07.01.2015 Views

silver 2003.pdf - Alice

silver 2003.pdf - Alice

silver 2003.pdf - Alice

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

R EPORTS<br />

High-Probability Uniquantal<br />

Transmission at Excitatory<br />

Synapses in Barrel Cortex<br />

R. Angus Silver, 1 *† Joachim Lübke, 2 Bert Sakmann, 3<br />

Dirk Feldmeyer 3 *<br />

The number of vesicles released at excitatory synapses and the number of<br />

release sites per synaptic connection are key determinants of information<br />

processing in the cortex, yet they remain uncertain. Here we show that the<br />

number of functional release sites and the number of anatomically identified<br />

synaptic contacts are equal at connections between spiny stellate and<br />

pyramidal cells in rat barrel cortex. Moreover, our results indicate that the<br />

amount of transmitter released per synaptic contact is independent of<br />

release probability and the intrinsic release probability is high. These properties<br />

suggest that connections between layer 4and layer 2/3 are tuned for<br />

reliable transmission of spatially distributed, timing-based signals.<br />

Synaptic weight is thought to be a primary<br />

determinant of neural computation. According<br />

to the quantal hypothesis (1, 2) the efficacy<br />

of a connection is determined by the<br />

product of the probability of release, the number<br />

of release sites, and the size of the<br />

postsynaptic response to a quantum of transmitter.<br />

Quantal parameters determine how<br />

reliably information is transmitted (3, 4),<br />

with higher rates of information transfer possible<br />

at connections with larger numbers of<br />

release sites and higher (sustained) release<br />

probabilities (3). However, basic synaptic<br />

properties, such as number of functional and<br />

anatomical release sites per connection and<br />

the release probability per site, remain uncertain<br />

at most cortical synapses. Moreover, the<br />

number of vesicles released at individual synaptic<br />

contacts is controversial. Correlation of<br />

functional release sites and anatomically<br />

1<br />

Department of Physiology, University College London,<br />

Gower Street, London WC1E 6BT, U.K. 2 Forschungszentrum<br />

Jülich GmbH, Institut für Medizin,<br />

D-52425 Jülich, Germany. 3 Max-Planck Institut für<br />

Medizinische Forschung, Abteilung Zellphysiologie,<br />

Jahnstrasse 29, D-69120 Heidelberg, Germany.<br />

*These authors contributed equally to the work.<br />

†To whom correspondence should be addressed. E-<br />

mail: a.<strong>silver</strong>@ucl.ac.uk<br />

Fig. 1. EPSP recordings from an L4 to L2/3 cell pair in rat barrel cortex under different release<br />

probability conditions at 36°C. (A) Current protocols for AP stimulation in the presynaptic<br />

spiny stellate cell (top trace) and for testing input resistance stability in postsynaptic pyramidal<br />

cell (second trace). Third trace shows the AP in the spiny stellate cell. The bottom trace shows<br />

voltage responses to current injection followed by individual EPSPs (gray lines) and the mean<br />

EPSP (black line) in the postsynaptic pyramidal cell. (B) Mean EPSPs recorded in different<br />

extracellular Ca 2 and Mg 2 concentrations at a potential of –73 mV. (C) Input resistance<br />

during different extracellular Ca 2 and Mg 2 concentrations.<br />

Fig. 2. Quantal parameters estimated<br />

from evoked EPSPs under low probability<br />

conditions. (A) The top trace<br />

shows the mean action potential<br />

evoked by a brief depolarizing current<br />

pulse at –72 mV; the bottom trace<br />

shows 14 individual EPSPs (gray<br />

lines), the mean of the 44 EPSP successes<br />

(black solid line), and the mean<br />

of the 279 failures (dashed line) recorded<br />

at –72 mV in 1 mM [Ca 2 ]<br />

and 5 mM [Mg 2 ]. (B) The amplitude<br />

histogram of EPSPs (bins) and scaled<br />

baseline noise (dashed line). The coefficient<br />

of variation of the quantal<br />

EPSP amplitude (CV Q<br />

)was calculated from the background-subtracted<br />

variance. (C) Relation between the variance of the EPSP peak<br />

amplitude (corrected for background variance) and mean peak EPSP<br />

amplitude for an individual L4 to L2/3 cell pair. Each data point shows<br />

a different probability condition. Error bars indicate the theoretical<br />

standard error in the estimate of the variance. Solid line shows the fit<br />

to a multinomial model giving Q 0.09 mV, N F<br />

5.25 and <br />

19800.<br />

www.sciencemag.org SCIENCE VOL 302 12 DECEMBER 2003 1981


R EPORTS<br />

identified synaptic contacts in inhibitory neurons<br />

of goldfish lead to the hypothesis that a<br />

maximum of one vesicle is released per active<br />

zone (“the single vesicle hypothesis”)<br />

(5). This hypothesis was subsequently supported<br />

at connections between hippocampal<br />

pyramidal cells and interneurons (6). However,<br />

recent studies indicate that the concentration<br />

of glutamate in the synaptic cleft changes<br />

with release probability, which implies that<br />

multiple vesicles are released at each synaptic<br />

contact per action potential (AP) (7, 8).<br />

We investigated the number of functional<br />

and anatomical release sites, the release probability,<br />

and the mode of release at synaptic<br />

connections between pairs of layer 4 (L4) and<br />

layer 2/3 (L2/3) excitatory neurons in rat<br />

barrel cortex. Paired whole-cell recordings<br />

were made from acute slices from 17- to<br />

23-day-old rats at near-physiological temperature<br />

(9, 10). In the presence of 2 mM [Ca 2 ]<br />

and N-methyl-D-aspartate (NMDA) receptor<br />

antagonists, single APs in L4 spiny neurons<br />

evoked unitary excitatory postsynaptic potentials<br />

(EPSPs) in L2/3 pyramidal cells with a<br />

mean peak amplitude of 0.54 0.06 mV at<br />

–70 mV (n 32 cells; Fig. 1A). When<br />

stimulated at low frequency (0.1 to 0.033<br />

Hz), EPSP amplitudes remained stable for<br />

long periods (2 to 3 hours). The amplitude<br />

of unitary EPSPs could be increased (in eight<br />

of nine cells) or reduced by altering the<br />

[Ca 2 ] and [Mg 2 ], which suggests that<br />

transmitter release probability was not maximal<br />

under our control conditions (Fig. 1B).<br />

Such alterations in the [Ca 2 ]/[Mg 2 ] ratio<br />

had little effect on the input resistance, and<br />

thus on the voltage response to a given synaptic<br />

conductance (Fig. 1C).<br />

We first determined the mean postsynaptic<br />

response to a single quantum of<br />

transmitter (Q) by evoking release under conditions<br />

of low release probability. In the presence<br />

of 1 mM extracellular Ca 2 , the proportion<br />

of presynaptic APs that failed to release<br />

transmitter was high (0.82 0.02; 0.8 to 1.0<br />

mM Ca 2 ; n 5), and thus, the vast majority<br />

of the EPSPs were single quantal events (Fig.<br />

2A) (10). Quantal EPSPs were identified on<br />

the basis of their characteristic shape, after<br />

each trace was digitally filtered. We checked<br />

that small-amplitude EPSPs were not missed<br />

by averaging the failures (Fig. 2A). On average,<br />

the mean amplitude of quantal EPSPs<br />

was 0.15 0.02 mV at –70 mV (n 5<br />

cells). The mean number of quanta contributing<br />

to the evoked EPSP (quantal content)<br />

under control conditions (2 mM Ca 2 ) was<br />

thus 3.6.<br />

The number of quantal events generated<br />

per synaptic contact can be determined by<br />

comparing the number of sites that can generate<br />

a quantal event (functional release sites,<br />

N F<br />

) and the number of anatomically identified<br />

synaptic contacts (N A<br />

). Because the<br />

quantal content does not provide information<br />

about N F<br />

or about the average release probability<br />

across these sites (P R<br />

), we estimated<br />

these quantal parameters by applying multiple-probability<br />

fluctuation analysis (MPFA)<br />

(11, 12), also known as variance-mean analysis<br />

(13). We first estimated the total nonuniformity<br />

in Q from amplitudes of quantal<br />

EPSPs under low P R<br />

conditions, using the<br />

same peak and baseline measurement windows<br />

used for MPFA (Fig. 2B) (10). This<br />

gave a coefficient of variation (standard deviation/mean;<br />

CV Q<br />

0.43 0.06; n 5<br />

cells) similar to that reported for other excitatory<br />

synapses (11, 14). The variability in<br />

quantal size at an individual release site (intrasite<br />

variability) has been measured directly<br />

at single site cerebellar synapses at physiological<br />

temperature (CV QI<br />

0.21) (15). If we<br />

assume the same value for intrasite quantal<br />

variability at L4 to L2/3 synapses (see verification<br />

below), the variation arising from<br />

differences in the mean quantal amplitude<br />

across sites (intersite variability) can be calculated<br />

(CV QII<br />

0.37) from the total quantal<br />

variability (10). These estimates of quantal<br />

Fig. 3. Relation between the number of anatomically<br />

identified synaptic contacts and functional<br />

release sites. (A) Camera lucida reconstruction of<br />

a cell pair, L4 stellate cell with L2/3 pyramidal<br />

cell. The soma and dendrites of the presynaptic L4<br />

cell are given in red, the axonal arbor in blue. The<br />

somatodendritic configuration of the postsynaptic<br />

L2/3 pyramidal cell is given in black, the axonal<br />

arbor in green. Lower inset shows the distribution<br />

of putative, light microscopically identified synaptic contacts. (B) Variance-mean plot and<br />

multinomial fit for the cell pair shown in (A), where Q and N F are the estimated quantal size<br />

and number of functional release sites, respectively. (C) Electron micrographs (EMs) of the<br />

synaptic contacts identified by light microscopy in (A), where the letters next to the blue dots<br />

in (A) indicate the position of each EM image shown in (C). Labeling in (C): b, presynaptic<br />

bouton with clearly visible synaptic vesicles; d, postsynaptic dendrites.<br />

1982<br />

12 DECEMBER 2003 VOL 302 SCIENCE www.sciencemag.org


R EPORTS<br />

variability were used to constrain a nonuniform<br />

quantal model, which allowed more<br />

accurate estimates of Q, N F<br />

, and P R<br />

.<br />

Time-stable epochs of EPSP amplitudes<br />

were identified for each release probability<br />

condition with a Spearman rank-order analysis<br />

(15). The mean and variance of the peak<br />

amplitude were calculated and corrected for<br />

deviations in driving force from –70 mV,<br />

with an assumed reversal potential for AMPA<br />

receptors of 0 mV (9). EPSP variance-mean<br />

plots were constructed after subtracting background<br />

variance and were fit with a multinomial<br />

model with fixed values of quantal variance<br />

(10); goodness of fit was assessed with<br />

the chi-square test. A variance-mean plot and<br />

multinomial fit are shown in Fig. 2C that<br />

gave a Q of 0.09 mV and an estimate of 5.3<br />

functional release sites for this cell pair. The<br />

mean Q varied little across cell pairs (0.14 <br />

0.02, n 9) and was not significantly different<br />

from that measured directly from quantal<br />

EPSPs under low P R<br />

conditions (P 0.55,<br />

unpaired t test, unequal variance). The N F<br />

across cells was 5.1 0.9 (n 9) and varied<br />

between 2.3 and 6.5 except for one outlying<br />

cell in which N F<br />

11.1 (the mean value<br />

excluding this cell was 4.4 0.5, n 8).<br />

Simulations suggest that part of the variance<br />

in N F<br />

for the eight clustered cells arose from<br />

uncertainty in estimating the variance from a<br />

limited number of EPSPs (10). Under control<br />

conditions, the P R<br />

at each site was high<br />

(0.79 0.04 in 2 mM Ca 2 , n 9), consistent<br />

with a previous study (16). An upper<br />

limit for the quantal variability at an individual<br />

site was calculated by dividing the variance<br />

remaining when P R<br />

was maximal<br />

(1.02 0.05, n 9) by N F<br />

. This gave a value<br />

of CV QI<br />

close to that assumed from measurements<br />

at single-site synapses (0.26 0.02<br />

versus 0.21) (15). Estimation of the coefficient<br />

of variation of release probability across<br />

sites gave a value (0.28 0.07 in 2 mM<br />

Ca 2 , n 9), similar to that obtained in<br />

hippocampal synapses in culture (17).<br />

Anatomical identification of axondendrite<br />

intersections with high-resolution<br />

light microscopy of biocytin-filled cell pairs<br />

(10) indicated between four and six putative<br />

synaptic contacts per connection located on<br />

the proximal dendritic segments of the<br />

postsynaptic neuron (blue dots, Fig. 3A, inset).<br />

Across cells, an average of 4.6 0.2<br />

(n 14) putative synaptic connections were<br />

found, a value not significantly different from<br />

Fig. 4. Fractional block of the EPSP by low-affinity glutamate antagonists is independent of release<br />

probability. (A) The upper trace shows the presynaptic APs in a L4 spiny neuron; middle traces show<br />

mean EPSPs recorded in 2 mM Ca 2 ,1mMMg 2 under control conditions and in 1 mM -DGG (red<br />

trace). The bottom trace shows mean EPSPs recorded in 1.25 mM Ca 2 under control conditions<br />

(blue) and in 1 mM -DGG (green). (B) Plot of EPSP amplitudes as a function of time through the<br />

experiment. (C) EPSPs in (A) normalized to the peak of the first control EPSPs. (D) Relation<br />

between percentage block of the EPSP and extracellular [Ca 2 ] for 0.5 mM kynurenic acid and<br />

1mM-DGG.<br />

the N F<br />

(P 0.58 unpaired t test, unequal<br />

variance). In four cell pairs, we were able to<br />

recover the anatomy after performing quantal<br />

analysis. A camera lucida reconstruction of a<br />

cell pair is shown in Fig. 3, for which we<br />

estimated an N F<br />

of 5.5 and an N A<br />

of 5 (Fig. 3,<br />

A and B). For cell pairs in which both quantal<br />

analysis and anatomical reconstruction were<br />

performed, N F<br />

and N A<br />

were not significantly<br />

different (P 0.21, paired t test, n 4; table<br />

S1). We further examined whether these<br />

close appositions were indeed synaptic contacts<br />

by making serial ultrathin sections<br />

through the entire dendritic domain of the<br />

postsynaptic neuron and examining them at<br />

the electron microscopic (EM) level. Synaptic<br />

contacts were identified on the basis of a<br />

clear presynaptic bouton-like structure containing<br />

closely packed synaptic vesicles together<br />

with closely apposing (20 nm) preand<br />

postsynaptic membranes (Fig. 3C). The<br />

presence of electron-dense biocytin reaction<br />

product both pre- and postsynaptically and<br />

the long duration of the recordings precluded<br />

the use of other anatomical identifiers such as<br />

the pre- and postsynaptic densities. Full EM<br />

analysis of two quantal analysis pairs of neurons<br />

and one additional cell confirmed identification<br />

of 13 out of 14 putative synaptic<br />

contacts, which gave a detection rate close to<br />

that observed at other cortical synapses (18,<br />

19). Synaptic contacts were located directly<br />

on dendritic shafts (Fig. 3C) and occasionally<br />

on spines. Applying the observed optical detection<br />

rate for synaptic contacts (93%)<br />

across the four pairs of neurons, we estimated<br />

N F<br />

/N A<br />

to be 0.90 0.16, which suggests a<br />

one-to-one relation between these synaptic<br />

properties (table S1).<br />

The all-or-none behavior of excitatory synapses<br />

could arise either presynaptically, from the<br />

release of a single quantum per AP, or postsynaptically,<br />

from saturation of postsynaptic receptors<br />

(20, 21) following multivesicular release<br />

(22, 23). We therefore examined whether the<br />

cleft glutamate concentration changes with P R<br />

by assaying the level of block of the EPSPs by<br />

low-affinity glutamate receptor antagonists (7,<br />

12, 24). The effect of 1 mM -d-glutamylglycine<br />

(-DGG) on the mean EPSPs recorded from a<br />

cell pair in normal Ca 2 (2 mM) and in the<br />

presence of a low Ca 2 (1.25 mM) concentration<br />

is shown in Fig. 4, A and B. Normalization to the<br />

first EPSP amplitude illustrates that the fractional<br />

block of the EPSP by -DGG was similar<br />

under these two conditions (Fig. 4C). On average,<br />

the fractional block of the EPSP by -DGG<br />

was not significantly different in 2 mM Ca 2<br />

and 1.25 mM Ca 2 (66 6% and 60 5%,<br />

respectively; P 0.43; n 5). Moreover, the<br />

paired-pulse ratio in 2 mM Ca 2 was unchanged<br />

in the presence of -DGG (0.59 to 0.52, P <br />

0.36; n 10), which indicates a similar fractional<br />

block on each pulse. We tested these results<br />

further by carrying out a series of experiments<br />

www.sciencemag.org SCIENCE VOL 302 12 DECEMBER 2003 1983


R EPORTS<br />

using kynurenic acid, a different rapidly equilibrating<br />

antagonist. As for -DGG, the fractional<br />

block of the EPSP by kynurenic acid was not<br />

significantly different in 2 mM Ca 2 and 1.25<br />

mM Ca 2 (55 3% and 53 6%, respectively;<br />

P 0.61; n 4; Fig. 4D). Our results contrast<br />

with the substantial increase in the level of block<br />

of excitatory postsynaptic currents (EPSCs) by<br />

-DGG and kynurenic acid at the climbing fiber,<br />

as the release probability was lowered (7) from a<br />

similar initial value (11). The fact that rapidly<br />

equilibrating competitive antagonists are equally<br />

effective at blocking the EPSP when P R<br />

was<br />

reduced by a factor of 4 suggests that the concentration<br />

of transmitter in the synaptic cleft,<br />

following release, is independent of P R<br />

at L4 to<br />

L2/3 synapses. These results indicate that each<br />

excitatory synaptic contact releases a single<br />

quantum of glutamate (probably one vesicle but<br />

could be a group of vesicles whose number is<br />

independent of P R<br />

) and operates independently,<br />

in an all-or-none manner.<br />

The high P R<br />

at L4 to L2/3 synapses is well<br />

suited to promote reliable transmission of the<br />

one (76%) or two spikes that whisker stimulation<br />

usually evokes in L4 cells in vivo (25),<br />

because short-term depression becomes significant<br />

during longer bursts of L4 activity [Fig.<br />

4A and (9)]. The depression of high P R<br />

synapses,<br />

together with the precise synaptic latency<br />

between L4 and L2/3 (1 to 3 ms) and rapid<br />

EPSP kinetics (9), may also be important for<br />

maintaining and strengthening L4 to L2/3 synapses,<br />

because this connection exhibits a spiketiming–based<br />

plasticity rule with a narrow coincidence<br />

window (26). Curiously, the small<br />

numbers of release sites, all-or-none uniquantal<br />

signaling and small quantal size are likely to<br />

endow an individual L4 to L2/3 connection<br />

with a relatively low capacity for transmitting<br />

information. These properties contrast with<br />

synapses in the retina, where large numbers of<br />

release sites (200) and graded transmission<br />

allow high information transmission rates (27).<br />

Moreover, transmission is finely tuned because<br />

only 10% of the 300 to 400 spiny stellate cells<br />

that innervate an individual L2/3 pyramidal cell<br />

fire synchronously, within 15 ms of single<br />

whisker stimulation (25), close to the minimum<br />

number required for a pyramidal cell to reach<br />

voltage threshold (9). Our results suggest that<br />

the functional and anatomical properties of L4<br />

to L2/3 synaptic connections are well adapted<br />

for efficient, timing-based distributed signaling.<br />

References and Notes<br />

1. B. Katz, The Release of Neural Transmitter Substances<br />

(Liverpool Univ. Press, Liverpool, 1969).<br />

2. D. Vere-Jones, Aust. J. Stat. 8, 53 (1966).<br />

3. A. Zador, J. Neurophysiol. 79, 1219 (1998).<br />

4. M. London, A. Schreibman, M. Häusser, M. E. Larkum,<br />

I. Segev, Nature Neurosci. 5, 332 (2002).<br />

5. H. Korn, A. Triller, A. Mallet, D. S. Faber, Science 213,<br />

898 (1981).<br />

6. A. I. Gulyas et al., Nature 366, 683 (1993).<br />

7. J. I. Wadiche, C. E. Jahr, Neuron 32, 301 (2001).<br />

8. T. G. Oertner, B. L. Sabatini, E. A. Nimchinsky, K.<br />

Svoboda, Nature Neurosci 5, 657 (2002).<br />

9. D. Feldmeyer, J. Lübke, R. A. Silver, B. Sakmann,<br />

J. Physiol. (London) 538, 803 (2002).<br />

10. Methods, table S1, and references are available as<br />

supporting online material on Science Online.<br />

11. R. A. Silver, A. Momiyama, S. G. Cull-Candy, J. Physiol.<br />

(London) 510, 881 (1998).<br />

12. J. D. Clements, R. A. Silver, Trends Neurosci 23, 105 (2000).<br />

13. C. A. Reid, J. D. Clements, J. Physiol. (London) 518,<br />

121 (1999).<br />

14. J. M. Bekkers, J. D. Clements, J. Physiol. (London) 516,<br />

227 (1999).<br />

15. R. A. Silver, S. G. Cull-Candy, T. Takahashi, J. Physiol.<br />

(London) 494, 231 (1996).<br />

16. M. A. Castro-Alamancos, B. W. Connors, Proc. Natl.<br />

Acad. Sci. U.S.A. 94, 4161 (1997).<br />

17. V. N. Murthy, T. J. Sejnowski, C. F. Stevens, Neuron<br />

18, 599 (1997).<br />

18. H. Markram, J. Lübke, M. Frotscher, A. Roth, B. Sakmann,<br />

J. Physiol. (London) 500, 409 (1997).<br />

19. J. Lübke, A. Roth, D. Feldmeyer, B. Sakmann, Cereb.<br />

Cortex 13, 1051 (2003).<br />

20. J. J. Jack, S. J. Redman, K. Wong, J. Physiol. (London)<br />

321, 65 (1981).<br />

21. P. Jonas, G. Major, B. Sakmann, J. Physiol. (London)<br />

472, 615 (1993).<br />

22. R. A. Silver, in Central Synapses: Quantal Mechanisms<br />

and Plasticity, D. S. Faber, H. Korn, S. J. Redman, S. M.<br />

Thompson, and J. S. Altman, Eds., 4th Human Frontier<br />

Science Workshop, Strasbourg, France, 22 to 24 April<br />

1997 (Human Frontier Science Program, Strasbourg,<br />

France, 1998), pp. 130–139.<br />

23. K. A. Foster, A. C. Kreitzer, W. G. Regehr, Neuron 35,<br />

1115 (2002).<br />

24. J. S. Diamond, C. E. Jahr, J. Neurosci. 17, 4672 (1997).<br />

25. M. Brecht, B. Sakmann, J. Physiol. (London) 543, 49<br />

(2002).<br />

26. D. E. Feldman, Neuron 27, 45 (2000).<br />

27. S. B. Laughlin, R. R. de Ruyter van Steveninck, J. C.<br />

Anderson, Nature Neurosci. 1, 36 (1998).<br />

28. Supported by the Max-Planck Society and the Wellcome<br />

Trust. R.A.S. is in receipt of a Wellcome Trust<br />

Senior Fellowship. We thank A. Roth and C. Saviane<br />

for help with simulations; D. Attwell, D. DiGregorio,<br />

M. Farrant, S. Mitchell, T. Nielsen, and I. Vida for<br />

comments on the manuscript; and N. Nestel and B.<br />

Joch for technical assistance.<br />

Supporting Online Material<br />

www.sciencemag.org/cgi/content/full/302/5652/1981/DC1<br />

Materials and Methods<br />

Table S1<br />

References<br />

23 May 2003; accepted 15 October 2003<br />

Anterior-Posterior Guidance of<br />

Commissural Axons by<br />

Wnt-Frizzled Signaling<br />

Anna I. Lyuksyutova, 2 Chin-Chun Lu, 1 Nancy Milanesio, 1<br />

Leslie A. King, 2 Nini Guo, 4 Yanshu Wang, 4 Jeremy Nathans, 4<br />

Marc Tessier-Lavigne, 5 * Yimin Zou 1,2,3 †<br />

Commissural neurons in the mammalian dorsal spinal cord send axons ventrally<br />

toward the floor plate, where they cross the midline and turn anteriorly toward the<br />

brain; a gradient of chemoattractant(s) inside the spinal cord controls this turning.<br />

In rodents, several Wnt proteins stimulate the extension of commissural axons after<br />

midline crossing (postcrossing). We found that Wnt4 messenger RNA is expressed<br />

in a decreasing anterior-to-posterior gradient in the floor plate, and that a directed<br />

source of Wnt4protein attracted postcrossing commissural axons. Commissural<br />

axons in mice lacking the Wnt receptor Frizzled3 displayed anterior-posterior<br />

guidance defects after midline crossing. Thus, Wnt-Frizzled signaling guides commissural<br />

axons along the anterior-posterior axis of the spinal cord.<br />

Axonal connections are patterned along the<br />

anterior-posterior (A-P) and dorsal-ventral<br />

(D-V ) neuraxes. Guidance molecules that<br />

play essential roles in the D-V guidance of<br />

axons have been identified, whereas the nature<br />

of the A-P guidance cues has remained<br />

an enigma (1, 2). The dorsal spinal cord<br />

commissural neurons form several ascending<br />

somatosensory pathways. During embryonic<br />

development, they project axons to the ventral<br />

midline (Fig. 1A). At the floor plate,<br />

commissural axons cross the midline, enter<br />

the contralateral side of the spinal cord, and<br />

make a sharp anterior turn toward the brain<br />

(Fig. 1B) (3). The initial ventral growth of the<br />

commissural axons is directed by a collaboration<br />

of two chemoattractants, netrin-1 (4–<br />

6) and Sonic hedgehog (Shh) (7), and chemorepellents<br />

of the bone morphogenetic protein<br />

(BMP) family (8). As the axons cross the<br />

midline, they lose responsiveness to these<br />

chemoattractants (9) but gain responsiveness<br />

to several chemorepellents, including Slit and<br />

semaphorin proteins (10), which guide axons<br />

from the D-V axis into the A-P axis (10).<br />

To determine why axons turn in an anterior<br />

direction, we studied the turning of commissural<br />

axons after midline crossing in “open-book” spi-<br />

1<br />

Department of Neurobiology, Pharmacology and Physiology,<br />

2 Committee on Developmental Biology, 3 Committee<br />

on Neurobiology, University of Chicago, Chicago, IL 60637,<br />

USA. 4 Departments of Molecular Biology and Ophthalmology,<br />

Howard Hughes Medical Institute, Johns Hopkins<br />

Medical School, Baltimore, MD 21205, USA. 5 Department<br />

of Biological Sciences, Howard Hughes Medical Institute,<br />

Stanford University, Stanford, CA 94305, USA.<br />

*Present address: Genentech Inc., South San Francisco,<br />

CA 94080, USA.<br />

†To whom correspondence should be addressed. E-<br />

mail: yzou@bsd.uchicago.edu<br />

1984<br />

12 DECEMBER 2003 VOL 302 SCIENCE www.sciencemag.org

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!