Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-45l2p Total loading time: 0 Render date: 2024-04-25T20:37:52.866Z Has data issue: false hasContentIssue false

2 - Organization of the principal pathways of prefrontal lateral, medial, and orbitofrontal cortices in primates and implications for their collaborative interaction in executive functions

Published online by Cambridge University Press:  11 September 2009

Jarl Risberg
Affiliation:
Lunds Universitet, Sweden
Jordan Grafman
Affiliation:
National Institute of Health, Bethesda, MD, USA
Get access

Summary

Overview

Ideas about the prefrontal cortex have changed drastically during the past century. Situated at the rostral pole of the brain, the prefrontal cortex was traditionally considered to be the seat of intelligence. Damage to the prefrontal cortex, however, does not affect overall intelligence in humans, but has detrimental effects on executive function. Converging lines of evidence indicate that the prefrontal cortex is an action-oriented region, guiding behavior by selecting relevant signals to accomplish the task at hand.

The prefrontal cortex is in a strategic position to exercise executive control. It receives information from most other cortical areas and subcortical structures and is connected with specialized oculomotor centers for searching the environment, and with motor control systems for action. These functions have been associated with the caudal parts of the lateral prefrontal cortex, namely areas 8 and the caudal part of area 46. These caudal lateral areas are connected with neighboring premotor as well as early-processing sensory association areas and with intraparietal visuomotor areas. However, the lateral prefrontal cortex is considerably more extensive, and includes, in addition, the more anteriorly situated areas around the principal sulcus (rostral area 46), area 9 on the dorsolateral surface, area 12 on the ventrolateral surface, and area 10 at the frontal pole, all of which have been associated with cognitive functions and executive control. Figure 2.1B shows the lateral prefrontal cortex and its relationship to the adjacent premotor/motor cortical system.

Type
Chapter
Information
The Frontal Lobes
Development, Function and Pathology
, pp. 21 - 68
Publisher: Cambridge University Press
Print publication year: 2006

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Adolphs, R., Tranel, D. & Damasio, A. R. (1998). The human amygdala in social judgment. Nature, 393, 470–4.CrossRefGoogle ScholarPubMed
Adolphs, R., Tranel, D., Damasio, H. & Damasio, A. (1994). Impaired recognition of emotion in facial expressions following bilateral damage to the human amygdala. Nature, 372, 669–72.CrossRefGoogle ScholarPubMed
Aggleton, J. P., Burton, M. J. & Passingham, R. E. (1980). Cortical and subcortical afferents to the amygdala of the rhesus monkey (Macaca mulatta). Brain Research, 190, 347–68.CrossRefGoogle Scholar
Alexander, G. E., Delong, M. R. & Strick, P. L. (1986). Parallel organization of functionally segregated circuits linking basal ganglia and cortex. Annual Review of Neuroscience, 9, 357–81.CrossRefGoogle ScholarPubMed
Alexander, M. P. & Freedman, M. (1984). Amnesia after anterior communicating artery aneurysm rupture. Neurology, 34, 752–7.CrossRefGoogle ScholarPubMed
Alexander, G. E. & Fuster, J. M. (1973). Effects of cooling prefrontal cortex on cell firing in the nucleus medialis dorsalis. Brain Research, 61, 93–105.CrossRefGoogle ScholarPubMed
Alheid, G. F., Heimer, L. & Switzer, R. C. III. (1990). Basal ganglia. In The Human Nervous System, ed. Paxinos, G.. San Diego: Academic Press, pp. 483–582.Google Scholar
Amaral, D. G. (2002). The primate amygdala and the neurobiology of social behavior: implications for understanding social anxiety. Biological Psychiatry, 51, 11–17.CrossRefGoogle ScholarPubMed
Amaral, D. G., Insausti, R., Zola-Morgan, S., Squire, L. R. & Suzuki, W. A. (1990). The perirhinal and parahippocampal cortices and medial temporal lobe memory function. In Vision, Memory, and the Temporal Lobe eds. Iwai, E. and Mishkin, M.. New York: Elsevier Science Publishing Co., Inc., pp. 149–161.Google Scholar
Amaral, D. G. & Price, J. L. (1984). Amygdalo-cortical projections in the monkey (Macaca fascicularis). The Journal of Comparative Neurology, 230, 465–96.CrossRefGoogle Scholar
Anderson, M. E. (2001). Pallidal and cortical detriments of thalamic activity. In Basal Ganglia and Thalamus in Health and Movement Disorders, ed. Kultas-Ilinsky, K. and Ilinsky, I. A.. New York: Kluwer Academic/Plenum Publishers, pp. 93–104.CrossRefGoogle Scholar
Asanuma, C., Andersen, R. A. & Cowan, W. M. (1985). The thalamic relations of the caudal inferior parietal lobule and the lateral prefrontal cortex in monkeys: Divergent cortical projections from cell clusters in the medial pulvinar nucleus. The Journal of Comparative Neurology, 241, 357–81.CrossRefGoogle ScholarPubMed
Bachevalier, J., Meunier, M., Lu, M. X. & Ungerleider, L. G. (1997). Thalamic and temporal cortex input to medial prefrontal cortex in rhesus monkeys. Experimental Brain Research, 115, 430–44.CrossRefGoogle ScholarPubMed
Bachevalie, J. & Mishkin, M. (1986). Visual recognition impairment follows ventromedial but not dorsolateral prefrontal lesions in monkeys. Behavioral Brain Research, 20, 249–61.CrossRefGoogle Scholar
Baimbridge, K. G., Celio, M. R. & Rogers, J. H. (1992). Calcium-binding proteins in the nervous system. Trends in Neuroscience, 15, 303–8.CrossRefGoogle Scholar
Baleydier, C. & Mauguiere, F. (1980). The duality of the cingulate gyrus in monkey. Neuroanatomical study and functional hypothesis. Brain, 103, 525–54.CrossRefGoogle ScholarPubMed
Barbas, H. (1986). Pattern in the laminar origin of corticocortical connections. The Journal of Comparative Neurology, 252, 415–22.CrossRefGoogle ScholarPubMed
Barbas, H. (1988). Anatomic organization of basoventral and mediodorsal visual recipient prefrontal regions in the rhesus monkey. The Journal of Comparative Neurology, 276, 313–42.CrossRefGoogle ScholarPubMed
Barbas, H. (1992). Architecture and cortical connections of the prefrontal cortex in the rhesus monkey. In Advances in Neurology, Vol. 57, ed. Chauvel, P., Delgado-Escueta, A. V., Halgren, E. and Bancaud, J.. New York: Raven Press, Ltd., pp. 91–115.Google Scholar
Barbas, H. (1993). Organization of cortical afferent input to orbitofrontal areas in the rhesus monkey. Neuroscience, 56, 841–64.CrossRefGoogle ScholarPubMed
Barbas, H. (1995). Anatomic basis of cognitive-emotional interactions in the primate prefrontal cortex. Neuroscience and Biobehavioral Reviews, 19, 499–510.CrossRefGoogle ScholarPubMed
Barbas, H. (1997). Two prefrontal limbic systems: Their common and unique features. In The Association Cortex: Structure and Function, ed. Sakata, H., Mikami, A. and Fuster, J. M.. Amsterdam: Harwood Academic Publ, pp. 99–115.Google Scholar
Barbas, H. (2000). Complementary role of prefrontal cortical regions in cognition, memory and emotion in primates. Advances in Neurology, 84, 87–110.Google Scholar
Barbas, H. & Blatt, G. J. (1995). Topographically specific hippocampal projections target functionally distinct prefrontal areas in the rhesus monkey. Hippocampus, 5, 511–33.CrossRefGoogle ScholarPubMed
Barbas, H. & Olmos, J. (1990). Projections from the amygdala to basoventral and mediodorsal prefrontal regions in the rhesus monkey. The Journal of Comparative Neurology, 301, 1–23.Google Scholar
Barbas, H., Ghashghaei, H., Dombrowsk, S. M. & Rempel-Clower, N. L. (1999). Medial prefrontal cortices are unified by common connections with superior temporal cortices and distinguished by input from memory-related areas in the rhesus monkey. The Journal of Comparative Neurology, 410, 343–67.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
Barbas, H., Ghashghaei, H., Rempel-Clower, N. & Xiao, D. (2002). Anatomic basis of functional specialization in prefrontal cortices in primates. In Handbook of Neuropsychology, ed. Grafman, J.. Amsterdam: Elsevier Science B. V., pp. 1–27.Google Scholar
Barbas, H., Henion, T. H. & Dermon, C. R. (1991). Diverse thalamic projections to the prefrontal cortex in the rhesus monkey. The Journal of Comparative Neurology, 313, 65–94.CrossRefGoogle ScholarPubMed
Barbas, H. & Hilgetag, C. C. (2002). Rules relating connections to cortical structure in primate prefrontal cortex. Neurocomputing, 44–46, 301–8.CrossRefGoogle Scholar
Barbas, H., Hilgetag, C. C., Saha, S., Dermon, C. R. & Suski, J. L. (2005a). Parallel organization of contralateral and ipsilateral prefrontal cortical projections in the rhesus monkey. BioMed Central Neuroscience, 6, 32.Google Scholar
Barbas, H., Medalla, M., Alade, O., et al. (2005b). Relationship of prefrontal connections to inhibitory systems in superior temporal areas in the rhesus monkey. Cerebral Cortex, 15, 1356–70.CrossRefGoogle Scholar
Barbas, H. & Mesulam, M. M. (1981). Organization of afferent input to subdivisions of area 8 in the rhesus monkey. The Journal of Comparative Neurology, 200, 407–31.CrossRefGoogle ScholarPubMed
Barbas, H. & Mesulam, M. M. (1985). Cortical afferent input to the principalis region of the rhesus monkey. Neuroscience, 15, 619–37.CrossRefGoogle ScholarPubMed
Barbas, H. & Pandya, D. N. (1987). Architecture and frontal cortical connections of the premotor cortex (area 6) in the rhesus monkey. The Journal of Comparative Neurology, 256, 211–18.CrossRefGoogle ScholarPubMed
Barbas, H. & Pandya, D. N. (1989). Architecture and intrinsic connections of the prefrontal cortex in the rhesus monkey. The Journal of Comparative Neurology, 286, 353–75.CrossRefGoogle ScholarPubMed
Barbas, H. & Rempel-Clower, N. (1997). Cortical structure predicts the pattern of corticocortical connections. Cerebral Cortex, 7, 635–46.CrossRefGoogle ScholarPubMed
Barbas, H., Saha, S., Rempel-Clower, N. & Ghashghaei, T. (2003). Serial pathways from primate prefrontal cortex to autonomic areas may influence emotional expression. BioMed Central Neuroscience, 4, 25.Google ScholarPubMed
Barris, R. W. & Schuman, H. R. (1953.) Bilateral anterior cingulate gyrus lesions. Syndrome of the anterior cingulate gyri. Neurology, 3, 44–52.CrossRefGoogle ScholarPubMed
Bauer, R. H. & Fuster, J. M. (1976). Delayed-matching and delayed-response deficit from cooling dorsolateral prefrontal cortex in monkeys. Journal of Comparative Physiology and Psychology, 90, 299–302.CrossRefGoogle ScholarPubMed
Bauman, M. D., Lavenex, P., Mason, W. A., Capitanio, J. P. & Amaral, D. G. (2004). The development of mother-infant interactions after neonatal amygdala lesions in rhesus monkeys. Journal of Neuroscience, 24, 711–21.CrossRefGoogle ScholarPubMed
Baxter, M. G., Parker, A., Lindner, C. C., Izquierdo, A. D. & Murray, E. A. (2000). Control of response selection by reinforcer value requires interaction of amygdala and orbital prefrontal cortex. Journal of Neuroscience, 20, 4311–19.CrossRefGoogle ScholarPubMed
Bechara, A., Damasio, H., Damasio, A. R. & Lee, G. P. (1999). Different contributions of the human amygdala and ventromedial prefrontal cortex to decision-making. Journal of Neuroscience, 19, 5473–81.CrossRefGoogle ScholarPubMed
Bechara, A., Tranel, D., Damasio, H. & Damasio, A. R. (1996). Failure to respond autonomically to anticipated future outcomes following damage to prefrontal cortex. Cerebral Cortex, 6, 215–25.CrossRefGoogle ScholarPubMed
Bisley, J. W. & Goldberg, M. E. (2003). The role of the parietal cortex in the neural processing of saccadic eye movements. Advances in Neurology, 93, 141–57.Google ScholarPubMed
Boussaoud, D., Ungerleider, L. G. & Desimone, R. (1990). Pathways for motion analysis: Cortical connections of the medial superior temporal and fundus of the superior temporal visual areas in the macaque. The Journal of Comparative Neurology, 296, 462–95.CrossRefGoogle ScholarPubMed
Broca, P. (1878). Anatomie compareé des enconvolutions cérébrales: Le grand lobe limbique et la scissure limbique dans la serie des mammifères. Revue Anthropologique, 1, 385–498.Google Scholar
Brodmann, K. (1905). Beitrage zur histologischen lokalisation der Grosshirnrinde. III. Mitteilung: Die Rindenfelder der niederen Affen. Journal of Psychology and Neurology, 4, 177–266.Google Scholar
Brodmann, K. (1909). Vergleichende Lokalizationslehre der Grosshirnrinde in ihren Prinizipien dargestelt auf Grund des Zellenbaues. Leipzig: Barth.Google Scholar
Buge, A., Escourolle, R., Rancurel, G. & Poisson, M. (1975). Akinetic mutism and bicingular softening. 3 anatomo-clinical cases. Revue Neurologique (Paris), 131, 121–31.Google ScholarPubMed
Carmichael, S. T., Clugnet, M.-C. & Price, J. L. (1994). Central olfactory connections in the macaque monkey. The Journal of Comparative Neurology, 346, 403–34.CrossRefGoogle ScholarPubMed
Carmichael, S. T. & Price, J. L. (1995a). Limbic connections of the orbital and medial prefrontal cortex in macaque monkeys. The Journal of Comparative Neurology, 363, 615–41.CrossRefGoogle Scholar
Carmichael, S. T. & Price, J. L. (1995b). Sensory and premotor connections of the orbital and medial prefrontal cortex of macaque monkeys. The Journal of Comparative Neurology, 363, 642–64.CrossRefGoogle Scholar
Carmichael, S. T. & Price, J. L. (1996). Connectional networks within the orbital and medial prefrontal cortex of macaque monkeys. The Journal of Comparative Neurology, 371, 179–207.3.0.CO;2-#>CrossRefGoogle ScholarPubMed
Castro-Alamancos, M. A. & Connors, B. W. (1997). Thalamocortical synapses. Progress in Neurobiology, 51, 581–606.CrossRefGoogle ScholarPubMed
Cavada, C., Company, T., Tejedor, J., Cruz-Rizzolo, R. J. & Reinoso-Suarez, F. (2000). The anatomical connections of the macaque monkey orbitofrontal cortex. A review. Cerebral Cortex, 10, 220–42.CrossRefGoogle ScholarPubMed
Chao, L. L. & Knight, R. T. (1997a). Age-related prefrontal alterations during auditory memory. Neurobiology of Aging, 18, 87–95.CrossRefGoogle Scholar
Chao, L. L. & Knight, R. T. (1997b). Prefrontal deficits in attention and inhibitory control with aging. Cerebral Cortex, 7, 63–9.CrossRefGoogle Scholar
Chao, L. L. & Knight, R. T. (1998). Contribution of human prefrontal cortex to delay performance. Journal of Cognitive Neuroscience, 10, 167–77.CrossRefGoogle ScholarPubMed
Chavis, D. A. & Pandya, D. N. (1976). Further observations on corticofrontal connections in the rhesus monkey. Brain Research, 117, 369–86.CrossRefGoogle ScholarPubMed
Chiba, T., Kayahara, T. & Nakano, K. (2001). Efferent projections of infralimbic and prelimbic areas of the medial prefrontal cortex in the Japanese monkey, Macaca fuscata. Brain Research, 888, 83–101.CrossRefGoogle ScholarPubMed
Cicerone, K. D. & Tanenbau, L. N. (1997). Disturbance of social cognition after traumatic orbitofrontal brain injury. Archives of Clinical Neuropsychology, 12, 173–88.CrossRefGoogle ScholarPubMed
Colby, C. L. & Goldberg, M. E. (1999). Space and attention in parietal cortex. Annual Review of Neuroscience, 22, 319–49.CrossRefGoogle ScholarPubMed
Conde, F., Lund, J. S., Jacobowitz, D. M., Baimbridge, K. G. & Lewis, D. A. (1994). Local circuit neurons immunoreactive for calretinin, calbindin D-28k or parvalbumin in monkey prefrontal cortex: distribution and morphology. The Journal of Comparative Neurology, 341, 95–116.CrossRefGoogle ScholarPubMed
Crowell, R. M. & Morawetz, R. B. (1977). The anterior communicating artery has significant branches. Stroke, 8, 272–3.CrossRefGoogle ScholarPubMed
D'Esposito, M., Alexander, M. P., Fischer, R., McGlinchey-Berroth, R. & O'Connor, M. (1996). Recovery of memory and executive function following anterior communicating artery aneurysm rupture. Journal of the International Neuropsychological Society, 2, 565–70.CrossRefGoogle ScholarPubMed
Damasio, A. R. (1994). Descarte's Error: Emotion, Reason, and the Human Brain. New York: G. P. Putnam's Sons.Google Scholar
Damasio, H., Grabowski, T., Frank, R., Galaburda, A. M. & Damasio, A. R. (1994). The return of Phineas Gage: clues about the brain from the skull of a famous patient. Science, 264, 1102–5.CrossRefGoogle ScholarPubMed
Davis, M. (1992). The role of the amygdala in fear and anxiety. Annual Revue of Neuroscience, 15, 353–75.CrossRefGoogle ScholarPubMed
Lima, A. D., Voigt, T. & Morrison, J. H. (1990). Morphology of the cells within the inferior temporal gyrus that project to the prefrontal cortex in the macaque monkey. The Journal of Comparative Neurology, 296, 159–72.CrossRefGoogle ScholarPubMed
De Olmos, J. (1990). Amygdaloid nuclear gray complex. In The Human Nervous System, ed. Paxinos, G.. San Diego: Academic Press, Inc., pp. 583–710.Google Scholar
DeFelipe, J., Hendry, S. H., Hashikawa, T., Molinari, M. & Jones, E. G. (1990). A microcolumnar structure of monkey cerebral cortex revealed by immunocytochemical studies of double bouquet cell axons. Neuroscience, 37, 655–73.CrossRefGoogle ScholarPubMed
DeFelipe, J., Hendry, S. H. & Jones, E. G. (1989a). Synapses of double bouquet cells in monkey cerebral cortex visualized by calbindin immunoreactivity. Brain Research, 503, 49–54.CrossRefGoogle Scholar
DeFelipe, J., Hendry, S. H. & Jones, E. G. (1989b). Visualization of chandelier cell axons by parvalbumin immunoreactivity in monkey cerebral cortex. Proceedings of the National Academy of Science, U.S.A., 86, 2093–7.CrossRefGoogle Scholar
Dermon, C. R. & Barbas, H. (1994). Contralateral thalamic projections predominantly reach transitional cortices in the rhesus monkey. The Journal of Comparative Neurology, 344, 508–31.CrossRefGoogle ScholarPubMed
Distler, C., Boussaoud, D., Desimone, R. & Ungerleider, L. G. (1993). Cortical connections of inferior temporal area TEO in macaque monkeys. The Journal of Comparative Neurology, 334, 125–50.CrossRefGoogle ScholarPubMed
Dombrowski, S. M. & Barbas, H. (1998). Distinction of prefrontal architectonic areas using stereologic procedures. Neuroscience Abstracts, 24, 1163.Google Scholar
Dombrowski, S. M., Hilgetag, C. C. & Barbas, H. (2001). Quantitative architecture distinguishes prefrontal cortical systems in the rhesus monkey. Cerebral Cortex, 11, 975–88.CrossRefGoogle ScholarPubMed
Douglas, R. J., Martin, K. A. & Whitteridge, D. (1991). An intracellular analysis of the visual responses of neurones in cat visual cortex. Journal of Physiology (London), 440, 659–96.CrossRefGoogle ScholarPubMed
Dum, R. P. & Strick, P. L. (1991). The origin of corticospinal projections from the premotor areas in the frontal lobe. Journal of Neuroscience, 11, 667–89.CrossRefGoogle ScholarPubMed
Felleman, D. J. & Essen, D. C. (1991). Distributed hierarchical processing in the primate cerebral cortex. Cerebral Cortex, 1, 1–47.CrossRefGoogle ScholarPubMed
Freedman, L. J., Insel, T. R. & Smith, Y. (2000). Subcortical projections of area 25 (subgenual cortex) of the macaque monkey. The Journal of Comparative Neurology, 421, 172–88.3.0.CO;2-8>CrossRefGoogle ScholarPubMed
Funahashi, S., Bruce, C. J. & Goldman-Rakic, P. S. (1990). Visuospatial coding in primate prefrontal neurons revealed by oculomotor paradigms. Journal of Neurophysiology, 63, 814–31.CrossRefGoogle ScholarPubMed
Funahashi, S., Bruce, C. J. & Goldman-Rakic, P. S. (1991) Neuronal activity related to saccadic eye movements in the monkey's dorsolateral prefrontal cortex. Journal of Neurophysiology, 65, 1464–83.CrossRefGoogle ScholarPubMed
Fuster, J. M. (1989). The Prefrontal Cortex. New York: Raven Press.Google Scholar
Fuster, J. M. (1993). Frontal lobes. Current Opinions in Neurobiology, 3, 160–5.CrossRefGoogle ScholarPubMed
Gabbott, P. L. & Bacon, S. J. (1996). Local circuit neurons in the medial prefrontal cortex (areas 24a, b,c, 25 and 32) in the monkey: II. Quantitative areal and laminar distributions. The Journal of Comparative Neurology, 364, 609–36.3.0.CO;2-7>CrossRefGoogle ScholarPubMed
Gabbott, P. L., Jays, P. R. & Bacon, S. J. (1997). Calretinin neurons in human medial prefrontal cortex (areas 24a, b,c, 32′, and 25). The Journal of Comparative Neurology, 381, 389–410.3.0.CO;2-Z>CrossRefGoogle Scholar
Galea, M. P. & Darian-Smith, I. (1994). Multiple corticospinal neuron populations in the macaque monkey are specified by their unique cortical origins, spinal terminations, and connections. Cerebral Cortex, 4, 166–94.CrossRefGoogle Scholar
Gehring, W. J. & Knight, R. T. (2002). Lateral prefrontal damage affects processing selection but not attention switching. Brain Research. Cognitive Brain Research, 13, 267–79.CrossRefGoogle Scholar
Ghashghaei, H. T. & Barbas, H. (2001). Neural interaction between the basal forebrain and functionally distinct prefrontal cortices in the rhesus monkey. Neuroscience, 103, 593–614.CrossRefGoogle ScholarPubMed
Ghashghaei, H. T. & Barbas, H. (2002). Pathways for emotions: Interactions of prefrontal and anterior temporal pathways in the amygdala of the rhesus monkey. Neuroscience, 115, 1261–79.CrossRefGoogle ScholarPubMed
Gilbert, C. D. & Kelly, J. P. (1975). The projections of cells in different layers of the cat's visual cortex. The Journal of Comparative Neurology, 163, 81–105.CrossRefGoogle ScholarPubMed
Goldman-Rakic, P. S. (1987). Motor control function of the prefrontal cortex. Ciba Foundation Symposium, 132, 187–200.Google ScholarPubMed
Goldman-Rakic, P. S. (1988). Topography of cognition: Parallel distributed networks in primate association cortex. Annual Review of Neuroscience, 11, 137–56.CrossRefGoogle ScholarPubMed
Goldman-Rakic, P. S., Isseroff, A., Schwartz, M. L. & Bugbee, N. M. (1983). The neurobiology of cognitive development. In Handbook of Child Psychology: Biology and Infancy Development, ed. Mussen, P.. New York: Wiley, pp. 281–344.Google Scholar
Goldman-Rakic, P. S., Lidow, M. S. & Gallager, D. W. (1990). Overlap of domaminergic, adrenergic, and serotoninergic receptors and complementarity of their subtypes in primate prefrontal cortex. Journal of Neuroscience, 10, 2125–38.CrossRefGoogle Scholar
Goldman-Rakic, P. S. & Porrino, L. J. (1985). The primate mediodorsal (MD) nucleus and its projection to the frontal lobe. The Journal of Comparative Neurology, 242, 535–60.CrossRefGoogle ScholarPubMed
Gower, E. C. (1989). Efferent projections from limbic cortex of the temporal pole to the magnocellular medial dorsal nucleus in the rhesus monkey. The Journal of Comparative Neurology, 280, 343–58.CrossRefGoogle ScholarPubMed
Grant, S. & Hilgetag, C.-C. (2004). Structural model explains laminar origins of projections in cat visual cortex. Neuroscience Abstracts, 30, 300.9.Google Scholar
Graybiel, A. M. (2000). The basal ganglia. Current Biolology, 10, R509–R511.CrossRefGoogle ScholarPubMed
Graybiel, A. M., Aosaki, T., Flaherty, A. W. & Kimura, M. (1994). The basal ganglia and adaptive motor control. Science, 265, 1826–31.CrossRefGoogle ScholarPubMed
Haber, S. & McFarland, N. R. (2001). The place of the thalamus in frontal cortical-basal ganglia circuits. The Neuroscientist, 7, 315–24.CrossRefGoogle ScholarPubMed
Hasegawa, R. P., Blitz, A. M. & Goldberg, M. E. (2004). Neurons in monkey prefrontal cortex whose activity tracks the progress of a three-step self-ordered task. Journal of Neurophysiology, 92, 1524–35.CrossRefGoogle ScholarPubMed
Heckers, S. (1997). Neuropathology of schizophrenia: cortex, thalamus, basal ganglia, and neurotransmitter-specific projection systems. Schizophrenia Bulletin, 23, 403–21.CrossRefGoogle ScholarPubMed
Heckers, S. & Konradi, C. (2002). Hippocampal neurons in schizophrenia. Journal of Neural Transmission, 109, 891–905.CrossRefGoogle Scholar
Heffner, H. E. & Heffner, R. S. (1986). Effect of unilateral and bilateral auditory cortex lesions on the discrimination of vocalizations by Japanese macaques. Journal of Neurophysiology, 56, 683–701.CrossRefGoogle ScholarPubMed
Heizmann, C. W. (1992). Calcium-binding proteins: Basic concepts and clinical implications. General Physiology and Biophysics, 11, 411–25.Google ScholarPubMed
Hendry, S. H. C., Jones, E. G., Emson, P. C., et al. (1989). Two classes of cortical GABA neurons defined by differential calcium binding protein immunoreactivities. Experimental Brain Research, 76, 467–72.CrossRefGoogle ScholarPubMed
Herzog, A. G. & Hoesen, G. W. (1976). Temporal neocortical afferent connections to the amygdala in the rhesus monkey. Brain Research, 115, 57–69.CrossRefGoogle ScholarPubMed
Hikosaka, O., Nakahara, H., Rand, M. K., et al. (1999). Parallel neural networks for learning sequential procedures. Trends in Neurosciences, 22, 464–71.CrossRefGoogle ScholarPubMed
Hikosaka, K. & Watanabe, M. (2000). Delay activity of orbital and lateral prefrontal neurons of the monkey varying with different rewards. Cerebral Cortex, 10, 263–71.CrossRefGoogle ScholarPubMed
Hilgetag, C. C., O'Neill, M. A. & Young, M. P. (1996). Indeterminate organization of the visual system. Science 271, 776–7.CrossRefGoogle ScholarPubMed
Hof, P. R., Glezer, I. I., Conde, F., et al. (1999). Cellular distribution of the calcium-binding proteins parvalbumin, calbindin, and calretinin in the neocortex of mammals: phylogenetic and developmental patterns. Journal of Chemical Neuroanatomy, 16, 77–116.CrossRefGoogle ScholarPubMed
Hof, P. R., Nimchinsky, E. A. & Morrison, J. H. (1995). Neurochemical phenotype of corticocortical connections in the macaque monkey: Quantitative analysis of a subset of neurofilament protein-immunoreactive projection neurons in frontal, parietal, temporal, and cingulate cortices, The Journal of Comparative Neurology, 362, 109–33.CrossRefGoogle ScholarPubMed
Hollerman, J. R., Tremblay, L. & Schultz, W. (2000). Involvement of basal ganglia and orbitofrontal cortex in goal-directed behavior. In Progress in Brain Research, ed. Uylings, H. B. M., Eden, C. G., Bruin, J. P. C., Feenstra, M. G. P. and Pennartz, C. M. A.. Paris: Elsevier Science, pp. 193–215.Google Scholar
Huerta, M. F., Krubitzer, L. A. & Kaas, J. H. (1987). Frontal eye field as defined by intracortical microstimulation in squirrel monkeys, owl monkeys, and Macaque monkeys II. Cortical connections. The Journal of Comparative Neurology, 265, 332–61.CrossRefGoogle ScholarPubMed
Hutchins, K. D., Martino, A. M. & Strick, P. L. (1988). Corticospinal projections from the medial wall of the hemisphere. Experimental Brain Research, 71, 667–72.CrossRefGoogle ScholarPubMed
Ilinsky, I. A., Jouandet, M. L. & Goldman-Rakic, P. S. (1985). Organization of the nigrothalamocortical system in the rhesus monkey. The Journal of Comparative Neurology, 236, 315–30.CrossRefGoogle ScholarPubMed
Insausti, R. & Munoz, M. (2001). Cortical projections of the non-entorhinal hippocampal formation in the cynomolgus monkey (Macaca fascicularis). European Journal of Neuroscience, 14, 435–51.CrossRefGoogle Scholar
Jacobsen, C. F. (1936). Studies of cerebral function in primates: I. The functions of the frontal association area in monkeys. Computational Psychology Monographs, 13, 3–60.Google Scholar
Jacobson, S., Butters, N. & Tovsky, N. J. (1978). Afferent and efferent subcortical projections of behaviorally defined sectors of prefrontal granular cortex. Brain Research, 159, 279–96.CrossRefGoogle ScholarPubMed
Jacobson, S. & Trojanowski, J. Q. (1975). Amygdaloid projections to prefrontal granular cortex in rhesus monkey demonstrated with horseradish peroxidase. Brain Research, 100, 132–9.CrossRefGoogle ScholarPubMed
Jacobson, S. & Trojanowski, J. Q. (1977). Prefrontal granular cortex of the rhesus monkey I. Intrahemispheric cortical afferents. Brain Research, 132, 209–33.CrossRefGoogle ScholarPubMed
Jones, E. G. (1985). The Thalamus. New York: Plenum Press.CrossRefGoogle Scholar
Jones, E. G. (2002). Thalamic organization and function after Cajal. Progress in Brain Research, 136, 333–57.CrossRefGoogle ScholarPubMed
Jones, E. G. & Powell, T. P. S. (1970). An anatomical study of converging sensory pathways within the cerebral cortex. Brain, 93, 793–820.CrossRefGoogle ScholarPubMed
Jones, E. G. & Wise, S. P. (1977). Size, laminar and columnar distribution of efferent cells in the sensory-motor cortex of monkeys. The Journal of Comparative Neurology, 175, 391–438.CrossRefGoogle ScholarPubMed
Jongen-Relo, A. L. & Amaral, D. G. (1998). Evidence for a GABAergic projection from the central nucleus of the amygdala to the brainstem of the macaque monkey: a combined retrograde tracing and in situ hybridization study. European Journal of Neuroscience, 10, 2924–33.CrossRefGoogle ScholarPubMed
Jürgens, U. & Müller-Preuss, P. (1977). Convergent projections of different limbic vocalization areas in the squirrel monkey. Experimental Brain Research, 29, 75–83.CrossRefGoogle ScholarPubMed
Kawaguchi, Y. & Kubota, Y. (1997). GABAergic cell subtypes and their synaptic connections in rat frontal cortex. Cerebral Cortex, 7, 476–86.CrossRefGoogle ScholarPubMed
Kennard, M. A. (1945). Focal autonomic representation in the cortex and its relation to sham rage. Journal of Neuropathology and Experimental Neurology, 4, 295–304.CrossRefGoogle Scholar
Kievit, J. & Kuypers, H. G. J. M. (1977). Organization of the thalamo-cortical connexions to the frontal lobe in the rhesus monkey. Experimental Brain Research, 29, 299–322.Google ScholarPubMed
Kling, A. & Steklis, H. D. (1976). A neural substrate for affiliative behavior in nonhuman primates. Brain, Behavior and Evolution, 13, 216–38.CrossRefGoogle ScholarPubMed
Knight, R. T., Staines, W. R., Swick, D. & Chao, L. L. (1999). Prefrontal cortex regulates inhibition and excitation in distributed neural networks. Acta Psychologica (Amsterdam), 101, 159–78.CrossRefGoogle ScholarPubMed
Koechlin, E., Basso, G., Pietrini, P., Panzer, S. & Grafman, J. (1999). The role of the anterior prefrontal cortex in human cognition. Nature, 399, 148–51.CrossRefGoogle ScholarPubMed
Kondo, H., Saleem, K. S. & Price, J. L. (2003). Differential connections of the temporal pole with the orbital and medial prefrontal networks in macaque monkeys. The Journal of Comparative Neurology, 465, 499–523.CrossRefGoogle ScholarPubMed
Kosaki, H., Hashikawa, T., He, J. & Jones, E. G. (1997). Tonotopic organization of auditory cortical fields delineated by parvalbumin immunoreactivity in macaque monkeys. The Journal of Comparative Neurology, 386, 304–16.3.0.CO;2-K>CrossRefGoogle ScholarPubMed
Kunzle, H. (1978). An autoradiographic analysis of the efferent connections from premotor and adjacent prefrontal regions (Areas 6 and 9) in Macaca fascicularis. Brain, Behavior and Evolution, 15, 185–234.CrossRefGoogle Scholar
LeDoux, J. (1996). The Emotional Brain. New York: Simon & Schuster.Google Scholar
Lund, J. S., Lund, R. D., Hendrickson, A. E., Hunt, A. B. & Fuchs, A. F. (1976). The origin of efferent pathways from the primary visual cortex, area 17, of the macaque monkey as shown by retrograde transport of horseradish peroxidase. The Journal of Comparative Neurology, 164, 287–304.CrossRefGoogle Scholar
Lynch, J. C. & Graybiel, A. M. (1983). Comparison of afferents traced to the superior colliculus from the frontal eye fields and from two sub-regions of area 7 of the rhesus monkey. Neuroscience Abstracts, 9, 750.Google Scholar
Mah, L., Arnold, M. C. & Grafman, J. (2004). Impairment of social perception associated with lesions of the prefrontal cortex. American Journal of Psychiatry, 161, 1247–55.CrossRefGoogle ScholarPubMed
Maioli, M. G., Squatrito, S., Galletti, C., Battaglini, P. P. & Sanseverino, E. R. (1983). Cortico-cortical connections from the visual region of the superior temporal sulcus to frontal eye field in the macaque. Brain Research, 265, 294–9.CrossRefGoogle ScholarPubMed
Malkova, L., Gaffan, D. & Murray, E. A. (1997). Excitotoxic lesions of the amygdala fail to produce impairment in visual learning for auditory secondary reinforcement but interfere with reinforcer devaluation effects in rhesus monkeys. Journal of Neuroscience, 17, 6011–20.CrossRefGoogle ScholarPubMed
Markowitsch, H. J. (1982). Thalamic mediodorsal nucleus and memory: A critical evaluation of studies in animals and man. Neuroscience and Biobehavioral Reviews, 6, 351–80.CrossRefGoogle ScholarPubMed
Mayberg, H. S. (2003). Modulating dysfunctional limbic-cortical circuits in depression: towards development of brain-based algorithms for diagnosis and optimised treatment. British Medical Bulletin, 65, 193–207.CrossRefGoogle ScholarPubMed
McFarland, N. R. & Haber, S. N. (2002). Thalamic relay nuclei of the basal ganglia form both reciprocal and nonreciprocal cortical connections, linking multiple frontal cortical areas. Journal of Neuroscience, 22, 8117–32.CrossRefGoogle ScholarPubMed
McGuire, P. K., Silbersweig, D. A., Wright, I., et al. (1995). Abnormal monitoring of inner speech: a physiological basis for auditory hallucinations. Lancet, 346, 596–600.CrossRefGoogle ScholarPubMed
McGuire, P. K., Silbersweig, D. A., Wright, I., et al. (1996). The neural correlates of inner speech and auditory verbal imagery in schizophrenia: relationship to auditory verbal hallucinations. British Journal of Psychiatry, 169, 148–59.CrossRefGoogle ScholarPubMed
Mesulam, M. M. (1981). A cortical network for directed attention and unilateral neglect. Annals of Neurology 10, 309–25.CrossRefGoogle ScholarPubMed
Mesulam, M. M. (1990). Large-scale neurocognitive networks and distributed processing for attention, language, and memory. Annals of Neurology, 28, 597–613.CrossRefGoogle Scholar
Mesulam, M. M. (1999). Spatial attention and neglect: parietal, frontal and cingulate contributions to the mental representation and attentional targeting of salient extrapersonal events. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 354, 1325–46.CrossRefGoogle ScholarPubMed
Middleton, F. A. & Strick, P. L. (1994). Anatomical evidence for cerebellar and basal ganglia involvement in higher cognitive function. Science, 266, 458–61.CrossRefGoogle ScholarPubMed
Middleton, F. A. & Strick, P. L. (2000). Basal ganglia and cerebellar loops: motor and cognitive circuits. Brain Research, Brain Research Reviews, 31, 236–50.CrossRefGoogle ScholarPubMed
Middleton, F. A. & Strick, P. L. (2002). Basal-ganglia 'projections' to the prefrontal cortex of the primate. Cerebral Cortex, 12, 926–35.CrossRefGoogle ScholarPubMed
Miller, E. K. & Cohen, J. D. (2001). An integrative theory of prefrontal cortex function. Annual Review of Neuroscience, 24, 167–202.CrossRefGoogle ScholarPubMed
Mitchell, I. J., Cooper, A. J. & Griffiths, M. R. (1999). The selective vulnerability of striatopallidal neurons. Progress in Neurobiology, 59, 691–719.CrossRefGoogle ScholarPubMed
Moga, M. M. & Gray, T. S. (1985). Peptidergic efferents from the intercalated nuclei of the amygdala to the parabrachial nucleus in the rat. Neuroscience Letters, 61, 13–18.CrossRefGoogle ScholarPubMed
Morecraft, R. J., Geula, C. & Mesulam, M.-M. (1992). Cytoarchitecture and neural afferents of orbitofrontal cortex in the brain of the monkey. The Journal of Comparative Neurology, 323, 341–58.CrossRefGoogle ScholarPubMed
Müller-Preuss, J. D., Newman, J. D. & Jürgens, U. (1980). Anatomical and physiological evidence for a relationship between the cingular vocalization area and the auditory cortex in the squirrel monkey. Brain Research, 202, 307–15.CrossRefGoogle ScholarPubMed
Müller-Preuss, P. & Ploog, D. (1981). Inhibition of auditory cortical neurons during phonation. Brain Research, 215, 61–76.CrossRefGoogle ScholarPubMed
Narr, K. L., Thompson, P. M., Szeszko, P., et al. (2004). Regional specificity of hippocampal volume reductions in first-episode schizophrenia. Neuroimage, 21, 1563–75.CrossRefGoogle ScholarPubMed
Nauta, W. J. H. (1961). Fibre degeneration following lesions of the amygdaloid complex in the monkey. Journal of Anatomy, 95, 515–31.Google ScholarPubMed
Nauta, W. J. H. (1979). Expanding borders of the limbic system concept. In Functional Neurosurgery, ed. Rasmussen, T. and Marino, R.. New York: Raven Press, pp. 7–23.Google Scholar
Nelson, M. D., Saykin, A. J., Flashman, L. A. & Riordan, H. J. (1998). Hippocampal volume reduction in schizophrenia as assessed by magnetic resonance imaging: a meta-analytic study. Archives of General Psychiatry, 55, 433–40.CrossRefGoogle ScholarPubMed
Nielsen, J. M. & Jacobs, L. L. (1951). Bilateral lesions of the anterior cingulate gyri. Bulletin of the Los Angeles Neurological Society, 16, 231–4.Google ScholarPubMed
Nishijo, H., Ono, T. & Nishino, H. (1988). Single neuron responses in amygdala of alert monkey during complex sensory stimulation with affective significance. Journal of Neuroscience, 8, 3570–83.CrossRefGoogle ScholarPubMed
Nitecka, L. & Ben Ari, Y. (1987). Distribution of GABA-like immunoreactivity in the rat amygdaloid complex. The Journal of Comparative Neurology, 266, 45–55.CrossRefGoogle ScholarPubMed
Ojima, H. (1994). Terminal morphology and distribution of corticothalamic fibers originating from layers 5 and 6 of cat primary auditory cortex. Cerebral Cortex, 4, 646–63.CrossRefGoogle ScholarPubMed
Öngur, D., An, X. & Price, J. L. (1998). Prefrontal cortical projections to the hypothalamus in macaque monkeys. The Journal of Comparative Neurology, 401, 480–505.3.0.CO;2-F>CrossRefGoogle ScholarPubMed
Pandya, D. N., Seltzer, B. & Barbas, H. (1988). Input-output organization of the primate cerebral cortex. In Comparative Primate Biology, Vol. 4, ed. Steklis, H. D. and Erwin, J.. New York: Alan R. Liss, pp. 39–80.Google Scholar
Pandya, D. N., Hoesen, G. W. & Domesick, V. B. (1973). A cingulo-amygdaloid projection in the rhesus monkey. Brain Research, 61, 369–73.CrossRefGoogle ScholarPubMed
Papez, J. W. (1937). A proposed mechanism of emotion. AMA Archives of Neurology and Psychiatry, 38, 725–43.CrossRefGoogle Scholar
Pare, D. & Smith, Y. (1993a). Distribution of GABA immunoreactivity in the amygdaloid complex of the cat. Neuroscience, 57, 1061–76.CrossRefGoogle Scholar
Pare, D. & Smith, Y. (1993b). The intercalated cell masses project to the central and medial nuclei of the amygdala in cats. Neuroscience, 57, 1077–90.CrossRefGoogle Scholar
Pare, D. & Smith, Y. (1994). GABAergic projection from the intercalated cell masses of the amygdala to the basal forebrain in cats. The Journal of Comparative Neurology, 344, 33–49.CrossRefGoogle ScholarPubMed
Penfield, W. & Jasper, H. (1954). Epilepsy and the Functional Anatomy of the Human Brain. Boston: Little, Brown and Company.Google Scholar
Peters, A., Palay, S. L. & Webster, H. D. (1991). The Fine Structure of the Nervous System. Neurons and their Supporting Cells. New York: Oxford University Press.Google Scholar
Peters, A. & Sethares, C. (1997). The organization of double bouquet cells in monkey striate cortex. Journal of Neurocytology, 26, 779–97.CrossRefGoogle ScholarPubMed
Petrides, M. (1995). Impairments on nonspatial self-ordered and externally ordered working memory tasks after lesions of the mid-dorsal part of the lateral frontal cortex in the monkey. Journal of Neuroscience, 15, 359–75.CrossRefGoogle ScholarPubMed
Petrides, M. (1996). Lateral frontal cortical contribution to memory. Seminars in the Neurosciences, 8, 57–63.CrossRefGoogle Scholar
Petrides, M. & Pandya, D. N. (1988). Association fiber pathways to the frontal cortex from the superior temporal region in the rhesus monkey. The Journal of Comparative Neurology, 273, 52–66.CrossRefGoogle ScholarPubMed
Petrovich, G. D., Canteras, N. S. & Swanson, L. W. (2001). Combinatorial amygdalar inputs to hippocampal domains and hypothalamic behavior systems. Brain Research. Brain Research Reviews, 38, 247–89.CrossRefGoogle ScholarPubMed
Petrovich, G. D., Setlow, B., Holland, P. C. & Gallagher, M. (2002). Amygdalo-hypothalamic circuit allows learned cues to override satiety and promote eating. Journal of Neuroscience, 22, 8748–53.CrossRefGoogle ScholarPubMed
Pitkanen, A. & Amaral, D. G. (1994). The distribution of GABAergic cells, fibers, and terminals in the monkey amygdaloid complex: an immunohistochemical and in situ hybridization study. Journal of Neuroscience, 14, 2200–24.CrossRefGoogle ScholarPubMed
Poremba, A., Malloy, M., Saunders, R. C., et al. (2004). Species-specific calls evoke asymmetric activity in the monkey's temporal poles. Nature, 427, 448–51.CrossRefGoogle ScholarPubMed
Porrino, L. J., Crane, A. M. & Goldman-Rakic, P. S. (1981). Direct and indirect pathways from the amygdala to the frontal lobe in rhesus monkeys. The Journal of Comparative Neurology, 198, 121–36.CrossRefGoogle ScholarPubMed
Preuss, T. M. & Goldman-Rakic, P. S. (1987). Crossed corticothalamic and thalamocortical connections of macaque prefrontal cortex. The Journal of Comparative Neurology, 257, 269–81.CrossRefGoogle ScholarPubMed
Preuss, T. M. & Goldman-Rakic, P. S. (1989). Connections of the ventral granular frontal cortex of macaques with perisylvian premotor and somatosensory areas: Anatomical evidence for somatic representation in primate frontal association cortex. The Journal of Comparative Neurology, 282, 293–316.CrossRefGoogle ScholarPubMed
Rakic, P. (1988). Specification of cerebral cortical areas. Science, 241, 170–6.CrossRefGoogle ScholarPubMed
Rakic, P. (2002). Neurogenesis in adult primate neocortex: an evaluation of the evidence. Nature Reviews. Neuroscience, 3, 65–71.CrossRefGoogle Scholar
Rauschecker, J. P. (1998). Parallel processing in the auditory cortex of primates. Audiology and Neuro-otology, 3, 86–103.CrossRefGoogle ScholarPubMed
Rauschecker, J. P., Tian, B. & Hauser, M. (1995). Processing of complex sounds in the macaque nonprimary auditory cortex. Science, 268, 111–14.CrossRefGoogle ScholarPubMed
Ray, J. P. & Price, J. L. (1993). The organization of projections from the mediodorsal nucleus of the thalamus to orbital and medial prefrontal cortex in macaque monkeys. The Journal of Comparative Neurology, 337, 1–31.CrossRefGoogle ScholarPubMed
Rempel-Clower, N. L. & Barbas, H. (1998). Topographic organization of connections between the hypothalamus and prefrontal cortex in the rhesus monkey. The Journal of Comparative Neurology, 398, 393–419.3.0.CO;2-V>CrossRefGoogle ScholarPubMed
Rempel-Clower, N. L. & Barbas, H. (2000). The laminar pattern of connections between prefrontal and anterior temporal cortices in the rhesus monkey is related to cortical structure and function. Cerebral Cortex, 10, 851–65.CrossRefGoogle ScholarPubMed
Robson, J. A. & Hall, W. C. (1975). Connections of layer VI in striate cortex of the grey squirrel (Sciurus carolinensis). Brain Research, 93, 133–9.CrossRefGoogle Scholar
Rockland, K. S. (1996). Two types of corticopulvinar terminations: round (type 2) and elongate (type1). The Journal of Comparative Neurology, 368, 57–87.3.0.CO;2-J>CrossRefGoogle Scholar
Romanski, L. M. & LeDoux, J. E. (1992). Equipotentiality of thalamo-amygdala and thalamo-cortico-amygdala circuits in auditory fear conditioning. Journal of Neuroscience, 12, 4501–9.CrossRefGoogle ScholarPubMed
Rouiller, E. M. & Welker, E. (1991). Morphology of corticothalamic terminals arising from the auditory cortex of the rat: a Phaseolus vulgaris-leucoagglutinin (PHA-L) tracing study. Hearing Research, 56, 179–90.CrossRefGoogle ScholarPubMed
Rouiller, E. M. & Welker, E. (2000). A comparative analysis of the morphology of corticothalamic projections in mammals. Brain Research Bulletin, 53, 727–41.CrossRefGoogle ScholarPubMed
Saha, S., Batten, T. F. & Henderson, Z. (2000). A GABAergic projection from the central nucleus of the amygdala to the nucleus of the solitary tract: a combined anterograde tracing and electron microscopic immunohistochemical study. Neuroscience, 99, 613–26.CrossRefGoogle ScholarPubMed
Sandell, J. H. & Schiller, P. H. (1982). Effect of cooling area 18 on striate cortex cells in the squirrel monkey. Journal of Neurophysiology, 48, 38–48.CrossRefGoogle ScholarPubMed
Sato, M. & Hikosaka, O. (2002). Role of primate substantia nigra pars reticulata in reward-oriented saccadic eye movement. Journal of Neuroscience, 22, 2363–73.CrossRefGoogle ScholarPubMed
Schall, J. D., Morel, A., King, D. J. & Bullier, J. (1995). Topography of visual cortex connections with frontal eye field in macaque: Convergence and segregation of processing streams. Journal of Neuroscience, 15, 4464–87.CrossRefGoogle ScholarPubMed
Schiller, P. H. & Tehovnik, E. J. (2001). Look and see: how the brain moves your eyes about. Progress in Brain Research, 134, 127–42.CrossRefGoogle Scholar
Schoenbaum, G., Chiba, A. A. & Gallagher, M. (1999). Neural encoding in orbitofrontal cortex and basolateral amygdala during olfactory discrimination learning. Journal of Neuroscience, 19, 1876–84.CrossRefGoogle ScholarPubMed
Schultz, W., Tremblay, L. & Hollerman, J. R. (2000.) Reward processing in primate orbitofrontal cortex and basal ganglia. Cerebral Cortex, 10, 272–84.CrossRefGoogle ScholarPubMed
Shao, Z. & Burkhalter, A. (1999). Role of GABAB receptor-mediated inhibition in reciprocal interareal pathways of rat visual cortex. Journal of Neurophysiology, 81, 1014–24.CrossRefGoogle ScholarPubMed
Shipp, S. (2005). The importance of being agranular: a comparative account of visual and motor cortex. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 360, 797–814.CrossRefGoogle ScholarPubMed
Simpson, J. R., Snyder, A. Z., Gusnard, D. A. & Raichle, M. E. (2001). Emotion-induced changes in human medial prefrontal cortex: I. During cognitive task performance. Proceedings of the National Academy of Science USA, 98, 683–7.CrossRefGoogle ScholarPubMed
Squire, L. R. (1992). Memory and the hippocampus: A synthesis from findings with rats, monkeys, and humans. Psychological Review, 99, 195–231.CrossRefGoogle ScholarPubMed
Squire, L. R. & Zola-Morgan, S. (1988). Memory: Brain systems and behavior. Trends in Neurosciences, 11, 170–5.CrossRefGoogle ScholarPubMed
Stefanacci, L. & Amaral, D. G. (2002). Some observations on cortical inputs to the macaque monkey amygdala: an anterograde tracing study. The Journal of Comparative Neurology, 451, 301–23.CrossRefGoogle Scholar
Steriade, M., Jones, E. G. & McCormick, D. A. (1997). Thalamus–Organisation and Function. Oxford: Elsevier Science.Google Scholar
Szeszko, P. R., Goldberg, E., Gunduz-Bruce, H., et al. (2003). Smaller anterior hippocampal formation volume in antipsychotic-naive patients with first-episode schizophrenia. American Journal of Psychiatry, 160, 2190–7.CrossRefGoogle ScholarPubMed
Takagi, S. F. (1986). Studies on the olfactory nervous system of the old world monkey. Progress in Neurobiology, 27, 195–250.CrossRefGoogle ScholarPubMed
Talland, G. A., Sweet, W. H. & Ballantine, T. (1967). Amnesic syndrome with anterior communicating artery aneurysm. Journal of Nervous and Mental Disease, 145, 179–92.CrossRefGoogle ScholarPubMed
Toni, I., Rowe, J., Stephan, K. E. & Passingham, R. E. (2002). Changes of cortico-striatal effective connectivity during visuomotor learning. Cerebral Cortex, 12, 1040–7.CrossRefGoogle ScholarPubMed
Turner, B. H., Mishkin, M. & Knapp, M. (1980). Organization of the amygdalopetal projections from modality- specific cortical association areas in the monkey. The Journal of Comparative Neurology, 191, 515–43.CrossRefGoogle ScholarPubMed
Van Hoesen, G. W. (1981). The differential distribution, diversity and sprouting of cortical projections to the amygdala of the rhesus monkey. In The Amygdaloid Complex, ed. Ben-Ari, Y.. Amsterdam: Elsevier/North Holland Biomedical Press, pp. 77–90.Google Scholar
Vogt, B. A. & Barbas, H. (1988). Structure and connections of the cingulate vocalization region in the rhesus monkey. In The Physiological Control of Mammalian Vocalization, ed. Newman, J. D.. New York: Plenum Publ. Corp., pp. 203–25.CrossRefGoogle Scholar
Vogt, B. A. & Pandya, D. N. (1987). Cingulate cortex of the rhesus monkey: II. Cortical afferents. The Journal of Comparative Neurology, 262, 271–89.CrossRefGoogle ScholarPubMed
Voytko, M. L. (1985). Cooling orbital frontal cortex disrupts matching-to-sample and visual discrimination learning in monkeys. Physiological Psychology, 13, 219–29.CrossRefGoogle Scholar
Wallis, J. D. & Miller, E. K. (2003). Neuronal activity in primate dorsolateral and orbital prefrontal cortex during performance of a reward preference task. European Journal of Neuroscience, 18, 2069–81.CrossRefGoogle ScholarPubMed
Webster, M. J., Bachevalier, J. & Ungerleider, L. G. (1994). Connections of inferior temporal areas TEO and TE with parietal and frontal cortex in macaque monkeys. Cerebral Cortex, 4, 470–83.CrossRefGoogle ScholarPubMed
Weiss, A. P., DeWitt, I., Goff, D., Ditman, T. & Heckers, S. (2005). Anterior and posterior hippocampal volumes in schizophrenia. Schizophrenia Research, 73, 103–12.CrossRefGoogle Scholar
Whalen, P. J., Rauch, S. L., Etcoff, N. L., et al. (1998). Masked presentations of emotional facial expressions modulate amygdala activity without explicit knowledge. Journal of Neuroscience, 18, 411–18.CrossRefGoogle ScholarPubMed
Wood, J. N. (2003). Social cognition and the prefrontal cortex. Behavioral and Cognitive Neuroscience Reviews, 2, 97–114.CrossRefGoogle ScholarPubMed
Xiao, D. & Barbas, H. (2002). Pathways for emotions and memory II: afferent input to the anterior thalamic nuclei from prefrontal, temporal, hypothalamic areas and the basal ganglia in the rhesus monkey. Thalamus and Related Systems, 2, 33–48.Google Scholar
Xiao, D. & Barbas, H. (2004). Circuits through prefrontal cortex, basal ganglia, and ventral anterior nucleus map pathways beyond motor control. Thalamus and Related Systems, 2, 325–43.CrossRefGoogle Scholar
Yakovlev, P. I. (1948). Motility, behavior and the brain: Stereodynamic organization and neurocoordinates of behavior. Journal of Nervous and Mental Disease, 107, 313–35.CrossRefGoogle ScholarPubMed
Yeterian, E. H. & Pandya, D. N. (1988). Corticothalamic connections of paralimbic regions in the rhesus monkey. The Journal of Comparative Neurology, 269, 130–46.CrossRefGoogle ScholarPubMed
Young, M. P. (1992). Objective analysis of the topological organization of the primate cortical visual system. Nature, 358, 152–5.CrossRefGoogle ScholarPubMed
Zald, D. H. & Kim, S. W. (1996). Anatomy and function of the orbital frontal cortex, I: anatomy, neurocircuitry; and obsessive-compulsive disorder. The Journal of Neuropsychiatry and Clinical Neurosciences, 8, 125–38.Google ScholarPubMed
Zola-Morgan, S. & Squire, L. R. (1993). Neuroanatomy of memory. Annual Review of Neuroscience, 16, 547–63.CrossRefGoogle Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×