The Intergovernmental Panel on Climate Change (IPCC, 2007) report

The Intergovernmental Panel on Climate Change (IPCC, 2007) reports an increasing trend in mean surface air temperature in Southeast Asia over the past several decades, with a 0.1–0.3 °C increase per decade recorded between 1951 and 2000. As mentioned earlier, all trematodes have check details complex life cycles and use snails as their first intermediate hosts. Asexual

multiplication of the trematodes in snails produces a large number of infective cercariae. The cercarial production rate in snails is fundamental for overall parasite transmission success and this process is relatively temperature dependent, in that an increase in temperature is coupled with an increase in cercarial output (Lo and Lee, 1996, Umadevi and Madhavi, 1997 and Mouritsen, 2002). Temperature-mediated changes in cercarial output also vary among trematode species, from small reductions to 200-fold increases

in response to a 10 °C rise in temperature (Poulin, 2006). In addition, geographical latitude may also affect the production of cercariae by snails. Within the latitude range of 20–55°, trematodes from lower latitudes showed more pronounced temperature-driven increases in cercarial output than those from higher latitudes. The net outcome of increasing temperature will be a greater number of cercarial infective stages in aquatic Selleckchem Akt inhibitor habitats. A few reports have mentioned the effects of climate and environment changes on O. viverrini and C. sinensis emergence, albeit indirectly ( Sithithaworn and Haswell-Elkins, 2003 and Andrews et al., 2008). By the nature of their life cycle, it is possible that climate change in SE Asia, including intense rainfall and flooding and warmer temperatures, may enhance liver fluke abundance and transmission. Schistosomiasis is caused by infection with species of found the blood-fluke Schistosoma. The life-cycle includes a single intermediate host, a freshwater snail, which for the endemic Chinese and Southeast Asian Schistosoma, is always of the family Pomatiopsidae. Three Schistosoma species are recognized as infecting humans in Southeast Asia, namely S. japonicum, S. malayensis and S. mekongi.

Phylogenetically all three species belong to the Schistosoma sinensium clade ( Attwood et al., 2008), so named because of the basal (ancestral) position of S. sinensium in the clade comprising these four Schistosoma species and S. ovuncatum. S. sinensium and its sister taxon S. ovuncatum are both transmitted by snails of the Triculinae and both are exclusively parasites of rodents; these two characters are regarded as plesiomorphic (ancestral) in Asian Schistosoma (see Davis, 1992 and Attwood et al., 2002). These taxa have also been referred to as the “Schistosoma japonicum-group” because all have a minutely spined egg as first described for S. japonicum (see Rollinson and Southgate, 1987). Schistosoma japonicum is often described as a “true zoonosis” ( Ross et al., 1997, Gan et al.

, 2004) and phosphorylation of tyrosine residues in the KCC2 C-te

, 2004) and phosphorylation of tyrosine residues in the KCC2 C-terminal domain, which triggers its lysosomal degradation (Lee et al., 2010).

The functional expression of GABAARs at the cell surface is first controlled at the level of assembly of subunits into heteropentameric complexes. A detailed understanding of this step is limited by the overabundance of different subunits coexpressed in individual neurons. Nevertheless, the use of concatenated subunit constructs selleck chemicals llc representative of the most abundant GABAAR subtype (α1β2γ2) established that assembly of heteropentamers follows strict rules, which ensure that the subunits assume a counterclockwise γ-β-α-β-α arrangement when viewed from the synaptic cleft (Baumann et al., 2001, Baumann et al., 2002 and Baur et al., 2006). Interestingly, corresponding analyses of αβδ receptors indicate that the δ subunit does not simply take the place of the γ2 subunit. Instead Tenofovir cost the optimal subunit arrangement of δ-containing receptors depends on the type of α subunit present (Sigel et al., 2009). Forced expression of subunits in heterologous cells can lead to homomeric assemblies and complexes between α and γ or β and γ subunits that are, however, in most cases retained in the endoplasmic reticulum (ER) (Connolly et al., 1996). Formation of such nonproductive dimers or oligomers renders assembly of functional

receptors rather inefficient, at least in heterologous cells (Gorrie et al., 1997). Unlike the α/γ or β/γ subunit combinations, coexpression of α and β subunits in heterologous cells results in formation of functional receptors that can reach the surface. Moreover, some evidence suggests that αβ receptors may exist naturally in small numbers and contribute to tonic inhibition mafosfamide of neurons (Brickley et al., 1999 and Mortensen and Smart, 2006). However, when α, β, and γ2 subunits are coexpressed the formation of receptors containing all three types of subunits is strongly favored over receptors composed of α and β subunits alone (Angelotti and Macdonald, 1993). Moreover, single

channel analyses of γ2 subunit knockout neurons indicate that receptors composed of α and β subunits alone are gated inefficiently by GABA and have much lower single channel conductances than naturally occurring receptors (Lorez et al., 2000). The assembly of complexes that are translocated to the cell surface involves the initial formation of αβ subunit heterodimers and is principally controlled by the N-terminal/luminal domain of subunits (Taylor et al., 1999, Taylor et al., 2000, Klausberger et al., 2000, Klausberger et al., 2001, Sarto et al., 2002, Bollan et al., 2003a, Ehya et al., 2003 and Sarto-Jackson et al., 2006). This process involves interaction with ER-associated chaperones such as calnexin and binding immunoglobulin protein (BiP) (Connolly et al., 1996 and Bradley et al., 2008).

Two promising candidates are the amygdala, which has been involve

Two promising candidates are the amygdala, which has been involved in encoding negative emotions, including when HIF activation losing money (Yacubian et al., 2006; De Martino et al., 2010; Schlund and Cataldo, 2010), and the anterior cingulate cortex, which is frequently coactivated with the AI when subjects experience negative affective states, also including monetary losses (Blair et al., 2006; Petrovic et al., 2008; Kahnt et al., 2009). We employed here the same task as that previously used for an fMRI study (Pessiglione et al., 2006), except

that pounds were changed into euros, that English was translated into French, and that, in order to shorten the experiment, subjects performed only two test sessions after one full training session. Subjects were provided with written instructions, which were reformulated orally when necessary, asking them to try and maximize their financial payoff (see Figure 1; Supplemental Information available online). Each session was an independent task containing three new pairs of cues to be learned. Cues were abstract visual Bortezomib datasheet stimuli taken from the Agathodaimon alphabet. Each pair of cues was presented 30 times for a total of 90 trials. The three cue pairs corresponded to the three conditions (gain, neutral, and loss), which were respectively associated with different pairs of outcomes (winning 1€

versus nothing, looking at 1€ versus nothing, and losing 1€ versus nothing). Within each pair, the two cues were associated to the two possible outcomes with reciprocal probabilities (0.8/0.2 and 0.2/0.8). Note that there

was no financial outcome associated with the neutral cue, for which the euro coin could be looked at, but not won or lost. On each trial, one pair was randomly presented and the two cues were displayed on a computer screen above and below a central fixation cross, their relative position being counterbalanced across trials. The subject was required to choose the upper stimulus by pressing the space bar (“go” response) or the lower stimulus by retaining from pressing any button (“no go” response) within a 4 s delay. Please note that, since the position on screen was counterbalanced, response (go versus no go) and value (good versus bad cue) were orthogonal. After the 4 s delay, the chosen cue was circled in red and then the outcome (either “nothing,” “gain,” Non-specific serine/threonine protein kinase “look,” or “loss”) was displayed on the screen. In order to win money, subjects had to learn by trial and error the cue–outcome associations, so as to choose the most rewarding cue in the gain condition and the less punishing cue in the loss condition. To identify ROI, we reanalyzed, using SPM8 software, the fMRI data acquired for a previous published study (Pessiglione et al., 2006) that investigated the effects of dopaminergic drugs on instrumental learning. Please refer to this publication for details about image acquisition and preprocessing.

Consequently, projection neurons respond to excitatory odors diff

Consequently, projection neurons respond to excitatory odors differently;

TCs readily increase firing rate while MCs additionally show graded phase-advance (Figure 7). In this study, identifying the key morphological features (Mori et al., 1983) such as the soma and dendritic position was essential in elucidating functional differences clearly. Other means of distinguishing projection neurons in the olfactory TSA HDAC supplier bulb such as the depth of recording might be correlated with cell types and thus potentially show similar trends; so far, however, such attempts have failed to distinguish classes of neurons not overlapping in functional measures, such as the sniff locking. One reason for this could be that larger tufted cells such as deep tufted cells (e.g., cells

12 and 13 in Figure S2) are easily confused MS-275 price with MCs if recording depth were the sole measure of identification. More data will be needed to extend this analysis to potential subgroups of TCs, such as superficial, middle, or deep TCs. Our data so far showed no tendency for further distinction in phase locking (Figure 2K). Our method described here to identify MCs and TCs based on the sniff locking would help with further investigation of how the two populations may differ, even in cases where morphology is unavailable. Importantly, this is likely to extend to the awake state, where we observed similar strength and diversity of phase preference (Figure S1). We have investigated the mechanistic basis of the observed phase locking using a newly developed modeling approach that generates a large number of models with randomly chosen connectivity and selects for those that are consistent with the experimental data. The simplicity

of the network models has made it possible to sample a vast fraction of connectivity space. This allowed us to extract features of the network that correlate with phase properties consistent with experimental data. While we do not claim that we unequivocally found the actual connectivity implemented in the olfactory bulb, several robust features emerged from this selection procedure. The first observation is the strong feed-forward inhibition, specifically strong PGo-MC connectivity (Figure 6C), which may underlie the GABAergic component crucial in separating MC activation away from the TC activation. Second, the models suggest that Rolziracetam MCs are predominantly driven weakly or indirectly and shaped by inhibition. The robustness of the TC phase in turn points toward OSN inputs strongly and directly exciting TCs. A number of recent investigations suggest that the excitatory pathway to MCs from OSNs is rather indirect (Gire and Schoppa, 2009; Najac et al., 2011; Gire et al., 2012). It is exciting to note the consistency of our modeling results with this view. However we cannot exclude an important role of direct transmission from OSNs to MCs, for example by glutamate spillover onto MCs at higher input strengths (Najac et al., 2011).

In experiment 4, we investigated whether amnesia following damage

In experiment 4, we investigated whether amnesia following damage that included PRC could be characterized

by a heightened susceptibility to perceptual interference. There were three conditions involving High Ambiguity stimuli (Low Interference 1, High Interference, Low Interference 2). The High Interference condition contained consecutive High Ambiguity Object trials, whereas every High Ambiguity Object trial in both Low Interference conditions was interspersed with two trials containing photographs of easily discriminable everyday objects (Figures 2E–2G). We predicted that the nature of the intervening stimuli would affect performance in individuals with PRC damage, with better performance under conditions of low interference. Analysis find more of eye movement patterns in healthy participants indicated that the High Ambiguity condition was associated with a greater degree of conjunctive processing than the other conditions. We performed a planned interaction comparison to determine if the High Ambiguity Object condition was associated with more conjunctive processing, relative to our size difficulty control: (High Ambiguity Objects – Low Ambiguity Objects) – (Difficult Size – Easy Size). This revealed that participants made more eye movement transitions within an individual object compared to transitions between the two objects in the High Ambiguity Enzalutamide solubility dmso condition relative

to the other conditions (t(15) = 4.08; p < 0.001) (Figure 3). Indeed, this ratio of within-item relative to between-item saccades was greater for High compared to Low Ambiguity discriminations (t(15) = 6.58, p < 0.001). We also performed an analysis of the temporal characteristics of these eye movements, which revealed a greater degree of temporal clustering in the High Ambiguity condition (see Supplemental Information, Figure S1C, available online). These results indicate that healthy participants analyzed the ambiguous objects as wholes, rather than by a serial comparison of single features. Experiments 2–4 investigated the neural substrates of this ability. In order to isolate brain regions associated with feature

ambiguity resolution, while controlling for general task difficulty, our planned comparison was the same interaction t contrast described above. Estimates of the mean BOLD Endonuclease signal for each of the four conditions were averaged across voxels within our two anatomically defined, bilateral regions of interest: the hippocampus and PRC. The planned comparison revealed feature ambiguity effects within the PRC (t(19) = 3.5, p < 0.001; Figure 4). This region showed reliably greater activity for High relative to Low Ambiguity discriminations (t(19) = 5.2, p < 0.001), but no difference in activity for Difficult relative to Easy Size discriminations (t(19) = 0.5, p = 0.3). By contrast, the comparison of High versus Low Ambiguity Objects was not significant in the hippocampus (t(19) = 1.4, p = 0.

Huntingtin protein levels were normalized to β-tubulin and expres

Huntingtin protein levels were normalized to β-tubulin and expressed as % vehicle treated controls. BACHD: Tissues were homogenized in RIPA lysis buffer. Fifteen micrograms of total protein lysate was resolved on a 4%–12% bis-tris gel (Invitrogen).

Proteins were transferred to a nitrocellulose membrane and probed with MAB2166 and anti-GAPDH (1:5,000 Abcam). All behavioral tests with the YAC128 mice were conducted independently at LGK-974 mouse Genzyme, and tests in BACHD and R6/2 animals were conducted independently at UCSD. The range of numbers of animals used in the various behavioral assays is listed in the figure legends. The exact animals numbers in each treatment group, in each of the various behavioral assays is included in the Supplemental Experimental Procedures. Metabolism inhibitor Also see Supplemental Experimental Procedures for detailed procedures for accelerating rotarod, elevated-plus maze, open-field test, light/dark analysis, body mass, and survival. Mean values were used for statistical analyses. Data are expressed as mean ± SEM. For two groups, unpaired two-tailed

t tests were used; for more than two group comparisons, one-way ANOVAs were used followed by the post hoc Tukey’s multiple comparison test; for more than two comparisons of two or more groups, two-way ANOVAs followed by Bonferroni’s post hoc tests were used (Prism GraphPad and Kalidagraph). p values for the ANOVAs are reported in the figure legends, and p values from the post hoc tests are included in the text when making paired comparisons. A chart listing the Terminal deoxynucleotidyl transferase post hoc comparisons of all rotarod analyses is provided

(Figure S5). Significance of survival was determined using the Kaplan-Meyer method. p < 0.05 was considered a statistically significant difference. Please see Supplemental Experimental Procedures for the following experimental procedures: Time resolved foerster resonance energy transfer (TR-FRET) and capillary gel electrophoresis. We thank W. Yang (University of California, Los Angeles) for the BACHD animal model and his technical support. We also thank G. Bates (Kings College London) for the R6/2 model and her technical support. We would like to thank S. Freier, A. Watt, and A. Salim (Isis Pharmaceuticals) for identification of the huntingtin ASOs and J. Matson for bioanalytical support on the mouse and monkey tissues. We thank J. Boubaker for her technical support. We thank R. Smith and D. Macdonald for their thoughtful discussions. H.B.K., E.V.W., C.M., G.H., and C.F.B. are employees of Isis Pharmaceuticals. L.M.S., S.H.C., and L.S.S. are employees of Genzyme Coropration. A.W. is an employee of Novartis Pharma AG. D.W.C. is a consultant to Isis Pharmaceuticals. This work was supported in part by CHDI Inc. and H.B.K. was supported by a postdoctoral fellowship from the Giannini Foundation.

We thank M E Hasselmo, E Kropff, T Solstad, and E A Zilli for

We thank M.E. Hasselmo, E. Kropff, T. Solstad, and E.A. Zilli for helpful discussions. This work was supported by a Marie Curie Fellowship, the Kavli Foundation, and a Centre of Excellence grant from the Research Council of Norway. “
“Comparative and pathological studies suggest the mammalian cerebral cortex to be the anatomical substrate of higher cognitive functions including language, episodic memory, and voluntary movement (Jones and Rakic, 2010, Kaas, 2008 and Rakic, 2009). The cerebral cortex has a uniform laminar structure that historically has been divided into six layers (Brodmann, 1909). The upper layers (1 to 4) form localized

BMS 354825 intracortical connections (Gilbert and Wiesel, 1979 and Toyama et al., 1974) and are thought to process information locally. The deep layers of the cortex, 5 and 6, Z-VAD-FMK clinical trial form longer-distance projections to subcortical targets (including the thalamus, striatum, basal pons, tectum, and spinal cord) and to the opposite hemisphere. Some layer 5 neurons are among the largest cells of the brain and exhibit the longest connections. Layer 6b in mouse neocortex is a distinct sublamina with characteristic connections, gene expression patterns, and physiological

properties (Hoerder-Suabedissen et al., 2009 and Kanold and Luhmann, 2010). Understanding how neurons and glia are organized into layers to assemble into functional microcircuits (Douglas and Martin, 2004) is one of the first steps that will be required to relate anatomical structures to cellular functions. Subclasses of pyramidal neurons

and interneurons populate specific layers, each characterized by a different depth in the cortex with a specific pattern of dendritic and axonal connectivity (Jones, 2000, Lorente de No, 1949 and Peters and Yilmaz, 1993). However, Fossariinae analyzing these laminar differences is difficult and often suffers from subjectivity (Zilles and Amunts, 2010). The currently available repertoire of markers that allow the distinction of cortical layers and of many neuronal and glial subtypes is rapidly improving because of developments in cell sorting and gene expression analysis (Doyle et al., 2008, Heintz, 2004, Miller et al., 2010, Molyneaux et al., 2007, Monyer and Markram, 2004, Nelson et al., 2006, Thomson and Bannister, 2003 and Winden et al., 2009). These molecular tags allow highly specific classes of neurons and glia to be monitored, modulated, or eliminated, thereby providing greater insights into cortical neurogenesis and the classification of lamina specific subclasses of cells. Laminar molecular markers were first identified by studying single protein-coding genes (Hevner et al., 2006, Molyneaux et al., 2007 and Yoneshima et al., 2006) but more recently, high-throughput in situ hybridization (Hawrylycz et al., 2010, Lein et al., 2007 and Ng et al., 2010) and microarrays (Oeschger et al., 2011, Arlotta et al.

To measure synchrony between FEF and V4, we used multitaper spect

To measure synchrony between FEF and V4, we used multitaper spectral methods to compute coherence between spikes from well isolated single units in the FEF and local field potentials (LFPs) in V4. First taking all types of FEF cells together, we found that spike-field coherence

in the gamma frequency range was significantly enhanced between FEF and V4 when attention was directed inside the joint RF (Figure 5A; coherence averaged between 35 and 60 Hz; paired t test p < 0.001). At the population level gamma band coherence increased by 13%. This result confirms and extends findings from our recent study based on multi-unit activity that demonstrated enhanced neural synchrony between FEF and V4 with attention (Gregoriou et al., 2009a). After subdividing the coherence spectra in FEF by cell class, the check details results showed that visual, visuomovement, and movement neurons display distinct FEF-V4 coherence Roxadustat price profiles. Coherence between the spikes of purely visual FEF neurons and LFPs in V4 showed a 16% enhancement with attention in the gamma range and this increase was statistically significant (Figure 5B; 35–60 Hz, paired t test, p < 0.001).

In agreement with our previous results we found that the distribution of the average (between 35 and 60 Hz) relative phase between FEF spikes and V4 LFPs had a median close to half a gamma TCL cycle (attend-in condition; median = 176°, Rayleigh test, p < 0.001). This phase shift corresponds to a time delay of ∼10 ms between spikes in the FEF and the phase of maximum depolarization in the V4 LFP, and we have previously suggested that a 10 ms time delay is needed to account for conduction and synaptic delays between the two areas (Gregoriou et al., 2009a). Spike-field coherence between FEF neurons with saccade-related activity (visuomovement and movement neurons) and V4 LFPs did not display any significant gamma band modulation with attention (Figures 5C and 5D; paired t test, visuomovement cells: p = 0.22, 7% increase; movement cells, p = 0.87; 1%

decrease with attention). For a distribution of attentional effects in gamma coherence see Figure S3. The attentional enhancement of gamma coherence was significantly different across the three FEF cell classes (Kruskal-Wallis, p < 0.001). Coherence between visual FEF cells and V4 LFPs was significantly enhanced relative to that between visuomovement or movement FEF cells and V4 (Tukey-Kramer, p < 0.001 for both pair comparisons), whereas attentional effects on FEF-V4 coherence were not significantly different for visuomovement and movement FEF cells (Tukey-Kramer, p = 0.69). We also confirmed that the absence of gamma coherence modulation with attention between FEF movement neurons and V4 cannot be attributed to low firing rate (see Supplemental Information).

A comparison of residues that constitute the 7-1a, 7-1b, and 7-2

A comparison of residues that constitute the 7-1a, 7-1b, and 7-2 epitopes of the Kolkata strains and the vaccine strains is

presented in Table 4. Twenty nine amino acid residues of this antigenic epitope of the VP7 proteins of circulating G1, G2, and G9 RVA strains were compared with the Rotarix-G1, RotaTeq-G1, RotaTeq-G2, and 116E-G9 vaccine strains. Kolkata G1 strains showed mismatches in 94, 100, 123, 291 and 217 positions in 7-1a and 7-2 domains with Rotarix-G1and RotaTeq-G1strains. Kolkata G2 strains also showed mismatches in 4 positions, 87, 291, 213 and 242 in respect to RotaTeq-G2 strains. When VP7 protein of G9 strains were compared with 116E-G9 vaccine strain, it revealed that circulating lineage III G9 strains also see more differ from 116E strain within antigenic domain at 87, 94, 100, 291, 242, 145 and 221 positions (Table 4). In low income countries of Asia (India, Bangladesh,

Pakistan, Vietnam, China) and Africa, high prevalence (30–40%) of RV has been reported among hospitalized children [17], [44], [45], [46], [47], [48] and [49]. In this study, the incidence was higher in hospitalized children (53.4%) and out-patients (47.5%) than previous reports. The this website children seeking treatment in outpatient departments may constitute a major source for dissemination of virus. Unlike developed countries where one or two genotypes predominate in a season [54] and [55], a large number of genotypes was observed (G9, G2, G1, G12) at >15% frequency in Kolkata. This agrees with the previous reports from India and Bangladesh Sitaxentan [17] and [44]. Although not demonstrated so far, emergence of new strains, which contributes to genetic diversity, may be one cause of lower vaccine efficacy

in developing countries. Selective pressure resulting from population immunity may drive emergence of strains able to evade vaccine immunity [13]. Moreover for improving efficacy, mass vaccination of children through national immunization program is required, whereas in countries like India, currently only a small proportion of children are vaccinated. Considering the socio-economic structure, high cost of vaccines and the large diversity of strains in low income countries, successful implementation of RV vaccines is still an unfulfilled goal [17], [25] and [50]. Thus to fulfill the lacunae of disease control by vaccination, continuous surveillance for RV is required to monitor incidence, circulating genotypes, emergence of new reassortant strains in population, which will also help in effective disease management and prevention of large scale outbreaks. In addition knowledge of currently circulating strains is needed prior to mass vaccination, for comparison and evaluation during post vaccination studies. As Kolkata has a tropical climate, seasonality of rotavirus infection (Fig.

We recorded from three different z planes separated by 10 μm, thu

We recorded from three different z planes separated by 10 μm, thus resulting in a temporal resolution of 10 Hz. To precisely register the electrophysiological with the optical recordings in time we recorded the trigger signal given by the camera at the beginning of every frame as a separate trace in the electrophysiological recording. Automated counting of trigger signals allowed then to determine exactly the beginning of every single frame of the calcium imaging in the electrophysiological trace. Image analysis was carried out automatically by custom made MATLAB software. As a first step in the analysis process, each set of three images from different z positions was collapsed into one maximum projection image. All maximum projection

images from one recording were collected in one stack. Next, an F0 image was generated in which see more each pixel represented the median of all pixel values at this position throughout the stack. The F0 image was top-hat filtered using a disk-shaped structure (radius, three pixels) to correct for uneven background brightness. All areas brighter than two times the SD of the F0 pixel values and larger than 200 pixels

were detected as dendrites. The F0 image was also used to generate a ΔF/F0 stack by subtracting it from each frame of the stack and dividing the result by F0. Next, we calculated the derivative in time of the ΔF/F0 stack using a 3D convolution filter. The derivative was eroded with a disk-shaped structure (radius, two pixels) and subsequently binarized using a threshold of 15% ΔF/F0 s−1. A calcium transient was defined as a minimum of ten connected pixels in the binarized stack. The spatial Selleck VX770 center of individual calcium transients was defined as the pixel coordinates where the largest increase in fluorescence occurred within each signal. Where Rutecarpine necessary, these positions were shift corrected across recordings. Next, calcium transients were assigned to specific dendritic sites whose locations along the dendrite were determined in an iterative process: in the first step the center of the first calcium transient in a given recording was estimated as described above. Subsequently, all transients that occurred within

4 μm distally or proximally from the center of the first transient along the dendrite were considered to be transients from the same site. The center coordinates of all these transients were averaged and this value was used as a refined locus for this site of activity for the second iteration. During the iterations, some transients were newly included or excluded from the population assigned to this particular site and its center location adapted slightly. The process was stopped after 20 iterations, when in fact no more changes occurred. Then the first transient in time that was not yet assigned to any site was used as a starting point for determining the next site of activity. Only sites where at least two transients occurred were considered for further analysis. On average 7.