While M/T responses were similar Bay 11-7085 for single- and multisite uncaging ( Figure S3), any further decorrelation of odor-evoked firing by local MOB circuits may

facilitate pattern separation by PCx. It also remains to be seen how PCx responses depend on temporal patterning of MOB output ( Bathellier et al., 2008, Cury and Uchida, 2010, Dhawale et al., 2010, Kashiwadani et al., 1999, Schaefer and Margrie, 2007, Spors et al., 2006 and Wesson et al., 2008; see Friedrich et al. [2004] and Perez-Orive et al. [2002] for work in other species). Temporal decoding mechanisms have been described for both individual pyramidal neurons ( Branco et al., 2010) and the PCx network ( Stokes and Isaacson, 2010). While our experiments focused on circuit connectivity, in the future photostimulation may also help evaluate the role of timing in cortical processing. Although glomerular pattern detection in PCx could potentially be explained by a simple linear feedforward mechanism, responses to coactive glomeruli were strongly supralinear.

By contrast, in the 6 hr and 1 day retention sessions, kif17+/+ m

By contrast, in the 6 hr and 1 day retention sessions, kif17+/+ mice showed a significant preference for the novel object, whereas kif17−/− mice exhibited decreased preference for the novel object ( Figure 6B; Movie S3). Next, we subjected kif17−/− mice to the Morris water maze ( Sakimura et al., 1995 and Silva

et al., 1992) to test their hippocampus-dependent spatial learning abilities. Both groups of mice swam LDN-193189 ic50 at a normal velocity ( Figure 6C). In the visible-platform test, kif17−/− mice performed as efficiently as kif17+/+ mice. However, in the hidden-platform test, kif17−/− mice displayed a longer latency to locate the platform than kif17+/+ mice ( Figure 6D; Movie S4). In the subsequent probe test, the navigation of kif17−/− mice was random, and their searching in the target quadrant was not as selective as that of the kif17+/+ mice ( Figures 6E and 6G). Furthermore, the kif17−/− mice crossed the platform less often than the kif17+/+ mice ( Figures 6F and 6G). We next tested contextual fear memory by assessing the freezing behavior of mice in the same environmental context (Bourtchuladze et al., 1994). Contextual fear memory is dependent on the hippocampus (Kim and Fanselow, 1992 and Phillips and LeDoux, 1992). No difference http://www.selleckchem.com/products/VX-770.html in freezing

between kif17+/+ and kif17−/− mice was indicated immediately after the foot shock. However, Urease kif17−/− mice exhibited far fewer freezing responses than kif17+/+ mice when they were tested at 1 hr, 24 hr, and 7 days after training ( Figures 6H–6M; Movie S5). These findings indicate that kif17−/− mice have a deficit in their context-dependent

fear memory. We also compared olfactory learning abilities between kif17+/+ and kif17−/− mice, because it is reported that OSM-3, a C. elegans homolog of KIF17, is an “accessory” intraflagellar transport (IFT) motor that is required for olfactory cyclic nucleotide-gated channel targeting ( Evans et al., 2006 and Jenkins et al., 2006). We did not find any abnormality in the olfactory learning of kif17−/− mice compared with kif17+/+ mice (data not shown). Together, our behavioral observations demonstrate a hippocampus-dependent memory disturbance in kif17−/− mice. To investigate how a downstream event induced by NMDA receptor activation is altered in kif17−/− neurons, we studied phosphorylation of cAMP-response element binding protein (CREB). Our previous results suggest a functional interaction between KIF17 and CREB ( Wong et al., 2002). We assessed the phosphorylation of CREB at S133 (pCREB), which can be triggered through NMDA receptor-mediated calcium influx ( Lonze and Ginty, 2002, West et al., 2002 and Zhu et al., 2002), in hippocampal cultures after glutamate-induced stimulation. Immunocytochemical analysis revealed similarly low basal pCREB levels in untreated kif17+/+ and kif17−/− neurons.

, 1982, McDonald, 1991 and Pan et al , 2010), but the specificity

, 1982, McDonald, 1991 and Pan et al., 2010), but the specificity and relevance of this input stream is not well understood. The amygdala provided considerably stronger synaptic input to the direct pathway (mean 4.3% ± 1.4% of total inputs

for D1R-Cre, versus 0.1% ± 0.1% for D2R-Cre mice, p = 0.02 by one-tailed Wilcoxon rank-sum test, z = 2.16, U = 18. Nonparametric statistical test used because D2R input is floored near zero). Amygdala inputs were manually registered via scaled rotation at 1/2 sampling density (Figure 6), selleck compound demonstrating biased input from both basolateral and central nuclei onto direct pathway MSNs. This observation mirrors the biased limbic cortical synaptic input to the direct pathway described in Figure 4. These results suggest that the limbic system, including both limbic cortex and amygdala, may convey affective value information to the striatum, biasing action selection

preferentially through the direct pathway, consistent with a role for the direct pathway in reinforcement (Kravitz et al., 2012 and Stuber et al., 2011). Although we targeted the direct versus indirect pathway independently of striosomal organization, our results regarding preferential innervation of the direct pathway from limbic structures parallels evidence in the striosomal literature; intriguingly, limbic cortices (Gerfen, 1984, Gerfen, 1989 and Jimenez-Castellanos and Graybiel, 1987), and the amygdala (Ragsdale and Graybiel, 1988) are thought to preferentially innervate striosomal compartments, which may themselves be preferentially populated with direct-pathway-like MSNs that project to the SNc, as well as Hormones antagonist to the SNr, GP, and EP (Fujiyama et al., 2011). These results, in conjunction with our own experiments, suggest that both target cell location within the striosome-matrix dichotomy

(Kincaid and Wilson, 1996) and neuronal cell type may interact to generate fine-scale Histamine H2 receptor organization within the dorsal striatum. We examined the strength of synaptic dopaminergic input from the substantia nigra onto striatal projection cells. Surprisingly, we observed that a relatively small proportion of total labeled inputs arose from the substantia nigra pars compacta (SNc), but that SNc similarly innervated both direct- and indirect-pathway MSNs. When using the monosynaptic rabies virus system, only 0.8% ± 0.3% of the brain-wide inputs arose from SNc onto either direct- or indirect-pathway MSNs (Figures 3 and 7C). Figure 7A shows a representative image of substantia nigra in a D1R-Cre mouse. As expected, dense striatonigral axon fibers from direct-pathway MSNs are detectable in substantia nigra pars reticulata (SNr), but relatively few retrogradely labeled neurons are visible in pars compacta. In D2R-Cre mice, few if any fibers were detected in SNr, as expected for targeting indirect-pathway MSNs ( Figure 7B). Again, relatively few retrogradely labeled neurons were detected in SNc.

The elemental analysis of the title compounds [9–12] is well comp

The elemental analysis of the title compounds [9–12] is well compatible with their proposed molecular formula ( Table 1). In compound 9, the two benzylic protons appeared as two singlets at 4.32 and 4.38 ppm. The two bridgehead protons are obtained as a singlet at 2.52 ppm. The multiplet centered at 2.80 ppm is due to H-7a proton and another multiplet centered at 1.25 ppm is assigned to H-7e proton. The multiplet centered at 1.60 ppm is attributed to H-6e and H-8e protons and the multiplet centered at 1.36 ppm

is due to H-6a and H-8a protons. Moreover, a broad singlet resonated at 3.57 ppm is unambiguously assigned to NH proton. The collection of signal observed in the range of 7.20 ppm–7.61 ppm are due to the protons of the two phenyl rings attached at C-2 and C-4 positions of the azabicyclo[3.3.1]nonane-9-one part of the compound. In the lower frequency region, two singlets are observed. Of the two singlets, the one at 1.45 ppm NLG919 molecular weight selleck compound is due to methyl protons attached at C-2 and C-6 positions of the tritertiarybutyl-cyclohexadienone part of the compound whereas the other singlet at 1.30 ppm is due to methyl protons attached at C-4 position of the tritertiarybutyl-cyclohexadienone part of the compound. A sharp singlet is observed at 6.70 ppm is due to the two methine protons at C-3 and C-5 of the cyclohexadienone part of the compound. In the 13C NMR spectrum

of compound 9, the signals of the benzylic carbons at C-2 & C-4 and the bridgehead carbons at C-1 & C-5 of the azabicyclo[3.3.1]nonane-9-one part of the compound appears at 65.6 ppm and 43.2 ppm respectively. Moreover, in the aliphatic region the signal appears at 26.6 ppm is assigned to carbons at C-6 and C-8 of the azabicyclo[3.3.1]nonane-9-one part of the compound Rolziracetam and the signal appears at 26.1 ppm is assigned to the carbon at C-7 of the azabicyclo[3.3.1]nonane-9-one part of the compound. 13C signals

resonated in the region from 126.8 ppm to 128.4 ppm are assigned to the carbons of the two phenyl rings attached at the C-2 and C-4 positions of the azabicyclo[3.3.1]nonane-9-one part of the compound. The signal at 141.4 ppm is assigned for the ipso carbons of the phenyl rings attached at C-2 and C-4 positions. In addition, the methyl and tertiary butyl carbon signals appear at 29.7 ppm & 21.6 ppm and 36.2 ppm & 34.5 ppm respectively are deputed for the tertiary butyl groups at C-2, C-6 and C-4 of the cyclohexadienone part of the compound. The C-2 and C-6 carbons of the cyclohexadienone part of the compound resonated at 151.3 ppm and the C-3 and C-5 methine carbons resonated at 142.5 ppm. Apart from the deputed signals, three un deputed signals which are resonated at 165.8, 181.1 and 84.0 ppm are due to the C N, C O and C–O carbons respectively. These assigned signals of the carbons proved the formation of the target compound.

1% for MST, 87 7% for V5/MT, 95 4% for V3A, 89 3% for V6, but onl

1% for MST, 87.7% for V5/MT, 95.4% for V3A, 89.3% for V6, but only 32.3% for V3B and 65.1% for VPS. The GLM’s beta estimates for “objective motion” and “retinal motion” (see Figure 7D) were near identical to those shown in Figure 3B, replicating the

results of experiment 2 also in conditions Inhibitor Library ic50 containing multiple velocities of objective motion and unmatched velocities between pursuit and objective motion. Overall, experiment 4 demonstrated that V5/MT and MST responded primarily to retinal motion during pursuit, whereas V3A and V6 were the only regions reporting velocity of objective planar motion also when pursuit velocities did not match those of objective planar motion. The ability to respond to objective (or head-centered) motion requires the multimodal integration of retinal visual motion signals with nonretinal motion signals of eye movements that together allow the brain to infer real motion (Gibson, 1954 and von Holst and Mittelstaedt, 1950). For planar motion, where efference copies can in principle fully Proteases inhibitor match—and thus cancel—retinal motion, the neural substrates involved in this integration have not been systematically investigated in humans before. We demonstrate here that area V3A has a highly specific

preference to planar motion in head-centered coordinates. We found it to be the only motion-responsive region that did not show any significant response to retinal

planar motion, while strongly responding to objective planar motion. V3A thus achieved a near-complete integration of visual with nonvisual planar motion cues related to eye movements, allowing it to discount pursuit-induced retinal motion from its response. This property allowed for a reliable, robust, and completely isolated localization of V3A in every subject examined, by contrasting two simple stimulus conditions. Amisulpride In addition to using a balanced stimulus design that excluded unwanted peripheral effects related to pursuit from affecting the results, an eccentricity-resolved analysis confirmed the key observations in all eccentricities of V3A, including its foveal and peri-foveal representations. In addition to V3A, V6 also responded to planar motion in head-centered coordinates, but its responses were additionally suppressed by retinal motion, leading to partial or full canceling of planar motion responses during fixation. V6 also showed a weak but significant capability to maintain significant responses to planar objective motion when stimuli contained added 3D expansion flow. Finally, V3A and V6 were the only regions reporting objective velocity differences when pursuit and retinal motion were kept the same.

Purkinje cells

express developmentally specific proteins

Purkinje cells

express developmentally specific proteins that delineate conserved parasagittal domains with connectivity to specific nuclei deeper in the cerebellum or brainstem (Gravel and Hawkes, 1990). Indeed, the nervous system has evolved mechanisms for stochastic expression of a variety of cell surface proteins that can determine precise connectivity, fine tune neuronal function, and contribute to the “individuality” of neurons of many types (Yagi, 2013). It can be argued that these expressed molecules are critically important for cellular function and that, therefore, they identify a cell type. However, to our minds, it makes more sense to recognize these mechanisms as capable of providing http://www.selleckchem.com/Akt.html fine-tuned BKM120 in vivo functional diversification within individual cells of a type and to use the molecular ground state as the operational criterion for identifying them as a single cell type. In this way,

one can both recognize the molecular individuality of single cells and maintain continuity with classical anatomical and electrophysiological studies. The practical issue to be addressed is the determination of the molecular ground state of an individual cell or group of cells. We and others have argued that the most objective methodology for this purpose is to profile gene expression. Expression profiles can be obtained from genetically targeted cell populations or randomly chosen single cells with the use of a variety of technologies. Although a discussion of the strengths and weaknesses of these approaches is not possible here, there are certain features of these two broad categories of approach ADP ribosylation factor that must be considered if one hopes to obtain a complete account

of cell types present in complex nervous systems. Strategies that employ genetic targeting allow repeated profiling of the same candidate cell type under a variety of different conditions (Heiman et al., 2008 and Doyle et al., 2008), and they can provide genetic accessibility to that cell type so that additional anatomical, electrophysiological, and functional data can be incorporated into a understanding of the roles it plays in the nervous system. These features allow both technical and biological replicates to be collected in order to improve the quality of the profiles obtained and their comparative analysis. They enable the interrogation of that cell type during development, and they facilitate the incorporation of a wide variety of independent experimental data sets into cell-type-centered databases.