6 ± 4 1, n = 93 boutons on 14 motoneurons) was similar to that fo

6 ± 4.1, n = 93 boutons on 14 motoneurons) was similar to that found in nonspinalized mice (p = 0.07; Figure 2F), excluding the possibility that YFP+ boutons contacting motoneurons derive primarily from supraspinal

neurons. Rabies virus trans-synaptic tracing has also identified dI3 INs as a source of synaptic input to motoneurons ( Stepien et al., 2010). Thus, glutamatergic dI3 INs project directly to motoneuron somata and dendrites ( Figure 2G). vGluT2+/YFP+ boutons were also detected in intermediate laminae of cervical and lumbar segments (12.8 ± 4.1 boutons/1,000 μm3, n = 5 sections from 2 spinal cords; Figure S2B). selleck screening library Some of these boutons were in apposition to other dI3 INs (Figure S2C). Thus, both motoneurons and INs are targets of dI3 INs. We determined whether dI3 INs receive direct input from primary sensory afferents. Expression of vGluT1 marks low-threshold cutaneous and proprioceptive primary afferent fibers and is excluded from spinal interneurons (Alvarez et al., 2004; MK-1775 cell line Oliveira et al., 2003; Todd et al., 2003). We used vGluT1 as a molecular marker of direct afferent input to dI3 INs (Figure 3A). We found that 88% of YFP+ dI3 INs (n = 46 out of 52 neurons) were contacted by vGluT1+ boutons (9.2 ± 3.7 boutons /dI3 IN soma and proximal dendrites, n = 18). In the early postnatal spinal cord, parvalbumin (PV) serves as a marker of proprioceptive afferents (Mentis et al., 2006).

Both vGluT1+/PV+ (n = 26) and vGluT1+/PVnull boutons (n = 85) were detected on dI3 INs at P1–P7 (n = 21, one to four optical heptaminol sections per neuron were analyzed; Figure 3B). Thus, proprioceptive and cutaneous sensory afferents converge on dI3 INs. Analysis of vGluT1 labeling in adult spinal cord tissue examined 7 days after thoracic spinalization (n = 2) revealed no diminution in the number of vGluT1+ boutons apposed to dI3 INs (n = 18 dI3 INs, 11.9 ± 8.0 boutons /dI3 IN, p = 0.2; Figure 3C), which was consistent with the view that these boutons derive from sensory afferents.

We used whole-cell patch-clamp recordings to assess the physiological connectivity between sensory afferents and dI3 INs. All dI3 INs in P5–P16 Isl1-YFP mice (n = 51, input resistance = 626 ± 356 MΩ) discharged repetitively. However, approximately one-sixth did not fire until after a delay of >50 ms because of the expression of a 4 AP-sensitive slowly inactivating potassium (ID-type) current ( Figures 4A and S3). Thus, transient synaptic excitation could elicit spike firing in most (approximately five-sixths) dI3 INs. Then, we assessed sensory input using electrical stimulation of L4 or L5 dorsal roots, and this revealed that 105 out of 114 (92%) dI3 INs had sensory-evoked excitatory responses (Figure 4B). Of these 105 dI3 INs, 31 (30%) responded with a single excitatory postsynaptic potential (EPSP) or action potential, and 35 (33%) responded with a pattern comprised of an early EPSP or action potential followed by a longer-lasting IPSP (Figure 4Bi).

, 2010), and some exciting insights have been made into signal or

, 2010), and some exciting insights have been made into signal or state-dependent activation of such players (e.g.,

Banerjee et al., 2009), a key question is how are individual genes targeted for specific regulation? Although multiple Akt cancer classes of sequence-specific RNA regulatory mechanisms contribute to shaping the functional landscape, and there are significant interactions between these molecular regulators, we will focus on microRNA (miRNA)-mediated control over the maturation and plasticity of neurons and their synaptic connections, highlighting primarily observations made in the past few years. miRNAs were first identified based on classical genetics as regulators of developmental timing in Caenorabditis elegans ( Lee et al., 1993; Reinhart et al., 2000). These short noncoding RNA were then found in other organisms by virtue of striking sequence conservation across species ( Pasquinelli et al., 2000). miRNA genes are transcribed

as RNA polymerase II or III transcripts (pri-miRNA) that are processed by specific Screening Library chemical structure nuclease cleavage (or RNA splicing for miRtrons) to produce short hairpin RNAs (pre-miRNA) that are transported out of the nucleus and then cleaved once more to generate mature miRNAs that can be loaded into protein complexes that allow binding to specific target mRNA ( Figure 1C; reviewed by Bartel and Chen, 2004). Mature miRNA-target mRNA pairs are formed by proteins in the Argonaute (Ago) family together with other components of the RNA-induced silencing complex (RISC; Du and Zamore, 2005). Although there are exceptions, miRNAs inhibit expression for most target genes by reducing steady-state message levels ( Guo et al., 2010), although this may occur after an initial blockade of translation ( Bazzini et al.,

2012; Djuranovic et al., 2012). Many rounds of transcriptome TCL sequence and expression analysis have uncovered a large number of miRNA genes spanning all multicellular organisms (see http://www.mirbase.org; Griffiths-Jones et al., 2006). Among animal species, the number of miRNA genes has expanded dramatically with increasing organismal complexity (i.e., numbers of differentiated cell types), contributing to speculation that despite high conservation in many miRNA families, diversification of other miRNA genes has contributed significantly to the evolution of different metazoan body plans (Sempere et al., 2006). For example, Cnidarian genomes contain tens of miRNA genes (e.g., 17 in Hydra and 49 in Nematostella), whereas Ecdysozoa have roughly 5- to 10-fold more (e.g., 223 in C. elegans and 240 in D. melanogaster), and Humans have over 1,500 (http://www.mirbase.org).

All the compounds were identified by spectral data In general, m

All the compounds were identified by spectral data. In general, mass spectrum showed the molecular

ion peak, which corresponds to the formula weight of the hydrazones. The elemental analyses of the compounds are in consistence with the molecular formula (Table 1). The electronic spectra of the hydrazones A1–A6 were taken in ethanol (10−3 mol−1). In the UV–Visible spectra of all these compounds the first band appeared around 257 nm was due to the π → π* transitions of the heterocyclic ring and the second one appeared around 350 nm was due to the n → π* transition of the >C]N–group. 8 FT-IR spectra showed the C]O peak around 1660 cm−1, C=N around 1560 cm−1 and the NH stretching vibrations around GSK1349572 in vitro 3064 cm−1. The 1H NMR spectrum showed the hydrazide (NH) protons as a singlet around 12.1 ppm, the imine protons (N]C–H) around 8.3 ppm, methoxy protons around 3.8 ppm and aromatic protons in the range 6.5–8.8. The 13C NMR spectrum showed the C]O signals around 162.5, C]N signals around 150.6 ppm, selleck inhibitor OCH3 signals around 55.5 ppm and aromatic carbon in the range 114.7–158.5 ppm. 9 Single crystals suitable for X-ray diffraction study for the hydrazone (A1) was grown from the slow evaporation of an ethanol solution at room

temperature. A pale yellow crystal of (A1) was mounted on a glass fiber and used for data collection. Crystal data was collected using graphite monochromatised Mo-Kα radiation (λ = 0.71073 Å). The structure was solved by direct method using SHELEX-97 and refined by full-matrix least-squares techniques against F2 using SHELEX-97. All the non-hydrogen atoms were refined anisotropically. A summary of pertinent crystal data along with further details of structure determination and refinement are given in Table 2. Selected bond lengths and bond angles are given in Table 3.The hydrazone crystallizes in an orthorhombic, chiral space group pbca. The single crystal

X-ray structure of A1 reveals the presence of two molecules in the unit cell. The C]N azomethine [N(3)–C(7)]-bond length 1.278 (3) Å in A1 has a double bond character. The existence of A1 in keto TCL form in solid state is evident from the [O(1)–C(6)] bond length 1.223 (3) Å and the side chain carbonyl [O(1)-C(6)] show a typical double bond character with bond length 1.223 (3) Å.10 and 11 In this compound, there is also an intermolecular hydrogen bond (Table 4) between the N(2)–H(4) and N(1)′ [N(2)–H(4)…N(1)′, 2.225 Å] and N(2)′–H(5) and N(1) [N(2)′–H(5)…N(1), 2.202 Å], stabilize the crystal structure forming a supramolecular architecture. ORTEP view and unit cell of A1 are given in Fig. 1 and Fig. 2 respectively.

05) Figure 5D shows the average deflection (n = 11,330 deflectio

05). Figure 5D shows the average deflection (n = 11,330 deflections, 28 electrodes from five sessions) and the power spectrum of the LFP trace over 400 ms windows centered on each deflection (100 ms before to 300 ms after the initial sharp voltage change). It revealed a large peak at alpha frequencies (12.2 Hz) and, indeed, the power spectra of the LFPs during strong deflection episodes showed a marked increase in alpha oscillations (∼12 Hz) (see Figure S4 for an example). We compared LFP oscillatory power on correct trials during baseline blocks with the postinjection blocks separately

for sessions or recording sites with and without deflections. In baseline blocks, there was a prominent alpha/beta band (10–30 Hz) during the fixation, delay, and saccade execution epochs (Figures 6A and 6B). During postinjection blocks after injection MK-8776 price of SCH23390 (n = 163 electrodes), but not saline (n = 84 electrodes), there was an increase in the power of oscillations

below 30 Hz compared to baseline blocks. The deflections have an alpha component (see above) so, naturally, sites with deflections (n = 95) showed an increase in alpha band (10–14 Hz) after SCH23390 and also in beta band (14–30 Hz) for novel and familiar associations over baseline blocks (Figures 6A and 6B, Wilcoxon test, p < 0.05, shaded areas; Figure 6C, last 20 trials/block). see more Importantly, this increase in low-frequency power in the LFPs was still observed in sites without deflections (n = 68, Figure 6B), indicating that the SCH23390-induced increase in low-frequency oscillations was not due to the deflections alone. This increase was more pronounced for novel than familiar associations in only sites without deflections (Figures 6B and 6C; Wilcoxon test, p < 0.05). The increase in alpha/beta oscillations was also observed in sessions without deflections on any recording site (Figure S5)

and was greater at sites where blockade of D1Rs impaired learning compared to areas where learning was intact (Figure S5). Our findings indicate that dopamine D1 receptors in the monkey lateral PFC are likely to be involved in learning new cue-response associations but less involved in performance of familiar associations. After the injection of a D1R antagonist, especially in the ventrolateral PFC, monkeys learned new cue-response associations much more slowly, whereas performance of highly familiar associations was intact. Two not mutually exclusive possibilities may account for this dissociation: (1) familiar associations are not dependent on prefrontal D1Rs, and (2) they are dependent on another brain area, such as the striatum, where they could be encoded as habits (Graybiel, 2008).

5 1 (congress mash) Laboratory wort filtration volume was measur

5.1 (congress mash). Laboratory wort filtration volume was measured according to the method of Evans et al. (2011). After returning the first 100 ml of wort collected during laboratory wort filtration, the volume of wort filtered in the next 25 min was measured as an index of mash filterability. The resulting bright worts were analysed for: hot wort extract using an Anton Proteases inhibitor Paar DMA 4500 density metre according to Analytica-EBC Method 4.5.1, free amino nitrogen (FAN) by the spectrophotometric

ninhydrin method (Analytica-EBC Method 4.10), wort viscosity according to Analytica-EBC Method 8.4 and EBC wort colour according to Analytica-EBC Method 4.7.1. All data, apart from malting and brewing data, were analysed using Genstat® Version 14.1 for Windows (VSN International Ltd., UK). Relationships between pathogen DNA and mycotoxins were analysed using single linear regression analysis. Multiple linear regression with groups was used to identify relationships between the DNA of Fusarium spp. and Microdochium spp. and quality parameters of barley grain such as TGW and SW. Where necessary DNA or mycotoxin data were

log10 transformed to normalise residual distributions. Unbalanced analysis of variance, using linear regression was carried out on fungal and mycotoxin data from 2010 to 2011 to determine the significance (P < 0.05) of sampling region and malting barley variety. It was not possible GSK126 chemical structure to include data from 2007 to 2009 in this analysis as 17-DMAG (Alvespimycin) HCl samples from these years were not randomly selected but on the basis of their known mycotoxin contents. Therefore descriptive statistics were used for the DNA, mycotoxin and malting/brewing data on all selected samples. The DNA of Fusarium spp. and Microdochium spp. and malting/brewing parameters of samples is

presented as mean with 95% confidence intervals and the mycotoxin data is presented as mean with 95th percentile and maximum values. The co-existence of the species of the FHB complex was explored using Principal Component Analysis (PCA) on the correlation matrix of eight variables. These variables were fungal biomass (log10 pg/ng of total DNA) of F. graminearum, F. culmorum, F. poae, F. tricinctum, F. avenaceum, F. langsethiae, M. majus and M. nivale. Malting and brewing quality data were entered retrospectively into a d-optimal factorial design space using experimental design software (Design Expert, v 7.0, Stat-Ease, Mn, USA). The malting and brewing quality parameters for the 54 barley samples were entered as responses and modelled against 15 factors, which were: the DNA contents of the individual species analysed in the samples for two Microdochium and six Fusarium species (QPCR data), the barley cultivar, harvest year and the concentrations of five mycotoxins analysed in the samples (NIV, DON, HT-2, T-2, ZON).

Several groups have documented shape learning in individual neuro

Several groups have documented shape learning in individual neurons in temporal cortex and proposed that such changes could occur as a consequence of competitive segregation of those neurons’ inputs by Hebbian mechanisms (Fukushima et al., 1988, Kourtzi and DiCarlo, 2006, Rolls and Tovee, 1995 and Sohal and Hasselmo, 2000). Polk and

Farah (1995) explicitly proposed that activity-dependent Hebbian mechanisms could drive the coarser segregation of neurons responsive to learned stimulus categories, like letters and words, from neurons responsive to other shapes. Here, we hypothesize that self-organizing buy Cilengitide segregation within cortical areas could underlie the formation of functional domains in the temporal lobe. In the same way that differential activity in the two eyes drives the segregation of ocular dominance columns within V1 or tactile experience

with differential whisker activity drives the organization of whisker barrels within each somatosensory cortical area, we propose that differential early experience with face parts being experienced conjunctively with other face parts, but disjunctively with other objects, and vice versa, could drive the segregation of category selective domains within cortical areas in inferotemporal cortex. We propose that intensive early experience with symbols drives the segregation of a domain selective for those learned Anticancer Compound Library concentration symbols, and by extension, we propose that intensive early experience with faces and other objects drives the segregation of face and shape domains. Figure 6 indicates that this segregation occurs independently several times along inferotemporal cortex, suggesting an underlying organizational principle of modular segregation within each cortical

area. This general organizational principle probably further almost involves interconnectivity between functionally related modules: modules in V1 are selectively interconnected with functionally related modules in V2 (Livingstone and Hubel, 1984), and Face-selective modules in different parts of IT are selectively interconnected (Moeller et al., 2008). By inspection of Figure 6, there is another peculiar similarity between the face/shape modular architecture in IT and other modules in the visual system, namely that the modular divisions within each area tend to run perpendicular to the areal border: ocular-dominance columns in old-world monkeys, orientation columns in new-world monkeys, and functional domains (cytochrome oxidase stripes) in V2 are all oriented perpendicular to the V1/V2 border (Blasdel and Campbell, 2001, Hubel and Freeman, 1977 and Tootell et al., 1983). This similarity is noteworthy because it is consistent with our hypothesis of a common rule-based organization.

DD remodeling occurs without retraction or extension of neurite p

DD remodeling occurs without retraction or extension of neurite processes. Instead, the DD ventral process switches from an axonal to a dendritic fate (and vice versa for the dorsal process). Many aspects of C. elegans larval development are controlled by cell intrinsic developmental timing genes, which are generically termed heterochronic genes

( Moss, 2007). In particular, the heterochronic gene lin-14 controls the timing of hypodermal development, whereby L2 hypodermal cell fates are expressed precociously during the L1 in lin-14 mutants ( Ambros and Horvitz, 1984). Similarly, lin-14 is expressed in DD neurons, and DD remodeling occurs earlier in lin-14 mutants, initiating during embryogenesis ( Hallam and Jin, 1998). Thus, LIN-14 dictates when DD remodeling is initiated. This study shows that heterochronic genes play a role in postmitotic neurons to pattern synaptic plasticity. Because lin-14 orthologs are not found in other organisms, it remains unclear Epacadostat in vitro if control of synaptic plasticity by heterochronic genes represents a conserved mechanism. DD plasticity

(like other forms of invertebrate plasticity) is generally considered to be genetically BLZ945 supplier hard wired, i.e., dictated by specific cell intrinsic genetic pathways. Thus, it also remains unclear if activity-induced refinement of vertebrate circuits and DD plasticity represent fundamentally distinct processes, which are mediated by distinct molecular mechanisms. about Here we show that a second heterochronic gene, hbl-1, regulates several aspects of DD plasticity. The hbl-1 gene encodes the transcription factor HBL-1 (Hunchback like-1) ( Fay et al., 1999). We show that convergent pathways regulate hbl-1 expression in D neurons, conferring cell and temporal specificity and activity dependence on D neuron plasticity. Thus, our results define a cell intrinsic genetic pathway that dictates a form of

activity-dependent synaptic refinement. The DD motor neurons are born during embryogenesis, and remodel their synapses during the L1. A second class of GABAergic motor neurons, the VD neurons, is born during the late L1 stage but does not undergo remodeling. VD neurons share many other characteristics with DD neurons, including similar cell body positions, similar axon morphologies, similar roles in controlling locomotion, and similar expression profiles (Jorgensen, 2005). Like DDs, VD neurons initially form ventral synapses; however, unlike the DDs, VD neurons retain these ventral synapses in the adult. VD and DD neurons also differ in that a transcriptional repressor (UNC-55) is expressed in the VD but not in the DD neurons, and this difference has been proposed to explain the disparity in their ability to undergo synaptic remodeling (Shan et al., 2005, Walthall, 1990, Walthall and Plunkett, 1995 and Zhou and Walthall, 1998). Prior studies suggested that VD neurons undergo ectopic remodeling in unc-55 mutants ( Shan et al.

, 2008) Slowly dividing NSCs with long-term self-renewal potenti

, 2008). Slowly dividing NSCs with long-term self-renewal potential are not located in close vicinity of periventricular vessels, but contact them via endfeet of their long basal processes (Beckervordersandforth et al., 2010 and Shen et al., 2008). These NSCs express VEGFR3, required for NSC maintenance and olfactory bulb neurogenesis (Calvo et al., 2011). Once Alpelisib mw activated to initiate division, NSCs and the continuously proliferating transit-amplifying progenitors (TAPs) become attracted to periventricular vessels via SDF1-CXCR4 signaling (Kokovay et al., 2010 and Tavazoie et al., 2008). The perivascular

ECM in the SEZ niche functions as a deposit of growth factors to support neural precursor proliferation. Thus, signals derived from SEZ vessels foster proliferation of neural precursors, while long-term self-renewing stem cells are located in the more

hypoxic niche to maintain quiescence (Mohyeldin et al., 2010). The development of the SGZ niche occurs primarily postnatally. SGZ vessels have an intact BBB and restrict access of systemic factors to NSCs. Proliferation of NSCs and neural progenitors is tightly coupled to SGZ angiogenesis, and proliferating ECs and neural precursors colocalize in the niche (Van der Borght et al., 2009). Stimuli like exercise increase hippocampal neurogenesis and angiogenesis by upregulating VEGF in this niche. However, besides direct neurogenic and angiogenic effects of VEGF, expansion of the vascular niche alone can also contribute, ROCK inhibitor since persistent vascular expansion in the SGZ induced by transient overexpression of VEGF increases neurogenesis even after cessation of VEGF expression (Licht et al., 2011).

Vessels provide a substrate for guidance of migrating neuroblasts in adult neurogenesis and facilitate long-range migration of neuroblasts out of the SEZ toward the olfactory bulb (OB) along the rostral migratory stream (RMS) (Figure 4C) (Saghatelyan, 2009). RMS vessels are aligned parallel to the route of neuroblast migration, and nearly all migrating cells are attached to vessels (Snapyan et al., 2009). ECs attract the neuroblasts by releasing BDNF; once attracted, neuroblasts release GABA, unless which triggers nearby astrocytes to take up BDNF, thereby ensuring navigation along RMS vessels. VEGF also regulates neuroblast migration along the RMS (Wittko et al., 2009). In acute brain insults, hypoxia triggers a neurovascular response that results in increased angiogenic and neurogenic activity at the border of the lesion. This adaptive response can last for months and is associated with functional recovery (Jin et al., 2010). This regenerative response relies on reciprocal neurovascular interactions. Indeed, after stroke, neuroblasts deviate from the RMS and are attracted to the growing vasculature by SDF-1α and Ang1 in the peri-infarct cortex (penumbra), where they start neurogenesis (Saghatelyan, 2009).

The available results suggest that this is the case, but more pre

The available results suggest that this is the case, but more precise experiments are needed. Recently, optogenetic methods have

been used to show that odor recognition can be disrupted by selectively interfering with information processes at particular phases of the sniff cycle (Smear et al., 2011). If the hypothesis we are proposing is correct, disrupting information at a particular theta phase should affect information represented at that phase, but not information represented at other phases. Theta is critical for the transmission of multi-item messages because it provides a phase reference that signifies the onset of the message. This phase reference must be shared by sender and receiver; the high observed theta coherence between communicating regions appears to satisfy this requirement. The role of gamma is to define an item in a multi-item Dabrafenib message. Selleck SRT1720 Gamma contributes to this in three ways: (1) it helps to form the message by allowing only the most excited cells to fire, (2) it synchronizes

spikes (clustered spiking can be effectively detected in downstream regions), and (3) it creates pauses between items that prevent errors in decoding the message. The communication of the multipart messages to downstream networks may be aided by coherence in the gamma band, but this is probably not required. We suspect that the small increases in gamma coherence that occur during communication are probably a result of effective communication rather than the cause. Because gamma cycles are not of the same duration, detection methods based on exact clocking are not plausible. Thus, although phase-dependent detectors ( Jensen, 2001) can be used to detect early versus late items, detection of the information in a specific gamma subcycle does not

appear possible. However, many useful functions do not require exact clocking. For instance, according to one model ( Fukai, 1999), the sequence of actions to be executed is sent from the hippocampus to PAK6 the striatum by a theta-gamma code; the striatum stores this sequence and then executes the actions in order, using other information to orchestrate the exact timing of each action. Another useful operation would be the recall of a sequence that contained a salient element such as reward. The detection of this element could be important to downstream networks even if the exact position of that element in the recalled sequence was uncertain. Finally, the entire recalled sequence may be processed (chunked) to represent a higher-level item. Network models that perform such chunking depend on the ordering of items rather than on exact timing (H. Sanders, B. Kolterman, D. Rinberg, A. Koulakov, and J. Lisman, 2012, Soc. Neurosci., abstract). As described above, when the hippocampus communicates with target regions, the theta in the two regions becomes high. Virtually nothing is known about how this coherence is produced.

Our results are consistent with those from previous fMRI

Our results are consistent with those from previous fMRI

experiments (Buracas and Boynton, 2007 and Murray, 2008) reporting additive offsets with attention as well as a voltage-sensitive dye experiment that reached a similar conclusion about selection (E. Seidemann, personal communication). Equal increases in responses at all contrasts may result when responses are averaged across populations of neurons for at least two reasons. First, if some neurons show enhancement primarily at low and intermediate contrasts (contrast-gain like changes) and other neurons show enhancement primarily at high contrasts (response-gain like changes), then the overall sum of activity (and, consequently, any population readout that depends on this sum) would be check details expected to show enhancement at all contrasts (i.e., an additive offset). Indeed, an electrophysiological study has reported that some Selleckchem PD-332991 neurons exhibit contrast-gain, others response-gain, and yet

others exhibit additive changes in the same experiment (Williford and Maunsell, 2006). Moreover, the normalization model of attention (Reynolds and Heeger, 2009) can yield contrast-gain or response-gain like changes in different neurons dependent on the locations and sizes of their receptive fields. These effects in individual neurons can appear as an additive offset change when averaged across neurons (unpublished simulations). Second, the majority

of single-unit electrophysiology experiments used stimulus parameters that were matched to the tuning properties of the individual units being recorded. But in fact, any either stimulus that is the target of attention will give rise to activity in many neurons whose receptive fields and tuning properties may only partially match with the stimulus. Small baseline shifts with attention (Luck et al., 1997, Reynolds et al., 2000 and Williford and Maunsell, 2006) in each of many neurons may sum to a large effect in the overall population output, evident in the fMRI responses. The behavioral performance improvements with attention may, for some stimuli and tasks, depend primarily on this component of the population responses that is correlated across neurons (not the response- and/or contrast-gain changes evident in each individual neuron’s responses). Our max-pooling selection rule exemplifies how such a baseline shift can lead to improved behavioral performance. Hence, it is possible to reconcile the attentional modulation effects that have been measured with fMRI with those measured electrophysiologically.