BRES : 44732
pp: 1210ðcol:fig: : NILÞ
Model7 brain research ] (]]]]) ]]]–]]]
121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180
Available online at www.sciencedirect.com
www.elsevier.com/locate/brainres
Research Report
Q2
CNQX facilitates inhibitory synaptic transmission in rat hypoglossal nucleus
Q1
Lei Hana,1, Shuhua Mua,1, Zhendan Hea, Zhiwei Wangc, Junle Qud, Wencai Yeb,n, Jian Zhanga,nn a
School of Medicine, Shenzhen University, 518060 Shenzhen, PR China Guangdong Province Key Laboratory of Pharmacodynamic Constituents of TCM and New Drugs Research, Jinan University, 510632 Guangzhou, PR China c Department of Neurosurgery, Shenzhen Shekou People's Hospital, 518060 Shenzhen, PR China d College of Optoelectronics Engineering, Shenzhen University, 518060 Shenzhen, PR China b
art i cle i nfo
ab st rac t
Article history:
6-cyano-7-nitroquinoxaline-2, 3-dione (CNQX) is a most commonly used antagonist of α-
Accepted 11 February 2016
amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptor in the central nervous system. During the past two decades, studies had demonstrated that CNQX could
Keywords: CNQX
partially activate AMPA receptors that are located on the hippocampal and cerebellar interneurons, thus subsequently leading to the facilitation of inhibitory transmission. However, whether CNQX could enhance inhibitory synaptic transmission in the hypoglos-
AMPA Inhibitory transmission Patch-clamp Hypoglossal nucleus
sal nucleus remains elusive. Here, using whole-cell patch-clamp recording in the brainstem slice, we showed that CNQX greatly increased both frequency and amplitude of spontaneous inhibitory postsynaptic currents in the hypoglossal motoneurons, whereas D(-)-2-Amino-5-phosphonopentanoic acid (D-AP5), N-methyl-D-aspartate (NMDA) receptor antagonist, had no effect on inhibitory synaptic transmission. Application of bicuculline and strychnine further identified that CNQX not only increased GABAergic sIPSCs but also glycinergic one in these motoneurons. Similar enhancement of inhibitory transmission was observed with application of DNQX, a quinoxaline derivative of CNQX, but not with application of GYKI 53655, a non-competitive antagonist of AMPA receptor. In the presence of tetradotoxin, the effect of CNQX on sIPSCs was abolished, suggesting that an increase in
Q4
Abbreviations: CNQX, 6-cyano-7-nitroquinoxaline-2, 3-dione; ACSF,
artificial cerebrospinal fluid; AMPA,
methyl-4-isoxazolepropionic acid; D-AP5,
D-(-)-2-Amino-5-phosphonopentanoic acid; DNQX,
GABA,
hypoglossal motoneurons; NMDA,
gamma-aminobutyric acid; HMs,
α-amino-3-hydroxy-5-
6,7-dinitroquinoxaline-2,3-dione;
N-methyl-D-aspartate; TARP,
transmembrane AMPA
receptor regulatory protein; TTX, tetrodotoin n Q3 correspondence to: School of Medicine, Shenzhen University, Nanhai Ave 3688, 518060 Shenzhen, P.R. China.author. nn Corresponding author. E-mail addresses:
[email protected] ( W. Ye),
[email protected] ( J. Zhang). 1 These two authors contributed equally to this article. http://dx.doi.org/10.1016/j.brainres.2016.02.020 0006-8993/& 2016 Published by Elsevier B.V.
Please cite this article as: Han, L., et al., CNQX facilitates inhibitory synaptic transmission in rat hypoglossal nucleus. Brain Research (2016), http://dx.doi.org/10.1016/j.brainres.2016.02.020
181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240
BRES : 44732
2
241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300
brain research ] (]]]]) ]]]–]]]
presynaptic interneuron spike firing rate induced by CNQX was responsible for the facilitation of sIPSCs. Taken together, these results demonstrated that the excitatory effect of CNQX on presynaptic interneurons triggered enhancement of both GABAergic and glycinergic synaptic transmission within the rat hypoglossal nucleus. & 2016 Published by Elsevier B.V.
1.
Introduction
from the other brain regions (Marchetti et al., 2002; Rekling et al., 2000). These neurons received excitatory glutamatergic inputs
The hypoglossal nucleus, where the overwhelming majority of
from the primary motor cortex and local reticular formation as
neurons are cholinergic hypoglossal motoneurons (HMs) (Davidoff
well as inhibitory inputs from interneurons in the reticular
and Schulze, 1988; Viana et al., 1990), is suggested to provide a
formation adjacent to the hypoglossal nucleus (Li et al., 1996,
fundamental function in respiration, swallowing, suckling, and
1997). Most of these inhibitory inputs are glycinergic in nature,
mastication (Sawczuk and Mosier, 2001). By coupling with each
whereas the other inhibitory components are GABAergic (O'Brien
other through gap junction, HMs can directly innervate tongue
and Berger, 2001). Regulation of inhibitory synaptic transmission
muscles to accomplish the physiological functions through hypo-
can affect respiratory rhythm by altering HMs membrane potential
glossal nerve. Coordination in the activity of these coupled HMs can generate rhythmic signal to guide genioglossus muscle,
and changing the output pattern of HMs. Previously, studies had investigated the isolation of spontaneous
thereby is crucial for maintaining various functions in the upper
inhibitory neuronal activity within the central nervous system (CNS)
airway. The output (i.e., firing rates and burst patterns) of HMs can
in the presence of 6-cyano-7-nitroquinoxaline-2, 3-dione (CNQX),
be modulated by either excitatory or inhibitory inputs projected
an α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA)
Fig. 1 – CNQX-induced potentiation of inhibitory synaptic transmission in hypoglossal motoneurons. (A) Left panel: sample trace of sIPSCs from a HM before and after perfusion of CNQX (10 μM). Noted that CNQX exerted a robustly increased spontaneous events. Right panel: photomicrograph of a recorded HM in the left panel, as labeled with neurobiotin. (B) Cumulative probability plot of inter-event interval and amplitude of sIPSCs from a sample neuron. The inter-event interval curve significantly shifted leftward with application of CNQX, whereas the amplitude curve significantly shifted rightward with application of CNQX. (C) Bar graph shows the effect of CNQX on the frequency and amplitude of sIPSCs. (D) No detectable inward or outward basal current was found in HM treated with CNQX. (E) CNQX did not change membrane potential and input resistance of HM. Voltage deflections were responses to periodic current injections of 100 pA for 100 ms at 0.5 Hz. Please cite this article as: Han, L., et al., CNQX facilitates inhibitory synaptic transmission in rat hypoglossal nucleus. Brain Research (2016), http://dx.doi.org/10.1016/j.brainres.2016.02.020
301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351
BRES : 44732
3
brain research ] (]]]]) ]]]–]]]
352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411
receptor antagonist. It was found that blockade of AMPA receptors that are expressed on the local interneurons in this network was expected to hyperpolarize these interneurons, which may slightly attenuate the inhibitory neurotransmitters release from its axon terminals, without facilitating the release of these neurotransmitters. However, since 1990s, a number of studies had demonstrated that CNQX and its quinoxalinedione analog, 6,7-dinitroquinoxaline-2,3dione (DNQX), commonly used as antagonist of AMPA receptors, was demonstrated to partially activate these AMPA receptors inversely, leading to subsequent depolarization of GABAergic interneurons, hence more gamma-aminobutyric acid (GABA) was triggered to release from their axon terminals in the hippocampus and cerebellum (Brickley et al., 2001; Hashimoto et al., 2004; Maccaferri and Dingledine, 2002; McBain et al., 1992; Menuz et al., 2007). In these limited neural networks, CNQX and DNQX were considered as weak partial agonists of AMPA receptors but not as competitive antagonists. Similarly, a great enhancement of inhibitory synaptic transmission was found upon application of CNQX in the hypoglossal nucleus. It is essential and interesting to study the novel function of CNQX in inhibitory transmission within this neural circuitry.
perfusion of CNQX (10 μM) robustly enhanced spontaneous inhibitory synaptic transmission in most recorded HMs (n¼ 97/118, left panel; Fig. 1A). The rest of HMs exhibited little response to the perfusion of CNQX. After recording, HM was visualized by the treatment of neurobiotin (right panel, Fig. 1A). The neurobiotin-labeled HMs demonstrated several processes arising from its huge soma. Further analysis of the changes in both frequency and amplitude was conducted. With application of CNQX, the cumulative probability curve of inter-event interval significantly shifted leftward (po0.001) while that of the amplitude of sIPSCs significantly shifted rightward, indicating that CNQX induced enhancement in both frequency and amplitude(po0.001; Fig. 1B). Bar graph shows that CNQX significantly increased the mean frequency of sIPSCs to 309.3735.1% of that of the baseline level (n¼ 97, po0.001; Fig. 1C) and the mean amplitude of sIPSCs to 181.1718.2% of the baseline level (n¼ 97, po0.001; Fig. 1C). The increased frequency of sIPSCs strongly indicated that CNQX facilitated the release of inhibitory neurotransmitter from presynaptic terminals that project to the HMs. More-
2.
Results
2.1. CNQX-induced facilitation of inhibitory synaptic transmission within the hypoglossal nucleus
over, we investigated the effect of CNQX on the basal current and input resistance of HMs. Administration of CNQX did not elicit detectable inward or outward basal current in HMs (98.5712.5% of the control level, p40.05; n¼ 7; Fig. 1D). Similarly, there was no change in basal membrane potential
In order to verify whether CNQX could facilitate inhibitory synaptic transmission within the hypoglossal nucleus, we first performed whole-cell patch-clamp recordings in HMs with application of CNQX. Under voltage-clamp, bath
(control:
70.270.4 mV;
CNQX:
69.870.7 mV;
p40.05;
n¼ 6) and input resistance (98.373.5% of the control level, p40.05; n¼ 6) when application of CNQX. Sample trace was illustrated in Fig. 1E. This result implies that AMPA receptors
Fig. 2 – NMDA receptor was independent on CNQX-potentiated sIPSCs in the hypoglossal nucleus. (A) Sample traces of sIPSCs recorded in HMs. D-AP5 (50 μM) had no obvious effects on sIPSCs but greatly facilitated this inhibitory transmission when CNQX (10 μM) was further added into the bath solution. (B) Cumulative probability distributions of the inter-event intervals and amplitudes of the sIPSCs as shown in the Fig. A. With perfusion of D-AP5, no significant differences can be detected in the distributions of inter-event intervals and amplitudes except after application of CNQX. (C) Pooled results of the effects of DAP5 and subsequent added CNQX on the frequency and amplitude of sIPSCs. N.S.: not significant. Please cite this article as: Han, L., et al., CNQX facilitates inhibitory synaptic transmission in rat hypoglossal nucleus. Brain Research (2016), http://dx.doi.org/10.1016/j.brainres.2016.02.020
412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471
BRES : 44732
4
472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531
brain research ] (]]]]) ]]]–]]]
on the HMs are not involved in the facilitation of inhibitory synaptic transmission.
2.2. NMDA receptor was not involved in the CNQXtriggered facilitation of sIPSCs Contrary to the traditional view of blocking AMPA receptors, CNQX can directly activate AMPA receptors on inhibitory interneurons. This hypothesis has been verified in both the hippocampus and cerebellum. An alternative hypothesis is that CNQX blocks AMPA receptor-driving inhibitory interneurons impinging upon inhibitory premotoneurons to HMs, which was also expected to enhance inhibitory inputs to HMs. It is usually accepted that N-methyl-D-aspartate (NMDA) receptors are mostly co-localized with AMPA receptors on functional dendritic spines. Thus, blockade of NMDA receptors can be formulated to facilitate inhibitory transmission within the hypoglossal nucleus according to the second hypothesis. In fact, bath perfusion of D-(-)-2-Amino-5-phosphonopentanoic acid (D-AP5, 50 μM), an antagonist of NMDA receptors, was demonstrated little effect on inhibitory transmission(frequency, 86.273.0% of control level, n ¼8, p40.05; amplitude, 88.973.1% of control level, n¼ 8, p40.05; Fig. 2) but significantly enhance that after co-application of D-AP5 and CNQX in both frequency (286.9748.3% of control level, n¼ 8, po0.01; Fig. 2) and amplitude (179.9724.5% of control level, n¼ 8, po0.05; Fig. 2). The differential effects of D-AP5 and CNQX on inhibitory transmission within the hypoglossal nucleus may postulate that the facilitation of inhibitory transmission by CNQX was not likely to the result from disinhibition of inhibitory premotoneurons to HMs.
2.3. Both GABAergic and glycinergic synaptic transmission were facilitated by CNQX It is known that HMs receive both GABAergic and glycinergic inhibitory innervations from surrounding inhibitory interneurons. To better understand whether CNQX could facilitate GABAergic or glycinergic transmission within the hypoglossal nucleus, bicuculline (GABAA receptor antagonist) and strychnine (glycine receptor antagonist) were used in the present experiment, respectively. Administration of bicuculline (10 μM) completely eliminated CNQX-induced enhancement of sIPSCs with pre-application of strychnine (1 μM) in HMs, confirming that CNQX enhanced GABAergic synaptic transmission within the hypoglossal nucleus (Fig. 3A). In these experiments, CNQX (10 μM) significantly increased the frequency of GABAA receptor-mediated sIPSCs (293.6755.3% of the baseline value; n¼ 8, po0.05; Fig. 3A) but exerted no effects in amplitude (119.1717.0% of the baseline value, n¼ 8, p40.05; Fig. 3A). These results indicated that CNQX remarkably increased the frequency but had little effects on amplitude of GABAA receptor-mediated sIPSCs. In the hypoglossal nucleus, glycine and GABA are reported to be co-released onto both neonatal and juvenile HMs in a certain number of inhibitory terminals (O'Brien and Berger, 1999, 2001). Accordingly, perfusion of strychnine (1 μM) could also block CNQX-triggered facilitation of sIPSCs with pre-incubation of bicuculline (10 μM) in HMs, indicating that CNQX induced an enhancement of glycine receptor-mediated
sIPSCs in HMs. Statistical analysis showed that CNQX(10 μM) significantly increased the frequency of glycine receptormediated sIPSCs (381.7784.2% of the baseline value; n¼ 8, po0.05; Fig. 3B) and the amplitude (142.9711.8% of the baseline value; n¼ 8, po0.05; Fig. 3B).The analysis of kinetics in postsynaptic currents indicated that bicuculline-insensitive current decayed much more quickly than GABAA receptormediated current recorded at the same postnatal stage (Fig. 3C). The decay time of bicuculline-insensitive events was 7.270.4 ms, similar to that of glycine receptor-mediated current reported in HMs (O’Brien and Berger, 2001). On the other hand, bar graph showed that the decay time of GABAA receptor-mediated current was two folds longer than that of bicuculline-insensitive current (n¼ 8, po0.001, unpaired t test; Fig. 3C).All these results illustrated that CNQX not only induced enhancement of GABAA receptor-mediated sIPSCs but also triggered the facilitation of glycine receptormediated sIPSCs in HMs.
2.4. DNQX but not GYKI 53655 mimicked the effects of CNQX on sIPSCs We then tested the effects of quinoxaline derivative DNQX on the sIPSCs in HMs. DNQX (10 μM) also greatly enhanced inhibitory synaptic transmission in partial of recorded HMs (n¼ 9, Fig. 4A), in agreement with the results of CNQX-induced facilitation of sIPSCs in HMs. With application of DNQX, the frequency of sIPSCs was significantly increased (211.1729.0% of the control level; n¼ 9, po0.01; Fig. 4A) whereas the amplitude was also greatly increased (124.377.6% of the baseline; n¼ 9, po0.01; Fig. 4A). However, GYKI 53655, a non-competitive AMPA receptor antagonist, could not mimic the effect of CNQX on sIPSCs in all recorded HMs. GYKI 53655 (10 μM) had little effect on either frequency (82.2713.9% of the control level, n¼ 9, p40.05; Fig. 4B) or amplitude (97.478.7% of the control level, n¼ 9, p40.05; Fig. 4B) of sIPSCs in HMs. Further administration of CNQX (10 μM) still could not facilitate inhibitory synaptic transmission in the presence of GYKI 53655 (10 μM) (frequency, 79.179.9% of control level, n¼ 9, p40.05; amplitude, 85.479.0% of control level, n¼ 9, p40.05; Fig. 4B). These results suggest that only quinoxaline derivative AMPA receptor competitive antagonist could favor inhibitory transmitters release from inhibitory neurons that impinging on HMs while little effect with noncompetitive antagonist can be observed, which is in agreement with Menuz et al. (2007). In addition, it appears that blockade of AMPA receptors by non-competitive AMPA receptor antagonist in advance prevented activation of AMPA receptors by CNQX, hence subsequently abolished enhancement of sIPSCs within the hypoglossal nucleus.
2.5. Action potential-dependent facilitation of sIPSCs was induced by CNQX sIPSCs are composed of both action potential-dependent and action potential-independent components. To elucidate the effects of CNQX on the two subtypes of IPSCs, tetrodotoxin (TTX, 1 μM) was applied to eliminate action potential-dependent IPSCs. Bath administration of TTX abolished CNQX-induced enhancement of sIPSCs in HMs, afterwards, both frequency and amplitude of sIPSCs were significantly dropped to
Please cite this article as: Han, L., et al., CNQX facilitates inhibitory synaptic transmission in rat hypoglossal nucleus. Brain Research (2016), http://dx.doi.org/10.1016/j.brainres.2016.02.020
532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591
BRES : 44732 brain research ] (]]]]) ]]]–]]]
592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651
5
Fig. 3 – CNQX-induced facilitation of GABAergic and glycinergic synaptic transmission in the hypoglossal nucleus. (A) Top panel: representative traces showed that bicuculline (10 μM) completely eliminated CNQX-potentiated GABAergic sIPSCs in HM. strychnine (1 μM) was present in ACSF all the time. Middle panel: the cumulative probability distributions of the interevent intervals and amplitudes of the sIPSCs from the experiment as shown in the top panel of Fig. A. Bottom panel: pooled data demonstrates the effect of CNQX on the frequency and amplitude of GABAergic sIPSCs. All data were normalized to the control level. (B) Top panel: sample traces showed that strychnine completely eliminated CNQX-potentiated glycinergic sIPSCs in HM. Bicuculline was present in ACSF all the time. Middle panel: cumulative probability distributions of the interevent intervals and amplitudes of the sIPSCs from the experiment shown in the top panel of Fig. A. Bottom panel: pooled data demonstrates the of CNQX on frequency and amplitude of glycinergic sIPSCs. All data were normalized to the control level. (C) Left panel: deactivation time course of GABAA receptor-mediated current (black line) and glycine receptor-mediated current (grey line). They were averaged and normalized from representative postsynaptic events in the middle sample trace of Figs. A and B, respectively. The bicuculline-insensitive current decayed much quicker than GABAA receptor-mediated current, which is consistent with the bar graph (right panel). N.S.: not significant.
12.870.5% of the control level (n¼ 7, po0.05; Fig. 5A) and 52.574.6% of the control level (n¼7, po0.05; Fig. 5A), respectively. However, we could not exclude the possibility that CNQX might enhance the action potential-independent miniature IPSCs
(mIPSCs). To examine whether CNQX induces enhancement of mIPSCs, TTX was pretreated for 10 min to isolate the miniature events. Under this condition, it was found that CNQX failed to induce any enhancement of mIPSCs (n¼7, p40.05 in both
Please cite this article as: Han, L., et al., CNQX facilitates inhibitory synaptic transmission in rat hypoglossal nucleus. Brain Research (2016), http://dx.doi.org/10.1016/j.brainres.2016.02.020
652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711
BRES : 44732
6
712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771
brain research ] (]]]]) ]]]–]]]
Fig. 4 – DNQX greatly facilitated inhibitory transmission but GYKI 53655 had little effect. (A) Top panel: representative traces of sIPSCs recorded from a HM. DNQX (10 μM) mimicked CNQX-induced facilitation of sIPSCs. Middle panel: cumulative curve of inter-event interval and amplitude of sIPSCs from sample neuron in the top panel of the Fig. A. Significant difference was found in the distributions of inter-event intervals and amplitudes when perfused with DNQX. Bottom panel: pooled results demonstrate the effect of DNQX on the frequency and amplitude of sIPSCs. (B) Top panel: representative traces of sIPSCs recorded from a HM. GYKI 53655 (10 μM) was ineffective to either frequency or amplitude of sIPSCs (middle trace). Preadministration of GYKI 53655 prevented the facilitation of sIPSCs by CNQX (bottom trace). Middle panel: cumulative curve of inter-event interval and amplitude of sIPSCs from a sample neuron in top panel of the Fig. B. No significant differences were found in the distributions of inter-event intervals and amplitudes even treated with CNQX. Bottom panel: pooled results demonstrate the effect of GYKI 53665 and subsequent CNQX on the frequency and amplitude of sIPSCs. N.S.: not significant.
frequency and amplitude; Fig. 5B). These results suggest that the effect of CNQX is selective to action potential-dependent sIPSCs but not to mIPSCs. In other words, CNQX increased firing rate of presynaptic neurons via activation of AMPA receptors located on their presynaptic interneurons to enhance action potentialdependent release of inhibitory neurotransmitters onto adjoining HMs.
3.
Discussion
3.1. Facilitation of glycinergic transmission in the hypoglossal nucleus by CNQX Our electrophysiological studies demonstrated that CNQX and its quinoxaline derivative DNQX could dramatically enhance
inhibitory inputs to HMs. Consistent with earlier findings in the hippocampus and cerebellum, CNQX was found to enhance GABAergic transmission in HMs, however, it not only facilitated GABAergic transmission in HMs, but also exhibited its capability in enhancement of glycinergic transmission in HMs. This is the first report to show CNQX-potentiated glycinergic synaptic transmission in the CNS. The absence of CNQX-triggered enhancement of glycinergic transmission in the other brain regions may largely attribute to specific distribution patterns of functional glycine receptors in the CNS. Synaptic glycine receptors are rich in the retina, spinal cord and brainstem but rare in the neocortex and hippocampus, whereas GABAA receptors are widely distributed throughout the CNS (Lynch, 2004; Pirker et al., 2000). Furthermore, mixed GABAergic/glycinergic inhibitory synapses had been functionally characterized in motoneurons of the
Please cite this article as: Han, L., et al., CNQX facilitates inhibitory synaptic transmission in rat hypoglossal nucleus. Brain Research (2016), http://dx.doi.org/10.1016/j.brainres.2016.02.020
772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831
BRES : 44732 brain research ] (]]]]) ]]]–]]]
832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891
7
Fig. 5 – Action potential-dependent facilitation of sIPSCs triggered by CNQX. (A) Top panel: representative traces of sIPSCs recorded from a HM. TTX (1 μM) dramatically suppressed the frequency and amplitude of sIPSCs enhanced by CNQX. Middle panel: the cumulative probability distributions of inter-event intervals and amplitudes of the sIPSCs as shown in the top panel of Fig. A. TTX significantly shifted the cumulative curves of both inter-event intervals and amplitudes that were pre-treated with CNQX. Bottom panel: pooled results of the effect of TTX on sIPSCs were facilitated by CNQX. (B) Pre-perfusion of TTX prevented CNQX-induced enhancement of inhibitory transmission as illustrated in the sample traces (top panel). The distributions of cumulative probability of inter-event intervals and amplitudes of sIPSCs from the experiment as shown in the top panel of Fig. B (middle panel). Pooled data of the effects of CNQX on mIPSCs (bottom panel). N.S.: not significant.
hypoglossal nucleus (Muller et al., 2006; O'Brien and Berger, 2001). Accordingly, it is not strange that CNQX is able to facilitate both GABAergic and glycinergic synaptic transmissions within the hypoglossal nucleus. Given that the synaptic strength is fundamentally dependent on the three main factors: the numbers of synaptic contacts, size of postsynaptic depolarization/hyperpolarization caused by neurotransmitter release from a single synaptic vesicle (termed as quantal size), and probability of neurotransmitter release at each synapse (Del Castillo and Katz, 1956). With application of CNQX, the first two issues appeared to be constant while the probability of glycine/GABA release increased similarly to the control. Based on this theory, the amplitude of GABAergic currents should increase significantly as well as glycinergic ones with CNQX. Interestingly, CNQX only triggered significant increment in the amplitude of glycinergic sIPSCs but not in that of GABAergic sIPSCs. One possible interpretation for
this discrepancy is due to the differences in number of spare postsynaptic glycine/GABAA receptors in the control condition. The CNQX-induced rise in the amplitude of glycinergic currents is a result from the activation of a large number of spare postynaptic glycine receptors, whereas little increment in amplitude of GABAergic currents is due to the few spare postynaptic GABAA receptors in HMs. Morphological and electrophysiological evidences further demonstrated that functional glycine receptors are predominant over GABAA receptors at the postsynaptic sites, which is potentially in agreement with this interpretation.
3.2. Possible underlying mechanisms for activation of AMPA receptor Based on pharmacological and structural analysis, previous study had revealed that the presence of transmembrane
Please cite this article as: Han, L., et al., CNQX facilitates inhibitory synaptic transmission in rat hypoglossal nucleus. Brain Research (2016), http://dx.doi.org/10.1016/j.brainres.2016.02.020
892 893 894 895 896 897 898 899 900 901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946 947 948 949 950 951
BRES : 44732
8
952 953 954 955 956 957 958 959 960 961 962 963 964 965 966 967 968 969 970 971 972 973 974 975 976 977 978 979 980 981 982 983 984 985 986 987 988 989 990 991 992 993 994 995 996 997 998 999 1000 1001 1002 1003 1004 1005 1006 1007 1008 1009 1010 1011
brain research ] (]]]]) ]]]–]]]
AMPA receptor regulatory protein (TARP) auxiliary subunits could determine whether quinoxaline derivative acted as a partial agonist or a antagonist (Menuz et al., 2007). These accessory proteins play a crucial role not only in AMPA receptor trafficking (Chen et al., 2000; Kato et al., 2007; Soto et al., 2009; Tomita et al., 2003), but also in the regulation of synaptic targeting and gating as well as the pharmacological profiles of AMPARs (Bedoukian et al., 2006; Cho et al., 2007; Jackson and Nicoll, 2011; Milstein et al., 2007; Tomita et al., 2005). With the help of TARPs, the combination of CNQX to the corresponding site on AMPA receptors caused the opening of ion channel. This finding explained the "excitatory" role of CNQX on inhibitory transmission in the hippocampus and cerebellum as well as the hypoglossal nucleus in our present study. However, another study claimed that the shift of CNQX antagonism potency could be wholly attributed to the increased agonist potency caused by TARPs, which only fits for the equilibrium conditions but not for glutamatergic synapse (nonequilibrium conditions) (Maclean and Bowie, 2011). CNQX is regarded as a non-competitive antagonist at glutamatergic synapses as they concluded, however, it is impossible to explain our experimental results collected in the non-equilibrium conditions. In the present study, CNQX appeared to serve as a weak partial agonist of AMPA receptors. CNQX alone can evoke a macroscopic inward current (about 40 pA) in the cerebellum Golgi cells (Menuz et al., 2007), suggesting that at least in some AMPA receptors on inhibitory neurons, CNQX serves as a weak partial agonist through overcoming CNQX-elicited desensitization of AMPA receptor. It is known that most neurons in the CNS express TARPs that co-localized with AMPA receptors (Lein et al., 2007; Tomita et al., 2003). This distribution pattern of TARPs therefore raises an interesting question that why the "excitatory" action of CNQX is only observed in a limited number of brain regions. The possible explanation is that splice variant (flip/ flop) of AMPA receptors determines the extent of desensitization of AMPA receptors. The flop variants of AMPA receptors are less desensitized by their agonists (Bertolino et al., 1993; Geiger et al., 1995; Partin et al., 1994). Accordingly, inhibitory neurons dominant in the expression of flop variant of AMPA receptors could be depolarized by CNQX, which elicits larger inward currents than its diminishment by desensitization. In brainstem nuclei such as the hypoglossal nucleus, both splice variants in all subunits of AMPA receptors can be detected (Paarmann et al., 2000). Thus, it can be speculated that CNQXtriggered activation of flop variant of AMPA receptors that localized at inhibitory synapse is responsible for the "excitatory" action on inhibitory interneurons that innervate HMs. As to the other AMPA receptor agonists, previous studies showed a high degree of variability in the effect of DNQX on inhibitory transmission, although the excitatory action of CNQX is consistent among all the previous reports. Administration of DNQX, however, showed no effects on sIPSCs in the CA3 of the hippocampus (McBain et al., 1992), whereas in the cerebellum (Brickley et al., 2001) and the hypoglossal nucleus, DNQX exerted an excitatory effect. This different phenomenon may be due to lower sensitivity of DNQX to AMPA receptors as compared with CNQX, which needs further studies to elucidate the underlying mechanisms.
GYKI 53655, a structurally different AMPA receptor antagonist, produced negligible alterations in sIPSCs in HMs, suggesting that only quinoxaline related AMPA receptor antagonist could act as a partial agonist of AMPA receptors. Moreover, the similar action of GYKI 53655 on sIPSCs to that of D-AP5 further excluded the possibility of disinhibition of inhibitory pre-motoneurons that innervate HMs. Preblockade of AMPA receptors by this non-competitive AMPA receptor antagonist abolished facilitation of sIPSCs induced by CNQX, indicating that activation of AMPA receptors by CNQX was the only cause for the enhancement of inhibitory transmission within the hypoglossal nucleus.
3.3. Direct depolarization of inhibitory neurons by CNQX appears to be responsible for facilitation of sIPSCs Previous electrophysiological data showed a CNQX-triggered inward current in inhibitory cerebellar Golgi cells and Purkinje cells, indicating a direct depolarization of these cerebellar inhibitory neurons by CNQX (Menuz et al., 2007). Similarly, in the hippocampus, CNQX could induce depolarization, thereby subsequently increase the firing rate of inhibitory neurons but not of pyramidal neurons (Hashimoto et al., 2004; Maccaferri and Dingledine, 2002). Thus, it is reasonable that increased firing rate of inhibitory neurons can account for the enhancement of inhibitory synaptic transmission, as a result of action potentialdependent GABA/glycine release onto the postsynaptic receptors of HMs. Therefore, blockade of action potential by TTX prevented CNQX-potentiated inhibitory transmission in HMs (Fig. 5). To confirm a direct excitatory action of CNQX on AMPA receptors, further experiments should be conducted in the inhibitory interneurons that are impinged onto HMs.
3.4. Dual role of CNQX in the CNS and its possible problems In the present study, we observed that CNQX could serve as a weak partial agonist of AMPA receptors for potentiating inhibitory strength, although CNQX is widely used as an antagonist of AMPA receptors to eliminate AMPA current. Therefore, it should be cautious to use CNQX as a regular blocker of AMPA receptors in the study of neural circuits, especially in the hippocampus, cerebellum, and hypoglossal nucleus, due to its suggested dual role. In most cases, CNQX serves as a non-competitive blocker of AMPA receptors in glutamatergic synapses for eliminating phasic glutamatergic events, however, in a small population of inhibitory neurons, CNQX works as a weak partial agonist of the AMPA receptors to induce a tonic depolarizing currents. Thus, it is still pharmacologically feasible to use CNQX for isolating GABAA/glycine receptor-mediated currents, however, it should be careful during the in vitro or in vivo chronic application of CNQX as a long-lasting activator of AMPA receptors, since the subsequent over-excitation may lead to neuron death.
Please cite this article as: Han, L., et al., CNQX facilitates inhibitory synaptic transmission in rat hypoglossal nucleus. Brain Research (2016), http://dx.doi.org/10.1016/j.brainres.2016.02.020
1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052 1053 1054 1055 1056 1057 1058 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070 1071
BRES : 44732 brain research ] (]]]]) ]]]–]]]
1072 1073 1074 1075 1076 1077 1078 1079 1080 1081 1082 1083 1084 1085 1086 1087 1088 1089 1090 1091 1092 1093 1094 1095 1096 1097 1098 1099 1100 1101 1102 1103 1104 1105 1106 1107 1108 1109 1110 1111 1112 1113 1114 1115 1116 1117 1118 1119 1120 1121 1122 1123 1124 1125 1126 1127 1128 1129 1130 1131
4.
Experimental procedure
Sprague Dawley (SD) rats of postnatal day (P) 4–17 were used in the experiments. All procedures and protocols on animal preparation were performed in accordance with the National Institute of Health Guide for the Care and Use of Laboratory Animals (NIH Publications no. 80-23). In this study, all efforts were made to minimize the numbers and suffering of the animals to produce statistically significant data.
4.1.
Slice preparations
SD rats were first decapitated quickly after isoflurane inhalation and the brains were dissected out from the skull subsequently. The brainstem was then removed to ice-cold artificial cerebrospinal fluid (ACSF) containing (in mM) 120 NaCl, 2 KCl, 2.5 CaCl2, 1.2 MgCl2, 1.2 KH2PO4, 26 NaHCO3 and 11 glucose, saturated with 95% O2 and 5% CO2, pH 7.3. Coronal brainstem slices (300 μm), containing the hypoglossal nucleus, were cut by a vibratome from the brainstem. The freshly obtained brainstem slices were then transferred to a pre-warmed (33 1C) water bath incubator where glass incubation chamber contained oxygenated ASCF (pH 7.3) for about 1 h before recording. After incubation, the brainstem slice was transferred to the recording chamber equipped with infrared differential interference contrast (IR-DIC) optics (Olympus, Japan). HMs were identified from their characteristic location and morphology which can be distinguished from the neurons beyond the hypoglossal nucleus.
4.2.
Whole-cell patch-clamp recording
Patch electrodes were pulled from borosilicate glass (Harvard Apparatus, USA) using a P-97 micropipette puller (Sutter Instruments, USA). The pipette resistance was 3–6 MΩ for recording HMs. Under voltage-clamp mode, internal solution composed of (in mM) 140 CsCl, 2 MgCl2, 2 Na2ATP, 1 EGTA, and 10 HEPES, pH 7.3 (adjusted by CsOH), 290 mOsm, was used for recording spontaneous inhibitory postsynaptic currents. To generate CNQX-induced facilitation of sIPSCs, both AMPA and NMDA receptor antagonists were not delivered to the slices before recording. Slice in recording chamber was continuously perfused with oxygenated ACSF at a rate of 2 ml/min. An Ag–AgCl wire immersed to the bath was used as a reference electrode for grounding. All whole-cell currents were recorded at 70 mV. To measure input resistance of HM, periodic current injections of 100 pA for 100 ms at 0.5 Hz was employed in the experiment under currentclamp mode. The membrane potential was held around 70 mV. Voltage deflections were responses to periodic hyperpolarizing current. All data were collected by a multiclamp 700A amplifier, Digidata 1322A analog/digital interface board and Clampex 10.3 software (Axon Instruments, USA). These signals were filtered at 3 KHz and digitized at 10 KHz, and stored for subsequent analysis. Neuron was discarded if the access resistance was over 20 MΩ or changed by more than 15% after recording. All recordings were performed at room temperature (20–25 1C).
4.3.
Morphology
9
1132 1133 1134 To visualize the morphology of the whole-cell recorded HMs, 1135 the pipette was filled with freshly prepared 0.3% of neuro1136 biotin (Vector, USA) in CsCl based-internal solution. After 1137 recording for at least more than 30 min, the slice was 1138 transferred to 4% paraformaldehyde fixative in 0.1 M PB and 1139 immersed overnight at41C.After rinsing with PBS for 3 times 1140 (each 10 min), the section was incubated in avidin conjugated 1141 with Texas red (Molecular Probes, 1:1,000 diluted by 0.3% 1142 Triton-PBS) for at least 4 h in room temperature. Finally, the 1143 reaction was stopped by rinsing the section with PBS (10 min) 1144 for 3times. The section was then mounted on gelatin-coated 1145 slide for further observation. Digital images were captured by 1146 a fluorescent microscope (Axioplan II, Zeiss, Germany) with 1147 CCD camera (Spot, Diagnostic Instrument, USA). 1148 1149 4.4. Drugs 1150 1151 The following drugs were obtained from (1) Sigma chemical 1152 company (Sigma-Aldrich, USA): strychnine and TTX; (2) Tocris 1153 Bioscience: CNQX, DNQX, D-AP5, bicuculline, and GIKI 53655. 1154 Drugs were dissolved as stock solutions in distilled water or 1155 DMSO. The final concentration of DMSO was less than 0.1%. All 1156 stock solution of the applied drugs was freshly added to ACSF on 1157 the day of experiment. Drugs were delivered by a gravity-feed 1158 system connected to the bath-recording chamber. 1159 1160 4.5. Data analysis 1161 1162 All data were processed off-line in Mini Analysis 6.0.3 1163 (Synaptosoft, USA), Clampfit 10.3 (Axon Instruments, USA), 1164 and SigmaPlot 12.0 (Jandel Scientific, Germany) software. 1165 IPSCs were analyzed by Mini Analysis program. Only single 1166 peak events were chosen for kinetic analysis. Numerical data 1167 were presented as the mean 7S.E.M. Statistical comparison 1168 of two cumulative probabilities was done by a Kolmogorov1169 Smirnov test. All the other comparisons were assessed using 1170 paired or unpaired Student's t-tests as appropriate. po0.05 1171 * was considered as significant difference. symbolizes po0.05, 1172 ** *** represents for po0.01 and represents for po0.001. 1173 1174 1175 Acknowledgments 1176 This work was supported by National Natural Science Foun- Q5 1177 1178 dation of China (81400064). 1179 1180 r e f e r e nc e s 1181 1182 1183 Bedoukian, M.A., Weeks, A.M., Partin, K.M., 2006. Different 1184 domains of the AMPA receptor direct stargazin-mediated trafficking and stargazin-mediated modulation of kinetics. J. 1185 Biol. Chem. 281, 23908–23921. 1186 Bertolino, M., Baraldi, M., Parenti, C., Braghiroli, D., DiBella, M., 1187 Vicini, S., Costa, E., 1993. Modulation of AMPA/kainate recep1188 tors by analogues of diazoxide and cyclothiazide in thin slices 1189 of rat hippocampus. Recept. Channels 1, 267–278. 1190 Brickley, S.G., Farrant, M., Swanson, G.T., Cull-Candy, S.G., 2001. 1191 CNQX increases GABA-mediated synaptic transmission in the
Please cite this article as: Han, L., et al., CNQX facilitates inhibitory synaptic transmission in rat hypoglossal nucleus. Brain Research (2016), http://dx.doi.org/10.1016/j.brainres.2016.02.020
BRES : 44732
10
1192 1193 1194 1195 1196 1197 1198 1199 1200 1201 1202 1203 1204 1205 1206 1207 1208 1209 1210 1211 1212 1213 1214 1215 1216 1217 1218 1219 1220 1221 1222 1223 1224 1225 1226 1227 1228 1229 1230 1231 1232 1233 1234 1235 1236 1237 1238 1239 1240 1241
brain research ] (]]]]) ]]]–]]]
cerebellum by an AMPA/kainate receptor-independent mechanism. Neuropharmacology 41, 730–736. Chen, L., Chetkovich, D.M., Petralia, R.S., Sweeney, N.T., Kawasaki, Y., Wenthold, R.J., Bredt, D.S., Nicoll, R.A., 2000. Stargazin regulates synaptic targeting of AMPA receptors by two distinct mechanisms. Nature 408, 936–943. Cho, C.H., St-Gelais, F., Zhang, W., Tomita, S., Howe, J.R., 2007. Two families of TARP isoforms that have distinct effects on the kinetic properties of AMPA receptors and synaptic currents. Neuron 55, 890–904. Davidoff, M.S., Schulze, W., 1988. Coexistence of GABA- and choline acetyltransferase (ChAT)-like immunoreactivity in the hypoglossal nucleus of the rat. Histochemistry 89, 25–33. Del Castillo, J., Katz, B., 1956. Localization of active spots within the neuromuscular junction of the frog. J. Physiol. 132, 630–649. Geiger, J.R., Melcher, T., Koh, D.S., Sakmann, B., Seeburg, P.H., Jonas, P., Monyer, H., 1995. Relative abundance of subunit mRNAs determines gating and Ca2þ permeability of AMPA receptors in principal neurons and interneurons in rat CNS. Neuron 15, 193–204. Hashimoto, Y., Miyakawa, H., Kudo, Y., Inoue, M., 2004. 6-Cyano7-nitroquinoxaline-2,3-dione (CNQX) increases GABAA receptor-mediated spontaneous postsynaptic currents in the dentate granule cells of rat hippocampal slices. Neurosci. Lett. 358, 33–36. Jackson, A.C., Nicoll, R.A., 2011. The expanding social network of ionotropic glutamate receptors: TARPs and other transmembrane auxiliary subunits. Neuron 70, 178–199. Kato, A.S., Zhou, W., Milstein, A.D., Knierman, M.D., Siuda, E.R., Dotzlaf, J.E., Yu, H., Hale, J.E., Nisenbaum, E.S., Nicoll, R.A., Bredt, D.S., 2007. New transmembrane AMPA receptor regulatory protein isoform, gamma-7, differentially regulates AMPA receptors. J. Neurosci. 27, 4969–4977. Lein, E.S., Hawrylycz, M.J., Ao, N., et al., 2007. Genome-wide atlas of gene expression in the adult mouse brain. Nature 445, 168–176. Li, Y.Q., Takada, M., Kaneko, T., Mizuno, N., 1996. GABAergic and glycinergic neurons projecting to the trigeminal motor nucleus: a double labeling study in the rat. J. Comp. Neurol. 373, 498–510. Li, Y.Q., Takada, M., Kaneko, T., Mizuno, N., 1997. Distribution of GABAergic and glycinergic premotor neurons projecting to the facial and hypoglossal nuclei in the rat. J. Comp. Neurol. 378, 283–294. Lynch, J.W., 2004. Molecular structure and function of the glycine receptor chloride channel. Physiol. Rev. 84, 1051–1095. Maccaferri, G., Dingledine, R., 2002. Complex effects of CNQX on CA1 interneurons of the developing rat hippocampus. Neuropharmacology 43, 523–529. Maclean, D.M., Bowie, D., 2011. Transmembrane AMPA receptor regulatory protein regulation of competitive antagonism: a problem of interpretation. J. Physiol. 589, 5383–5390. Marchetti, C., Pagnotta, S., Donato, R., Nistri, A., 2002. Inhibition of spinal or hypoglossal motoneurons of the newborn rat by glycine or GABA. Eur. J. Neurosci. 15, 975–983.
McBain, C.J., Eaton, J.V., Brown, T., Dingledine, R., 1992. CNQX increases spontaneous inhibitory input to CA3 pyramidal neurones in neonatal rat hippocampal slices. Brain Res. 592, 255–260. Menuz, K., Stroud, R.M., Nicoll, R.A., Hays, F.A., 2007. TARP auxiliary subunits switch AMPA receptor antagonists into partial agonists. Science 318, 815–817. Milstein, A.D., Zhou, W., Karimzadegan, S., Bredt, D.S., Nicoll, R. A., 2007. TARP subtypes differentially and dose-dependently control synaptic AMPA receptor gating. Neuron 55, 905–918. Muller, E., Le Corronc, H., Triller, A., Legendre, P., 2006. Developmental dissociation of presynaptic inhibitory neurotransmitter and postsynaptic receptor clustering in the hypoglossal nucleus. Mol. Cell. Neurosci. 32, 254–273. O’Brien, J.A., Berger, A.J., 1999. Cotransmission of GABA and glycine to brain stem motoneurons. J. Neurophysiol. 82, 1638–1641. O’Brien, J.A., Berger, A.J., 2001. The nonuniform distribution of the GABA(A) receptor alpha 1 subunit influences inhibitory synaptic transmission to motoneurons within a motor nucleus. J. Neurosci. 21, 8482–8494. Paarmann, I., Frermann, D., Keller, B.U., Hollmann, M., 2000. Expression of 15 glutamate receptor subunits and various splice variants in tissue slices and single neurons of brainstem nuclei and potential functional implications. J. Neurochem. 74, 1335–1345. Partin, K.M., Patneau, D.K., Mayer, M.L., 1994. Cyclothiazide differentially modulates desensitization of alpha-amino-3hydroxy-5-methyl-4-isoxazolepropionic acid receptor splice variants. Mol. Pharmacol. 46, 129–138. Pirker, S., Schwarzer, C., Wieselthaler, A., Sieghart, W., Sperk, G., 2000. GABA(A) receptors: immunocytochemical distribution of 13 subunits in the adult rat brain. Neuroscience 101, 815–850. Rekling, J.C., Funk, G.D., Bayliss, D.A., Dong, X.W., Feldman, J.L., 2000. Synaptic control of motoneuronal excitability. Physiol. Rev. 80, 767–852. Sawczuk, A., Mosier, K.M., 2001. Neural control of tongue movement with respect to respiration and swallowing. Crit. Rev. Oral Biol. Med. 12, 18–37. Soto, D., Coombs, I.D., Renzi, M., Zonouzi, M., Farrant, M., CullCandy, S.G., 2009. Selective regulation of long-form calciumpermeable AMPA receptors by an atypical TARP, gamma-5. Nat. Neurosci. 12, 277–285. Tomita, S., Chen, L., Kawasaki, Y., Petralia, R.S., Wenthold, R.J., Nicoll, R.A., Bredt, D.S., 2003. Functional studies and distribution define a family of transmembrane AMPA receptor regulatory proteins. J. Cell Biol. 161, 805–816. Tomita, S., Adesnik, H., Sekiguchi, M., Zhang, W., Wada, K., Howe, J.R., Nicoll, R.A., Bredt, D.S., 2005. Stargazin modulates AMPA receptor gating and trafficking by distinct domains. Nature 435, 1052–1058. Viana, F., Gibbs, L., Berger, A.J., 1990. Double- and triple-labeling of functionally characterized central neurons projecting to peripheral targets studied in vitro. Neuroscience 38, 829–841.
Please cite this article as: Han, L., et al., CNQX facilitates inhibitory synaptic transmission in rat hypoglossal nucleus. Brain Research (2016), http://dx.doi.org/10.1016/j.brainres.2016.02.020
1242 1243 1244 1245 1246 1247 1248 1249 1250 1251 1252 1253 1254 1255 1256 1257 1258 1259 1260 1261 1262 1263 1264 1265 1266 1267 1268 1269 1270 1271 1272 1273 1274 1275 1276 1277 1278 1279 1280 1281 1282 1283 1284 1285 1286 1287 1288 1289 1290