-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy pathreferences.bib
4574 lines (4268 loc) · 394 KB
/
references.bib
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
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
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
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
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
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
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
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
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
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
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
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
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
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
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
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
@article{kass_statistical_2005,
title = {Statistical issues in the analysis of neuronal data},
volume = {94},
doi = {10.1152/jn.00648.2004},
number = {1},
journal = {Journal of neurophysiology},
author = {Kass, Robert E. and Ventura, Valérie and Brown, Emery N.},
year = {2005},
note = {Publisher: American Physiological Society},
keywords = {⛔ No INSPIRE recid found},
pages = {8--25},
}
@misc{kohn_utah_2016,
title = {Utah array extracellular recordings of spontaneous and visually evoked activity from anesthetized macaque primary visual cortex ({V1}).},
url = {http://crcns.org/data-sets/vc/pvc-11},
language = {en},
urldate = {2022-12-19},
publisher = {CRCNS.org},
author = {Kohn, A. and Smith, M.A.},
year = {2016},
doi = {10.6080/K0NC5Z4X},
keywords = {Macaque, Neuroscience, Primary visual cortex, ⛔ No INSPIRE recid found},
}
@misc{stringer_mouselandrastermap_2020,
title = {{MouseLand}/rastermap: {A} multi-dimensional embedding algorithm},
url = {https://github.com/MouseLand/rastermap},
urldate = {2022-12-19},
author = {Stringer, Carsen},
year = {2020},
keywords = {⛔ No INSPIRE recid found},
}
@article{simoncelli_characterization_2003,
title = {Characterization of {Neural} {Responses} with {Stochastic} {Stimuli}},
url = {http://pillowlab.princeton.edu/pubs/simoncelli03c-preprint.pdf},
language = {en},
author = {Simoncelli, Eero P and Paninski, Liam and Pillow, Jonathan and Schwartz, Odelia},
year = {2003},
keywords = {⛔ No DOI found, ⛔ No INSPIRE recid found},
}
@inproceedings{deneve_bayesian_2004,
title = {Bayesian inference in spiking neurons},
volume = {17},
url = {https://papers.nips.cc/paper/2004/hash/cdd96eedd7f695f4d61802f8105ba2b0-Abstract.html},
abstract = {We propose a new interpretation of spiking neurons as Bayesian integra- tors accumulating evidence over time about events in the external world or the body, and communicating to other neurons their certainties about these events. In this model, spikes signal the occurrence of new infor- mation, i.e. what cannot be predicted from the past activity. As a result, firing statistics are close to Poisson, albeit providing a deterministic rep- resentation of probabilities. We proceed to develop a theory of Bayesian inference in spiking neural networks, recurrent interactions implement- ing a variant of belief propagation.},
urldate = {2022-12-19},
booktitle = {Advances in {Neural} {Information} {Processing} {Systems}},
publisher = {MIT Press},
author = {Deneve, Sophie},
year = {2004},
keywords = {⛔ No DOI found, ⛔ No INSPIRE recid found},
}
@article{ben-yishai_traveling_1997,
title = {Traveling {Waves} and the {Processing} of {Weakly} {Tuned} {Inputs} in a {Cortical} {Network} {Module}},
volume = {77},
author = {Ben-yishai, Rani and Hansel, David},
year = {1997},
note = {00000},
keywords = {\#nosource, orientation selectivity, population vector, primary visual cortex, ⛔ No INSPIRE recid found},
pages = {57--77},
}
@misc{nadafian_bio-plausible_2020,
title = {Bio-plausible {Unsupervised} {Delay} {Learning} for {Extracting} {Temporal} {Features} in {Spiking} {Neural} {Networks}},
url = {https://arxiv.org/abs/2011.09380},
abstract = {The plasticity of the conduction delay between neurons plays a fundamental role in learning. However, the exact underlying mechanisms in the brain for this modulation is still an open problem. Understanding the precise adjustment of synaptic delays could help us in developing effective brain-inspired computational models in providing aligned insights with the experimental evidence. In this paper, we propose an unsupervised biologically plausible learning rule for adjusting the synaptic delays in spiking neural networks. Then, we provided some mathematical proofs to show that our learning rule gives a neuron the ability to learn repeating spatio-temporal patterns. Furthermore, the experimental results of applying an STDP-based spiking neural network equipped with our proposed delay learning rule on Random Dot Kinematogram indicate the efficacy of the proposed delay learning rule in extracting temporal features.},
publisher = {arXiv},
author = {Nadafian, Alireza and Ganjtabesh, Mohammad},
month = nov,
year = {2020},
keywords = {Computer Science - Neural and Evolutionary Computing, Quantitative Biology - Neurons and Cognition, ⛔ No INSPIRE recid found},
}
@article{perrinet_sparse_2004,
series = {New {Aspects} in {Neurocomputing}: 10th {European} {Symposium} on {Artificial} {Neural} {Networks} 2002},
title = {Sparse spike coding in an asynchronous feed-forward multi-layer neural network using matching pursuit},
volume = {57},
issn = {0925-2312},
url = {https://www.sciencedirect.com/science/article/pii/S0925231204000670},
doi = {10.1016/j.neucom.2004.01.010},
abstract = {In order to account for the rapidity of visual processing, we explore visual coding strategies using a one-pass feed-forward spiking neural network. We based our model on the work of Van Rullen and Thorpe Neural Comput. 13 (6) (2001) 1255, which constructs a retinal representation using an orthogonal wavelet transform. This strategy provides a spike code, thanks to a rank order coding scheme which offers an alternative to the classical spike frequency coding scheme. We extended this model to efficient representations in arbitrary linear generative models by implementing lateral interactions on top of this feed-forward model. This method uses a matching pursuit scheme—recursively detecting in the image the best match with the elements of a dictionary and then subtracting it—and which may similarly define a visual spike code. In particular, this transform could be used with large and arbitrary dictionaries, so that we may define an over-complete representation which may define an efficient sparse spike coding scheme in arbitrary multi-layered architectures. We show here extensions of this method of computing with spike events, introducing an adaptive scheme leading to the emergence of V1-like receptive fields and then a model of bottom-up saliency pursuit.},
language = {en},
urldate = {2022-12-13},
journal = {Neurocomputing},
author = {Perrinet, Laurent and Samuelides, Manuel and Thorpe, Simon},
month = mar,
year = {2004},
keywords = {Natural images statistics, Parallel asynchronous processing, Sparse coding, Ultra-rapid categorization, Vision, Wavelet Hansform, ⛔ No INSPIRE recid found},
pages = {125--134},
}
@article{perrinet_coding_2004,
title = {Coding static natural images using spiking event times: do neurons cooperate?},
volume = {15},
doi = {10.1109/TNN.2004.833303},
number = {5},
journal = {IEEE Transactions on neural networks},
author = {Perrinet, Laurent and Samuelides, Manuel and Thorpe, Simon},
year = {2004},
note = {Publisher: IEEE},
keywords = {Biomembranes, Central nervous system, Filters, Fires, Image coding, Image reconstruction, Neurons, Retina, Statistics, Wavelet transforms, ⛔ No INSPIRE recid found},
pages = {1164--1175},
}
@article{konishi_coding_2003,
title = {Coding of auditory space},
volume = {26},
issn = {0147-006X},
doi = {10.1146/annurev.neuro.26.041002.131123},
abstract = {Behavioral, anatomical, and physiological approaches can be integrated in the study of sound localization in barn owls. Space representation in owls provides a useful example for discussion of place and ensemble coding. Selectivity for space is broad and ambiguous in low-order neurons. Parallel pathways for binaural cues and for different frequency bands converge on high-order space-specific neurons, which encode space more precisely. An ensemble of broadly tuned place-coding neurons may converge on a single high-order neuron to create an improved labeled line. Thus, the two coding schemes are not alternate methods. Owls can localize sounds by using either the isomorphic map of auditory space in the midbrain or forebrain neural networks in which space is not mapped.},
language = {eng},
journal = {Annual Review of Neuroscience},
author = {Konishi, Masakazu},
year = {2003},
pmid = {14527266},
keywords = {Animals, Auditory Pathways, Auditory Perception, Behavior, Animal, Brain, Brain Mapping, Forms and Records Control, Models, Neurological, Nerve Net, Neural Inhibition, Sound Localization, Space Perception, Strigiformes, ⛔ No INSPIRE recid found},
pages = {31--55},
}
@article{carr_circuit_1990,
title = {A circuit for detection of interaural time differences in the brain stem of the barn owl},
volume = {10},
copyright = {© 1990 by Society for Neuroscience},
issn = {0270-6474, 1529-2401},
url = {https://www.jneurosci.org/content/10/10/3227},
doi = {10.1523/JNEUROSCI.10-10-03227.1990},
abstract = {Detection of interaural time differences underlies azimuthal sound localization in the barn owl Tyto alba. Axons of the cochlear nucleus magnocellularis, and their targets in the binaural nucleus laminaris, form the circuit responsible for encoding these interaural time differences. The nucleus laminaris receives bilateral inputs from the cochlear nucleus magnocellularis such that axons from the ipsilateral cochlear nucleus enter the nucleus laminaris dorsally, while contralateral axons enter from the ventral side. This interdigitating projection to the nucleus laminaris is tonotopic, and the afferents are both sharply tuned and matched in frequency to the neighboring afferents. Recordings of phase-locked spikes in the afferents show an orderly change in the arrival time of the spikes as a function of distance from the point of their entry into the nucleus laminaris. The same range of conduction time (160 mu sec) was found over the 700-mu m depth of the nucleus laminaris for all frequencies examined (4-7.5 kHz) and corresponds to the range of interaural time differences available to the barn owl. The estimated conduction velocity in the axons is low (3-5 m/sec) and may be regulated by short internodal distances (60 mu m) within the nucleus laminaris. Neurons of the nucleus laminaris have large somata and very short dendrites. These cells are frequency selective and phase-lock to both monaural and binaural stimuli. The arrival time of phase-locked spikes in many of these neurons differs between the ipsilateral and contralateral inputs. When this disparity is nullified by imposition of an appropriate interaural time difference, the neurons respond maximally. The number of spikes elicited in response to a favorable interaural time difference is roughly double that elicited by a monaural stimulus. Spike counts for unfavorable interaural time differences fall well below monaural response levels. These findings indicate that the magnocellular afferents work as delay lines, and the laminaris neurons work as co- incidence detectors. The orderly distribution of conduction times, the predictability of favorable interaural time differences from monaural phase responses, and the pattern of the anatomical projection from the nucleus laminaris to the central nucleus of the inferior colliculus suggest that interaural time differences and their phase equivalents are mapped in each frequency band along the dorsoventral axis of the nucleus laminaris.},
language = {en},
number = {10},
urldate = {2022-12-16},
journal = {Journal of Neuroscience},
author = {Carr, C. E. and Konishi, M.},
month = oct,
year = {1990},
pmid = {2213141},
note = {Publisher: Society for Neuroscience
Section: Articles},
keywords = {⛔ No INSPIRE recid found},
pages = {3227--3246},
}
@article{bienenstock_model_1995,
title = {A model of neocortex},
volume = {6},
issn = {0954-898X},
url = {https://doi.org/10.1088/0954-898X_6_2_004},
doi = {10.1088/0954-898X_6_2_004},
abstract = {Prompted by considerations about (i) the compositionality of cognitive functions, (ii) the physiology of individual cortical neurons, (iii) the role of accurately timed spike patterns in cortex, and (iv) the regulation of global cortical activity, we suggest that the dynamics of cortex on the 1-ms time scale may be described as the activation of circuits of the synfire-chain type (Abeles 1982, 1991). We suggest that the fundamental computational unit in cortex may be a wave-like spatio-temporal pattern of synfire type, and that the binding mechanism underlying compositionality in cognition may be the accurate synchronization of synfire waves that propagate simultaneously on distinct, weakly coupled, synfire chains. We propose that Hebbian synaptic plasticity may result in a superposition of synfire chains in cortical connectivity, whereby a given neuron participates in many distinct chains. We investigate the behaviour of a much-simplified model of cortical dynamics devised along these principles. Calculations and numerical experiments are performed based on an assumption of randomness of stored chains, in the style of statistical physics. It is demonstrated that: (i) there exists a critical value for the total length of stored chains; (ii) this storage capacity is linear in the network's size; (iii) the behaviour of the network around the critical point is characterized by the self-regulation of the number of synfire waves coactive in the network at any given time.},
number = {2},
urldate = {2022-12-16},
journal = {Network: Computation in Neural Systems},
author = {Bienenstock, Elie},
month = jan,
year = {1995},
note = {Publisher: Taylor \& Francis
\_eprint: https://doi.org/10.1088/0954-898X\_6\_2\_004},
keywords = {⛔ No INSPIRE recid found},
pages = {179--224},
}
@article{lazar_time_2004,
title = {Time encoding with an integrate-and-fire neuron with a refractory period},
volume = {58-60},
issn = {09252312},
url = {https://linkinghub.elsevier.com/retrieve/pii/S0925231204000177},
doi = {10.1016/j.neucom.2004.01.022},
abstract = {Time encoding is a formal method of mapping amplitude information into a time sequence. We show that under simple conditions, bandlimited stimuli encoded with an integrate-and-ÿre neuron with an absolute refractory period can be recovered loss-free from the neural spike train at its output. We provide an algorithm for perfect recovery and derive conditions for its convergence. c 2003 Elsevier B.V. All rights reserved.},
language = {en},
urldate = {2022-12-19},
journal = {Neurocomputing},
author = {Lazar, Aurel A.},
month = jun,
year = {2004},
keywords = {⛔ No INSPIRE recid found},
pages = {53--58},
}
@article{golding_dendritic_2002,
title = {Dendritic spikes as a mechanism for cooperative long-term potentiation},
volume = {418},
copyright = {2002 Macmillan Magazines Ltd.},
issn = {1476-4687},
url = {https://www.nature.com/articles/nature00854},
doi = {10.1038/nature00854},
abstract = {Strengthening of synaptic connections following coincident pre- and postsynaptic activity was proposed by Hebb as a cellular mechanism for learning1. Contemporary models assume that multiple synapses must act cooperatively to induce the postsynaptic activity required for hebbian synaptic plasticity2,3,4,5. One mechanism for the implementation of this cooperation is action potential firing, which begins in the axon, but which can influence synaptic potentiation following active backpropagation into dendrites6. Backpropagation is limited, however, and action potentials often fail to invade the most distal dendrites7,8,9,10. Here we show that long-term potentiation of synapses on the distal dendrites of hippocampal CA1 pyramidal neurons does require cooperative synaptic inputs, but does not require axonal action potential firing and backpropagation. Rather, locally generated and spatially restricted regenerative potentials (dendritic spikes) contribute to the postsynaptic depolarization and calcium entry necessary to trigger potentiation of distal synapses. We find that this mechanism can also function at proximal synapses, suggesting that dendritic spikes participate generally in a form of synaptic potentiation that does not require postsynaptic action potential firing in the axon.},
language = {en},
number = {6895},
urldate = {2022-12-16},
journal = {Nature},
author = {Golding, Nace L. and Staff, Nathan P. and Spruston, Nelson},
month = jul,
year = {2002},
note = {Number: 6895
Publisher: Nature Publishing Group},
keywords = {Humanities and Social Sciences, Science, multidisciplinary, ⛔ No INSPIRE recid found},
pages = {326--331},
}
@article{mel_synaptic_2017,
title = {Synaptic plasticity in dendrites: complications and coping strategies},
volume = {43},
issn = {1873-6882},
shorttitle = {Synaptic plasticity in dendrites},
doi = {10.1016/j.conb.2017.03.012},
abstract = {The elaborate morphology, nonlinear membrane mechanisms and spatiotemporally varying synaptic activation patterns of dendrites complicate the expression, compartmentalization and modulation of synaptic plasticity. To grapple with this complexity, we start with the observation that neurons in different brain areas face markedly different learning problems, and dendrites of different neuron types contribute to the cell's input-output function in markedly different ways. By committing to specific assumptions regarding a neuron's learning problem and its input-output function, specific inferences can be drawn regarding the synaptic plasticity mechanisms and outcomes that we 'ought' to expect for that neuron. Exploiting this assumption-driven approach can help both in interpreting existing experimental data and designing future experiments aimed at understanding the brain's myriad learning processes.},
language = {eng},
journal = {Current Opinion in Neurobiology},
author = {Mel, Bartlett W. and Schiller, Jackie and Poirazi, Panayiota},
month = apr,
year = {2017},
pmid = {28453975},
keywords = {Dendrites, Humans, Learning, Models, Neurological, Neuronal Plasticity, Synapses, ⛔ No INSPIRE recid found},
pages = {177--186},
}
@article{jeffress_place_1948,
title = {A place theory of sound localization},
volume = {41},
issn = {0021-9940},
doi = {10.1037/h0061495},
abstract = {The author presents a place theory of sound localization based upon the time difference of stimulation of the 2 ears. The hypothesis depends upon the known slow rate of conduction in small nerve fibers and the phenomenon of spatial summation. It assumes that some secondary fibers of the auditory tract divide, sending branches homolaterally and contralaterally. There is a further assumption that these neurons make synaptic connection with other fibers on each side, the latter neurones synapsing with both contralateral and homolateral neurones. Then, if the sound is in a median plane, the summation effect would be maximal in a central group of the tertiary fibers on each side. If the sound source is shifted, the summation effect would result in a shifting of the transmission through synapses in the tertiary zone. This provides a spatial change in the pattern of nerve discharge as a consequence of a temporal change in the binaural stimulation. The anatomical location of such a center is suggested either in the inferior colliculus or the medial geniculate body. Possible experimental procedures are suggested. (PsycINFO Database Record (c) 2016 APA, all rights reserved)},
journal = {Journal of Comparative and Physiological Psychology},
author = {Jeffress, Lloyd A.},
year = {1948},
note = {Place: US
Publisher: American Psychological Association},
keywords = {Auditory Localization, Ear (Anatomy), Neurons, Temporal Frequency, ⛔ No INSPIRE recid found},
pages = {35--39},
}
@incollection{middlebrooks_chapter_2015,
series = {The {Human} {Auditory} {System}},
title = {Chapter 6 - {Sound} localization},
volume = {129},
url = {https://www.sciencedirect.com/science/article/pii/B9780444626301000068},
abstract = {The auditory system derives locations of sound sources from spatial cues provided by the interaction of sound with the head and external ears. Those cues are analyzed in specific brainstem pathways and then integrated as cortical representation of locations. The principal cues for horizontal localization are interaural time differences (ITDs) and interaural differences in sound level (ILDs). Vertical and front/back localization rely on spectral-shape cues derived from direction-dependent filtering properties of the external ears. The likely first sites of analysis of these cues are the medial superior olive (MSO) for ITDs, lateral superior olive (LSO) for ILDs, and dorsal cochlear nucleus (DCN) for spectral-shape cues. Localization in distance is much less accurate than that in horizontal and vertical dimensions, and interpretation of the basic cues is influenced by additional factors, including acoustics of the surroundings and familiarity of source spectra and levels. Listeners are quite sensitive to sound motion, but it remains unclear whether that reflects specific motion detection mechanisms or simply detection of changes in static location. Intact auditory cortex is essential for normal sound localization. Cortical representation of sound locations is highly distributed, with no evidence for point-to-point topography. Spatial representation is strictly contralateral in laboratory animals that have been studied, whereas humans show a prominent right-hemisphere dominance.},
language = {en},
urldate = {2022-12-16},
booktitle = {Handbook of {Clinical} {Neurology}},
publisher = {Elsevier},
author = {Middlebrooks, John C.},
editor = {Aminoff, Michael J. and Boller, François and Swaab, Dick F.},
month = jan,
year = {2015},
doi = {10.1016/B978-0-444-62630-1.00006-8},
keywords = {HRTF, Spatial hearing, auditory cortex, auditory motion, distance perception, interaural level difference, interaural time difference, precedence effect, superior colliculus, superior olivary complex, ⛔ No INSPIRE recid found},
pages = {99--116},
}
@article{rinberg_speed-accuracy_2006,
title = {Speed-{Accuracy} {Tradeoff} in {Olfaction}},
volume = {51},
issn = {0896-6273},
url = {https://www.sciencedirect.com/science/article/pii/S0896627306005538},
doi = {10.1016/j.neuron.2006.07.013},
abstract = {The basic psychophysical principle of speed-accuracy tradeoff (SAT) has been used to understand key aspects of neuronal information processing in vision and audition, but the principle of SAT is still debated in olfaction. In this study we present the direct observation of SAT in olfaction. We developed a behavioral paradigm for mice in which both the duration of odorant sampling and the difficulty of the odor discrimination task were controlled by the experimenter. We observed that the accuracy of odor discrimination increases with the duration of imposed odorant sampling, and that the rate of this increase is slower for harder tasks. We also present a unifying picture of two previous, seemingly disparate experiments on timing of odorant sampling in odor discrimination tasks. The presence of SAT in olfaction provides strong evidence for temporal integration in olfaction and puts a constraint on models of olfactory processing.},
language = {en},
number = {3},
urldate = {2022-12-16},
journal = {Neuron},
author = {Rinberg, Dmitry and Koulakov, Alexei and Gelperin, Alan},
month = aug,
year = {2006},
keywords = {SYSNEURO, ⛔ No INSPIRE recid found},
pages = {351--358},
}
@incollection{cleland_construction_2014,
title = {Construction of {Odor} {Representations} by {Olfactory} {Bulb} {Microcircuits}},
volume = {208},
isbn = {978-0-444-63350-7},
url = {https://linkinghub.elsevier.com/retrieve/pii/B9780444633507000073},
abstract = {Like other sensory systems, the olfactory system transduces specific features of the external environment and must construct an organized sensory representation from these highly fragmented inputs. As with these other systems, this representation is not accurate per se, but is constructed for utility, and emphasizes certain, presumably useful, features over others. I here describe the cellular and circuit mechanisms of the peripheral olfactory system that underlie this process of sensory construction, emphasizing the distinct architectures and properties of the two prominent computational layers in the olfactory bulb. Notably, while the olfactory system solves essentially similar conceptual problems to other sensory systems, such as contrast enhancement, activity normalization, and extending dynamic range, its peculiarities often require qualitatively different computational algorithms than are deployed in other sensory modalities. In particular, the olfactory modality is intrinsically high dimensional, and lacks a simple, externally defined basis analogous to wavelength or pitch on which elemental odor stimuli can be quantitatively compared. Accordingly, the quantitative similarities of the receptive fields of different odorant receptors (ORs) vary according to the statistics of the odor environment. To resolve these unusual challenges, the olfactory bulb appears to utilize unique nontopographical computations and intrinsic learning mechanisms to perform the necessary high-dimensional, similarity-dependent computations. In sum, the early olfactory system implements a coordinated set of early sensory transformations directly analogous to those in other sensory systems, but accomplishes these with unique circuit architectures adapted to the properties of the olfactory modality.},
language = {en},
urldate = {2022-12-16},
booktitle = {Progress in {Brain} {Research}},
publisher = {Elsevier},
author = {Cleland, Thomas A.},
year = {2014},
doi = {10.1016/B978-0-444-63350-7.00007-3},
keywords = {⛔ No INSPIRE recid found},
pages = {177--203},
}
@article{kashiwadani_synchronized_1999,
title = {Synchronized {Oscillatory} {Discharges} of {Mitral}/{Tufted} {Cells} {With} {Different} {Molecular} {Receptive} {Ranges} in the {Rabbit} {Olfactory} {Bulb}},
volume = {82},
issn = {0022-3077},
url = {https://journals.physiology.org/doi/full/10.1152/jn.1999.82.4.1786},
doi = {10.1152/jn.1999.82.4.1786},
abstract = {Individual glomeruli in the mammalian olfactory bulb represent a single or a few type(s) of odorant receptors. Signals from different types of receptors are thus sorted out into different glomeruli. How does the neuronal circuit in the olfactory bulb contribute to the combination and integration of signals received by different glomeruli? Here we examined electrophysiologically whether there were functional interactions between mitral/tufted cells associated with different glomeruli in the rabbit olfactory bulb. First, we made simultaneous recordings of extracellular single-unit spike responses of mitral/tufted cells and oscillatory local field potentials in the dorsomedial fatty acid–responsive region of the olfactory bulb in urethan-anesthetized rabbits. Using periodic artificial inhalation, the olfactory epithelium was stimulated with a homologous series ofn-fatty acids or n-aliphatic aldehydes. The odor-evoked spike discharges of mitral/tufted cells tended to phase-lock to the oscillatory local field potential, suggesting that spike discharges of many cells occur synchronously during odor stimulation. We then made simultaneous recordings of spike discharges from pairs of mitral/tufted cells located 300–500 μm apart and performed a cross-correlation analysis of their spike responses to odor stimulation. In ∼27\% of cell pairs examined, two cells with distinct molecular receptive ranges showed synchronized oscillatory discharges when olfactory epithelium was stimulated with one or a mixture of odorant(s) effective in activating both. The results suggest that the neuronal circuit in the olfactory bulb causes synchronized spike discharges of specific pairs of mitral/tufted cells associated with different glomeruli and the synchronization of odor-evoked spike discharges may contribute to the temporal binding of signals derived from different types of odorant receptor.},
number = {4},
urldate = {2022-12-16},
journal = {Journal of Neurophysiology},
author = {Kashiwadani, Hideki and Sasaki, Yasnory F. and Uchida, Naoshige and Mori, Kensaku},
month = oct,
year = {1999},
note = {Publisher: American Physiological Society},
keywords = {⛔ No INSPIRE recid found},
pages = {1786--1792},
}
@article{vidyasagar_multiple_1996,
title = {Multiple mechanisms underlying the orientation selectivity of visual cortical neurones},
volume = {19},
issn = {0166-2236},
url = {https://www.sciencedirect.com/science/article/pii/S016622369620027X},
doi = {10.1016/S0166-2236(96)20027-X},
abstract = {For over three decades, the mechanism of orientation selectivity of visual cortical neurones has been hotly debated. While intracortical inhibition has been implicated as playing a vital role, it has been difficult to observe it clearly. On the basis of recent findings, we propose a model in which the visual cortex brings together a number of different mechanisms for generating orientation-selective responses. Orientation biases in the thalamo-cortical input fibres provide an initial weak selectivity either directly in the excitatory input or by acting via cortical interneurones. This weak selectivity of postsynaptic potentials is then amplified by voltage-sensitive conductances of the cell membrane and excitatory and inhibitory intracortical circuitry, resulting in the sharp tuning seen in the spike discharges of visual cortical cells.},
language = {en},
number = {7},
urldate = {2022-12-16},
journal = {Trends in Neurosciences},
author = {Vidyasagar, T. R. and Pei, X. and Volgushev, M.},
month = jul,
year = {1996},
keywords = {⛔ No INSPIRE recid found},
pages = {272--277},
}
@article{ben-yishai_traveling_1996,
title = {Traveling {Waves} and the {Processing} of {Weakly} {Tuned} {Inputs} in a {Cortical} {Network} {Module}},
abstract = {Recent studies have shown that local cortical feedback can have an important effect on the response of neurons in primary visual cortex to the orientation of visual stimuli. In this work, we study the role of the cortical feedback in shaping the spatiotemporal patterns of activity in cortex. Two questions are addressed: one, what are the limitations on the ability of cortical neurons to lock their activity to rotating oriented stimuli within a single receptive field? Two, can the local architecture of visual cortex lead to the generation of spontaneous traveling pulses of activity? We study these issues analytically by a population-dynamic model of a hypercolumn in visual cortex. The order parameter that describes the macroscopic behavior of the network is the time-dependent population vector of the network. We first study the network dynamics under the influence of a weakly tuned input that slowly rotates within the receptive field. We show that if the cortical interactions have strong spatial modulation, the network generates a sharply tuned activity profile that propagates across the hypercolumn in a path that is completely locked to the stimulus rotation. The resultant rotating population vector maintains a constant angular lag relative to the stimulus, the magnitude of which grows with the stimulus rotation frequency. Beyond a critical frequency the population vector does not lock to the stimulus but executes a quasi-periodic motion with an average frequency that is smaller than that of the stimulus. In the second part we consider the stable intrinsic state of the cortex under the influence of isotropic stimulation. We show that if the local inhibitory feedback is sufficiently strong, the network does not settle into a stationary state but develops spontaneous traveling pulses of activity. Unlike recent models of wave propagation in cortical networks, the connectivity pattern in our model is spatially symmetric, hence the direction of propagation of these waves is arbitrary. The interaction of these waves with an external-oriented stimulus is studied. It is shown that the system can lock to a weakly tuned rotating stimulus if the stimulus frequency is close to the frequency of the intrinsic wave.},
language = {en},
journal = {Journal of Computational Neuroscience},
author = {Ben-Yishai, Rani and Hansel, David and Sompolinsky, Haim},
year = {1996},
keywords = {⛔ No DOI found, ⛔ No INSPIRE recid found},
}
@article{chagnac-amitai_horizontal_1989,
title = {Horizontal spread of synchronized activity in neocortex and its control by {GABA}-mediated inhibition},
volume = {61},
issn = {0022-3077},
url = {https://journals.physiology.org/doi/abs/10.1152/jn.1989.61.4.747},
doi = {10.1152/jn.1989.61.4.747},
abstract = {1. Suppression of GABAA receptor-mediated inhibition disrupts the neural activity of neocortex and can lead to synchronized discharges that mimic those of partial epilepsy. We have studied the role of GABAA-mediated inhibition in controlling the synchronization and horizontal (tangential) spread of cortical activity. 2. Slices of rat SmI were maintained in vitro and focally stimulated in layer VI while recording with a horizontal array of extracellular electrodes. Inhibition was slightly suppressed by adding low concentrations of the GABAA antagonists bicuculline or bicuculline methiodide to the bathing medium. Under control conditions neural activity was narrowly confined to a vertical strip of cortex. The horizontal spread of activity expanded about twofold in the presence of antagonist concentrations (less than or equal to 0.5 microM) that were expected to suppress GABAA function by no more than 10-20\%. 3. At antagonist concentrations between 0.4 and 1.0 microM, evoked epileptiform activity appeared. These threshold-dose epileptiform events showed wide variations in size and duration (even at the same recording site), very variable distances of horizontal propagation, specific sites of propagation failure, reversals of propagation direction, and directional asymmetries in their probability of propagation. This contrasts with activity observed previously (Ref. 9) in high bicuculline concentrations (greater than or equal to 10 microM): large, stereotyped events that propagate reliably without decrement or reflection. 4. Intracellular recordings were obtained from pyramidal neurons in layers II/III in the presence of less than or equal to 1 microM bicuculline. Inhibitory postsynaptic potentials (IPSPs) were observed during both primary evoked responses and propagating epileptiform events and were often comparable in size and duration to those in untreated cortex. Epileptiform field potentials were always correlated with synaptic activity in single cells, but the pattern and type of PSPs varied with the form of the field potentials. Large amplitude epileptiform events coincided with an overwhelming inhibition of upper layer neurons. 5. We conclude that 1) the horizontal spread of normal cortical activity is strongly constrained by GABAA-mediated IPSPs, 2) a relatively small reduction in the efficacy of inhibition leads to a large increase in the spread of excitation, 3) initiation and propagation of synchronized epileptiform activity can occur even in the presence of robust cortical inhibition, and 4) the character of epileptiform activity is strongly affected by the influences of inhibition.},
number = {4},
urldate = {2022-12-16},
journal = {Journal of Neurophysiology},
author = {Chagnac-Amitai, Y. and Connors, B. W.},
month = apr,
year = {1989},
note = {Publisher: American Physiological Society},
keywords = {⛔ No INSPIRE recid found},
pages = {747--758},
}
@article{ballard_dual_2011,
title = {Dual {Roles} for {Spike} {Signaling} in {Cortical} {Neural} {Populations}},
volume = {5},
issn = {1662-5188},
url = {https://www.frontiersin.org/articles/10.3389/fncom.2011.00022},
abstract = {A prominent feature of signaling in cortical neurons is that of randomness in the action potential. The output of a typical pyramidal cell can be well fit with a Poisson model, and variations in the Poisson rate repeatedly have been shown to be correlated with stimuli. However while the rate provides a very useful characterization of neural spike data, it may not be the most fundamental description of the signaling code. Recent data showing γ frequency range multi-cell action potential correlations, together with spike timing dependent plasticity, are spurring a re-examination of the classical model, since precise timing codes imply that the generation of spikes is essentially deterministic. Could the observed Poisson randomness and timing determinism reflect two separate modes of communication, or do they somehow derive from a single process? We investigate in a timing-based model whether the apparent incompatibility between these probabilistic and deterministic observations may be resolved by examining how spikes could be used in the underlying neural circuits. The crucial component of this model draws on dual roles for spike signaling. In learning receptive fields from ensembles of inputs, spikes need to behave probabilistically, whereas for fast signaling of individual stimuli, the spikes need to behave deterministically. Our simulations show that this combination is possible if deterministic signals using γ latency coding are probabilistically routed through different members of a cortical cell population at different times. This model exhibits standard features characteristic of Poisson models such as orientation tuning and exponential interval histograms. In addition, it makes testable predictions that follow from the γ latency coding.},
urldate = {2022-12-16},
journal = {Frontiers in Computational Neuroscience},
author = {Ballard, Dana and Jehee, Janneke},
year = {2011},
keywords = {⛔ No INSPIRE recid found},
}
@article{feller_dynamic_1997,
title = {Dynamic {Processes} {Shape} {Spatiotemporal} {Properties} of {Retinal} {Waves}},
volume = {19},
issn = {0896-6273},
url = {https://www.sciencedirect.com/science/article/pii/S089662730080940X},
doi = {10.1016/S0896-6273(00)80940-X},
abstract = {In the developing mammalian retina, spontaneous waves of action potentials are present in the ganglion cell layer weeks before vision. These waves are known to be generated by a synaptically connected network of amacrine cells and retinal ganglion cells, and exhibit complex spatiotemporal patterns, characterized by shifting domains of coactivation. Here, we present a novel dynamical model consisting of two coupled populations of cells that quantitatively reproduces the experimentally observed domain sizes, interwave intervals, and wavefront velocity profiles. Model and experiment together show that the highly correlated activity generated by retinal waves can be explained by a combination of random spontaneous activation of cells and the past history of local retinal activity.},
language = {en},
number = {2},
urldate = {2022-12-16},
journal = {Neuron},
author = {Feller, Marla B. and Butts, Daniel A. and Aaron, Holly L. and Rokhsar, Daniel S. and Shatz, Carla J.},
month = aug,
year = {1997},
keywords = {⛔ No INSPIRE recid found},
pages = {293--306},
}
@article{javanshir_advancements_2022,
title = {Advancements in {Algorithms} and {Neuromorphic} {Hardware} for {Spiking} {Neural} {Networks}},
volume = {34},
issn = {0899-7667},
url = {https://doi.org/10.1162/neco_a_01499},
doi = {10.1162/neco_a_01499},
abstract = {Artificial neural networks (ANNs) have experienced a rapid advancement for their success in various application domains, including autonomous driving and drone vision. Researchers have been improving the performance efficiency and computational requirement of ANNs inspired by the mechanisms of the biological brain. Spiking neural networks (SNNs) provide a power-efficient and brain-inspired computing paradigm for machine learning applications. However, evaluating large-scale SNNs on classical von Neumann architectures (central processing units/graphics processing units) demands a high amount of power and time. Therefore, hardware designers have developed neuromorphic platforms to execute SNNs in and approach that combines fast processing and low power consumption. Recently, field-programmable gate arrays (FPGAs) have been considered promising candidates for implementing neuromorphic solutions due to their varied advantages, such as higher flexibility, shorter design, and excellent stability. This review aims to describe recent advances in SNNs and the neuromorphic hardware platforms (digital, analog, hybrid, and FPGA based) suitable for their implementation. We present that biological background of SNN learning, such as neuron models and information encoding techniques, followed by a categorization of SNN training. In addition, we describe state-of-the-art SNN simulators. Furthermore, we review and present FPGA-based hardware implementation of SNNs. Finally, we discuss some future directions for research in this field.},
number = {6},
urldate = {2022-12-16},
journal = {Neural Computation},
author = {Javanshir, Amirhossein and Nguyen, Thanh Thi and Mahmud, M. A. Parvez and Kouzani, Abbas Z.},
month = may,
year = {2022},
keywords = {⛔ No INSPIRE recid found},
pages = {1289--1328},
}
@misc{fang_incorporating_2021,
title = {Incorporating {Learnable} {Membrane} {Time} {Constant} to {Enhance} {Learning} of {Spiking} {Neural} {Networks}},
url = {http://arxiv.org/abs/2007.05785},
doi = {10.48550/arXiv.2007.05785},
abstract = {Spiking Neural Networks (SNNs) have attracted enormous research interest due to temporal information processing capability, low power consumption, and high biological plausibility. However, the formulation of efficient and high-performance learning algorithms for SNNs is still challenging. Most existing learning methods learn weights only, and require manual tuning of the membrane-related parameters that determine the dynamics of a single spiking neuron. These parameters are typically chosen to be the same for all neurons, which limits the diversity of neurons and thus the expressiveness of the resulting SNNs. In this paper, we take inspiration from the observation that membrane-related parameters are different across brain regions, and propose a training algorithm that is capable of learning not only the synaptic weights but also the membrane time constants of SNNs. We show that incorporating learnable membrane time constants can make the network less sensitive to initial values and can speed up learning. In addition, we reevaluate the pooling methods in SNNs and find that max-pooling will not lead to significant information loss and have the advantage of low computation cost and binary compatibility. We evaluate the proposed method for image classification tasks on both traditional static MNIST, Fashion-MNIST, CIFAR-10 datasets, and neuromorphic N-MNIST, CIFAR10-DVS, DVS128 Gesture datasets. The experiment results show that the proposed method outperforms the state-of-the-art accuracy on nearly all datasets, using fewer time-steps. Our codes are available at https://github.com/fangwei123456/Parametric-Leaky-Integrate-and-Fire-Spiking-Neuron.},
urldate = {2022-12-15},
publisher = {arXiv},
author = {Fang, Wei and Yu, Zhaofei and Chen, Yanqi and Masquelier, Timothee and Huang, Tiejun and Tian, Yonghong},
month = aug,
year = {2021},
note = {arXiv:2007.05785 [cs]},
keywords = {Computer Science - Computer Vision and Pattern Recognition, Computer Science - Machine Learning, Computer Science - Neural and Evolutionary Computing, ⛔ No INSPIRE recid found},
}
@article{vinje_sparse_2000,
title = {Sparse {Coding} and {Decorrelation} in {Primary} {Visual} {Cortex} {During} {Natural} {Vision}},
volume = {287},
url = {https://www.science.org/doi/10.1126/science.287.5456.1273},
doi = {10.1126/science.287.5456.1273},
abstract = {Theoretical studies suggest that primary visual cortex (area V1) uses a sparse code to efficiently represent natural scenes. This issue was investigated by recording from V1 neurons in awake behaving macaques during both free viewing of natural scenes and conditions simulating natural vision. Stimulation of the nonclassical receptive field increases the selectivity and sparseness of individual V1 neurons, increases the sparseness of the population response distribution, and strongly decorrelates the responses of neuron pairs. These effects are due to both excitatory and suppressive modulation of the classical receptive field by the nonclassical receptive field and do not depend critically on the spatiotemporal structure of the stimuli. During natural vision, the classical and nonclassical receptive fields function together to form a sparse representation of the visual world. This sparse code may be computationally efficient for both early vision and higher visual processing.},
number = {5456},
urldate = {2022-12-15},
journal = {Science},
author = {Vinje, William E. and Gallant, Jack L.},
month = feb,
year = {2000},
note = {Publisher: American Association for the Advancement of Science},
keywords = {⛔ No INSPIRE recid found},
pages = {1273--1276},
}
@article{decharms_primary_1996,
title = {Primary cortical representation of sounds by the coordination of action-potential timing},
volume = {381},
issn = {1476-4687},
url = {http://www.nature.com/articles/381610a0},
doi = {10.1038/381610a0},
abstract = {CORTICAL population coding could in principle rely on either the mean rate of neuronal action potentials, or the relative timing of action potentials, or both. When a single sensory stimulus drives many neurons to fire at elevated rates, the spikes of these neurons become tightly synchronized1,2, which could be involved in 'binding' together individual firing-rate feature representations into a unified object percept3. Here we demonstrate that the relative timing of cortical action potentials can signal stimulus features themselves, a function even more basic than feature grouping. Populations of neurons in the primary auditory cortex can coordinate the relative timing of their action potentials such that spikes occur closer together in time during continuous stimuli. In this way cortical neurons can signal stimuli even when their firing rates do not change. Population coding based on relative spike timing can systematically signal stimulus features, it is topographically mapped, and it follows the stimulus time course even where mean firing rate does not.},
language = {en},
number = {6583},
urldate = {2022-12-15},
journal = {Nature},
author = {deCharms, R. Christopher and Merzenich, Michael M.},
month = jun,
year = {1996},
note = {Number: 6583
Publisher: Nature Publishing Group},
keywords = {Humanities and Social Sciences, Science, multidisciplinary, ⛔ No INSPIRE recid found},
pages = {610--613},
}
@article{maunsell_functional_1983,
title = {Functional properties of neurons in middle temporal visual area of the macaque monkey. {I}. {Selectivity} for stimulus direction, speed, and orientation},
volume = {49},
issn = {0022-3077, 1522-1598},
url = {https://www.physiology.org/doi/10.1152/jn.1983.49.5.1127},
doi = {10.1152/jn.1983.49.5.1127},
abstract = {1. Recordings were made from single units in the middle temporal visual area (MT) of anesthetized, paralyzed macaque monkeys. A computer-driven stimulator was used to make quantitative tests of selectivity for stimulus direction, speed, and orientation. The data were taken from 168 units that were histologically identified as being in MT. 2. The results confirm previous reports of a high degree of direction selectivity in MT. The response above background to stimuli moving in a unit's preferred direction was, an average, 10.9 times that to stimuli moving in the opposite direction. There was a marked tendency for nearby units to have similar preferred directions. 3. Most units were also sharply tuned for the speed of stimulus motion. For some cells the response fell to less than half-maximal at speeds only a factor of two from the optimum; on average, responses were greater than half-maximal only over a 7.7-fold range of speed. The distribution of preferred speeds for different units was unimodal, with a peak near 32 degrees/s; the total range of preferred speeds extended from 2 to 256 degrees/s. Nearby units generally responded best to similar speeds of motion. 4. Most units in MT showed selectivity for stimulus orientation when tested with stationary, flashed bars. However, stationary stimuli generally elicited only brief responses; when averaged over the duration of the stimulus, the responses were much less than those to moving stimuli. The preferred orientation was usually, but not always, perpendicular to the preferred direction of movement. 5. A comparison of the results of the present study with a previous quantitative investigation in the owl monkey shows a striking similarity in response properties in MT of the two species. 6. The presence of both direction and speed selectivity in MT of the macaque suggests that this area is more specialized for the analysis of visual motion than has been previously recognized.},
language = {en},
number = {5},
urldate = {2022-12-15},
journal = {Journal of Neurophysiology},
author = {Maunsell, J. H. and Van Essen, D. C.},
month = may,
year = {1983},
keywords = {⛔ No INSPIRE recid found},
pages = {1127--1147},
}
@article{montemurro_phase--firing_2008,
title = {Phase-of-{Firing} {Coding} of {Natural} {Visual} {Stimuli} in {Primary} {Visual} {Cortex}},
volume = {18},
issn = {0960-9822},
url = {http://www.cell.com/current-biology/abstract/S0960-9822(08)00168-1},
doi = {10.1016/j.cub.2008.02.023},
language = {English},
number = {5},
urldate = {2022-12-15},
journal = {Current Biology},
author = {Montemurro, Marcelo A. and Rasch, Malte J. and Murayama, Yusuke and Logothetis, Nikos K. and Panzeri, Stefano},
month = mar,
year = {2008},
pmid = {18328702},
note = {Publisher: Elsevier},
keywords = {SYSNEURO, ⛔ No INSPIRE recid found},
pages = {375--380},
}
@article{butts_temporal_2007,
title = {Temporal precision in the neural code and the timescales of natural vision},
volume = {449},
issn = {1476-4687},
url = {http://www.nature.com/articles/nature06105},
doi = {10.1038/nature06105},
abstract = {In mammalian visual system, spikes evoked by visual stimuli have millisecond-scale timing even though the relevant timescales of visual processing themselves are much slower. It has therefore long been debated whether spike timing itself carries some form of the neural code. Now experiments in the lateral geniculate nucleus of cats, the part of the brain that is the primary processor of visual information, show that spike timing precision is not absolute for all classes of visual stimuli. Rather, the degree of precision is relative to the timescale of the stimulus, and this relatively high level of precision is required to construct an accurate representation of the stimulus.},
language = {en},
number = {7158},
urldate = {2022-12-15},
journal = {Nature},
author = {Butts, Daniel A. and Weng, Chong and Jin, Jianzhong and Yeh, Chun-I. and Lesica, Nicholas A. and Alonso, Jose-Manuel and Stanley, Garrett B.},
month = sep,
year = {2007},
note = {Number: 7158
Publisher: Nature Publishing Group},
keywords = {Humanities and Social Sciences, Science, multidisciplinary, ⛔ No INSPIRE recid found},
pages = {92--95},
}
@article{gouras_graded_1960,
title = {Graded potentials of bream retina},
volume = {152},
issn = {0022-3751},
url = {https://www.ncbi.nlm.nih.gov/pmc/articles/PMC1363334/},
abstract = {Images
null},
number = {3},
urldate = {2022-12-14},
journal = {The Journal of Physiology},
author = {Gouras, P.},
month = jul,
year = {1960},
pmid = {13828605},
pmcid = {PMC1363334},
keywords = {⛔ No INSPIRE recid found},
pages = {487--505},
}
@article{bryant_spike_1976,
title = {Spike initiation by transmembrane current: a white-noise analysis.},
volume = {260},
copyright = {© 1976 The Physiological Society},
issn = {1469-7793},
shorttitle = {Spike initiation by transmembrane current},
url = {https://onlinelibrary.wiley.com/doi/abs/10.1113/jphysiol.1976.sp011516},
doi = {10.1113/jphysiol.1976.sp011516},
abstract = {1. Those features of a transmembrane current correlated with spike initiation were examined in Aplysia neurones using a Gaussian white-noise stimulus. This stimulus has the advantages that it presents numerous wave forms in random order without prejudgement as to their efficacies, and that it allows straightforward statistical calculations. 2. Stimulation with a repeating segment of Gaussian white-noise current revealed remarkable invariance in the firing times of the tested neurones and indicated a high degree of reliability of their response. 3. Frequencies (less than 5 Hz) involved in spike triggering propagated faithfully for up to several millimetres, justifying intrasomatic current injection to examine spike initiation at the trigger locus. 4. Examination of current wave forms preceding spikes indicated that a wide variety could be effective. Hence, a statistical analysis was performed, including computation of probability densities, averages, standard deviations and correlation coefficients of pairs of current values. Each statistic was displayed as a function of time before the spike. 5. The average current trajectory preceding a spike was multiphasic and depended on the presence and polarity of a d.c. bias. An early relatively small inward- or outward-going phase was followed by a large outward phase before the spike. The early phase tended to oppose the polarity of the d.c. bias. 6. The late outward phase of the average current trajectory reached a maximum 40–75 msec before triggering the action potential (AP) and returned to near zero values at the moment of triggering. The fact that the current peak occurs in advance of the AP may be partially explained by a phase delay between the transmembrane current and potential. The failure of the average current trajectory to return to control values immediately following the peak argues for a positive role of the declining phase in spike triggering. 7. Probability densities preceding spikes were Gaussian, indicating that the average was also the most probable value. Although the densities were broad, confirming that spikes were preceded by a wide variety of current wave forms, their standard deviations were reduced significantly with respect to controls, suggesting preferred status of the average current trajectory in spike triggering. 8. The matrix of correlation coefficients between current pairs suggested that spikes tended to be preceded by wave forms that in part kept close to the average current trajectory and in part preserved its shape. 9. The average first and second derivatives of spike-evoking epochs revealed that current slope and acceleration, respectively, were most crucial in the last 200 msec before spike triggering, and that these dynamic stimulus components were more important for a cell maintained under a depolarizing, rather than a hyperpolarizing bias. 10...},
language = {en},
number = {2},
urldate = {2022-12-13},
journal = {The Journal of Physiology},
author = {Bryant, H L and Segundo, J P},
year = {1976},
note = {\_eprint: https://onlinelibrary.wiley.com/doi/pdf/10.1113/jphysiol.1976.sp011516},
keywords = {⛔ No INSPIRE recid found},
pages = {279--314},
}
@article{perrinet_feature_2004,
series = {Decoding and interfacing the brain: from neuronal assemblies to cyborgs},
title = {Feature detection using spikes: {The} greedy approach},
volume = {98},
issn = {0928-4257},
shorttitle = {Feature detection using spikes},
url = {https://www.sciencedirect.com/science/article/pii/S0928425705000161},
doi = {10.1016/j.jphysparis.2005.09.012},
abstract = {A goal of low-level neural processes is to build an efficient code extracting the relevant information from the sensory input. It is believed that this is implemented in cortical areas by elementary inferential computations dynamically extracting the most likely parameters corresponding to the sensory signal. We explore here a neuro-mimetic feed-forward model of the primary visual area (VI) solving this problem in the case where the signal may be described by a robust linear generative model. This model uses an over-complete dictionary of primitives which provides a distributed probabilistic representation of input features. Relying on an efficiency criterion, we derive an algorithm as an approximate solution which uses incremental greedy inference processes. This algorithm is similar to ‘Matching Pursuit’ and mimics the parallel architecture of neural computations. We propose here a simple implementation using a network of spiking integrate-and-fire neurons which communicate using lateral interactions. Numerical simulations show that this Sparse Spike Coding strategy provides an efficient model for representing visual data from a set of natural images. Even though it is simplistic, this transformation of spatial data into a spatio-temporal pattern of binary events provides an accurate description of some complex neural patterns observed in the spiking activity of biological neural networks.},
language = {en},
number = {4},
urldate = {2022-12-13},
journal = {Journal of Physiology-Paris},
author = {Perrinet, Laurent},
month = jul,
year = {2004},
keywords = {Distributed probabilistic representation, Inverse linear model, Matching pursuit, Neuronal representation, Over-complete dictionaries, Sparse spike coding, Spike-event computation, ⛔ No INSPIRE recid found},
pages = {530--539},
}
@article{perrinet_sparse_2002,
title = {Sparse {Image} {Coding} {Using} an {Asynchronous} {Spiking} {Neural} {Network}},
abstract = {In order to explore coding strategies in the retina, we use a wavelet-like transform which output is sparse, as is observed in biological retinas [4]. This transform is defined in the context of a one-pass feed-forward spiking neural network, and the output is the list of its neurons' spikes: it is recursively constructed using a greedy matching pursuit scheme which first selects higher contrast energy values.},
author = {Perrinet, Laurent and Samuelides, Manuel},
month = aug,
year = {2002},
keywords = {⛔ No DOI found, ⛔ No INSPIRE recid found},
}
@book{dilorenzo_spike_2013,
title = {Spike {Timing}: {Mechanisms} and {Function}},
isbn = {978-1-4398-3815-0},
shorttitle = {Spike {Timing}},
abstract = {Neuronal communication forms the basis for all behavior, from the smallest movement to our grandest thought processes. Among the many mechanisms that support these functions, spike timing is among the most powerful and—until recently—perhaps the least studied. In the last two decades, however, the study of spike timing has exploded. The heightened interest is due to several factors. These include the development of physiological tools for measuring the activity of neural ensembles and analytical tools for assessing and characterizing spike timing. These advances are coupled with a growing appreciation of spike timing’s theoretical importance for the design principles of the brain. Spike Timing: Mechanisms and Function examines the function of spike timing in sensory, motor, and integrative processes, providing readers with a broad perspective on how spike timing is produced and used by the nervous system. It brings together the work and ideas of leaders in the field to address current thinking as well as future possibilities. The first section of the book describes the foundation for quantitative analysis and theory. It examines the information contained in spike timing, how it can be quantified, and how neural systems can extract it. The second section explores how input-output relationships are reflected in spike timing across a range of sensory systems. Drawing together multiple perspectives, including theoretical and computational studies as well as experimental studies in a range of model systems, the book provides a firm background for investigators to consider spike timing as it applies to their own work. It also offers a glimpse of future advances related to mechanisms of spike timing and its role in neural function, such as the development of novel computational technologies.},
language = {en},
publisher = {CRC Press},
author = {DiLorenzo, Patricia M. and Victor, Jonathan D.},
month = may,
year = {2013},
note = {Google-Books-ID: KTHUIMUpQCUC},
keywords = {Computers / Software Development \& Engineering / Systems Analysis \& Design, Medical / Biotechnology, Science / Life Sciences / Biophysics, Science / Life Sciences / Neuroscience, Science / Physics / General, Technology \& Engineering / Biomedical, ⛔ No INSPIRE recid found},
}
@inproceedings{arnold_conduction_2021,
title = {Conduction delay plasticity can robustly learn spatiotemporal patterns embedded in noise},
doi = {10.1109/IJCNN52387.2021.9533934},
abstract = {Noise and temporal dynamics are ubiquitous in neural systems yet the computational consequences of these two phenomena interacting are not well studied. Temporal dynamics in spiking networks are often considered only implicitly as part of membrane time constants or synaptic transfer functions. We explicitly model temporal structure using plastic conduction delays between neuron's and characterise the influence of different kinds of noise on learning including temporal jitter, dropout, pattern size, and pattern presentation frequency. We simplify the conduction delay plasticity (CDP) rule called synaptic delay variance learning (SDVL) and demonstrate it is robust to several kinds of noise including; internal pattern jitter, number of pattern afferents, pattern presentation rate, and pattern spike dropout. In particular, after unsupervised training the simplified version of SDVL can achieve an accuracy of up to 99.7 percent averaged over 100 trials. These results demonstrate that learning algorithms based on explicitly modelling temporal structure in inputs can be functional and robust for unsupervised learning of spatiotemporal patterns across a range of noise conditions.},
booktitle = {2021 {International} {Joint} {Conference} on {Neural} {Networks} ({IJCNN})},
author = {Arnold, Joshua and Stratton, Peter and Wiles, Janet},
month = jul,
year = {2021},
note = {ISSN: 2161-4407},
keywords = {Biological systems, Conduction delay, Delays, Jitter, Noise measurement, Spatiotemporal phenomena, Training, Transfer functions, delay learning, noise robust, plasticity, spiking neural network, ⛔ No INSPIRE recid found},
pages = {1--10},
}
@article{rubinsky_spatio-temporal_2008,
title = {Spatio-temporal motifs ‘remembered’ in neuronal networks following profound hypothermia},
volume = {21},
issn = {0893-6080},
url = {https://www.sciencedirect.com/science/article/pii/S0893608008001202},
doi = {10.1016/j.neunet.2008.06.008},
abstract = {Surgical procedures using hypothermic temperatures have been linked to complications such as seizures, impaired mental development and impaired memory. Although there is some evidence that the profound hypothermia ({\textless}12 ∘C) used in these procedures may be contributing to these neurological impairments, skepticism remains because of lack of evidence from experimental studies isolating the effects of hypothermia on neuronal networks. In order to attain a better understanding of profound hypothermia effects on neurons during surgical procedures, we applied cold to a cultured in-vitro neuronal network. The typical pattern of activity of such cultures is in the form of synchronized bursts, in which most of the recorded neurons fire action potentials in a short time period. In most cases, the bursting activity shows one or more repeating precise spatio-temporal patterns (motifs) that are sustained over long periods of time. In this experimental study, neuronal networks grown on microelectrode arrays (MEA) are subjected to profound hypothermia for an hour and the collective dynamics of the network as a whole are assessed. We show, by using a similarity analysis that compares changes in the time delays between neuronal activation at different burst motifs, that neuronal networks survive total inhibition by profound hypothermia and retain their intrinsic synchronized burst motifs even with substantial generalized neuronal degeneration. By applying multiple sessions of cold, we also show a marked monotonic reduction in the rate of burst firing and in the number of spikes of each neuron after each session.},
language = {en},
number = {9},
urldate = {2022-11-19},
journal = {Neural Networks},
author = {Rubinsky, Liel and Raichman, Nadav and Lavee, Jacob and Frenk, Hanan and Ben-Jacob, Eshel},
month = nov,
year = {2008},
keywords = {Low temperature, Microelectrode arrays, Neuronal cultures, Synchronized bursting events, ⛔ No INSPIRE recid found},
pages = {1232--1237},
}
@article{coull_distinction_2022,
title = {The distinction between temporal order and duration processing, and implications for schizophrenia},
volume = {1},
doi = {10.1038/s44159-022-00038-y},
number = {5},
journal = {Nature Reviews Psychology},
author = {Coull, Jennifer T. and Giersch, Anne},
year = {2022},
note = {Publisher: Nature Publishing Group},
keywords = {⛔ No INSPIRE recid found},
pages = {257--271},
}
@article{furber_overview_2013,
title = {Overview of the {SpiNNaker} {System} {Architecture}},
volume = {62},
issn = {0018-9340},
url = {http://ieeexplore.ieee.org/document/6226357/},
doi = {10.1109/TC.2012.142},
number = {12},
urldate = {2022-11-13},
journal = {IEEE Transactions on Computers},
author = {Furber, Steve B. and Lester, David R. and Plana, Luis A. and Garside, Jim D. and Painkras, Eustace and Temple, Steve and Brown, Andrew D.},
month = dec,
year = {2013},
keywords = {⛔ No INSPIRE recid found},
pages = {2454--2467},
}
@article{merolla_million_2014,
title = {A million spiking-neuron integrated circuit with a scalable communication network and interface},
volume = {345},
issn = {0036-8075, 1095-9203},
url = {https://www.science.org/doi/10.1126/science.1254642},
doi = {10.1126/science.1254642},
abstract = {Modeling computer chips on real brains
Computers are nowhere near as versatile as our own brains. Merolla
et al.
applied our present knowledge of the structure and function of the brain to design a new computer chip that uses the same wiring rules and architecture. The flexible, scalable chip operated efficiently in real time, while using very little power.
Science
, this issue p.
668
,
A large-scale computer chip mimics many features of a real brain.
,
Inspired by the brain’s structure, we have developed an efficient, scalable, and flexible non–von Neumann architecture that leverages contemporary silicon technology. To demonstrate, we built a 5.4-billion-transistor chip with 4096 neurosynaptic cores interconnected via an intrachip network that integrates 1 million programmable spiking neurons and 256 million configurable synapses. Chips can be tiled in two dimensions via an interchip communication interface, seamlessly scaling the architecture to a cortexlike sheet of arbitrary size. The architecture is well suited to many applications that use complex neural networks in real time, for example, multiobject detection and classification. With 400-pixel-by-240-pixel video input at 30 frames per second, the chip consumes 63 milliwatts.},
language = {en},
number = {6197},
urldate = {2022-11-13},
journal = {Science},
author = {Merolla, Paul A. and Arthur, John V. and Alvarez-Icaza, Rodrigo and Cassidy, Andrew S. and Sawada, Jun and Akopyan, Filipp and Jackson, Bryan L. and Imam, Nabil and Guo, Chen and Nakamura, Yutaka and Brezzo, Bernard and Vo, Ivan and Esser, Steven K. and Appuswamy, Rathinakumar and Taba, Brian and Amir, Arnon and Flickner, Myron D. and Risk, William P. and Manohar, Rajit and Modha, Dharmendra S.},
month = aug,
year = {2014},
keywords = {⛔ No INSPIRE recid found},
pages = {668--673},
}
@book{furber_spinnaker_2020,
title = {{SpiNNaker}: {A} {Spiking} {Neural} {Network} {Architecture}},
isbn = {978-1-68083-652-3 978-1-68083-653-0},
shorttitle = {{SpiNNaker}},
url = {https://nowpublishers.com/article/BookDetails/9781680836523},
urldate = {2022-11-13},
publisher = {Now Publishers},
editor = {Furber, Steve and Bogdan, Petrut},
year = {2020},
doi = {10.1561/9781680836523},
keywords = {⛔ No INSPIRE recid found},
}
@article{davies_loihi_2018,
title = {Loihi: {A} {Neuromorphic} {Manycore} {Processor} with {On}-{Chip} {Learning}},
volume = {38},
issn = {0272-1732, 1937-4143},
shorttitle = {Loihi},
url = {https://ieeexplore.ieee.org/document/8259423/},
doi = {10.1109/MM.2018.112130359},
number = {1},
urldate = {2022-11-13},
journal = {IEEE Micro},
author = {Davies, Mike and Srinivasa, Narayan and Lin, Tsung-Han and Chinya, Gautham and Cao, Yongqiang and Choday, Sri Harsha and Dimou, Georgios and Joshi, Prasad and Imam, Nabil and Jain, Shweta and Liao, Yuyun and Lin, Chit-Kwan and Lines, Andrew and Liu, Ruokun and Mathaikutty, Deepak and McCoy, Steven and Paul, Arnab and Tse, Jonathan and Venkataramanan, Guruguhanathan and Weng, Yi-Hsin and Wild, Andreas and Yang, Yoonseok and Wang, Hong},
month = jan,
year = {2018},
keywords = {⛔ No INSPIRE recid found},
pages = {82--99},
}
@article{fields_myelin_2020,
title = {Myelin makes memories},
volume = {23},
issn = {1546-1726},
doi = {10.1038/s41593-020-0606-x},
language = {eng},
number = {4},
journal = {Nature Neuroscience},
author = {Fields, R. Douglas and Bukalo, Olena},
month = apr,
year = {2020},
pmid = {32094969},
keywords = {Animals, Memory, Memory Consolidation, Mice, Myelin Sheath, ⛔ No INSPIRE recid found},
pages = {469--470},
}
@article{madadi_asl_dendritic_2018,
title = {Dendritic and {Axonal} {Propagation} {Delays} {May} {Shape} {Neuronal} {Networks} {With} {Plastic} {Synapses}},
volume = {9},
issn = {1664-042X},
doi = {10.3389/fphys.2018.01849},
abstract = {Biological neuronal networks are highly adaptive and plastic. For instance, spike-timing-dependent plasticity (STDP) is a core mechanism which adapts the synaptic strengths based on the relative timing of pre- and postsynaptic spikes. In various fields of physiology, time delays cause a plethora of biologically relevant dynamical phenomena. However, time delays increase the complexity of model systems together with the computational and theoretical analysis burden. Accordingly, in computational neuronal network studies propagation delays were often neglected. As a downside, a classic STDP rule in oscillatory neurons without propagation delays is unable to give rise to bidirectional synaptic couplings, i.e., loops or uncoupled states. This is at variance with basic experimental results. In this mini review, we focus on recent theoretical studies focusing on how things change in the presence of propagation delays. Realistic propagation delays may lead to the emergence of neuronal activity and synaptic connectivity patterns, which cannot be captured by classic STDP models. In fact, propagation delays determine the inventory of attractor states and shape their basins of attractions. The results reviewed here enable to overcome fundamental discrepancies between theory and experiments. Furthermore, these findings are relevant for the development of therapeutic brain stimulation techniques aiming at shifting the diseased brain to more favorable attractor states.},
language = {eng},
journal = {Frontiers in Physiology},
author = {Madadi Asl, Mojtaba and Valizadeh, Alireza and Tass, Peter A.},
year = {2018},
pmid = {30618847},
pmcid = {PMC6307091},
keywords = {living systems, mathematical modeling, propagation delays, spike-timing-dependent plasticity, synchronization, ⛔ No INSPIRE recid found},
pages = {1849},
}
@article{neftci_surrogate_2019,
title = {Surrogate {Gradient} {Learning} in {Spiking} {Neural} {Networks}: {Bringing} the {Power} of {Gradient}-{Based} {Optimization} to {Spiking} {Neural} {Networks}},
volume = {36},
issn = {1053-5888, 1558-0792},
shorttitle = {Surrogate {Gradient} {Learning} in {Spiking} {Neural} {Networks}},
url = {https://ieeexplore.ieee.org/document/8891809/},
doi = {10.1109/MSP.2019.2931595},
number = {6},
urldate = {2022-11-14},
journal = {IEEE Signal Processing Magazine},
author = {Neftci, Emre O. and Mostafa, Hesham and Zenke, Friedemann},
month = nov,
year = {2019},
keywords = {⛔ No INSPIRE recid found},
pages = {51--63},
}
@article{fields_new_2015,
title = {A new mechanism of nervous system plasticity: activity-dependent myelination},
volume = {16},
copyright = {2015 Nature Publishing Group, a division of Macmillan Publishers Limited. All Rights Reserved.},
issn = {1471-0048},
shorttitle = {A new mechanism of nervous system plasticity},
url = {https://www.nature.com/articles/nrn4023},
doi = {10.1038/nrn4023},
abstract = {The precise timing of impulse transmission along axons is crucial for synaptic plasticity and brain oscillations, and is partly determined by myelin thickness. In this Opinion article, R. Douglas Fields discusses how electrical activity influences myelin thickness and thus conduction velocity and circuit properties.},
language = {en},
number = {12},
urldate = {2021-01-07},
journal = {Nature Reviews Neuroscience},
author = {Fields, R. Douglas},
month = dec,
year = {2015},
note = {Number: 12
Publisher: Nature Publishing Group},
keywords = {biology, delay-learning, myelination, ⛔ No INSPIRE recid found},
pages = {756--767},
}
@article{steadman_disruption_2020,
title = {Disruption of {Oligodendrogenesis} {Impairs} {Memory} {Consolidation} in {Adult} {Mice}},
volume = {105},
issn = {0896-6273},
url = {https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7579726/},
doi = {10.1016/j.neuron.2019.10.013},
abstract = {The generation of myelin-forming oligodendrocytes persists throughout life and is regulated by neural activity. Here we tested whether experience-driven changes in oligodendrogenesis are important for memory consolidation. We found that water maze learning promotes oligodendrogenesis and de
novo myelination in the cortex and associated white matter tracts. Preventing these learning-induced increases in oligodendrogenesis without affecting existing oligodendrocytes impaired memory consolidation of water maze, as well as contextual fear, memories. These results suggest that de
novo myelination tunes activated circuits, promoting coordinated activity that is important for memory consolidation. Consistent with this, contextual fear learning increased the coupling of hippocampal sharp wave ripples and cortical spindles, and these learning-induced increases in ripple-spindle coupling were blocked when oligodendrogenesis was suppressed. Our results identify a non-neuronal form of plasticity that remodels hippocampal-cortical networks following learning and is required for memory consolidation.,
, Experience-dependent de
novo myelination may fine-tune activated circuits by promoting brain synchrony, important for memory consolidation. Steadman et al. find that blocking this form of adaptive myelination prevents learning-induced increases in coordinated activity and impairs memory consolidation.},
number = {1},
urldate = {2022-11-14},
journal = {Neuron},
author = {Steadman, Patrick E. and Xia, Frances and Ahmed, Moriam and Mocle, Andrew J. and Penning, Amber R.A. and Geraghty, Anna C. and Steenland, Hendrik W. and Monje, Michelle and Josselyn, Sheena A. and Frankland, Paul W.},
month = jan,
year = {2020},
pmid = {31753579},
pmcid = {PMC7579726},
keywords = {⛔ No INSPIRE recid found},
pages = {150--164.e6},
}
@article{pan_preservation_2020,
title = {Preservation of a remote fear memory requires new myelin formation},
volume = {23},
issn = {1546-1726},
doi = {10.1038/s41593-019-0582-1},
abstract = {Experience-dependent myelination is hypothesized to shape neural circuit function and subsequent behavioral output. Using a contextual fear memory task in mice, we demonstrate that fear learning induces oligodendrocyte precursor cells to proliferate and differentiate into myelinating oligodendrocytes in the medial prefrontal cortex. Transgenic animals that cannot form new myelin exhibit deficient remote, but not recent, fear memory recall. Recording population calcium dynamics by fiber photometry, we observe that the neuronal response to conditioned context cues evolves over time in the medial prefrontal cortex, but not in animals that cannot form new myelin. Finally, we demonstrate that pharmacological induction of new myelin formation with clemastine fumarate improves remote memory recall and promotes fear generalization. Thus, bidirectional manipulation of myelin plasticity functionally affects behavior and neurophysiology, which suggests that neural activity during fear learning instructs the formation of new myelin, which in turn supports the consolidation and/or retrieval of remote fear memories.},
language = {eng},
number = {4},
journal = {Nature Neuroscience},
author = {Pan, Simon and Mayoral, Sonia R. and Choi, Hye Sun and Chan, Jonah R. and Kheirbek, Mazen A.},
month = apr,
year = {2020},
pmid = {32042175},
pmcid = {PMC7213814},
keywords = {Animals, Cell Proliferation, Conditioning, Classical, Fear, Memory, Long-Term, Mice, Mice, Transgenic, Myelin Sheath, Oligodendrocyte Precursor Cells, Oligodendrocyte Transcription Factor 2, Prefrontal Cortex, ⛔ No INSPIRE recid found},
pages = {487--499},
}
@article{wan_impaired_2020,
title = {Impaired {Postnatal} {Myelination} in a {Conditional} {Knockout} {Mouse} for the {Ferritin} {Heavy} {Chain} in {Oligodendroglial} {Cells}},
volume = {40},
issn = {0270-6474},
url = {https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7531557/},
doi = {10.1523/JNEUROSCI.1281-20.2020},
abstract = {To define the importance of iron storage in oligodendrocyte development and function, the ferritin heavy subunit (Fth) was specifically deleted in oligodendroglial cells. Blocking Fth synthesis in Sox10 or NG2-positive oligodendrocytes during the first or the third postnatal week significantly reduces oligodendrocyte iron storage and maturation. The brain of Fth KO animals presented an important decrease in the expression of myelin proteins and a substantial reduction in the percentage of myelinated axons. This hypomyelination was accompanied by a decline in the number of myelinating oligodendrocytes and with a reduction in proliferating oligodendrocyte progenitor cells (OPCs). Importantly, deleting Fth in Sox10-positive oligodendroglial cells after postnatal day 60 has no effect on myelin production and/or oligodendrocyte quantities. We also tested the capacity of Fth-deficient OPCs to remyelinate the adult brain in the cuprizone model of myelin injury and repair. Fth deletion in NG2-positive OPCs significantly reduces the number of mature oligodendrocytes and myelin production throughout the remyelination process. Furthermore, the corpus callosum of Fth KO animals presented a significant decrease in the percentage of remyelinated axons and a substantial reduction in the average myelin thickness. These results indicate that Fth synthesis during the first three postnatal weeks is important for an appropriate oligodendrocyte development, and suggest that Fth iron storage in adult OPCs is also essential for an effective remyelination of the mouse brain., SIGNIFICANCE STATEMENT To define the importance of iron storage in oligodendrocyte function, we have deleted the ferritin heavy chain (Fth) specifically in the oligodendrocyte lineage. Fth ablation in oligodendroglial cells throughout early postnatal development significantly reduces oligodendrocyte maturation and myelination. In contrast, deletion of Fth in oligodendroglial cells after postnatal day 60 has no effect on myelin production and/or oligodendrocyte numbers. We have also tested the consequences of disrupting Fth iron storage in oligodendrocyte progenitor cells (OPCs) after demyelination. We have found that Fth deletion in NG2-positive OPCs significantly delays the remyelination process in the adult brain. Therefore, Fth iron storage is essential for early oligodendrocyte development as well as for OPC maturation in the demyelinated adult brain.},
number = {40},
urldate = {2022-11-14},
journal = {The Journal of Neuroscience},
author = {Wan, Rensheng and Cheli, Veronica T. and Santiago-González, Diara A. and Rosenblum, Shaina L. and Wan, Qiuchen and Paez, Pablo M.},
month = sep,
year = {2020},
pmid = {32868463},
pmcid = {PMC7531557},
keywords = {⛔ No INSPIRE recid found},
pages = {7609--7624},
}
@article{xue_demyelination_2021,
title = {Demyelination of the {Optic} {Nerve}: {An} {Underlying} {Factor} in {Glaucoma}?},
volume = {13},
issn = {1663-4365},
shorttitle = {Demyelination of the {Optic} {Nerve}},
url = {https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8593209/},
doi = {10.3389/fnagi.2021.701322},
abstract = {Neurodegenerative disorders are characterized by typical neuronal degeneration and axonal loss in the central nervous system (CNS). Demyelination occurs when myelin or oligodendrocytes experience damage. Pathological changes in demyelination contribute to neurodegenerative diseases and worsen clinical symptoms during disease progression. Glaucoma is a neurodegenerative disease characterized by progressive degeneration of retinal ganglion cells (RGCs) and the optic nerve. Since it is not yet well understood, we hypothesized that demyelination could play a significant role in glaucoma. Therefore, this study started with the morphological and functional manifestations of demyelination in the CNS. Then, we discussed the main mechanisms of demyelination in terms of oxidative stress, mitochondrial damage, and immuno-inflammatory responses. Finally, we summarized the existing research on the relationship between optic nerve demyelination and glaucoma, aiming to inspire effective treatment plans for glaucoma in the future.},
urldate = {2022-11-13},
journal = {Frontiers in Aging Neuroscience},
author = {Xue, Jingfei and Zhu, Yingting and Liu, Zhe and Lin, Jicheng and Li, Yangjiani and Li, Yiqing and Zhuo, Yehong},
month = nov,
year = {2021},
pmid = {34795572},
pmcid = {PMC8593209},
keywords = {⛔ No INSPIRE recid found},
pages = {701322},
}
@article{nave_axonal_2006,
title = {Axonal regulation of myelination by neuregulin 1},
volume = {16},
issn = {0959-4388},
doi = {10.1016/j.conb.2006.08.008},
abstract = {Neuregulins comprise a family of epidermal growth factor-like ligands that interact with ErbB receptor tyrosine kinases to control many aspects of neural development. One of the most dramatic effects of neuregulin-1 is on glial cell differentiation. The membrane-bound neuregulin-1 type III isoform is an axonal ligand for glial ErbB receptors that regulates the early Schwann cell lineage, including the generation of precursors. Recent studies have shown that the amount of neuregulin-1 type III expressed on axons also dictates the glial phenotype, with a threshold level triggering Schwann cell myelination. Remarkably, neuregulin-1 type III also regulates Schwann cell membrane growth to adjust myelin sheath thickness to match axon caliber precisely. Whether this signaling system operates in central nervous system myelination remains an open question of major importance for human demyelinating diseases.},
language = {eng},
number = {5},
journal = {Current Opinion in Neurobiology},
author = {Nave, Klaus-Armin and Salzer, James L.},
month = oct,
year = {2006},
pmid = {16962312},
keywords = {Animals, Axons, Cell Differentiation, Cell Lineage, Humans, Myelin Sheath, Neuregulin-1, Neuroglia, Protein Isoforms, Receptor, ErbB-2, Receptor, ErbB-3, Signal Transduction, Stem Cells, ⛔ No INSPIRE recid found},
pages = {492--500},
}
@article{baraban_ca2_2018,
title = {Ca2+ activity signatures of myelin sheath formation and growth in vivo},
volume = {21},
issn = {1097-6256},
url = {https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5742537/},
doi = {10.1038/s41593-017-0040-x},
abstract = {During myelination, individual oligodendrocytes initially over-produce short myelin sheaths that are either retracted or stabilised. By live imaging oligodendrocyte Ca2+ activity in vivo, we find that high-amplitude long-duration Ca2+ transients in sheaths prefigure retractions, mediated by calpain. Following stabilisation, myelin sheaths grow along axons, and we find that higher frequency Ca2+ transient activity in sheaths precedes faster elongation. Our data implicate local Ca2+ signalling in regulating distinct stages of myelination.},
number = {1},
urldate = {2022-11-13},
journal = {Nature neuroscience},
author = {Baraban, Marion and Koudelka, Sigrid and Lyons, David A},
month = jan,
year = {2018},
pmid = {29230058},
pmcid = {PMC5742537},
keywords = {⛔ No INSPIRE recid found},
pages = {19--23},
}
@article{kuhn_oligodendrocytes_2019,
title = {Oligodendrocytes in {Development}, {Myelin} {Generation} and {Beyond}},
volume = {8},
issn = {2073-4409},
url = {https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6912544/},
doi = {10.3390/cells8111424},
abstract = {Oligodendrocytes are the myelinating cells of the central nervous system (CNS) that are generated from oligodendrocyte progenitor cells (OPC). OPC are distributed throughout the CNS and represent a pool of migratory and proliferative adult progenitor cells that can differentiate into oligodendrocytes. The central function of oligodendrocytes is to generate myelin, which is an extended membrane from the cell that wraps tightly around axons. Due to this energy consuming process and the associated high metabolic turnover oligodendrocytes are vulnerable to cytotoxic and excitotoxic factors. Oligodendrocyte pathology is therefore evident in a range of disorders including multiple sclerosis, schizophrenia and Alzheimer’s disease. Deceased oligodendrocytes can be replenished from the adult OPC pool and lost myelin can be regenerated during remyelination, which can prevent axonal degeneration and can restore function. Cell population studies have recently identified novel immunomodulatory functions of oligodendrocytes, the implications of which, e.g., for diseases with primary oligodendrocyte pathology, are not yet clear. Here, we review the journey of oligodendrocytes from the embryonic stage to their role in homeostasis and their fate in disease. We will also discuss the most common models used to study oligodendrocytes and describe newly discovered functions of oligodendrocytes.},
number = {11},
urldate = {2022-11-13},
journal = {Cells},
author = {Kuhn, Sarah and Gritti, Laura and Crooks, Daniel and Dombrowski, Yvonne},
month = nov,
year = {2019},
pmid = {31726662},
pmcid = {PMC6912544},
keywords = {⛔ No INSPIRE recid found},
pages = {1424},
}
@article{cullen_periaxonal_2021,
title = {Periaxonal and nodal plasticities modulate action potential conduction in the adult mouse brain},
volume = {34},
issn = {2211-1247},
url = {https://www.cell.com/cell-reports/abstract/S2211-1247(20)31630-2},
doi = {10.1016/j.celrep.2020.108641},
language = {English},
number = {3},
urldate = {2022-11-13},
journal = {Cell Reports},
author = {Cullen, Carlie L. and Pepper, Renee E. and Clutterbuck, Mackenzie T. and Pitman, Kimberley A. and Oorschot, Viola and Auderset, Loic and Tang, Alexander D. and Ramm, Georg and Emery, Ben and Rodger, Jennifer and Jolivet, Renaud B. and Young, Kaylene M.},
month = jan,
year = {2021},
pmid = {33472075},
note = {Publisher: Elsevier},
keywords = {action potential, computational modeling, conduction velocity, myelin, node of Ranvier, oligodendrocyte, periaxonal space, plasticity, spatial learning, transcranial magnetic stimulation, ⛔ No INSPIRE recid found},
}
@article{gibson_neuronal_2014,
title = {Neuronal {Activity} {Promotes} {Oligodendrogenesis} and {Adaptive} {Myelination} in the {Mammalian} {Brain}},
volume = {344},
issn = {0036-8075},
url = {https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4096908/},
doi = {10.1126/science.1252304},
abstract = {Myelination of the central nervous system requires the generation of functionally mature oligodendrocytes from oligodendrocyte precursor cells (OPCs). Electrically active neurons may influence OPC function and selectively instruct myelination of an active neural circuit. In this work, we use optogenetic stimulation of the premotor cortex in awake, behaving mice to demonstrate that neuronal activity elicits a mitogenic response of neural progenitor cells and OPCs, promotes oligodendrogenesis, and increases myelination within the deep layers of the premotor cortex and subcortical white matter. We further show that this neuronal activity–regulated oligodendrogenesis and myelination is associated with improved motor function of the corresponding limb. Oligodendrogenesis and myelination appear necessary for the observed functional improvement, as epigenetic blockade of oligodendrocyte differentiation and myelin changes prevents the activity-regulated behavioral improvement.},
number = {6183},
urldate = {2022-11-13},
journal = {Science (New York, N.Y.)},
author = {Gibson, Erin M. and Purger, David and Mount, Christopher W. and Goldstein, Andrea K. and Lin, Grant L. and Wood, Lauren S. and Inema, Ingrid and Miller, Sarah E. and Bieri, Gregor and Zuchero, J. Bradley and Barres, Ben A. and Woo, Pamelyn J. and Vogel, Hannes and Monje, Michelle},
month = may,
year = {2014},
pmid = {24727982},
pmcid = {PMC4096908},
keywords = {⛔ No INSPIRE recid found},
pages = {1252304},
}
@article{spencer_compensation_2018,
title = {Compensation for {Traveling} {Wave} {Delay} {Through} {Selection} of {Dendritic} {Delays} {Using} {Spike}-{Timing}-{Dependent} {Plasticity} in a {Model} of the {Auditory} {Brainstem}},
volume = {12},
issn = {1662-5188},
doi = {10.3389/fncom.2018.00036},
abstract = {Asynchrony among synaptic inputs may prevent a neuron from responding to behaviorally relevant sensory stimuli. For example, "octopus cells" are monaural neurons in the auditory brainstem of mammals that receive input from auditory nerve fibers (ANFs) representing a broad band of sound frequencies. Octopus cells are known to respond with finely timed action potentials at the onset of sounds despite the fact that due to the traveling wave delay in the cochlea, synaptic input from the auditory nerve is temporally diffuse. This paper provides a proof of principle that the octopus cells' dendritic delay may provide compensation for this input asynchrony, and that synaptic weights may be adjusted by a spike-timing dependent plasticity (STDP) learning rule. This paper used a leaky integrate and fire model of an octopus cell modified to include a "rate threshold," a property that is known to create the appropriate onset response in octopus cells. Repeated audio click stimuli were passed to a realistic auditory nerve model which provided the synaptic input to the octopus cell model. A genetic algorithm was used to find the parameters of the STDP learning rule that reproduced the microscopically observed synaptic connectivity. With these selected parameter values it was shown that the STDP learning rule was capable of adjusting the values of a large number of input synaptic weights, creating a configuration that compensated the traveling wave delay of the cochlea.},
language = {eng},
journal = {Frontiers in Computational Neuroscience},
author = {Spencer, Martin J. and Meffin, Hamish and Burkitt, Anthony N. and Grayden, David B.},
year = {2018},
pmid = {29922141},
pmcid = {PMC5996126},
keywords = {auditory brainstem, cochlear nucleus, dendritic delay, octopus cells, spike-timing dependent plasticity, ⛔ No INSPIRE recid found},
pages = {36},
}
@book{mead_analog_1989,
title = {Analog {VLSI} {Implementation} of {Neural} {Systems}},
isbn = {978-0-7923-9040-4},
abstract = {This volume contains the proceedings of a workshop on Analog Integrated Neural Systems held May 8, 1989, in connection with the International Symposium on Circuits and Systems. The presentations were chosen to encompass the entire range of topics currently under study in this exciting new discipline. Stringent acceptance requirements were placed on contributions: (1) each description was required to include detailed characterization of a working chip, and (2) each design was not to have been published previously. In several cases, the status of the project was not known until a few weeks before the meeting date. As a result, some of the most recent innovative work in the field was presented. Because this discipline is evolving rapidly, each project is very much a work in progress. Authors were asked to devote considerable attention to the shortcomings of their designs, as well as to the notable successes they achieved. In this way, other workers can now avoid stumbling into the same traps, and evolution can proceed more rapidly (and less painfully). The chapters in this volume are presented in the same order as the corresponding presentations at the workshop. The first two chapters are concerned with fmding solutions to complex optimization problems under a predefmed set of constraints. The first chapter reports what is, to the best of our knowledge, the first neural-chip design. In each case, the physics of the underlying electronic medium is used to represent a cost function in a natural way, using only nearest-neighbor connectivity.},
language = {en},
publisher = {Springer Science \& Business Media},
author = {Mead, Carver and Ismail, Mohammed},
month = aug,
year = {1989},
note = {Google-Books-ID: 9e29dOiXeiMC},
keywords = {Computers / CAD-CAM, Computers / Computer Vision \& Pattern Recognition, Technology \& Engineering / Electrical, Technology \& Engineering / Electronics / Circuits / General, Technology \& Engineering / Electronics / General, Technology \& Engineering / Imaging Systems, ⛔ No INSPIRE recid found},
}
@article{benjamin_neurogrid_2014,
title = {Neurogrid: {A} {Mixed}-{Analog}-{Digital} {Multichip} {System} for {Large}-{Scale} {Neural} {Simulations}},
volume = {102},
issn = {1558-2256},
shorttitle = {Neurogrid},
doi = {10.1109/JPROC.2014.2313565},
abstract = {In this paper, we describe the design of Neurogrid, a neuromorphic system for simulating large-scale neural models in real time. Neuromorphic systems realize the function of biological neural systems by emulating their structure. Designers of such systems face three major design choices: 1) whether to emulate the four neural elements-axonal arbor, synapse, dendritic tree, and soma-with dedicated or shared electronic circuits; 2) whether to implement these electronic circuits in an analog or digital manner; and 3) whether to interconnect arrays of these silicon neurons with a mesh or a tree network. The choices we made were: 1) we emulated all neural elements except the soma with shared electronic circuits; this choice maximized the number of synaptic connections; 2) we realized all electronic circuits except those for axonal arbors in an analog manner; this choice maximized energy efficiency; and 3) we interconnected neural arrays in a tree network; this choice maximized throughput. These three choices made it possible to simulate a million neurons with billions of synaptic connections in real time-for the first time-using 16 Neurocores integrated on a board that consumes three watts.},
number = {5},
journal = {Proceedings of the IEEE},
author = {Benjamin, Ben Varkey and Gao, Peiran and McQuinn, Emmett and Choudhary, Swadesh and Chandrasekaran, Anand R. and Bussat, Jean-Marie and Alvarez-Icaza, Rodrigo and Arthur, John V. and Merolla, Paul A. and Boahen, Kwabena},
month = may,
year = {2014},
note = {Conference Name: Proceedings of the IEEE},
keywords = {Analog circuits, Computer architecture, Electronic circuits, Integrated circuit modeling, Nerve fibers, Neural networks, Neuroscience, Random access memory, Synchronous digital hierarchy, application specific integrated circuits, asynchronous circuits, brain modeling, computational neuroscience, interconnection networks, mixed analog-digital integrated circuits, neural network hardware, neuromorphic electronic systems, ⛔ No INSPIRE recid found},
pages = {699--716},