These results suggest a facilitatory effect of microstimulation o

These results suggest a facilitatory effect of microstimulation on contraversive saccades. In contrast, when delivered before saccade onset at “blindly” sampled sites, caudate microstimulation increases RT for contraversive saccades and, to a lesser extent,

decreases RT for ipsiversive saccades on a pro-/antisaccade task (Watanabe and Munoz, 2010, 2011). These results suggest a suppressive effect of microstimulation on contraversive saccades. In light of our observations, these previous reports may have resulted from Selleckchem PI3K Inhibitor Library differential activation of distinct functional groups of neurons. More specifically, microstimulation that preferentially activates neurons participating in saccade generation facilitates generation of contraversive saccades. In contrast, microstimulation that preferentially

activates neurons participating in perceptual-decision formation or other cognitively demanding forms of saccade selection facilitates selection of ipsilateral saccade targets. The former effect dominates for evoked saccades and for simple saccade tasks with targeted microstimulation sites. Both effects are in place for pro-/antisaccade tasks with blindly sampled microstimulation sites and for the dots task. The dots task enables the dissociation of perceptual decision-making and saccade effects, with manipulations of stimulus strength (Petrov et al., 2011). In contrast to the microstimulation find more effects on choice bias, we did not observe a consistent effect on discrimination threshold. This result is consistent with our interpretation of caudate

Tolmetin response properties in the context of the DDM (Ding and Gold, 2010). According to that framework, discrimination threshold is determined by the decision bounds and a constant of proportionality used to convert the evidence to a log likelihood ratio-related quantity (Gold and Shadlen, 2002; Ratcliff, 1978). The decision bounds govern the speed-accuracy tradeoff and in our previous study were not encoded in caudate: unlike in LIP and FEF, evidence-accumulation activity in caudate did not converge at a DDM-like bound just prior to saccade onset on the RT dots task (Ding and Gold, 2010, 2012; Roitman and Shadlen, 2002). The constant of proportionality may already be incorporated in the inputs from MT and thus not influenced by caudate microstimulation. However, despite this consistency with our previous recording study, the lack of an effect on discrimination threshold is not consistent with previous computational modeling and fMRI studies that posit a role for the basal ganglia pathway in mediating the appropriate speed-accuracy tradeoff (Bogacz et al., 2010; Brown et al., 2004; Forstmann et al., 2008; Frank, 2006; Gurney et al., 2004; Lo and Wang, 2006; Rao, 2010; van Veen et al., 2008). This discrepancy might reflect a sampling bias in the present study favoring sites with the kind of task-modulated neural activity we described previously.

We utilized a two-virus system, in combination with Cre-expressin

We utilized a two-virus system, in combination with Cre-expressing mouse lines (Gong et al., 2007), to target genetically specified projection neuron subtypes in the striatum and specifically label their monosynaptic inputs (Haubensak et al., 2010 and Wall et al., 2010). The first

virus is a Cre-dependent adenoassociated virus (AAV) that expresses TVA and rabies glycoprotein; these proteins are necessary for infection and monosynaptic spread of a modified rabies virus, respectively. The second virus is a monosynaptic rabies virus that has been modified in two ways: first, the native rabies glycoprotein in the viral membrane has been replaced with an avian sarcoma leucosis virus envelope protein (EnvA), preventing infection of AZD6738 mammalian neurons in the absence of its binding partner, TVA. Second, the glycoprotein gene from the rabies virus genome has been deleted, preventing new particles from spreading retrogradely in the absence of another source of glycoprotein. Once TVA from the AAV is expressed in Cre+ neurons, the rabies virus

specifically infects these cells. Since the Cre-dependent AAV provides Cre+ cells with a source of rabies glycoprotein, newly formed rabies virus particles can spread retrogradely from these Cre+ cells to their directly connected inputs. These input cells do not contain Cre (and thus

do not express Selleckchem Rapamycin TVA or rabies glycoprotein), preventing the rabies virus from spreading beyond this step. This technique L-NAME HCl effectively restricts rabies virus infection to only Cre+ cells and their direct, monosynaptic inputs. We injected either D1R-Cre mice, D2R-Cre mice, or wild-type C57 control mice with 180 nl of helper virus (Figure 1A), followed 3 weeks later with 180 nl of modified rabies virus injected at the same location, but along a different injection tract (Figure 1B), to avoid potential double-labeling of dopamine receptor-expressing cells along the injection tract. We then waited one week for the rabies virus to replicate and spread monosynaptically before tissue processing and analysis (Figure 1C). We mounted every second section and stained against dsRed to amplify mCherry expression from the rabies virus, and counterstained with a fluorescent Nissl marker (Neurotrace 500/525). We then scanned each slide on a semiautomatic fluorescence slide scanner and counted labeled somata to determine the numbers of retrogradely labeled cells in each brain region. Mice with fewer than 50 input cells originating outside of striatum were excluded from analysis to prevent small number bias, yielding a final data set comprising inputs from 9 D1R-Cre mice and 10 D2R-Cre mice.

We next examined whether depleting ApNRX also blocks the 5-HT-ind

We next examined whether depleting ApNRX also blocks the 5-HT-induced synaptic growth that accompanies LTF. Again, we found a significant decrease in the number of new presynaptic varicosities when presynaptic ApNRX was downregulated by the injection of antisense

oligonucleotides (Figures 6C and 6D; % increase in varicosity numbers: no injection –10.4 ± 4.8, n = 8; antisense alone –12.2 ± 3.8, n = 5; sense alone −6.8 ± 6.3, n = 4; 5-HT 35.2 ± 8.2, n = 13; 5-HT+antisense 0.6 ± 5.0, n = 17, p < 0.01 versus 5-HT; 5-HT+sense 36.8 ± 7.8, n = 9). We have hypothesized that the cytoplasmic tail of ApNRX is necessary to recruit new molecular components important for the learning-related C59 wnt mouse assembly of the presynaptic active zone and its consequent 5-HT-induced remodeling and growth associated with LTF at the Aplysia sensory-to-motor

neuron synapse. To test this idea, we generated a C-terminal deletion construct of ApNRX (ApNRXΔC) that lacks the cytoplasmic tail. We expressed this construct in the sensory neurons and found that there was no obvious Ivacaftor in vitro difference in the expression of ApNRXΔC-GFP compared to ApNRX-GFP (data not shown). To bring the expression level to a steady state, we waited two days after injecting ApNRX-GFP or ApNRXΔC-GFP into presynaptic sensory neurons and treated sensory-to-motor neuron cocultures with five pulses of 5 min of 5-HT (10 μM) and measured EPSPs before and 24 hr after 5-HT treatment. We found that the overexpression of ApNRXΔC in the presynaptic sensory neuron making functional connections with the postsynaptic motor neuron in culture leads to a significant reduction of LTF at 24 hr, but that overexpression of wild-type ApNRX did not enhance LTF ( Figure 6E;

% initial EPSP amplitude: 5-HT 70.4 ± 9.4, n = 51; 5-HT + ApNRXΔC overexpression 3.9 ± 5.6, n = 14, p < 0.001 versus 5-HT; 5-HT + ApNRX overexpression 71.8 ± 13.7, n = 10). Overexpression of wild-type ApNRX PD184352 (CI-1040) or ApNRXΔC had no effect on basal transmission (% initial EPSP amplitude: no expression −12.2 ± 4.2, n = 28; ApNRXΔC overexpression alone –12.4 ± 11.6, n = 5; ApNRX overexpression alone −5.5 ± 13.2, n = 6). Unlike its effect on LTF, ApNRXΔC overexpression had no effect on STF induced by one pulse of 5-HT (10 μM) ( Figure 6F; % initial EPSP amplitude: no injection –8.5 ± 4.0, n = 19; ApNRXΔC overexpression alone 1.5 ± 5.7, n = 4; ApNRX overexpression alone 7.3 ± 12.8, n = 6; 5-HT 61.0 ± 7.3, n = 26; 5-HT + ApNRXΔC overexpression 77.0 ± 12.0, n = 9; 5-HT + ApNRX overexpression 78.9 ± 16.4, n = 11). These results with ApNRXΔC provide additional support for the notion that ApNRX is an important regulatory component of long-term memory storage in Aplysia perhaps via intracellular signaling cascades mediated by the cytoplasmic domain.

016) and between the ball and chair conditions (p = 0 015), respe

016) and between the ball and chair conditions (p = 0.015), respectively. Similarly, there were significant differences in the left foot COP speed in the AP direction between the ball and air-cushion conditions (p = 0.019) and between the ball and chair conditions (p = 0.028), respectively. The purpose of this study was to determine if active sitting would result in increased trunk motion and alterations of foot COP. Three sitting surfaces were introduced in this study: stability ball, air-cushion, and a hard surface (chair). Subjects performed a 30-min sitting on each of the sitting surfaces. Trunk motion and foot COP data were

collected and analyzed. Our findings indicate that the average T_COM and T_AVEL significantly increased with increased NVP-AUY922 order seating surface compliance. In addition, there were differences in the average speeds of the right and left foot COP in the AP direction between the ball and air cushion conditions and the ball and chair conditions.

We had hypothesized Roxadustat cell line that average T_COMs and T_AVELs would be influenced by sitting compliance. This hypothesis was supported. We found there were greater average T_COMs in the AP and longitudinal directions of spinal motion as surface compliance increased. There was also increased T_AVEL associated with the ball condition around the AP and longitudinal axes. This finding is in agreement with previous research,6 which reported that sitting on unstable surfaces induces greater spinal motion. It was reported that hypomobility of spine TCL due to a lack of mechanical stimulus yields adaptive changes that are related to reduced nutrient transport.5 There is a strong correlation between reduced or disrupted disc nutrition and occurrence of disc degeneration.14 Thus, increasing T_COM through active sitting may help prevent spinal hypomodiblity

and improve spine health. It is worth noting that the air-cushion condition resulted in greater trunk motion in the ML direction than the chair condition. It is possible that the subtle trunk motion in the ML direction during active sitting using an air-cushion could introduce dynamic mechanic-stimulus to lateral aspects of vertebrates. The potential risk of prolonged asymmetric intervertebral disc compression could be offset and the risk of disc herniation could be lowered. We had hypothesized that sitting on a stability ball or an air-cushion would increase trunk T_ANG range of motion. This hypothesis was not supported. There were no significant differences in T_ANG among the three sitting conditions. As the subjects were required to focus on a TV screen during the testing, it was essential to maintain a stable upper body during sitting so that the video viewing task would not be interfered. In this study, subjects were able to maintain an upright trunk position without experiencing increased range of motion when sitting on unstable surfaces.

The blockade of KV channels transformed this decremental pattern

The blockade of KV channels transformed this decremental pattern of trunk spike invasion (Figures 5F–5I). Direct

electrical recording revealed that KV channel blockade decreased the threshold current required to initiate apical dendritic trunk spikes and allowed these spikes to propagate with little decrement into the tuft (25 μM quinidine; n = 30; Figures 5F, 5G, and 6D). Furthermore, quinidine (25 μM), barium (20–50 μM), and the IA channel blocker 4-AP (3 mM) dramatically enhanced trunk spike invasion into terminal tuft branches as assessed by Ca2+ imaging (3°–5° branches; distance from nexus = 313 ± 14 μm; Figures 5H and 5I). In this set of experiments, we carefully adjusted the amplitude and/or time course of positive current steps used to evoke dendritic trunk spikes, to generate spikes of amplitude, duration, and Ca2+ signaling similar Bcl-2 expression to those recorded under control conditions at the nexus site of generation (Figure S7). We next explored

how KV channels shape the forward propagation of voltage from tuft sites to the nexus. Quinidine (25 μM) did not alter the intense distance-dependent attenuation of subthreshold voltage responses in the tuft (n = 30; Figures 6A and 6B). In contrast, quinidine reduced the threshold current required for the initiation of both tuft and trunk spikes (Figures 6C and 6D) and converted short-duration tuft-generated Na+ spikes into sustained local plateau potentials S3I-201 (Figures 6C and 6E). Similarly, quinidine and barium (50 μM) significantly enhanced both the peak amplitude and area of tuft spikes generated by two-photon glutamate uncaging recorded at the nexus (quinidine: 349 ± 27 μm from nexus, n = 7; barium: 197 ± 39 μm, n = 5; Figures 6F and 6G). Taken together, these data indicate that KV channels regulate the spread of tuft regenerative activity. Interactions between

active integration compartments in pyramidal neurons facilitate correlation-based neuronal computations (Larkum et al., 2004, Larkum et al., 1999, Takahashi and Magee, 2009 and Williams, 2005), which we have shown to be exploited in L5B pyramidal neurons during behavior to produce an object localization signal (Xu et al., 2012). To investigate how KV channels shape all such interactive integration, we paired patterns of ongoing AP firing in L5B pyramidal neurons, evoked by injection of barrages of simulated EPSCs at the soma (Williams, 2005), with subthreshold apical dendritic trunk depolarization (also generated by simEPSCs; Figure 7A). Under control conditions the rate of AP firing was progressively increased by barrages of dendritic simEPSCs of increasing frequency, due to the recruitment of dendritic trunk electrogenesis (Larkum et al., 2004, Larkum et al., 1999 and Williams, 2005) (Figures 7A–7C).

The nature of the

CR evaluation, therefore, is “absolute

The nature of the

CR evaluation, therefore, is “absolute.” Determining if a person’s blood pressure is normal based on his/her systolic and diastolic pressures is a good example of a CR evaluation. When the measurement interest is on “the more (e.g., number of pull-ups a student can do), or less (e.g., how fast a student can finish a one-mile run/walk click here test), the better”, the NR evaluation is more appropriate. Constructing an NR evaluation is relatively easy as long as a large, current and representative sample of a population can be obtained and regularly updated. With such a sample, norms (e.g., percentiles and percentile ranks) can be computed and derived. There are, however, several major limitations often associated with the NR evaluation framework. First, it is difficult to update

norms regularly due to cost, time, and manpower constraints. As an example, the PPFA’s norms were based on the 1985 National School Population Fitness Survey and there have been no major national fitness studies in the USA since the 1980s. As a result, these outdated values likely do not reflect current norms (e.g., an 80th percentile from the 1980s may now be equivalent to the 90th percentile), but rather how the present values compare to the previous norms, making them inaccurate in its original evaluation framework and the key “percentage” information no longer exists. Second, the interpretation under the NR evaluation depends on the “normal” Selleckchem CX 5461 status of the reference population. The designations of “average” or “above average” have limited meaning if the majority of a population is not normal (e.g., obese, unfit or unhealthy). Third, the selection of a percentile associated with health outcome measures (e.g., 85th or 95th percentiles as the cutoff values for “overweight” or “obese”) is often arbitrary with little scientific foundation. It is likely that other percentiles (say 83th vs. 97th) may

be the more appropriate values for when connecting these cut-off values with outcome variables of interest (e.g., health outcomes such as metabolic syndrome). Fourth, the employment of the NR evaluation framework tends to reward children and youth who are already fit while potentially discouraging those who are not fit. If rewards are based on achieving the 85th percentile (as with the PPFA) only highly fit youth may be motivated to try to achieve it. Less fit youth may be less motivated because they know their chances of achieving the standard are very low. If unfit students are less motivated during physical fitness testing they may come to perceive physical education classes, especially physical fitness testing, as a punitive, rather than enjoyable. The problem of the “17% in the 95th percentile” statement noted earlier is a good example of the first three limitations of the NR evaluation.

This simultaneous

encoding of alternative competing motor

This simultaneous

encoding of alternative competing motor goals is also fundamentally different from the representation of two sequential movement goals. Previous experiments showed that in the parietal cortex, during the planning of a multicomponent (double-step) movement, two neural populations GSK 3 inhibitor were activated, each of which was selective for one of the single movement components (Medendorp et al., 2006 and Baldauf et al., 2008). Double-step experiments do not induce a decision process between mutually exclusive action goals, and rather suggest that multiple components of a complex movement can be planned at once. Our finding of simultaneous encoding of alternative competing motor goals does complement previous observations in effector-selection experiments, which showed that alternative eye or hand movements to the same spatial target, instructed (Calton et al., 2002) or freely chosen (Cui and Andersen, 2007), can elicit simultaneous movement planning activity in LIP and PRR. The advantage of the goal-selection scheme over the rule-selection scheme for decision

making could be that—by computing all associated motor goal alternatives and their implicit action plans during the ambiguous state of planning—a more comprehensive cost-benefit calculation of each choice can be achieved. When the striker in our introductory example has to decide between aiming for the position of the goal keeper versus the opposite corner, then it is not enough to consider the likelihood of the Panobinostat in vitro keeper to jump or stay. Also the costs associated with the striker’s action alternatives are relevant, e.g., the striker might be poor at aiming for right-side goals, or the ball might be in an immediate position that eases aiming for one corner but not the other. Our results imply that the decision process in our rule-selection experiment selected between competing motor-goal alternatives, not between

different transformation rules or target stimuli, and that this competition likely happened in the sensorimotor areas that are involved in planning the respective movements. Note, we do note rule out the possibility that in Levetiracetam parallel a competition between the two potential rules takes place in rule-encoding frontal cortical areas (White and Wise, 1999, Wallis et al., 2001, Wallis and Miller, 2003 and Genovesio et al., 2005). The rule-competition could then, in the extreme case, just be mirrored by probabilistic motor goal representations in downstream sensorimotor areas. Because of the observed response normalization in our data (see below), we believe that if at all there was a rule-competition in our task then it was paralleled by a goal-competition in the sensorimotor areas, which would make sense for economical reasons, as discussed in the previous paragraph (Cisek and Kalaska, 2010).

Before transection, the IA response was scattered across the late

Before transection, the IA response was scattered across the lateral horn (Figure 2B2). After transection, IA response appeared most intense in the ventral lateral horn near the lateral horn entry site of vlpr dendrites (Figure 2B3, white arrow). This change of spatial pattern was evident when we superimposed the IA response before and after transection on the same lateral horn (Figure 2D1). By contrast, the spatial patterns of IA response in the control hemisphere appeared similar before and after mACT

transection (compare Figures 2B1 LY294002 in vivo and 2B4; Figure 2D2). We used two approaches to quantitatively analyze the changes of IA response before and after mACT transection. In the first approach, we defined a region of interest (ROI) based on the spatial pattern of the after-transection IA response for each imaging plane (see Supplemental Experimental Procedures). In the control hemisphere, this ROI encompasses the activated regions of both iPNs and vlpr neurons. In the experimental hemisphere, however, this

ROI would correspond to activated regions of vlpr neurons only, since iPN input was eliminated after mACT transection. We then quantified ΔF/F signals within the ROI for the IA responses before and MAPK Inhibitor Library supplier after transection. In the experimental hemisphere, the after-transection response was significantly increased compared to that before transection (Figure 2E1), suggesting that most after-transection responses in the ROI (i.e., vlpr neuronal responses) were newly gained as a consequence of mACT transection. This difference Histone demethylase was highly significant across individual flies (Figure 2F1). To rule out the contribution

of olfactory adaptation or potential nonspecific deterioration of fly physiology during the imaging procedure, we used the lateral horn IA response in the control hemisphere from the same fly as an internal control. The magnitude of the IA response in the lateral horn remained unchanged in the example fly (Figure 2E2). Although across flies there was a slight increase in the control hemisphere after transection compared with before (Figure 2F2; see Supplemental Experimental Procedures for a likely cause), when we used calibrated responses (IA responses within ROI of the experimental hemisphere divided by that of the control hemisphere from the same fly), IA response increase was highly significant across individual flies after mACT transection (Figure 2G). In the second approach, we analyzed the correlation of spatial patterns of IA response before and after mACT transection (see Supplemental Experimental Procedures). The control hemisphere showed a high correlation (Figure 2H, right column), consistent with the resemblance of their spatial activity patterns before and after transection. By contrast, the experimental hemisphere exhibited a significantly smaller correlation coefficient (Figure 2H, left column) compared to the control hemisphere.

, 1996 and Jankowska et al , 1979) Consistent with this idea, PS

, 1996 and Jankowska et al., 1979). Consistent with this idea, PSDCs also receive inputs from nonprimary sensory neuron sources, which include GABA and glycinergic interneurons as well as inputs from corticospinal and spinocervical tracts,

providing opportunities for presynaptic and postsynaptic modulation of LTMR inputs onto PSDCs (Bannatyne et al., 1987, Maxwell, 1988 and Maxwell Caspase pathway et al., 1995). Therefore, we speculate that PSDC output neurons are main carriers of integrated information emanating from both glabrous and hairy skin and pertaining to a variety of stimulus modalities. While PSDC neurons respond to a wide variety of sensory stimuli, SCT projection neurons are mainly concerned with hair follicle movement and therefore represent a main dorsal horn output for hairy skin innervating LTMRs. Nearly everything that we know about CP-673451 research buy the morphological and physiological characteristics of SCT neurons come from studies performed in the cat. In comparison to PSDC neurons, we know considerably more about the physiological properties of SCT neurons, due in part to the fact that SCT neuron somata are larger and therefore easier to identify and record. Like PSDC neurons, SCT neurons can also be easily identified in physiological

recording experiments by antidromic activation of their axonal tracts, in this case, the dorsal lateral funiculus or the LCN (Taub and Bishop, 1965). SCT neurons respond maximally to hair follicle deflection, with a single impulse in a hair follicle afferent capable of evoking a large excitatory postsynaptic potential. Furthermore, SCT response properties are similar to primary hair follicle afferents, suggesting direct

excitatory inputs from hairy skin LTMRs (Brown et al., 1987). Unlike PSDCs, SCT neurons do not receive SA-LTMR input from hairy skin, any LTMR input from glabrous skin, or Pacinian corpuscle (RAII-LTMR) inputs (Brown, 1981b and Hongo and Koike, 1975). Based on their response Vasopressin Receptor properties to electrical and natural stimulations, SCT neurons can be categorized into three main groups: low-threshold, wide-dynamic range, and high-threshold SCT neurons, presumably reflecting the types of LTMR inputs that they receive. Low-threshold SCTs make up 30% of the total population and are excited solely by hair movement. Wide-dynamic range SCT neurons respond to both hair movement as well as pressure or pinch stimuli and receive inputs from axons with varied conduction velocities. This subgroup represents about 70% of the total SCT population and it is thought to receive monosynaptic input from both hairy skin Aβ- as well as Aδ-LTMRs.

3% ± 3 8% of the wild-type cells but only in 25 3% ± 5 4% of AC K

3% ± 3.8% of the wild-type cells but only in 25.3% ± 5.4% of AC KO neurons (p < 0.01; Figures 3I–3J). Consistently, rhodamine-kabiramide, which binds directly to the barbed ends of actin filaments ( Petchprayoon et al., 2005), localized in a distal-proximal gradient in wild-type cells but displayed dispersed staining throughout SB203580 manufacturer the soma of the KO neurons ( Figures 3K

and 3L). Taken together, these data indicate that the barbed end orientation toward the leading edge is disrupted in the AC KO neurons. Despite the abnormal barbed end distribution, Abi, an essential component of the WAVE complex, showed strong staining at the membrane of AC KO neurons (data not shown), indicating that WAVE-Arp2/3-mediated actin nucleation can occur at the appropriate location in the absence of AC. High-resolution electron microscopy tomography revealed that in wild-type stage 1 neurons, actin filaments were radially oriented in tight bundles in filopodia or in a meshwork of filaments largely oriented toward the cell edge of lamellipodial veils (Figures 4A–4C, Movie S2). In contrast, AC KO neurons

had a disorganized dense actin filament network with a large population of individual filaments oriented circumferentially, parallel to the cell edge. Furthermore, there was no actin bundling observed selleck inhibitor in AC KO neurons ( Figures 4A–4C, Movie S3). Thus, AC proteins not only regulate the F-actin quantity, but also the

STK38 configuration of the neuronal actin network. The increased levels of F-actin and disordered filaments in AC KO neurons suggested a defect in actin turnover. Therefore, we examined actin dynamics by live-cell imaging using Lifeact-GFP. Stage 1 wild-type neurons had a very dynamic actin network, forming and retracting filopodia within minutes ( Figure 5A, Movie S4). In contrast, AC KO neurons had an immobile actin network ( Figure 5B, Movie S4). Kymograph analysis showed that while wild-type neurons had an average retrograde flow rate of 4.46 ± 0.97 μm/min, AC KO neurons had an average rate of 0.14 ± 0.4 μm/min, a more than 30-fold reduction (p < 0.001; Figures 5A–5C). There was a similar reduction in protrusion frequency and distance in AC KO neurons compared to wild-type neurons ( Figures 5D and 5E). Consistently, photobleaching GFP-actin in the peripheral actin network of AC KO neurons showed an over 40-fold higher half-fluorescence recovery (t1/2) compared to wild-type neurons ( Figures 5F and 5G). Having found such severe changes in actin organization and dynamics, we wanted to test whether AC KO cells could recover and form normal actin structures. To this end, we altered Cofilin expression levels in a temporally controlled manner by fusing Cofilin to a destabilization domain (DD), which targets proteins for rapid proteasome-mediated degradation ( Banaszynski et al., 2006).